143
Thermal and electrical properties of semiconductors measured by means of photopyroelectric and photocarrier radiometry techniques. DISSERTATION zur Erlangung des Grades „Doktor der Naturwissenschaften“ an der Fakultät für Physik und Astronomie der Ruhr-Universität Bochum von Michal Pawlak aus Bydgoszcz Bochum 2009

Thermal and electrical properties of semiconductors ...€¦ · Thermal and electrical properties of semiconductors measured by means of photopyroelectric and photocarrier radiometry

  • Upload
    vuthuy

  • View
    221

  • Download
    0

Embed Size (px)

Citation preview

Thermal and electrical properties of semiconductors measured by

means of photopyroelectric and photocarrier radiometry

techniques.

DISSERTATION

zur

Erlangung des Grades „Doktor der Naturwissenschaften“

an der Fakultät für Physik und Astronomie der Ruhr-Universität Bochum

von Michał Pawlak

aus

Bydgoszcz

Bochum 2009

ii

1. Gutachter: Prof. Dr. Josef Pelzl 2. Gutachter: Prof. Dr. Andreas Wieck Datum der Disputation: 18.01.2010

iii

Acknowledgements

First and foremost, I would like to thank Prof. Dr. J.Pelzl for his kind supervision,

valuable guidance, helpful comments on this research and also for his full support during this

two very hard years I have had.

From the group of many researchers that I was fortune to work with, I would like to

thank prof A. Mandelis who I have participated with in developing of the photocarrier

radiometry in laboratories in Bochum as well in Toronto, Canada, as well prof. H.Meczynska

for her support and valuable advices. During my visit in Toronto I worked also with Dr.

J.Tolev and I want to thank him for his support and a lot of valuable guidance in photocarrier

and photothermal radiometry.

Also, I would like thank Prof A. Wieck, head of the GRK 348 program, for opportunity

to participate in interesting lectures and projects related with micro- and nanoelectronics.

Many thanks go to my colleagues from Ruhr University Bochum (Mr B.K.Bein, Dr. D.

Dietzel, Dr. R. Meckenstock, Dr D.Spoddig, Mr. J.Gibkes, Mr M. Mueller, Dr

S.Chotikaprakhan), for their helpful advices and discussions about Germany culture during

my stay in Bochum.

Many thanks go to my colleagues from Nicolaus Copernicus University in Torun,

Poland. Especially I would like thank prof F.Friszt, prof. S.Legowski for their valuable

comments, Dr J.Zakrzewski for familiarizing me with photoacoustic science and

Mrs.A.Marasek, Dr J.Szatkowski and Dr K.Strzalkowski for produced the CdMgSe single

crystals used in this work.

iv

Many thanks go to my colleagues from University of Toronto (Dr Tolev, Dr. X.Gou,

Dr J.Garcia) and also to Prof. Mihai Chirtoc (from University of Reims, France) for useful

advice during the construction of the PPE experimental setup.

Also, I would like to thank prof. M.Malinski from Technical University of Koszalin,

Poland for his advices and helpful comments.

I want to thank my parents and Grandfather for their support throughout all of my life

especially during last two years. Finally, I would like to thank Malgosia for her love and

encouragement, as well as her patience during the many hours I spent completing this work.

Bochum, October 2009 Michal Pawlak

v

Abbreviations

1-D one dimensional

3-D three dimensional

BNC Bayonet Neill-Concelman

CDW carrier density wave

CdSe cadmium selenide

DSC differential scanning calorimetry

IPPE inverse PPE

GaAs Gallium arsenide

GaN Gallium nitride

Ge-Si Germanium silicon alloys

Mg Magnesium

MgSe Magnesium selenide

NIR near infrared

PAS photoacoustic spectroscopy

PPE photopyroelectric technique

PPT photothermal piezoelectric technique

PZT lead zirconate titanate

SCL space charge layer

Si Silicon

Si -Si02 silicon-silicon dioxide interface

TW thermal wave

J0 Bessel function of first kind of order 0

J1(λr) Bessel function of first kind of order 1

PL Photoluminescence

vi

Nomenclature

Symbol Unity Name

Nt [m-2] charged interface state density energy Et.

W0 [m] space charge layer width

( )νβ hI ,0 [W m-2] optical intensity

W∆ [m] effective SCL width

0W [m] dc component of the SCL

mW [m] modulated component of the SCL

*D [m2/s] ambipolar diffusion coefficient

τ [s] bulk recombination lifetime

0sqψ [J] interface potential energy

Tri(ω) [s] complex interface lifetime

τri [s] charged interface recombination lifetime

( )trF ,rr

[W/m2] heat flow

ρ [kg/m3] mass density

C [J/kg·K] specific heat

k [W/mK] thermal conductivity

( )trQ ,r

[J] heat source

( )trT ,r

[K] temperature distribution (or field)

Dt [m2/s] thermal diffusivity

T0 [K] ambient temperature

( )rTdc

r [K] steady temperature distribution

( )trTac ,r

[K] temporal temperature distribution

σt [m-1] thermal wave number

vii

e [Ws1/2/m2K] thermal effusivity

µth [m] thermal diffusion length

Eph [J] photon energy

EG [J] energy band gap

h [J s] Planck constant

v [s-1] wave frequency

),,( λtrQIB

r [J/m3] intraband heat release rate per unit volume

Gη [m-3] quantum efficiency for photogenerated carriers

( )λα [m-1] optical absorption coefficient at the excitation

wavelength λ

( )tN0 photon deposition rate per volume

jL [m] diffusion length, where j=n (electron) or p

(hole)

FSη front surface non-radiative quantum efficiency

BSη rear surface non-radiative quantum efficiency

( )λ,,0 tN [m-3] photogenerated electron density distribution in

a one dimensional geometry

N0 [m-3] equilibrium density

SFS [m/s] front recombination velocity

SBS [m/s] rear recombination velocity

N(r,t) [m-3] concentration of the excess electron-hole pairs

( )trg ,r

[m-3] carrier generation rate per unit volume from an

external source of excitation

µe [m2/Vs] electron mobility

µh [m2/Vs] hole mobility

σe [m-1] carrier density-wave wave number

ηQ quantum yield of the photogenerated carriers

t [s] Time

ω [s-1] angular frequency

f [s-1] Frequency

ϕ [rad] Phase

σ [J/K] Stefan-Boltzmann constant

viii

τR [s] radiative recombination lifetime

τNR [s] non-radiative recombination lifetimes

q [C] Charge

B [m3/s] radiative recombination probability

W [m] laser beam of a spot size

EC [J] energy of the conduction

EV [J] energy of the valance band edge

p [C/m2K] pyroelectric coefficient of the detector

A [m2] transducer area

WCdSe [mK/W] thermal resistivities of CdSe

WMgSe [mK/W] thermal resistivities of MgSe

CCd-Mg [mK/W] nonlinear parameter

R Reflectivity

c [m/s] speed of light

m* [kg] effective mass

n reflactive index

ix

Contents

Acknowledgements………………………………………………………………………..…iii

Abbreviations and Nomenclature……………………………………………………..……..v

Contents…………………………………………………………………………………........ix

Chapter 1: Introduction……………………………………………………………… ……...1

1.1 Motivation and objectives………………………………………………………………….1

1.2 Organization of the thesis………………………………………………………………….2

Chapter 2: Physical Principles of the thermal waves………………………………………3

2.1 Historical …………………………………………………………………………………..3

2.2 Heat conduction equation ……………………………………………………………….…4

2.3 Review of the photothermal methods……………………………………………………...7

Chapter 3: Physical basics of the carrier density waves in semiconductors…………..…11

3.1 De-excitation processes in semiconductors………………................................................11

3.2 Ambipolar diffusion equation…………………………….................................................14

3.3 One dimensional excess CDW field in Cartesian geometries………..…………………...15

3.4 Three dimensional CDW field in cylindrical geometries………………………………...16

3.5 Recombination processes in semiconductors……………………………………………..18

Chapter 4: Experimental methods, signal generation mechanisms and instrumentation

of PCR and PPE methods…………………………………………………………………..21

4.1 The Photocarrier Radiometry (PCR) signal generation mechanism and instrumentation..21

4.1.1. Introduction…………………………………………………………………………….21

4.1.2. Contribution to the PCR ……………………………………………………………….22

4.1.3. Instrumentation and normalization of PCR signals……………………………………24

4.1.4. The one dimensional Photocarrier Radiometry Signal ………………………………..25

4.1.5. The three dimensional PCR signal……………………………………………………..30

4.1.6. PCR Dimensionality criterion………………………………………………………….31

4.1.7 Photo-carrier radiometry microscope…………………………………………………...33

4.2 Photopyroelectric effect (PPE)……………………………………………………………36

x

4.2.1. Experimental setup …………………………………………………………………….36

4.2.2. The PPE signal generation …………………………………………………………….37

4.2.3. Normalization of the PPE signal……………………………………………………….40

Chapter 5: Thermal properties of Cd1-xMgxSe single crystals measured by means of

photopyroelectric technique………………………………………………………………..50

5.1. Materials…………………………………………………………………………………50

5.2. Experimental results and computational algorithm……………………………………...52

5.2.1 Thermal diffusivity – PPE phases………………………………………………………52

5.2.2 Thermal conductivity – Normalized PPE amplitude…………………………………...55

5.2.3 Discussion………………………………………………………………………………56

Chapter 6: Influence of the space charge layer (SCL) on the charge carrier transport

properties measured by means of the photocarrier radiometry (PCR)…………………61

6.1 Theory of optically modulated p-type SiO2-Si interface energetics in the presence of

charged interface states ………………………………………………………………………61

6.2 The expression of the PCR signal including effects due to an existing SCL……………..64

6.3 Numerical simulations of an influence of the existence of the SCL on the electronic

transport properties…………………………………………………………………………...65

6.3.1 Numerical simulation of the PCR signal dependence on the electrical transport

properties in the presence of SCL……………………………………………………………65

6.3.2 Numerical simulation of the PCR signal dependence on the existence of SCL width…68

6.4 Experimental conditions and materials……………………………………………...……71

6.4.1 Experimental methodology……………………………………………………………..71

6.4.2. Experimental set up…………………………………………………………………….72

6.4.3. Materials………………………………………………………………………………..73

6.5 Experimental results………………………………………………………………………73

6.5.1 Effect of chemical etching on the PCR signal………………………………………….73

6.5.2 The perturbation effects of the primary modulated laser beam on the PCR signal…….75

6.5.3 The effect of polishing on the PCR signal……………………………………………...76

6.6 Determination of carrier transport properties in SCL and the depth profile

reconstruction…………………………………………………………………………………77

6.7 Summary………………………………………………………………………………….82

xi

Chapter 7: Non-linear dependence of photocarrier radiometry signals from p-Si wafers

on optical excitation intensity and its effect on charge carrier transport properties…...84

7.1 Introduction……………………………………………………………………………….84

7.2 Experimental methodology and materials………………………………………………...87

7.2.1 Low resolution PCR system…………………………………………………………….87

7.2.2 High resolution PCR system……………………………………………………………89

7.2.3 Materials………………………………………………………………………………..90

7.3. Numerical simulations of the PCR signal as a function of the non-linear coefficient β and

photo-injected carriers………………………………………………………………………..90

7.4 Experimental results and discussion……………………………………………………..93

7.4.1 Laser power dependencies ……………………………………………………………..93

7.4.2. Modulation frequency dependence at the low resolution system……………………...99

7.4.3. Modulation frequency dependence at 532 nm…..……………………………………108

7.5 Summary…………………………………………………………………………….…..112

Chapter 8: Conclusion and outlook ………………………………………………………113

Bibliography………………………………………………………………………………..116

Curriculum Vitae and conference contributions………………………………………...121

Appendix A: Current controller……… …………………………………………………..128

xii

1

Chapter 1: Introduction

1.1 Motivation and objectives

Nowadays, the trend is increasing to develop semiconductor nano-devices. To obtain

high quality appropriate substrates are required. A good choice of the substrate in electronic

or opto-electronic nano-devices is detrimental task of the design process. Hence, monitoring

of quality of the substrate during technological process is very important task. Whereas the

development of characterization strategies capable of evaluating the effects of the bulk

substrate Si properties on the performance of microelectronic devices is an issue of growing

importance, as the electronic properties of the bulk can seriously affect the electrical

characteristics of the device [Schroder, 1997]. For these reasons on-line techniques able to

measure the electrical properties are required.

On the other hand, the thermal management is also a very important aspect. For

instance, thermal conductivity is an important parameters that determines the maximum

power at operation of semiconductor devices. For semiconductors used in thermoelectric

energy conversion the thermal conductivity is one of the most important parameters

2

determining the efficiency of the device [Tritt, 2004]. For miniaturized semiconducting

devices thermal management of the energy dissipation has become a key problem. In this

context, the thermal diffusivity is a very important physical parameter in device modeling. It

is a parameter specific for each material, which dependents on the composition and structural

characteristic of the sample.

1.2 Organization of the thesis

The following chapters in this thesis are organized as follows:

Chapter 2 introduces the basic concepts of thermal waves and reviews the different

experimental techniques.

Chapter 3 describes on the basic concepts of carrier density waves in semiconductors.

Different recombination processes in semiconductors are discussed.

Chapter 4 reports on the experimental set-ups which are constructed during the work, such as

photopyroelectric (PPE) and photocarrier radiometry (PCR). This chapter shows also

numerical simulations of the photopyroelectric as well photoradiometry signals. Changes

depending on the experimental conditions and the constructed experimental set-ups are

discussed.

Chapter 5 discusses the results of the frequency-dependent and temperature-dependent

measurements on Cd1-xMgxSe single crystals by means of photopyroelectric technique. For

this purpose the photopyroelectric cell with Peltier element was constructed.

Chapter 6 is devoted to the development of photocarrier radiometry. The two-laser beam PCR

system was constructed and the experimental verification of the Mandelis theory [Mandelis,

2005a] is presented.

Chapter 7 reports on the photocarrier radiometry experimental results of the frequency-

dependent and intensity-dependent measurements on silicon wafers. In this chapter the

nonlinear parameter is introduced to take into account the nonlinearities phenomena.

Chapter 8: summarizes the results of this thesis and presents some propositions for interesting

directions of future work.

3

Chapter 2: Physical principles of the thermal

waves

2.1 Historical

Thermal wave (TW) is an temperature distribution oscillating in time and space

representing a continuous energy dissipation [Mandelis, 2001]. Thermal waves were used first

by A. Angström in the mid 19th century. In 1861 Ångström [Ångström, 1861] had reported

determination of the thermal diffusivity of the long copper bar by means of detection and

interpretation of the periodical heating from the investigated material. If the thermal wave is

excited by the photons the thermal response is named: photothermal effect. Graham Bell and

his co-workers were the first who observed that sunlight modulated by chopper incident on a

strongly absorbing substance causes audible sound emitted from the substance [Bell, 1880]. It

was almost a century later when Rosencwaig and Gersho explained Bell’s photoacoustic

experiment in the frame of the thermal waves [Rosencwaig and Gersho, 1976]. Since then,

4

much more attention to the photothermal effect was attributed and a lot of new experimental

techniques based on the effect were developed.

2.2 Heat conduction equation

When temperature gradient exists in a material then a heat transfer from places with

higher temperature to places with lower temperature is observed. There are three distinct

methods of this transfer: conduction, convection and radiation. In most solid state problems

conduction is the most important process of the thermal energy transfer. Mathematically the

heat conduction process is described by heat diffusion equation, which is simply an

expression of the energy conservation principle. In the case of the isotropic homogeneous

solid the general form of the heat conduction equation in Cartesian co-ordinate is given by

[Carslaw and Jaeger, 1959]:

( ) ( ) ( )trQtrFt

trTC ,,

, rrrrr

+⋅∇−=∂

∂ρ , (2.1)

where ( )trF ,rr

is the heat flow which is defined in Fournier’s law:

( ) ( )trTktrF ,,rrr

∇−= . (2.2)

Here, kC,,ρ are mass density, specific heat and thermal conductivity, respectively. The

negative sign indicates the direction of heat flow from hot to cold areas. By appropriate

boundary conditions and the strength and localization of the heat source( )trQ ,r

, the

temperature distribution (or field)( )trT ,r

can be evaluated from solution of the heat diffusion

equation. Inserting (2.2) to (2.1) and ordering particular terms equation (2.1) becomes

( ) ( )k

trQ

t

trT

DtrT

t

,,1),(2

rrr −=

∂∂−∇ , (2.3)

where Dt is a thermal diffusivity of the solid and is defined by

C

kDt ρ

= . (2.4)

The heat sources caused increase the temperature inside material:

( )trTrTTtrT acdc ,)(),( 0

rrr ++= , (2.5)

5

where T0, ( )rTdc

r and ( )trTac ,

r are the ambient temperature, steady and temporal temperature

distribution due to the heat sources in material, respectively. Assuming that the temporal

Fournier transform of ( )trT ,r

exist one can write [Mandelis, 2001]

( ) ( )∫∞

∞−

⋅⋅−⋅= dtetrTr ti ωωθ ,,rr

(2.6)

and taking the Fournier transform of Eq. (2.3) yields [Mandelis,2001]

( ) ( ) dtetrQk

dtet

trT

DdtetrT titi

t

ti

∫∫∫∞

∞−

−−∞

∞−

⋅⋅−∞

∞−

−=∂

∂−⋅∇ ωωω ,1,1

),(2 rr

rr (2.7)

leads to the transformed equation

( ) ( )k

rQrr t

ωωθσωθ ,,),( 22

rrr −=−∇ , (2.8)

where the definition

( ) ( )tt

t Di

D

i

21

ωωωσ +== [m-1] (2.9)

was used. Mandelis [Mandelis, 2001] has proven that in the special case where the heat

source is harmonically modulated at angular frequency ω0 the equation is valid

( ) ( ) ( ).,,, 0ωωθ ω rTdetrtrT ti rrr ≡= ∫∞

∞−

(2.10)

Therefore the heat diffusion can be re-written as follows:

( ) ( ) ( )k

rQrTrT t

ωωωσω ,,),( 22

rrr −=−∇ (2.11)

where simple changing the symbol ω0 back to ω. In one-dimensional geometry the heat

conduction equation can be written

( ) ( ) ( )k

xQxTxT

dx

dt

ωωωσω ,,),( 2 −=− . (2.11a)

Mandelis [Mandelis, 2001] used the Green function method and homogeneous Neumann

boundary condition at 00 =x to solve the heat conduction equation (2.11a):

( )tixt tek

DFxT ⋅⋅+⋅−⋅

⋅⋅= ωσ

ωω 0),( (2.12)

where 00 IF ⋅=η . In fact the thermal wave field is given by the real part of (2.12)

( )

⋅−−⋅⋅⋅

⋅⋅= ⋅− xte

k

DFxT t

xt t )(4

cos),( 0 ωσπωω

ω ωσ (1.13)

6

From the structure of the thermal wave formula one can deduce that the physical meaning of

the earlier defined parameter σt (the real part of the definition (2.9)) is related with wave-like

behavior and can be named the thermal wave number [m-1]. In addition the ratio tD

kis

another an important thermal parameter: a thermal effusivity e which is the relevant parameter

for time-varying heating or cooling processes of surfaces and heat transport across composite

layered bodies and can be also written as

Cke ρ= (2.14)

The thermal diffusivity-(2.4)-describes the rate at which heat distributed in a material. High

values of the thermal effusivity lead to low surface temperature oscillations while high values

of the thermal diffusivity contribute to a relatively deeper penetration of the thermal wave.

Main features of the thermal waves can be deducted from (2.13). As compared to

normal-wavelike behavior, the thermal waves are very heavily damped with a decay length

which is the reciprocal of the real part of the thermal wave-number (2.9)

ωµ t

th

D2= (2.15)

µth is known as thermal diffusion length [m]. The depth to which the thermal waves can

penetrate increase with the square root of the thermal diffusivity of a material (if D is high

then waves reach deeper region in a material) and with the reciprocal of the square root of the

modulation frequency of the heating (if the frequency is low then waves penetrate in deeper

region of the material). This profilometric feature gives the thermal waves methodology great

attention in science and technology. Quantitatively, along the distance of the thermal diffusion

length the thermal wave is damped by 386.0/1 =e of its beginning value. This parameter

therefore defines the range of effective use of the thermal wave technique [Almond and Patel,

1996].

The phase lag between thermal wave field described by (2.13) and the optical

modulation heating is given by

42

πωϕ +⋅=∆ xDt

(2.16)

∆φ increases linearly with the propagation distance x of the thermal wave. The phase lag

shows also that the thermal waves are highly dispersive, because the high frequency thermal

waves propagate faster than low frequency thermal waves.

7

2.3 Review of the Photothermal Methods

The periodic heating of the sample modifies also other physical properties of the

sample. These resulting modifications which oscillate at the same frequency as the heating

can be used to detect the thermal wave propagation in the sample. Figure 2.1 illustrates

schematically the different physical properties and parameters used to detect the thermal wave

response.

Figure 2.1 When a modulated laser beam strikes a surface, it generates a thermal wave field,

which, in turns, causes a refractive index gradient to appear, IR emission, acoustic wave

generation or propagate through the material [Mandelis, 2000]

Based on the parameters shown in Fig.2.1 a variety of experimental techniques have been

developed to measure the photothermal effect: The most important ones are sketched in

Fig.2.2.

8

Figure 2.2 The schematic representation of the different configuration in photothermal

techniques [Pelzl and Bein, 1990]

The photoacoustic effect relies on measurements of the pressure fluctuation, induced in the

gas volume by the heat flux across the solid/gas interface, by means of a microphone mounted

inside the cell (Figure 2.2a). The first theoretical explanation of the photoacoustic signal

generation in a solid state was given by [Rosencwaig and Gersho, 1976]. They found that

photoacoustic signal is proportional to the average of the local modulated temperature rise

resulting from optical heating. Based on their work the photoacoustic spectroscopy (PAS) was

established and it was found that PAS enables to be used on a broad range of materials such

as solids [Murphy and Aamodt, 1977, Pelzl and Bein, 1992], gases [Meyer and Sigrist, 1990,

Harren et al., 2000], semiconductors. The limitation of this method lies in problems with

enclose the sample in a photoacoustic cell.

9

The photopyroelectric effect is bound up with generate the electrical potential when a

material, which has a pyroelectric feature, is heated or cooled. This effect is used in a

photopyroelectric technique (PPE) where the sample is heated by absorption of a modulated

light. Then a direct thermal contact is performed by placing a pyroelectric sensor at the rear

(normal PPE) or at the front (IPPE-inverse PPE) of the sample. The temperature changes from

the sample reach the pyroelectric sensor where are converted to current and measurements by

means of lock in detection. A theoretical explanation was given by [Mandelis and Zver, 1985]

and [Chirtoc and Mihailescu, 1989].

The piezoelectric effect (and detection) relies on generation of a voltage in response of

applied mechanical stress. In the photothermal piezoelectric technique (PPT) sensor is

connected with a sample by means of a metallic hemi-sphere which can collect stresses

generated in a sample [Zakrzewski, 2003]. The photothermal piezoelectric technique was

used in investigation of the optical and thermal properties of semiconductor.

Infrared emission (photothermal radiometry) relies on the Stefan-Boltzmann law which

connects the energy of the emitted radiation E with emitter’s temperature T. Applying this law

to the photothermal methodology where only an ac component of the temperature distribution

can be monitored the Stefan-Boltzman law becomes

acTTdE 34εσ= (2.17)

where σ is the Stefan-Boltzmann constant, and ε is the emissivity of the material. The

photothermal methods used this phenomena is called photothermal radiometry or infrared

radiometry was purposed by [Nordal and Kanstad, 1979]. This technique is non-contact and

non-destructive hence is applied to broad range of materials.

Mirage effect (photothermal beam deflection) is based on the changes of the refractive

index of the surrounding gas due to the thermal waves propagating from solid state into the

gas. The thermal wave is excited by pump laser within a solid while a second laser beam laser

beam probe the gradient of the refractive index perpendicular and parallel to the sample

surface. This technique was first proposed by Boccara et al. (1980) who used position

sensitive detectors such as quadrant or lateral diodes to measure the deflection angle down to

10-8 radians.

10

Photothermally modulated optical reflection relies on the changes of the optical reflection

by the thermal waves. A second laser beam can be used to measure changes of the reflection

index of the surface [Rosencwaig, 1985]. The measured signal provides a relationship

between the temperature dependence of the optical reflectivity [Gruss et al., 1997, Schaub,

2001] and electrical properties in the case of semiconductors [Fournier, 1992, Kiepert et al.

1999, Dietzel, 2001, Fotsing, 2003, Dietzel et al., 2003a]. This technique is used in industry

for inspection of wafers due to the fact that is rapid, non-contact and non-destructive.

11

Chapter 3: Physical basics of the carrier density

waves in semiconductors

3.1 De-excitation processes in semiconductors

When the energy of incident photons, νhEph = , is greater than the energy band gap of a

semiconductor Gph EE > , then electrons from the valance band to the conduction band are

excited (Fig.3.1.)

Figure 3.1 n-type semiconductor energy-band gap diagram showing excitation and

recombination processes. Energy emission processes include nonradiative intraband and

interband decay accompanied by phonon emission, as well as, direct band-to-band

recombination radiative emissions of energy ( )Gh λν and band – to –defect/impurity-state

recombination IR emissions of energy ( )DIRh λν [Mandelis, 2003].

12

The energy difference GEh −ν is deposed in the kinetics energy of the electrons. This

energy is gradually lost on collisions with other carriers and lattice phonons until thermal

equilibrium is achieved. This process is named the thermalization (process (a) in Fig 3.1) and

as a result creates a heat source and can be accurately described by the optical absorption

distribution [Mandelis, 1998]:

( ) ( )( )rEhvtNtrQ gGIB

rr λαηλ −⋅−= exp)(),,( 0

3m

W, (3.1)

where ),,( λtrQIB

r is the intraband heat release rate per unit volume of the optically excited

semiconductor, Gη is the quantum efficiency for photogenerated carriers, ( )λα is the optical

absorption coefficient at the excitation wavelength λ, and ( )tN0 is the photon deposition rate

per volume. Assuming an average intraband relaxation time to be in order of 10-12 [s]

[Bandeira et al.,1982], effects due to intraband thermalization may be neglected on the time

scale of the conventional frequency domain photothermal response of the semiconductor as is

seen in Figure 3.2. Figure 3.2 shows discussed de-excitation processes as a function of time.

Figure 3.2 De-exctitation processes in semiconductor as a function of time.

After the thermalization electron (hole) to the bottom (the top) of the conduction band

(the valance band) electrons and holes create electron-hole pairs. The energy of these pairs

can be changed to other existing type of energy (as e.g. heat) on several processes depending

on the type of the energy bandgap (direct or indirect), defects and/or concentration of these

pairs. Photo-exited carriers can diffuse (process 3 in Figure 3.2) through a distance called a

13

diffusion length jjj DL τ= (where j=n (electron) or p (hole)) and then recombine

radiatively or non-radiatively through the energy bandgap (processes b and d in Figure 3.1

and process 4 in Figure 3.2) with respect to theirs lifetime τ which can be written

NRR τττ111 += , (3.2)

where τR and τNR are lifetimes related with radiative and non-radiative processes, respectively.

In a semiconductor with a direct bandgap (e.g. CdSe) the emission of photons is a result of the

radiative recombination of photo-excited pairs (in fact in CdSe electrons and holes create an

excition). In case of semiconductors with an indirect energy bandgap (as silicon or

germanium) the emission of photons is accompanied with phonons (process d on Figure 3.1).

It is also non-zero probability that excess electron-hole pairs can recombine through non-

radiative bulk interband transition whence generate a heat source [Quimby et al., 1980]. The

heat release rate per unit volume due to non-radiative recombination is given according

Bandiera et al. [Bandiera, 1982], Thielemann and Rheinlaender [Thielemann et al., 1985] by:

( ) ( ) ( )( )rEtNtrQ GNRGBB

rr λαηηλ −= exp;, 0 , only relevant when GEh ≥ν

3m

W (3.3)

where NRη is the non-radiative quantum efficiency.

Besides above discussed processes photo-excited carriers can nonradiatively recombine

at the semiconductor surface generating another heat source. This surface heat release rates

per unit area is given by Flasier and Cahen [Flasier et al, ] and Bandeira [Bandiera et al.,1982]

for a front surface:

( ) ( )[ ] GFSFSFS ESNtNtQ 0,,0,,0 −= ληλ (3.4a)

and for a rear surface:

( ) ( )[ ] GBSBSBS ESNtLNtLQ 0,,,, −= ληλ (3.4b)

where FSη ( BSη ) is the front (rear) surface non-radiative quantum efficiency, ( )λ,,0 tN is the

photogenerated electron density distribution in a one dimensional geometry, N0 is the

equilibrium density, and SFS [m/s] (SBS) is the front (rear) recombination velocity. Both

surface recombination velocities are parameters which characterize the density of the surface

defect states. When S=0 the surface retains bulk properties. When S>0 the surface acts as a

sink for photogenerated carriers. Values of S in the range 1000 <≤ S [cm/s] generally

indicate good passivation for silicon surface [Guidotii et al. 1989].

14

3.2 Ambipolar diffusion equation

At the point r in a semiconductor, the time rate of change in the concentration N(r,t) of

the excess electron-hole pairs is governed by both diffusion and recombination, and can be

described by an ambipolar diffusion equation of the form [Sze, 1981]:

( ) ( ) ( ) ( ) ( )trgtrNtrBNtrN

t

trNtrND ,,,

,,),( 322* rrr

rrrr

=−−−∂

∂−∇ γτ

(3.5)

where ( )trg ,r

is the carrier generation rate per unit volume from an external source of

excitation and *D is an ambipolar diffusion coefficient defined by

( )he

Bhe q

Tk

Dµµ

µµ

+

=2

* (3.6)

where µe and µh are an electron and a hole mobility, respectively. The others parameters from

Eq. (3.5) will be discussed in Section 3.6.

The ambipolar diffusion equation can be linearized when the excess carrier density is

sufficiently small. For example in silicon with the concentration up to 17101×=N [cm-3]

Guidotti et al. [Guidotti et al.,1989] found that τ

γ NNBN <<32 ~ and eq. (3.5) can be written

as

( ) ( ) ( )trgtrN

t

trNtrND ,

,,),(2* r

rrrr

=−∂

∂−∇τ

(3.7)

Assuming that the temporal Fournier transform of ( )trN ,r

exist one can write

( ) ( )∫∞

∞−

⋅⋅−⋅= dtetrNrN ti ωω ,,rr

(3.8)

and taking the Fournier transform of Eq. (3.7) yields

( ) ( ) ( ) dtetrgdtetrN

dtet

trNdtetrND titititi

∫∫∫∫∞

∞−

−∞

∞−

−−∞

∞−

⋅⋅−∞

∞−

−=−∂

∂−⋅∇ ωωωω

ττ,

,,1),(*2 r

rrrr

(3.9)

Follow by Mandelis [Mandelis, 2001] an integration by parts in the second term on the left-

handside (l.h.s) and using the boundary condition for ( )trN ,r

at ±∞→t , results in the

transformed equation

( ) ( ) ( )ωωωσω ,1

,),(*

2 rqD

rNrN e

rrrr−=⋅−∇ (3.10)

15

where the real part of the *

1

D

ie τ

ωτσ += is carrier density-wave wave number, [m-1]. The

CDW wave number is completely different from the thermal wave one. In the case of the

CDW wave number the real and the imaginary parts are unequal. This inequality brings on

that CDW arise only when condition 1≥ωτ is fulfilled.

3.3 One dimensional excess CDW field in Cartesian geometries

Mandelis [Mandelis, 2001] considered the excess CDW field in a semiconductor of the

thickness L. He assumed that the excess carrier distribution is generated according to the

Beer-Lambert Law ( ) ( )tixQ eeh

Ixq ωα

ναη

ω += − 12

, 0 , where β is an absorption coefficient, ηQ is

the quantum yield of the photogenerated carriers, hν is the incident photon energy and I0 is the

optical intensity. Figure 3.3 shows such geometry in one dimension.

Figure 3.3 Illustration of the concept of one-dimensional carrier density wave.

In this one-dimensional geometry, the boundary conditions can be written as

( ) ( ),,0, 10

* ωω NSxNdx

dD

x=

= (3.11a)

16

( ) ( ),,, 1* ωω LNSxN

dx

dD

Lx=−

= (3.11b)

where S1 and S2 are the surface recombination velocities on the two bounding surfaces x=0

and x=L, respectively.

Mandelis [Mandelis, 2001] used a Green Function Formalism to solve Eq. (3.10). The

resulting excess CDW field is

( )( ) ( )

Γ−Γ−+

Γ−ΓΓ−Γ

−= −−−

−−−

+−LxL

L

Lx

L

L

e

Q eee

ee

e

e

Dh

IxN e

e

e

e

e

eασ

σ

σασ

σ

ασ γγγγσαν

αηω )2(

212

212

12

121222*

0

2),(

(3.12)

where

1

11 SD

SD

e

e

+−=Γ ∗

σσ

(3.13a),

2

22 SD

SD

e

e

−+=Γ ∗

σσ

(3.13b),

1

11 SD

SD

e ++= ∗

σαγ (3.13c)

and 2

22 SD

SD

e −−= ∗

σαγ (3.13d)

3.4 Three dimensional CDW field in cylindrical geometries

Mandelis [Mandelis, 2001] deduced that the full three-dimensional cylindrically

symmetric photo-excited carrier-density-wave field in a cylindrical domain of infinite lateral

dimensions, which is generated by a Gaussian source, such as TEM00 laser beam of spot size

W, is given by the solution of a one-dimensional carrier density wave field generated by a

uniform source producing the same incident optical flux under the same boundary condition,

according to the operational transformation

( ) ( )[ ] ( ) λλλωλξσωλ

drJezNWzrNW

eeDD ∫∞

−→=

0

02

12

3

2

,,2,, (3.14)

where ( ) 22, ee σλωλξ += .

Based on the presented theorem the three dimensional expression of the excess carrier density

wave field in an electronic laterally infinite solid of thickness L can be easily obtained by

means of Eq. (3.12). The three dimensional geometry is presented in figure 3.4 where a free

17

carrier density flux is generated by a normally incident Gaussian laser beam of a spot size W.

The incident photons with energies hν cause the absorption and CDW generation which is

occurred according to the Beer-Lambert Law:( ) ( )tiz

W

rQ eeh

IrQ ωα

ναη

ω +=−−

12

,2

2

0 .

Figure 3.4 Illustration of the concept of three-dimensional carrier density wave.

Putting Eq.(3.12) to the Eq.(3.14) one can obtain the expression for three dimensional excess

CDW

( ) ( )

( )( )

∫∞

−−−−

−−−

+−

−−

+

−−

=0

0

22

4)2(

212

212

12

1212

*

0 ...

2),,(

22

λλλ

ξαν

αηω

λ

αξξ

ξαξ

σ

αξ

drJ

eee

egg

egge

eGG

eGggG

Dh

IzrN e

W

zxL

L

Lz

L

L

Qe

e

e

e

e

e

(3.15)

where J0(λr) is the Bessel function of first kind of order 0 and

1

11 SD

SDG

e

e

+−= ∗

ξξ

(3.16a),

2

22 SD

SDG

e

e

−+= ∗

ξξ

(3.16b),

1

11 SD

SDg

e ++= ∗

ξα

(3.16c)

and 2

22 SD

SDg

e −−= ∗

ξα

. (3.16d)

18

3.5 Recombination processes in semiconductors

Photoexcitated carriers in semiconductors can recombine by one of three mechanisms

[W.M.Bullis and H.R.Huff, 1996]: a Shockley-Read-Hall (SRH) recombination which is

related to multi-phonon release [W.Shockley and W.Read, 1952 and R.Hall 1952]; photon

release (radiative recombination) and Auger recombination, in which the recombination

energy is carried away by a third carrier.

Assuming that trapping is negligible, so the number of excess holes ),( trPr

is equal the

number of excess electrons ),( trNr

, the total bulk recombination rate Rtotal, will be the sum of

individual rates and the average carrier lifetime can be expressed as

111

1−−− ++

=∆=AugerradSRHtotalR

N

ττττ (3.17)

The SRH lifetime can be reproduced from [W.Shockley and W.Read, 1952] for one existing

dominant center:

( ) ( )NPN

PPPNNN NPSRH ∆++

∆+++∆++=00

100100 τττ (3.18)

where N0 and P0 are the equilibrium electron and hole densities, respectively. Here N1 and P1

are the equilibrium carrier densities related with energy of the defect center ED coincides with

the Fermi level and can be written as

( )

kTNN

TC EE

C

−−

= exp1 (3.19a)

( )

kTNP

VD EE

V

−−

= exp1 (3.19b)

where NC and NV are the densities of states in the conduction and valance band, respectively.

EC and EV are the energies of the conduction and valance band edge, respectively, k is the

Boltzmann constant, and T is the temperature. The time constant for capture of an electron

(hole) by an empty (full) defect state given by

thNDN N υσ

τ 10 = (3.20a)

thPDP N υσ

τ 10 = (3.20b)

19

where ND is the density of defect states. Nσ and Pσ are the capture cross sections for electrons

and holes by the defect, and thυ is the thermal velocity of carriers. The thermal velocity of

carriers is depend on temperature and is expressed by

m

kTth π

υ 8= (3.21)

where m is a mass of an electron.

The second mechanism of recombining carriers is related with a radiative recombination

lifetimes. In semiconductors with direct band gap transition of excited electrons from

minimum of conduction band to maximum of valance band (band-to-band) is more probable

than in the case of semiconductor with indirect energy band gap such as silicon. The

probability of this transition for both types of the semiconductors, with direct and indirect

energy band gap, can be described by the radiative recombination probability B which is

related to recombination lifetime in formula [Varshni, 1967]

( )NPNBrad ∆++=

00

1τ (3.22)

The coefficient B for semiconductors with indirect energy band gap is in order of 10-15

[Varshni, 1967; Gerlach et al., 1972; Augustine et al. 1992] while for direct energy band gap

semiconductor is 10-9 [cm3/s].

The recombination lifetime expression in the case of Auger recombination can be

written as [Schroder, 1998]

( ) ( )20

20

20

20 22

1

PPNNCNNPPC NPAuger ∆+∆++∆+∆+

=τ (3.23)

where CP and CN are the Auger coefficient for holes and electrons, respectively. In highly

doped silicon those coefficients were found by Dziewoir and Schmid [Dziewoir and Schmid,

1977]: 31108.2 −×=NC 32109.9 −×=PC [cm6/s].

For demonstration purposes one can calculate recombination lifetimes for p-type silicon is

shown in Figure 3.5 as a function of injection level for assumed defect level in the middle of

the band gap and the following parameter values: 31108.2 −×=NC 32109.9 −×=PC [cm6/s],

15104 −×=B [cm3/s], 14101 −×== NP σσ [cm2], 71007.1 ×=thν [cm/s],

191086.2 ×=CN , 19101.3 ×=VN , 1210=TN [cm-3], 562.0=TE [eV]and 1510== AA Np [cm-3]

which corresponds to a resistivity of approximately 15 Ωcm.

20

1015 1016 1017 1018 1019 102010-11

10-9

10-7

10-5

10-3

10-1

101

SRH radiative Auger total

Rec

ombi

natio

n Li

fetim

e [s

]

Carrier Injection level [cm-3]

Figure 3.5 Calculated recombination lifetimes in silicon as a function of injection level

for a p-type Si wafer with a doping density 1510== AA Np [cm-3]. Assumed

parameter values are given in the text.

It is clear seen that the radiative recombination lifetime is much longer than the SRH or Auger

recombination time constant and has negligible effects on the overall recombination rate. For

low injection level (of an order 1710 [cm-3]) the recombination processes are dominated by

the SRH recombination time constant while at higher injection level by the Auger

recombination time constant.

21

Chapter 4: Experimental methods, signal

generation mechanisms and instrumentation of

PCR and PPE methods

4.1 The Photocarrier Radiometry signal generation mechanism and instrumentation

4.1.1. Introduction

Mandelis et al. [Mandelis et al., 2003] proposed a new technique for the measurement of

the carrier-density – wave and named it the Photocarrier Radiometry (PCR). A modification

of this technique is the room- or high temperature near infrared photoluminescence (NIR-PL).

In the past the NIR-PL has been associated with the presence of impurities or defects and

band-to-band recombination [King and Hall, 1994, Haynes 1956; Varshni 1967].

22

4.1.2. Contribution to the PCR signal

The PCR is associated with room- or high temperature near infrared photoluminescence.

In Figure 4.1 one can see that the photoluminescence spectrum of silicon at room temperature

has two peaks [King and Hall, 1994] The first at 1.09 eV (≈ 1.14 µm) is associated with band

– to – band transitions [Haynes 1956; Varshni 1967]. The second one is an approximately at

0.73 eV (≈ 1.6 µm) and is observed only for silicon wafers grown by Czochralski method and

is associated with oxygen dependent defects complexes [Kitgawara 1992; King and Hall,

1994].

Figure 4.1 Measured photoluminescence spectra at T=30, 130 and 300 K for

Czochralski-grown Si annealed at T=450 C. The data were obtained using a Ge

photodetector [King and Hall, 1994].

The PCR signal theory was discussed thoroughly by Mandelis et al. [Mandelis et al., 2003].

They authors considered an elementary slice of thickness dz, centered at depth z in a

semiconductor slab supported by a backing, but not necessarily in contact with the backing,

Fig 4.2.

23

Figure 4.2 Cross-sectional view of contributions to front-surface radiative emissions

of IR photons from (a) a semiconductor strip of thickness dz at depth z; (b) reentrant

photons from the back surface due to reflection from a backing support material; (c)

emissive IR photons from the backing at thermodynamic temperature T. The carrier-

wave depth profile results in a depth dependent IR absorption/emission coefficient

due to free carrier absorption of the infrared photon fields, both ac and dc [Mandelis,

2003].

A modulated laser beam at angular frequency ω=2πf and wavelength λvis impinges on the

front surface of the semiconductor. The super-bandgap radiation is absorbed within a (short)

distance from the surface and excited carriers are subjected to several de-excitation processes

discussed in Chapter 3. At thermal and electronic equilibrium, a detailed consideration of all

IR emission, absorption, and reflection processes [Mandelis, 2003] yields an expression for

the total IR emissive power at the fundamental modulation frequency across the front surface

of the material in the presence of a backing support which acts both as reflector of

semiconductor-generated IR radiation with spectrum centered at λ. Instrumental filtering of all

thermal infrared emission contributions and bandwith matching to the IR photodetector

allows for all Planck-mediated (8-12 µm) terms to be eliminated and the PCR signal can be

written follow as.[Mandelis, 2003]

( ) ( )[ ] ( )[ ] ( )[ ] ( ) ( )∫ ∫++−=2

1 0

11 ,,111λ

λ

λωεληλλλλωL

fceRRbPCR dzzWRRRdS (4.1)

24

where R1 is the front surface reflectivity, Rb is the backing support material reflectivity, εfc is

the IR emission coefficient due to the free photoexcited carrier wave, WeR(λ) is the spectral

power per unit wavelength, the product to the recombination transition rate from band do

band or from bandedge to defect or impurity state, as the case may be, multiplied by the

energy difference between initial and final states, ηR is the quantum yield for IR emission

upon carrier recombination into one of these states.

For a semiconductor which is in thermal and electronic equilibrium with its environment

Kirchhoff’s theorem is fulfilled:

( ) ( )λωαλωε ,,,, zz fcIRfc = , (4.2)

where αfc is absorption coefficient due to the free photo-excited carriers.

For relatively low carrier densities the absorption coefficient depends on the free carrier

density as [Smith, 1978]

( ) ( )λωµεπ

λλωα ,,4

,,2*3

02

2

zNnmc

qzfcIR ∆= . (4.3)

Putting (4.3) into Eq. (4.1) one can write the expression for the PCR signal in one dimension

( ) ( ) ( )∫∆=L

PCR dzzNFS0

21 ,, ωλλω , (4.4)

where ( ) ( )[ ] ( )[ ] ( )[ ] ( ) ( ) λλληλλλλλλ

λ

dCWRRRF eRRb∫ ++−=2

1

1121 111, .

4.1.3. Instrumentation and normalization of PCR signals

Four photocarrier radiometric systems have been used. The first was related with study

of the effect of the space charge layer on the PCR signal system and the results are presented

in Chapter 6. The second and third systems were constructed in the Center for Advanced

Diffusion Wave Technologies, University of Toronto, Canada [Shaungnessy, 2005] and were

used for the study of the influence of the optical excitation intensity on the PCR signal

(Chapter 7). The last one is the PCR microscope constructed in order to monitor the ion

implanted process in silicon wafers. The common part of all these systems is an InGaAs p-i-n

photodetector (Thornlabs model PDA 400) with the following parameters: spectral bandwidth

of 700-1800nm; an active area of 0.8 mm2, and an adjustable transimpedance gain; the unit

was used at the intermediate gain setting (1.5 ×105 V/A) at which it has a noise equivalent

power (NEP) of 3.8 ×10-12 W Hz-1/2 (at 1310 nm) and a frequency bandwidth of 700 kHz) a

long-pass filter (Spectrogon model LP-1000: a steep cut-on (5% at 1010 nm, 78% at 1060

25

nm) and a transmission range 1042 – 2198 nm is placed in front of the detector in order to

ensure that any diffuse reflections of the excitation source do not contribute to the signal). The

spectral responsivity is shown in Fig 4.3.

Fig 4.3 Spectral responsivity of the PCR detector.

All frequency dependent measurements were normalized by the corresponding wide-

bandwidth instrumental transfer functions. The transfer functions were obtained by measuring

the amplitude and phase of modulated laser radiation scattered from a microscopically rough

metallic surface positioned at the focal plane of the parabolic mirror, and partly transmitted

through the filter.

4.1.4. The one dimensional Photocarrier Radiometry (PCR) Signal

The PCR signal in one dimensional geometry (as this from Figure 3.3) can be written

with help of Eq. (4.4) and Eq. (3.12)

( ) ( ) ( ) ( ) ( ) ( )∫ =∆=L

DDD MEFdxxNFS0

1121211 ,,, ωωλλωλλω (4.5)

where

( ) ( )( )22*

01 2

1

eD Dh

RIE

σανηαω

−−= (4.6)

( ) ( ) ( ) ( )LeL

LLL

D ee

eeeM

e

eeα

σ

ασασ

ασγγω −

−+−−

−−Γ−Γ

+Γ−+Γ= 12

12

)(1221

1 (4.7).

26

For silicon wafers, this equation has been applied with superband-gap radiation of absorption

coefficient α(hν) > 103 cm-1, such that the semiconductor material is entirely opaque to the

incident radiation, and thus e-αL ≈ 0. The quantities in (4.6) and (4.7) were defined in Chapter

3 (Eqs.: 3.13 a-d). Using MATLAB program and equation (4.5) simulations of the electronic

parameters on the PCR signal were performed.

10 1 10 2 10 3 10 4 10 510 18

10 19

10 20

10 21

10 1 10 2 10 3 10 4 10 5

-80

-60

-40

-20

0 τ

n=5 µ s

τn=50 µ s

τn=100 µ s

τn=500 µ s

τn=1000 µ s

PC

R A

mpl

itude

[a.u

.]

F re qu e nc y [H z ]

τn= 5 µ s

τn= 50 µ s

τn= 100 µ s

τn= 500 µ s

τn= 1000 µ s

PC

R P

hase

[deg

.]

F req u en c y [H z ]

Figure 4.4 The PCR Amplitude (a) and Phase (b) of p-type silicon wafer versus modulation

frequency with the different values of minority bulk recombination lifetime. Parameter

settings: S1=300cm/s, S2=105 cm/s, Dn* = 30 cm2/s, L = 550 µm, α(λ=514nm) = 7.76×103 cm-1.

Figure 4.4 shows a behavior of the PCR amplitude (a) and phase (b) for a silicon wafer with

the minority recombination lifetime τn in the variations in the range 5 µs ≤≤ nτ 1 ms. A

decrease in τn diminishes the PCR amplitude and shifts the position of the turning point

(“knee”) to higher frequencies as the density of the carrier-wave over one period decreases

with decreasing recombination time [Mandelis, 2001]. The PCR phases exhibit zero delay

with respect to the modulation source at low frequencies, such that ωτn << 1 but they begin to

lag behind the source phase as soon as this condition is not valid. As τn decreases, the

foregoing condition becomes violated at progressively higher frequencies, whence the shift of

the PCR phases in Fig. 4.4b [Mandelis et al., 2003].

27

101 102 103 104 105

-80

-60

-40

-20

0

101 102 103 104 105

1019

1020

D*

n=5 [cm2/s]

D*

n=10 [cm2/s]

D*

n=20 [cm2/s]

D*

n=30 [cm2/s]

D*

n=45 [cm2/s]

PC

R P

hase

[deg

.]

Frequency [Hz]

D*

n=5 [cm2/s]

D*

n=10 [cm2/s]

D*

n=20 [cm2/s]

D*

n=30 [cm2/s]

D*

n=45 [cm2/s]

PC

R A

mpl

itude

[a.u

.]

Frequency [Hz]

Figure 4.5 The PCR Amplitude (a) and Phase (b) of p-type silicon wafer versus

modulation frequency with the different values of ambipolar diffusivity. Parameter

settings: τn = 100 µs, S1 = 300 cm/s, S2 = 105 cm/s, L = 550 µm, α(λ=514nm) =

7.76×103 cm-1

Figure 4.5a shows the change in the PCR amplitudes affected by altering the

ambipolar diffusivity Dn*. If Dn

* is controlled by the bulk of the semiconductor, then an

increase of this quantity will decrease the PCR amplitude. This behavior can be explained by

the fact that the CDW “centroid” – center of the charge carriers - shifts away from detection

point at surface therefore the contribution of the recombining carrier density wave to the PCR

signal generated at the surface (or/and subsurface) is smaller. At low frequency the PCR

phase doesn’t show any lag, until 1>>ωτ is fulfilled, then an onset of the PCR phase lag is

observed. The PCR phase lag exhibits a shift to higher frequencies with increasing Dn*. High

frequencies can affect the position of the CDW centroid shifting it to smaller depth. This

effect is observed on the PCR amplitude and the PCR phase.

28

101 102 103 104 105

1019

1020

101 102 103 104 105-100

-80

-60

-40

-20

0

S1=0 [cm/s]

S1=0.1 [cm/s]

S1=1 [cm/s]

S1=10 [cm/s]

S1=1000 [cm/s]

P

CR

Am

plitu

de [a

.u.]

Frequency [Hz]

S1=0 [cm/s]

S1=0.1 [cm/s]

S1=1 [cm/s]

S1=10 [cm/s]

S1=1000 [cm/s]

PC

R P

hase

[deg

.]

Frequency [Hz]

Figure 4.6 The PCR Amplitude (a) and Phase (b) of p-type silicon wafer versus

modulation frequency with the different values of front recombination velocity.

Parameter settings: τn = 100 µs, S2 = 105 cm/s, Dn* = 30 cm2/s, L = 550 µm,

α(λ=514nm) = 7.76×103 cm-1.

Figure 4.6 shows the effect of changing the recombination velocity S1 on the PCR

frequency scans. The PCR amplitudes decrease (fig. 4.6a) as the value of S1 increases. This

behavior is quite similar to the case observed for decreasing of the minority recombination

lifetime τn, although the “knee” shift to higher frequencies is not as pronounced. A similar

behavior is observed for the PCR phases (fig. 4.6b), where the phases lag move to higher

frequencies. Additionally, the phase lag shows gradual decrease with increasing S1 due to the

sub-surface ac diffusion length (or “centroid”) of the CDW which is no longer controlled by

the bulk recombination lifetime τn alone but it becomes controlled by an effective lifetime, τeff,

defined as follows [Mandelis 2005]:

29

speff τττ111 += (4.8)

where τs is the interface lifetime related to the interface recombination velocity S1. This time

constant begins to influence the effective lifetime (and hence the phase saturation level) at S1

values such as τs ~ τn.

Figure 4.7 shows the effect of changing recombination velocity S2 on PCR frequency scans.

101 102 103 104 105-100

-80

-60

-40

-20

0

101 102 103 104 105

1019

1020

S2=0 [cm/s]

S2=0.1 [cm/s]

S2=1 [cm/s]

S2=10 [cm/s]

S2=1000 [cm/s]

PC

R P

hase

[deg

.]

Frequency [Hz]

S2=0 [cm/s]

S2=0.1 [cm/s]

S2=1 [cm/s]

S2=10 [cm/s]

S2=1000 [cm/s]

PC

R A

mpl

itude

[a.u

.]

Frequency [Hz]

Figure 4.7 The PCR Amplitude (a) and Phase (b) of p-type silicon wafer versus

modulation frequency with the different values of rear recombination velocity.

Parameter settings: τn = 100 µs, S1 = 300 cm/s, Dn* = 30cm2/s, L = 550 µm,

α(λ=514nm) = 7.76×103 cm-1.

30

4.1.5. The three dimensional PCR signal

In Chapter 3 (Section 3.4) it was shown that the photo-excited carrier density wave field

in three dimensional cylindrical geometries can be described by: (see Eq. (3.14) )

( ) ( )[ ] ( ) λλλωλξσωλ

drJezNWzrNW

eeDD ∫∞

−→=

0

02

12

3

2

,,2,, (4.9)

where ( ) ( )[ ]2

21

23 ,,2,,

~

−→=

W

eccDD ezNWzNλ

ωλξσωλ is the Hankel transform of Eq. (3.14).

In order to account for contributions over the thickness of the wafer the Hankel transform of

the carrier density field has to be integrated over the depth:

( ) ( )∫=L

DD dzzNN0

33 ,,~

,~ ωλωλ (4.10)

The finite area of the detector must be taken into consideration to account for carrier diffusion

out of the field of view of the collection optics/detector assembly [Shaughnessy, 2005].

Assuming a disc of radius a2 and area A as the effective detector size and using the relation

[Ikari et al., 1999]:

( ) ( ) ( ) ( )∫ =2

0

2132

03 ,~1

,~1

a

DD aJNa

dJNA

λωλλπ

ρρλρωλ (4.11)

where J1is the Bessel function of first kind of order 1. The PCR signal can be expressed in

final form as the inverse Hankel transform of (4.9) integrated over the detector area

( ) ( ) ( )∫∞

=0

2132

3, ,~ λλωλ

πω daJN

a

CS DDPCR (4.12)

where

( ) ( ) ( )ωλωλωλ ,,,~

333 DDD MEN = (4.13)

( ) ( )( )22*

40

3 2

1,

22

e

W

D Dh

eRIE

ξαναηωλ

λ

−−=

(4.14)

( ) ( ) ( ) ( )[ ] ( )α

ωλωλξ

ωλα

ξξ L

L

e

L

D

eeCC

eM e

e −−

− −−+−= 1,,

1, 213 (4.15)

( )( )

−−= −

−−

L

L

e

e

eGG

eggGGC ξ

ξα

ωλ2

12

21211 , (4.16)

31

( )( )

L

L

e

e

eGG

eGgGgC ξ

ξα

ωλ2

12

22112 , −

−−

−−= (4.17)

Coefficients in (4.12) - (4.17) were defined in Chapter 3.

4.1.6. The photocarrier radiometry (PCR) dimensionality criterion

Figure 4.8 shows a frequency-scan simulations in two sets of linear PCR signals from

silicon based on 1-D (full lines) and 3-D (squares and rhombs) theoretical models (Eqs 4.5

and 4.10, respectively) with laser wavelength 830 nm, spotsize 4 mm and recombination

lifetimes τ = 20 µs and 800 µs. The optical absorption coefficient was taken to be αP = 635

cm-1. The carrier transport parameters were assumed equal for both sets of curves. The

amplitude curves are normalized to unity at f = 10 Hz.

0.1

1

0.01 0.1 1 10 100-100

-80

-60

-40

-20

0

1-D Simulation τ - 800 µs

1-D Simulation τ - 20 µs

Log(

Am

plitu

de)

[a.u

.]

1-D Simulation τ - 800 µs

1-D Simulation τ - 20 µs

Pha

se [d

egre

es]

Log (Frequency [kHz])

Figure 4.8 PCR frequency scan simulations with short and long carrier recombination

lifetimes using 1-D and 3-D theoretical model. 3-D simulations with τ = 20 µs () and

800 µs (); 1-D simulations with τ = 20 µs (—) and 800 µs (—) coincide with the

corresponding 3-D simulation. Other transport parameters: D = 15 cm2/s, S1 = 200

cm/s, S2 = 105 cm/s, αp = 659 cm-1. Laser beam spotsize: 4 mm.

32

It is clear that the simulations using the 1-D and the 3-D equations coincide, as expected, for

the chosen large spotsizes compared to the maximum carrier-wave diffusion length, LD(ω) = |

σn(ω) |-1 at the lowest frequency f = 10 Hz where LD(10 Hz) =173.2 µm for τ = 20 µs and

1,095 µm for τ = 800 µs. In practice, the use of 1-D theory to explain PCR data is warranted

when a change of the beam spotsize on the semiconductor surface does not produce

measurable change in the PCR phase. This is an important dimensionality criterion, therefore,

the dependence of the PCR phase on laser spotsize using the 1-D or the 3-D theory with

various lifetimes (or diffusion lengths of the photo-excited free carrier density-wave) at two

frequencies is presented in Fig. 4.9.

10 100 1000

-20

-15

-10

-5

0

10 100 1000

-80

-70

-60

-50

-40

-30

-20

-10

0

830 µm387 µm 24 µm

PC

R P

hase

[deg

ree]

Log(Laser Spotsize [µm])

Lifetimes: τ

1 - 1 µs;

τ2 - 20 µs;

τ3 - 50 µs;

τ4 - 200 µs;

ba

830 µm 387 µm24 µm

PC

R P

hase

[deg

ree]

Log(Laser Spotsize [µm])

Figure 4.9 Laser beam spotsize dependence of PCR phases for frequency 1 kHz (a)

and 100 kHz (b) with broad range of lifetimes and otherwise same other transport

parameters: D = 10 cm2/s, S1 = 500 cm/s, S2 = 105 cm/s, αp = 659 cm-1.

It is clearly seen that, for PCR signals with τ = 1 µs, 1-D theory can be used with spotsizes

2W ≥ 387 µm at 1 kHz and 100 kHz. However, with τ = 20 µs, 50 µs, and 200 µs, the

condition 2W ≥ 830 µm is required at 100 kHz. In the latter cases all lower frequency ranges

require a 3-D theoretical approach.

33

4.1.7 Photo-carrier radiometry microscope

In this section the capability of the PCR technique to monitor a quality of ion implanted

wafers is presented. In order to present that PCR signal is sensitive to the change of carrier

transport properties the photocarrier radiometry microscope was constructed and it is

presented on Figure 4.10. As an excitation source of carrier density waves a 808 nm laser

diode (0.5mm beam radius) was used. The power of the laser diode was typically 200 mW.

The diode laser beam was focused onto the sample surface using lens. The position of the

laser beam is coincident with the focal point of an off-axis paraboloidal mirror that collects

a portion of infrared radiation from the samples. The collected light is then focused onto

the detector by means of lens. Sample was placed onto aluminum holder (acted as a

mechanical support and signal amplifier by redirecting the forward emitted IR photons

back toward the detector [Mandelis, 2003]. The x-y position scans were realized by means

of homemade x-y motor stage. All instruments, data acquisition, and data storage are

controlled by a computer running Pascal program with a graphical user interface and real-

time display of experimental data.

34

Figure 4.10. The photocarrier radiometry microscope.

Typical result obtained using modulation frequency 10 kHz for an ion implanted wafer is

presented at Fig. 4.11.

Figure 4.11: The PCR amplitude and phase as a function of x-y-scan of the silicon

wafer implemented with 6.3 1016 doses of protons [cm-2]. Sample preparation: The

35

energy of the protons was 1 MeV with an implantation depth of 18 µm, the beam

was focused on 5x5 mm2 area on the silicon wafer surface.

The squares on Figure 4.11 depict the ion implanted regions. Figure 4.12 shows the PCR

phase as a function of coordinate.

Figure 4.12: The PCR phase a a function of x-scan of the silicon wafer

implemented with 6.3 1016 doses of protons [cm-2] at the difference frequencies.

Sample preparation: The energy of the protons was 1 MeV with an implantation

depth of 18 µm, the beam was focused on 5x5 mm2 area on the silicon wafer

surface.

From Fig. 4.12 one can see that for 1 kHz the PCR technique is unable to detect any

changes in the PCR phase so in the carrier transport properties. Whereas above 1 kHz

changes in the PCR phases are clearly seen. The PCR phase lag was observed in an ion

implanted region. This is can be explained because the photo-carrier diffusion length is too

36

large to detect any inhomogeneous in the free carrier density depth in the implantation

region.

4.2 The photopyroelectric (PPE) signal generation mechanism and instrumentation

4.2.1. Experimental Setup

The PPE measurements were performed in the back detection configuration, where the heat

is generated on the front side of the sample and the temperature oscillations are measured with

the pyroelectric detector contacted to the back side of the sample. The experimental setup

constructed for the back detection configuration is presented in Fig. 4.13.

Figure 4.13: PPE experimental set up

The thermal wave are excited by an argon ion laser with output power 200 mW and

operating wavelength λ = 514 nm. The laser beam of 1.89 mm diameter was intensity modulated

by means of an acousto–optical modulator in the frequency range 1 Hz to 10 Hz and focused onto

the sample. The front surface of the sample was covered by an optically opaque 20 µm to 30 µm

graphite coating. Samples were attached to a pyroelectric detector by means of a grease layer

37

(Apiezon T grease). As the grease layer was very thin, its contribution to the PPE signal could be

neglected. A 0.98 mm thick lead zirconate titanate PZT crystal was used as a pyroelectric

detector. The PPE signal detection was performed by means of a lock-in amplifier (Stanford

830). The detector was placed on a cooper plate with a drilled hole (inside was air). The sample –

detector – copper support assembly was placed in an aluminum chamber. A schematic of the

PPE chamber is presented in Fig. 4.14.

Fig. 4.14: PPE chamber: 1 Peltier-element, 2 aluminum support, 3 cylindrical cooper

support, 4 the PZT detector, 5 sample with optically opaque cover-layer, 6 quartz

window, 7. BNC connector for PPE Signal

The temperature was varied in the range from 20° C up to 40° C by means of a Peltier

element which was driven by a homemade current controller (Appendix A).

4.2.2. The PPE signal generation mechanism.

The average temperature oscillation Tp at angular frequency ω0 in a pyroelectric detector

leads to variations of the surface charge density Q due to the pyroelectric effect and can be

written according to B. R. Holeman [Holeman, 1972] as:

38

( ) ( ) ,0p0 ωω TpQ = (4.18)

where p is the pyroelectric coefficient of the detector. Time-dependent variations of the surface

charge causes a current flow through the detector of the thickness Lp [Mandelis and Zver, 1985;

Rombouts et al., 2005];

( ) ( ) ( ) ( ) ,eed

dd),(

1

d

d

dt

d00

p

0p00pp

0p00

titi

L

ipAt

xxTL

pAt

TpA

QAI ωω ωθωω

ωωω =

=== ∫ (4.19)

where A is the transducer area and ( ) xxTL L

d),(1

p

0pp

0p ∫= ωωθ . Tp(ω0,x) is the temperature field in

the pyroelectric detector.

Figure 4.15 shows a schematic of the sample’s model.

Figure 4.15 Schematic of the sample’s model

39

In the used arrangement the front surface was covered by a 20 µm to 30 µm thick graphite layer

which prevents exciting light to penetrate into the sample. The rear surface was connected to the

detector which monitored the thermal wave transmitted through the sample. The distribution of

the thermal wave is the solution of one-dimensional thermal transport equations as a result of heat

conduction through the sample. Similar theoretical models were considered by Chirtoc and

Mihalescu [Chirtoc and Mihalescu, 1989] and Mandelis and Zver [Mandelis and Zver, 1985]. In

both works the influence of the thermal interface between the rear surface of the sample and the

detector was neglected. Experimentally a good thermal contact was achieved with a very thin

grease layer. As in the presented experiments the thermal waves were generated by surface

heating the contribution to the heat transport problem of the thermally thin graphite surface layer

can be neglected. Also, the thermal contact of the sample to the detector by the grease layer is

considered to be ideal. In some cases, however, the thermal diffusivity Dt of the sample can be

underestimated due to the influence of the grease layer as demonstrated by Salazar

[Salazar,2003]. He calculated the error of the Dt estimation in the presence of about a 2 µm to 3

µm thick grease layer. He found that the error is large for thin and good thermal conductors at

high frequencies and decreases with increasing thickness and decreasing thermal diffusivity of a

material and modulation frequencies. Although in our measurements we used a different grease,

one can deduce that the investigated samples as well as a glassy carbon are rather poor thermal

conductors. Furthermore, this effect is additionally reduced because measurements were

performed at modulated low frequencies. When the sample and the detector are both thermally

thick and optically opaque the temperature field can be obtained using a formula of Chirtoc and

Mihalescu [Chirtoc and Mihalescu, 1989] and Mandelis and Zver [Mandelis and Zver, 1985]:

( ) ,22

exp2

exp

1

ss

0s

s

0

0

sp

ps

p

ps0

+−

+

=Θ LD

iLD

Dk

Dkk

Dp

ωπω

ω

ηω (4.20)

where ηs is the nonradiative conversion efficiency of the absorbing layer: The sample is

characterized by a thickness Ls, a thermal diffusivity Ds and a thermal conductivity ks. The

40

detector is characterized by a thermal diffusivity Dp and a thermal conductivity kp. The PPE

signal is then given by

( ) ( ) ( ) ,22

exp2

exp1 s

s

0s

s

0

spp

ps00

+−

+= L

DiL

Dbk

DApII

ωπωηω (4.21)

where I0 is the intensity of the optical excitation: The phase of the PPE-signal is given by

.22 s

s

fmLD

f −−=−−= πππϕ (4.22)

The coefficient m can be easily determined from experimental data. Thus, with the known sample

thickness the thermal diffusivity can be deduced by the relation

.2

s2

s m

LD

π= (4.23)

The PPE amplitude can be written as

( )( ) ,2

lns

s2

0s

s2

0 fmBfD

LB

D

LBI −=−=

−+= πωω (4.24)

where

+

=

sp

ps

p

ps0

1

ln

Dk

Dkk

DApIB

η.

4.2.3. Normalization of the PPE Signal

Figure 4.16 shows the PPE amplitude and the PPE phase from the detector alone as a

function of the modulation frequency in the temperature range from 26 °C to 36 ºC. Error bars

for PPE phases were approximately 1.5°.

41

2 4 6 8 10

90

92

94

96

2 4 6 8 101.60E-010

1.65E-010

1.70E-010

1.75E-010

1.80E-010

Frequency [Hz]

26 0C 27 0C 28.5 0C 31 0C 33 0C 36.50C

PP

E p

hase

[deg

]

PP

E a

mpl

itude

[A]

Frequency [Hz]

Fig. 4.16: PPE amplitudes (a) and phases (b) of the detector alone at different

temperatures (in °C).

From the experimental data in Fig. 4.16b, one can see that in the investigated range of the

temperature the PPE phases remains constant within error bars. This means that the thermal

properties of the detector can be assumed constant under our measurement conditions. We had

also observed small changes in the PPE amplitudes, Fig. 4.16a, but these can be caused by the

temperature-dependence of the pyroelectric coefficient and/or thermal effusivity ep (ep = kp(Dp)-

1/2) of the PZT detector as well as long term fluctuations of the laser intensity. These effects can

be minimized by an appropriate normalization procedure. Detenclos et al. [Delenclos et al., 2001]

normalized the PPE signal from an investigated material to the one obtained with the detector

alone or to the signal obtained with a reference sample. They considered the PPE signal for the

sample and the detector both thermally thick and optically opaque and pointed out that the

normalized signal is not influenced by the temperature-dependence of the pyroelectric

coefficient, hence, only a knowledge of thermal effusivity of the detector is required. In fact, PPE

amplitudes were normalized to the reference sample instead of the detector alone as the

absorption of laser light at the detector electrode is different from that in the graphite layer

[Delenclos et al., 2001]. In addition, it is also possible that the heating spot (laser beam spot)

42

interacts (energy exchange) with silver contacts on the surface of the detector, and this could lead

to a worse signal-to-noise ratio (SNR) than in the case of normalization to a reference material.

As a reference sample, a 0.98 mm thick piece of a glassy carbon (type G) was used. The

specific heat capacity C of the glassy carbon in the temperature range from 26 °C to 80 °C was

determined from differential scanning calorimetry (DSC) measurements and these results are

presented in Table 4.1.

Table 4.1 presents the specific heat capacity of the glassy carbon (thickness L=0,98mm)

in different temperatures estimated from the PPE phases and amplitudes and result of

the DSC measurements.

Temperature [°C] The specific heat capacity

C [J/kg°C]

26.85 1054.568

36.85 1064.533

46.85 1151.807

56.85 1268.092

66.85 1397.104

76.85 1510.841

86.85 1612.371

The error limit of the differential scanning calorimetry (DSC) measurements was 3% to

5%. One can see that in the covered temperature range (from 20 to 40° C) the specific heat

capacity C is about 1050 J⋅kg-1⋅K-1 within the error limit.

Figures 4.17a and 4.17b present the PPE signal phases and amplitudes of the glassy

carbon, respectively, at room temperature as a function the square root of the modulation

frequency. It is worthwhile to note that the error bars were approximately 0.5°.

43

1.5 2.0 2.5 3.0 3.5

-1.0

-0.5

0.0

0.5

1.0 1.5 2.0 2.5 3.0 3.5

e-16

e-15

experimental data the best linear fitting

Sqrt(Frequency) / Hz

PP

E P

hase

/ de

g

experimental data the best linear fitting

Sqrt (Frequency) /Hz

ln(A

mpl

itude

)/au

Fig.4.17 The PPE Phase and Amplitude of the glassy carbon (thickness L=0.98µm) as a

function of a square frequency, respectively. The linear fitting of the experimental PPE

Phase to the Eq. 4.22 gives m=-0,860±0,003, whereas the linear fitting of experimental

PPE amplitude to the Eq. 4.24 gives m=0,857±0,001.

Using Eqs. (4.22) and (4.23) the thermal diffusivity from the as measured PPE phases of

the glassy carbon was estimated to Ds = 4.22x10-6 m2⋅s-1. The same value of the thermal

diffusivity was obtained from the measured PPE amplitudes by Eq. (4.24) and Eq.(4.23). Using

the literature value of the mass-density ρ = 1.42⋅103 kg⋅m-3 [http://www.htw-gmbh.de/] the

thermal conductivity of the glassy carbon type G was calculated to ks = 6.3 W⋅m-1⋅K-1 which is in

excellent agreement with the value deduced from the data sheet of the producer of the glassy

carbon [http://www.htw-gmbh.de/]. This demonstrates the reliability of the present experimental

setup and measurement procedure for the experimental determination of the thermal diffusivity.

Table 4.2 presents temperature dependence of the thermal diffusivity of the glassy carbon

calculated using Eq. (4.22) - (4.23).

44

Table 4.2 presents the thermal diffusivity of the glassy carbon (thickness L=0,98mm) at

different temperatures determined from the PPE phases and amplitudes.

Temperature

[°C]

mAmpl mPhase DAmpl x 10-6

[m2/s]

DPhase x 10-6

[m2/s]

22.2 -0.841 -0.845 4.266 4.226

24.7 -0.842 -0.845 4.256 4.226

27.4 -0.843 -0.845 4.246 4.226

29.8 -0.845 -0.847 4.226 4.206

33.1 -0.845 -0.848 4.226 4.196

36.0 -0.847 -0.849 4.206 4.186

39.0 -0.849 -0.852 4.186 4.156

As compared to the black glassy carbon, the investigated semiconductor samples have smaller

absorption coefficients β; thus, the light can penetrate deeper into the sample generating heat also

in the subsurface regions. For this reason a thin black graphite layer was deposited on the surface.

The optically opaque cover layer at the present experimental conditions (514 nm laser) avoids the

super-bandgap excitation that creates photocarriers which can act as scattering centers for

phonons. The scattering centres alter the thermal transport properties (decreasing k) of

investigated semiconductors and complicate the experimental data interpretation. Figure 4.18a

shows the frequency dependence of the PPE amplitude and phase in the presence of the graphite

cover layer on the glassy carbon sample.

45

1.0 1.5 2.0 2.5 3.0 3.5

e-16

e-15

1.0 1.5 2.0 2.5 3.0 3.5-1.5

-1.0

-0.5

0.0

0.5

Sqrt (Frequency) /Hz

ln(A

mpl

itude

)/au

experimental data the best linear fitting

PP

E P

hase

/ ra

d

Sqrt(Frequency) / Hz

Figure 4.18 The PPE phase and amplitude of the glassy carbon (thickness L=0.98mm)

with a graphite layer as a function of a square frequency, respectively. The linear fitting

of the experimental PPE phase to the Eq. 4.22 gives m=-0,860±0,003, whereas the linear

fitting of experimental PPE amplitude to the Eq. 4.24 gives m=0,863±0,003.

It is clearly seen that the graphite layer does not affect the PPE phase. The same value of the

coefficient m was obtained for PPE amplitudes, Fig. 4.18b. Therefore, the thermal diffusivity can

be calculated from the as measured PPE phases and amplitudes in the presence of the graphite

coating in the same way as for the glassy carbon sample alone.

Formula (4.20) describes the thermal wave field in the sample when the sample and the

detector are both thermally thick and optically opaque. Experimentally this condition is fulfilled

in a certain frequency range for a given sample and detector geometry (thickness). The lower

frequency limit – the threshold frequency –for the applicability of the simplified formulaes such

as equ. 4.20. is very difficult to determine in the experiment. In order to illustrate the effect of the

threshold frequency one can proceed from equ.3.8, the full-expression for the thermal wave

[Mandelis and Zver, 1985]

( ) ( ) [ ]( ) [ ]( ) (

×−−−+−

= − ...1111

2 220

0 bpL

bpL

sss

ss

ppp bebe

kL

Ipppp σσ

σβηβ

σωθ

( ) ( )( ) ( )( )[ ][ ]) ( ) ×

−+−−+++−+

−−−

22111112

ppp

LppLL

sgsL

sgsssg k

eeebrebrrb

ss

ssssss

σβηβ β

βσσ

46

]( ) ]( )[ ( ) ]( ) [ ][[([ ×−++−++−−−+− −− pppppppp L

ps

L

rpsbp

L

bp

L eberbbebe σσσσ 11111111

( )( ) ) ( )( ) ( )( ) [ ]]×−−−+++−−− −−− pppppppp LLbsbp

Lbsbp

p

Lpbpps eebbebb

rerbb βσσβ 11111

11

( ) [ ]( ) [ ] ( )([ ( ) [ ] ( )+−−+−−−−+−++ − 11111111 psL

ppsbpL

bpLL

sg berbbebeeb ppppppssσσσσ

[ ]( )( ) ) ( )( ) ( )( ) ×+−+−+−−+− −−− pppppppp Lpsbp

Lpsbp

p

Lpbpps

Lebbebb

rerbbe

σσβσ 11111

11

[ ]]( ) ( )( )( ) ( )( )[ ]( +−−++++÷−− −−−ssppppsspp LL

psbpL

psbpsgL

sgL eebbebbbebe σσσσβ 1111111

( )( )( ) ( )( )[ ] )sspppp LLpsbp

Lpsbpsg eebbebbb σσσ −−+−+−+− 11111 (4.25)

where ( ) jj ai+= 1σ

nn

mmmn ak

akb =

( )ssLX σexp=

( )spLY σexp=

( )( )psp LLZ += σexp

( )220

2 sss

ss

k

IE

σβηβ−

=

( ) ( )( )spsppp

pp Lk

IF ββ

σβηβ

−−−

= exp2 22

0

Substituting Eq. (4.25) into Eq.(4.21) one obtains a full expression for the PPE signal.

Figure 4.19 presents the frequency dependence of the numerical simulations of the PPE

amplitudes with thermal properties of the detector as parameters. The normalized amplitudes and

phases of the PPE signal were simulated numerically using MATLAB program and Eq. (4.25). The

glassy carbon was used as a reference sample.

=

thj D

a2

j

jjr

σβ

=

47

0 1 2

-0.6

-0.4

-0.2 α

p=10 -7[m/s],k

p=0.9[W /mK]

αp=10 -7[m/s],k

p=0.9[W /mK]

αp=5x10 -7[m/s],k

p=1.13[W /mK]

αp=5x10 -7[m/s],k

p=1.13[W /mK]

αp=8x10 -7[m/s],k

p=1.6[W /mK]

αp=8x10 -7[m/s],k

p=1.6[W /mK]

ln(N

orm

aliz

ed P

PE

Am

plitu

de[a

.u.])

Frequency1/2[Hz1/2]

Fthreshold

Figure 4.19: Numerical simulations based on equ. (4.25) (lines) and equ (4.20)(scatters) of

normalized PPE amplitudes as function the square root of the modulation frequency for

different thermal properties of PZT detector. The others parameters are: Lp=0.98 mm,

Ls=1.325 mm, Lr=0.98 mm, Ds=4.8 10-6 [m/s], Dr=4.2 10-6 [m/s], Dg=2.2 10-5 [m/s],ks=9

[W/mK], kr=6.3 [W/mK], kr=0.026 [W/mK],αs=106[1/m], αr=106 [1/m].

It is clearly seen that the threshold frequency Fthreshold (indicated in Figure 4.19) is independent of

the thermal properties of the PZT detector. Changes of the magnitude of the normalized PPE

amplitude are unfavorable when the thermal conductivity has to be determined. From Figure 4.19

it is evident that the normalization procedure doesn’t solve this problem in the used range of the

thermal properties. In fact, comparing Fig. (4.16) and Fig. (4.19) one can deduce that in limited

temperature range around room temperature the thermal properties of the detector don’t change

much and so their influence on the normalized PPE amplitudes can be neglected.

48

Marinelli [Marinelli, 1992] found that from the as-measured PPE phase it is possible to

estimate an absolute value of the thermal diffusivity. The numerical simulations of the as-

measured PPE phases are presented in figure 4.20.

0 1 2-25

-20

-15

-10

-5

0 1 2 3 4-55

-50

-45

-40

-35

-30

-25

-20

-15

-10

-5

0 PP

E P

hase

[deg

.]

Frequency1/2 [Hz1/2]

PP

E P

hase

[deg

.]

Frequency1/2 [Hz1/2]

Fthreshold

Figure 4.20 Numerical simulations based on equ. (4.25) (lines) and equ (4.20) (scatters) of

PPE phasess as a function of the square root of the modulation frequency for different

thermal properties of PZT detector. The others parameters are: Lp=0.98 mm, Ls=1.325

mm, Lr=0.98 mm, Ds=4.8 10-6 [m/s], Dr=4.2 10-6 [m/s], Dg=2.2 10-5 [m/s],ks=9 [W/mK],

kr=6.3 [W/mK], kr=0.026 [W/mK], αs=106[1/m], αr=106 [1/m].

It is clearly seen that the as measured PPE Phases are not affected by any changes in

thermal properties of the detector in the investigated frequency range when the sample and the

detector are both thermally thick and optically opaque. Below the threshold frequency the PPE

phases described by Eq. (4.25) deviates from their equivalent curves plotted from Eq. (4.20). For

49

this reason the m coefficient for glassy carbon were determined in the frequency range from 2 to

10 Hz.

Comparing Fig. (4.19) and Fig.(4.20) one can conclude that the threshold frequency is

more less the same for normalized PPE amplitudes and as measured PPE phases.

50

Chapter 5: Thermal properties of Cd1-xMgxSe

single crystals measured by means of

photopyroelectric technique.

5.1. Materials: Cd1-xMgxSe single crystals

Cd1-xMgxSe single crystals were grown by the high-pressure Bridgman method without

seed under an argon overpressure [F. Firszt et al., 1995]. The mixture of CdSe and metallic Mg

was put into a graphite crucible. The purity of CdSe and Se reaction components was 6N that of

Mg was 99.8 %. The temperature of the heating zone was kept at (1880 ± 0.5) K. The crucible

was held at that temperature for 2 h and then moved out from the heating zone with lowering

speed 4.2 mm⋅h-1. The obtained crystals were cylinders with raw dimensions 8 mm to 10 mm in

diameter and 40 mm to 50 mm in length. The crystals have a longitudinal gradient of magnesium

concentration (Mg content increases from the tip to the end of the crystal). For a crystal with x =

0.3, the Mg concentration gradient is about 0.01 cm-1. The crystals exhibit the wurtzite structure.

The lattice constant of Cd1-xMgxSe crystals decreases with increasing Mg content [F. Firszt et al. ,

1998]. The crystals were cut perpendicular to the growth direction into 0.9 mm to 1.5 mm thick

plates. Next the plates were mechanically polished and chemically etched in a mixture of

51

K2Cr2O7, H2SO3, and H2O in the proportion 3:2:1. Then they were treated in CS2 and hot 50 %

NaOH solution and finally rinsed in water and ethyl alcohol.

Table 5.1: Basic information of Cd1-x MgxSe single crystals.

Sample Technological

Number

Magnesium concentration,

x (mole fraction)

Thickness, L (mm)

1 A423_10 0.00 1.325

2 A557_III 0.00 0.923

3 A558_2 0.06 1.043

4 A440A_VI 0.14 1.306

5 A559_3 0.15 0.944

6 A433_XIV 0.33 1.174

7 A424A_X 0.36 1.294

52

5.2. Experimental results and computational algorithm

5.2.1 Thermal diffusivity – PPE Phases

Figure 5.1 shows experimental phase lags of investigated mixed semiconductors as a

function of the square root of the modulation frequency.

1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2

-3.5

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

0.0

Sample 1 Sample 2 Sample 3 Sample 4 Sample 5 Sample 6 Sample 7 Sample 1 Sample 2 Sample 3 Sample 4 Sample 5 Sample 6 Sample 7

PP

E p

hase

, rad

ians

Frequency1/2, Hz1/2

Fig. 5.1: Experimental PPE phase lags (scatters) of the investigated crystals with linear

fittings (lines) vs. square root of the modulation frequency. The fitting parameters are

collected in Table 5.2.

The thermal diffusivities reported in Table 5.2 were calculated using Eq. (4.23) by fitting the as–

measured PPE phases, from Fig. 5.1, with Eq. (4.22).

53

Table 5.2 Results of linear fits to Eq. (4.22) and calculated values of the thermal

diffusivities of Cd1-xMgxSe mixed crystals from Eq. (4.23).

Sample Magnesium

concentration, x (mole

fraction)

Thickness,

L (mm)

m Thermal diffusivity D

(10-6 m2⋅s-1)

1 0.00 1.325 -1.08 4.73

2 0.00 0.923 -0.77 4.48

3 0.06 1.043 -1.26 2.15

4 0.14 1.306 -1.79 1.67

5 0.15 0.944 -1,34 1.58

6 0.33 1.174 -1.85 1.26

7 0.36 1.294 -2.12 1.17

From Table 5.2 it is clearly seen that with increasing magnesium concentration the thermal

diffusivity decreases markedly. Figure 5.2 shows experimental data and theoretical curves of the

temperature dependence of thermal diffusivities for the investigated crystals. The fits assume a

linear temperature variation Dt(T)=aT.

54

20 22 24 26 28 30 32 34 36 38 40 42 441.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Sample 1,a=-0.04 Sample 2,a=-0.04 Sample 3,a=-0.01 Sample 4,a=-0.006 Sample 5,a=-0.006 Sample 6,a=-0.006 Sample 7,a=-0.006

T

herm

al d

iffus

ivity

, x 1

0-6 m

2 ·s-1

Temperature, °C

Fig. 5.2: Temperature dependence of the thermal diffusivity for the investigated crystals

fitted with a linear temperature dependence Dt(T)=aT with a given in the figure.

It is clearly seen that with increasing temperature the thermal diffusivity for all investigated

crystals decrease. The steepness of this slope decreases with increasing magnesium

concentration.

55

5.2.2 Thermal conductivity – normalized PPE amplitude

The obtained values of thermal diffusivities for investigated crystals from raw (as-

measured) PPE phases and amplitudes were substituted in the normalized PPE amplitude An(f)

defined by

onglassycarb

samplen Amp

AmpfA =)( ,

(ratio of the amplitude of the sample and that of the reference sample ) in order to determine

thermal conductivities. The nonlinear data-fitting procedure relied on minimizing the following

expression in a least-squares sense:

( ) ( )( )∑=

−N

iinis fAfkF

1

2,2

1 (5.1)

Only one parameter (thermal conductivity ks) was applied. F(ks,fi) is the theoretical PPE

amplitude (and normalized by theoretical response of the PPE amplitude of the glassy carbon)

described by Eq. (4.21) and N is the number of experimental points. The nonlinear data-fitting

was based on the built-in MATLAB function LSQCURVEFIT. The following parameters were

used during the fitting procedure for a pyroelectric detector kp=1.13 W⋅m-1⋅K-1 and Dp=4.95x10-7

m2⋅s-1 and for the glassy carbon ks=6.3 W⋅m-1⋅K-1 and Ds=4.22⋅10-6 m2⋅s-1. It was assumed that the

temperature dependence of the thermal conductivity of the glassy carbon results only from that of

the thermal diffusivity as determined from the PPE phases. Figure 5.3 shows the best fits to the

normalized PPE amplitude for sample 1.

56

4 6 8 100.30

0.35

0.40

0.45

0.50 N

orm

aliz

ed P

PE

Am

plitu

de, a

.u.

23 oC 23 oC, k = 9.94 W·m-1·K-1 25.1 oC 25.1 oC, k = 9.6 W·m-1·K-1

27.7 oC 27.7 oC, k = 9.28 W·m-1·K-1

31 oC 31 oC, k = 9.14 W·m-1·K-1

33.0 oC 33 oC, k = 8.94 W·m-1·K-1

36.3 oC 36.3 oC, k = 8.39 W·m-1·K-1

Frequency, Hz

Fig. 5.3: Best fittings (lines) to the normalized PPE amplitude (scatters) of sample 1 at

different temperatures.

5.2.3 Discussion

It is clearly seen that with increasing temperature the thermal conductivity for sample 1

decreases. For the temperature 27.7 °C (300.7 K) the thermal conductivity of sample 1 (CdSe) is

9.28 W⋅m-1⋅K-1. This value is in good agreement with the thermal conductivity of CdSe (9 W⋅m-

1⋅K-1) crystal obtained by Slack [Slack, 1972]. Additionally, for single crystals, the thermal

conductivity may depend on crystallographic directions as in the case of Zn0 [Slack, 1972].

Although our samples were not oriented, this effect can be neglected because measurements were

performed at high temperatures. The frequency-dependent normalized amplitude obtained at

different temperatures from the other samples have been fitted in the same way. The best results

for all crystals of the fitting procedure are collected in Table 5.3 and Fig. 5.4. Figure 5.4 presents

the temperature dependence of the thermal conductivity for Cd1-xMgxSe mixed crystals with

different magnesium concentrations.

57

25 30 35 40 45 50

3

6

9 Sample 1 n=-0.32 Sample 3 n=-0.28 Sample 4 n=-0.23 Sample 5 n=-0.29 Sample 6 n=-0.29 Sample 7 n=-0.3

The

rmal

Con

duct

ivity

, W·m

-1·K

-1

Temperature, C

Fig. 5.4: Temperature dependence of the thermal conductivity (scatters ) for Cd1-xMgxSe

mixed crystals with different Mg concentrations and the best fits (lines) to k(T)=aTn on

log-log scale.

The dynamics of these changes can be expressed by coefficient n extracted from k(T)=aTn on log-

log scales. Values of coefficient n are presented in Fig. 5.4. Obviously, n is nearly constant for all

investigated crystals within the error limit of the fitting (±0.05). The electron contribution to the

thermal conductivity can be calculated for a degenerate semiconductor (because carrier

concentrations of investigated crystals are high) from a formula given by [Bhandari and Rowe,

1998],

,

2

e LTe

kk σ

= (5.2)

where T is temperature, e is the elementary charge, and k is the Boltzmann constant. The

electrical conductivity σ for samples 1, 4 and 6 were taken from Hall measurements [Perzynska

58

et al., 2000] and the Lorenz factor was assumed to be 32π . The obtained results are collected in

Table 5.3.

Table 5.3 Thermal conductivity k, electrical conductivity σ and electron contribution to

thermal conductivity ke of Cd1-xMgxSe mixed crystals at room temperature as results of

the best fits using MATLAB

Sample Magnesium

concentration, x

(mole fraction)

Thickness L

(mm)

ks (W⋅m-1⋅K-1) Electrical

Conductivity σ

(Ω-1 ⋅m-1)

ke (W⋅m-1⋅K-1)

1 0.00 1.325 9.28 1000 0.007

2 0.00 0.923 - - -

3 0.06 1.043 5.34 - -

4 0.14 1.306 3.82 1200 0.009

5 0.15 0.944 3.93 - -

6 0.33 1.174 3.13 500 0.004

7 0.36 1.294 3.00 - -

Since the electron contribution to the thermal conductivity is very small indicating that the heat is

mainly carried by phonons. This behavior is expected for CdSe crystals at room temperature

[Slack, 1972]. Therefore, the thermal conductivity can be related only to the lattice contribution

and the thermal resistivity can be described as the inverse of the (lattice) thermal conductivity.

Figure 5.5 presents the thermal resistivity W as a function of magnesium concentration at room

temperature. The thermal resistivity W increases with increasing magnesium concentration. This

behavior is expected when large numbers of magnesium atoms are added to the host lattice and

they act as scattering centers for phonons. A similar behavior of the thermal resistivity was found

by Adachi [Adachi, 1983] in III-V semiconductors. He gave the expression of the thermal

resistivity as a function of concentration x [Adachi, 1983]:

( ) ( )xxCWxxWxWCdMgSe −+−+= 11)( Mg-CdMgSeCdSe , (5.4)

where WCdSe and WMgSe are the thermal resistivities of CdSe and MgSe, respectively. The

coefficient CCd-Mg is called a nonlinear parameter and it is a contribution arising from the lattice

59

disorder generated in ternary Cd1-xMgxSe system by random distribution of Cd and Mg atoms in

one of the two sublattice sites [S. Adachi, 1983]. In Fig. 5.5 the best fitting (line) to Eq. (5.4) is

also shown.

0.0 0.2 0.4 0.6 0.8 1.0

0.10

0.15

0.20

0.25

0.30

0.35

0.40

0.45

0.50

0.55

Mg concentration, x

The

rmal

res

istiv

ity, m

·K·W

-1

Fig. 5.5: Reciprocal thermal conductivity (thermal resistivity) of the Cd1-x MgxSe mixed

crystals as a function of magnesium concentration at room temperature.

The nonlinear parameter was found to be 1.59 W⋅m-1⋅K-1 while the thermal conductivity of the

MgSe is 7.69 W⋅m-1⋅K-1. Although MgSe does not exist in nature, the value obtained from fitting

is rather unexpected. This can be explained because of the Adachi model is a simplification of the

Abeles model [Abeles, 1963]. Adachi had taken into account only strain scattering while for Ge-

Si alloys mass-defect scattering could be important. Similarly for the mass difference can be

important. Tsen et al. [Tsen et al., 1997] reported that the total electron-longitudinal optical

phonon scattering rate in GaN is about one order of magnitude larger than that in GaAs. They

attributed this enormous increase in the electron-longitudinal optical phonon scattering rate to the

60

much larger ionicity in GaN. In Cd1-xMgxSe mixed crystals an increasing Mg content in the solid

solution will increase the iconicity. Supplementary Hall measurements on the mixed crystals

indicate a high carrier concentration; hence, this scattering mechanism cannot be neglected [K.

Perzynska et al., 2000].

.

61

Chapter 6: Influence of the space charge layer

(SCL) on the charge carrier transport properties

measured by means of the photocarrier

radiometry (PCR)

In Chapter 4 an introduction to the PCR technique was presented. It was shown that the

PCR signal is sensitive to the carrier transport properties and its linear function of the optical

excitation intensity. In this chapter the influence of the space charge layer on the PCR signal is

presented. Part of the following results were published by A.Mandelis, J.Batista, M.Pawlak,

J.Gibkes and J.Pelzl in J. Phys. IV France 125 (2005) 565-567 and Journal of Applied Physics 97

083507 (2005).

62

6.1 Theory of optically modulated p-type SiO2-Si interface energetics in the presence of

charged interface states [Mandelis, 2005a]

The Si02-Si interface energy diagram of p-type Si in the presence of the positively charged

interface state density Nt (m-2) acting as traps of free minority carriers (electrons) is presented in

Figure 6.1.

Figure 6.1 Band-structure energetics at a p-type Si-SiO2 interface with a positively

charged interface state (trap) density Nt assumed to be at energy Et. The band bending is

modulated by the external optical field. [Mandelis, 2005a]. Other parameters are defined

in text.

At equilibrium in the dark [Sze, 1981] the energy bands at the interface are bent with a total

interface potential energy 0sqψ , measured with respect to the intrinsic Fermi level, causing the

maximal value of the space charge layer W0. Additional W0 serves as the reference value for the

non equilibrium configuration under optical incidence of intensity ( )να hI ,0 (W m-2). The

effective SCL width can be defined by mWWW −=∆ 0 and mW were changed with the incident

modulated laser intensity0I . 0W and mW were the dc and modulated components of the SCL:

timeWWIW ω−= 00 )( . (6.1)

63

The modulated excitation generates a free-carrier density wave mostly within depth αµ 1= [m-1]

from the surface [Mandelis, 2001] that usually includes the very thin ( mµ5.0~ in Si) space

charge layer. This is a coherent excitation of minority carriers which is characterized by a

frequency-dependent (ac) diffusion length

( ) ( )n

nnn i

DL

τωτω⋅⋅+

⋅=

1

*

, (6.2)

where *nD is the ambipolar diffusion coefficient and nτ is the bulk lifetime of the CDW.

Mandelis [Mandelis, 2005a] proved that these parameters are composite (effective) quantities

involving interface as well as bulk values. Additionally, he pointed out that the interface lifetime

value is affected by the details of the trapping dynamics. The minority carriers (electrons) within

an ac diffusion length from the edge of the SCL can be swept into it and slide down the energy-

band slope of the SCL edge under depletion or inversion conditions for the majority carriers

(hole). This increases the charge density within the SCL, which in turns, affects the occupation of

the interface affects the occupation of the interface state Et. Under interface illumination with

intensity tieI ωλ)(0 there is a non-zero probability that a fraction of the occupied interface states,

Nt0, will absorb photons from the incident radiation and will eject trapped electrons into the

conduction bandedge, which will increase the degree of band-bending [Mandelis 2005a]. The

SCL acts as a thin spatial region (~ 0.5 µm) in which recombination is essentially absent due to

efficient separation of the local electron-hole pair and the completely ionized impurity states at,

or above, room temperature. As a consequence, the recombination lifetime in this region has been

set equal to infinity, Fig. 6.2.

64

Figure 6.2 Optical source depth profile and carrier-density-wave transport parameters in

p-type semiconductor Si with a transparent surface oxide and interface state density Nt

[Mandelis, 2005a]

6.2 The expression of the PCR signal including effects due to an existing SCL

The expression of an influence the SCL width under optical modulation on the PCR signal

was given by Mandelis [Mandelis, 2005a]

[ ]

( )LLmemL

een

Wm

riQIRSCLPCR

ee

eeeWW

eD

eWT

h

IRRRCIS

σσσ

αα

σαασ

γσα

αα

ων

αληααω

−−−

∆−

+Γ−+−+

Γ−Γ−+

+−

−≅

2212

122*

012100,

])(1[)1()()(

1

)1(1)(

2

)](1[);,();,(

(6.3)

( )( )[ ] ] ( )[ ] ( )[ ] Wm

Lm

Lm eWWWeWeWWW e ∆−−− ∆−−∆−−+−+Γ∆−−∆+ αασ αα

αα 11

2

11 2

2

65

where Tri(ω) is a complex interface lifetime defined as (Appendix A in

[Mandelis,2005a])ri

riri i

Tωτ

τω+

≡1

)( ,where τri is a charged interface recombination lifetime. In

case of no charged surface states, then 0=∆W and 0)( 00 == IWW m for 00 >I , equation (6.3)

reduces to (4.5). It is worthwhile to emphasize that the conventional expression of the PCR signal

(4.5) exhibits a linear dependence on 0I while the structure of equation (6.3) depicts – in case of

the presence of a nonzero density of charged interface states – that there will be a nonlinear low-

intensity range bound up with SCL oscillation from maximum value of 0W (in the dark) and 0

(full illumination).

6.3 Numerical simulations of an influence of the existence of the SCL on the electronic

transport properties

6.3.1 Numerical simulation of the PCR signal dependence on the electrical transport

properties in the presence of SCL

For the simulation purpose it was assumed that exponential dependence of the SCL width

W0 (=1µm) on IAC,

)exp(0 acm BIAWWW −=−≡∆ , (6.5)

where A,B are constants (set 0.1x10-5 and 2x10-4, respectively). The results of numerical

simulations for the fixed modulation frequency (200 Hz) of the effects of changing bulk minority

recombination lifetime are shown in Fig. 6.3, and those of changing interface recombination

velocity are shown in Fig. 6.4.

66

0,1 1

1010

1011

1012

1013

0,1 1

1010

1011

1012

1013

τ=1µs linear range

PC

R a

mlit

ude[

a.u.

]

Intensity [W/cm2]

τ=10µs linear range

1 2 3-0,8

-0,7

-0,6

-0,5

-0,4

-0,3

-0,2

-0,1

0,0

Intensity [W/cm2]

τ=1µs τ=10µs

PC

R p

hase

[deg

.]

Figure 6.3: Simulated PCR amplitudes (a) and phases (b) as a function of optical

intensity for p-type Si with recombination lifetime as a parameter. Other parameters: Dn*

= 30 cm2/s, , S1 = 300 cm/s, S2 = 105 cm/s, L = 550 µm, α(λ=514nm) = 7.76×103 cm-1.

67

0,1 1

1011

1012

1013

0,1 1

1011

1012

1013

S1=10 cm/s

linear range

PC

R a

mlit

ude[

a.u.

]

Intensity [W/cm2]

S1=100 cm/s

linear range

0 1 2 3-2,5

-2,4

-2,3

-2,2

-2,1

-2,0

-1,9

-1,8

-1,7

-1,6

-1,5

-1,4

-1,3

Intensity [W/cm2]

S1=10cm/s

S1=100cm/s

PC

R p

hase

[deg

.]

Figure 6.4 Simulated PCR amplitudes (a) and phases (b) as a function of optical intensity

for p-type Si with surface recombination as a parameter. Other parameters: Dn* = 30

cm2/s,τ = 100 µs, S2 = 105 cm/s, L = 550 L = 550 µm,α(λ=514nm)=7.76×103 cm-1.

68

In both cases a similar shape of the PCR amplitudes is observed. Additionally, if the SCL exists

the PCR signal deviates from the linear dependence on the optical excitation intensity. The

straight line in Fig.6.3a and 6.4a follows the linear dependence on the optical excitation intensity

above around intensity 1.5 W/cm2 (the intensity for which the PCR is a linear function of the

intensity is named Ilinear) where the SCL is completely vanished (calculated value of the effective

for 2 W/cm2 is ∆W=2x10-8 m). In both cases the shapes of the amplitude curves, A(Ilinear), are

similar. The phases, φ(I linear), remain flat, because the interface recombination lifetime was

assumed constant. From the PCR Phases one can see that this effect can be more important for

wafers with a long lifetime.

6.3.2. Numerical simulation of the PCR Signal dependence on the existence of SCL width

Assuming negligible Wm for low intensity of the modulated laser compared to W0 an

influence of the dc 0W and modulated components mW of the SCL width on the PCR signal the

numerical simulations using MATLAB are shown in Figure 6.5 and Figure 6.6, respectively.

69

10 100 1000 10000 1000001018

1019

1020

10 100 1000 10000 100000-100

-80

-60

-40

-20

0

PC

R a

mpl

itude

[a.u

.]

Frequency [Hz]

W0=0µm

W0=0.2µm

W0=0.4µm

W0=0.6µm

W0=1µm

(a)

W0=0µm

W0=0.2µm

W0=0.4µm

W0=0.6µm

W0=1µm

PC

R p

hase

[deg

.]

Frequency[Hz]

(b)

Fig.6.5: PCR amplitudes (a) and phases (b) vs. modulation frequency simulations for p-

type Si with constant transport parameters and with the SCL width as a parameter.

Other parameters: Wm = 0, Dn* = 30 cm2/s,τn = 100 µs, S1 = 300 cm/s, S2 = 105 cm/s, L = 550

µm, α(λ=514nm) = 7.76×103 cm-1.

Figure 6.5 shows that increasing W0 results in a monotonically decreasing amplitude and

essentially no change in phase occurs. This behavior is expected, since an increased degree of

band-bending is the result of increased interface charge density in the model and thus higher trap

density and loss of free carriers. The process takes place right at the Si-SiO2 interface where free

minority carriers de-excite mostly non-radiatively in trap states over the space-charge barrier

[Johnson, 1958] and therefore are not available to contribute to the PCR signal through radiative

NIR emissions. Furthermore, there is no measurable phase shift because the interface

70

recombination lifetime τs was assumed constant. This simplification turns out not to be true

experimentally, but the simulation points out the important fact that it is not the value of W0 itself

that causes a phase shift, but rather the change in this value has an effect on the transport

properties at the Si-SiO2 interface.

101 102 103 104 105

1019

1020

101 102 103 104 105-100

-80

-60

-40

-20

0

Wm=0µm

Wm=0.2µm

Wm=0.4µm

Wm=0.6µm

Wm=1µm

PC

R a

mpl

itude

[a.u

.]

Frequency [Hz]

(a)

Wm=0µm

Wm=0.2µm

Wm=0.4µm

Wm=0.6µm

Wm=1µm

PC

R p

hase

[deg

.]

Frequnecy[Hz]

(b)

Figure 6.6 PCR amplitudes (a) and phases (b) vs. modulation frequency simulations for

p-type Si with constant transport parameters and with the optically modulated SCL

width, Wm, as a parameter. W0 = 1 m, Dn* = 30 cm2/s, τ= 100 µs, S1 = 300 cm/s, S2 = 105

cm/s, L = 550 µm, α(λ=514nm) = 7.76×103 cm-1.

Figure 6.6 corresponds to the case where the maximum value, W0, of the SCL width is

fixed, but is subject to oscillating amplitude, Wm, which increases with, e.g., increasing intensity

of the modulated laser source. As expected, the PCR amplitude increases as the modulated band

curvature tends to offset the effect of the CDW-inhibiting band bending through more effective

71

neutralization of interface charges over the modulation cycle. As a result a higher density of free

minority carriers can survive over one illumination period and contribute radiatively to the

increased PCR signal. As in the case of Fig. 6.6b, the phase does not show any change over the

entire range of Wm values used in this simulation. When the changes in the SCL widths, Figs. 6.5

and 6.6, and in transport parameters, Figs. 4.4-4.7, are combined in a frequency plot for fixed

values of the transport parameters, it is found that the former simply shift the PCR amplitude

accordingly, while the phase remains fixed. This conclusion proves to be very helpful in

interpreting experimental PCR measurements of the SCL. It is worthwhile to emphasis that

changing of the physical parameters responsible for the transport properties of the carriers

(electronic parameters) and the space charge layer can shift the PCR amplitude. Unlike to the

PCR Phase which is only sensitive for altering electronic parameters.

6.4 Experimental Conditions and Materials

6.4.1 Experimental Methodology

In chapter 4 it was shown that the PCR signal is not only sensitive to the electrical and

optical properties of a material but also to experimental conditions like the beam size. The effect

of the beam size on the PCR signal was discussed in Section 4.1.6. In some cases the optical

excitation intensity, another experimental condition, can affect the linear dependence of the PCR

signal on the optical excitation intensity. In case of the presence of a nonzero density of charged

interface states – that there will be a nonlinear low-intensity range of the PCR signal dependence

on the optical excitation intensity bound up with SCL oscillation from maximum value of 0W (in

the dark) and 0 (full illumination).

For monitoring effects of the SCL on the PCR signal two-laser system was proposed. A

fixed (low) intensity modulated (“ac”) laser produces a fixed density of free electron pairs (EHP)

carrier waves acting as PCR probe. During that time a coincident unmodulated (“dc”) laser

source with variable intensity,dcI , substantially exceeding that of the modulated laser but for the

lowest values, can change the occupation of the surface states up to the flatband condition when

FBdc II = and acts as optical bias. In this configuration the dc laser induces a change in degree of

72

steady band bending at the surface from the dark (maximum) value, ( ) max0 0 WIW dc == ,of the

SCL width up to flatband (minimum) value,( ) 00 =FBIW . In this ideal case, the modulated laser

is non-perturbing, acting only as a PCR signal carrier with negligible effect on the SCL width,

since Wm(IAC)<< W0(IDC). A variation in dc laser intensity within the range 0 DC FBI I≤ ≤ is

expected to act as a variable optical bias by means of the steady-state excess recombination

events of minority carriers into impurity states. The enhanced steady bulk recombination further

affects (increases) the steady SCL minority charge density Qsi, driving the SiO2-Si interface into

deeper depletion or inversion. This change alters the (thermo)dynamically coupled interface

charge state coverage [Mandelis,2005a] causing a concomitant change in the degree of band-

bending between maximum and zero (flatbands). Because these interface changes perturb the

small CDW generated by the modulated laser, the entire process can be described according to

Eq. (6.3), with an effective steady value of W0(IDC) for each IDC value and a fixed Wm(IAC).

6.4.2. Experimental set up

Figure 6.7 presents the designed and constructed two-laser PCR system.

Figure 6.7 Two laser PCR experimental set up.

73

As an excitation source of carrier density waves a modulated (by means of a chopper) low-

power laser a He-Ne laser (632.8 nm; 0.4 mm beam radius) or a 830 nm GaAlAs laser diode

(0.3mm beam radius) was used. The non-modulated optical bias laser was an Ar-ion laser (514

nm; 1.89 mm beam radius). The power of the modulated laser was typically PAC ~ 0.5 - 4 mW,

whereas the power of the non-modulated laser was varied up to 350 mW. The laser beams were

focused onto the sample surface using lens. The position of the laser beams is coincident with the

focal point of an off-axis paraboloidal mirror that collects a portion of infrared radiation from the

samples. The collected light is then focused onto the detector by means of lens. Samples were

placed onto aluminum holder (acted as a mechanical support and signal amplifier by redirecting

the forward emitted IR photons back toward the detector [Mandelis, 2003]. All instruments, data

acquisition, and data storage are controlled by a computer running Pascal program with a

graphical user interface and real-time display of experimental data. Taking into account the

reflectivity of Si at 514 nm at normal incidence (R = 0.38) and the laser beam radius r = 1890

µm, the effective maximum photon flux at 350 mW was

Fp,max = 1.103×1018 photons/cm2s. (6.4)

6.4.3. Materials

Samples were four and six-inch 5-10 Ω−cm, 550-µm thick, p-type Si wafers which were

oxidized with a gate oxide of ca. 1000 Å. In addition, n-type wafers were also oxidized with ca.

5000 Å oxide.

6.5 Experimental results

6.5.1 Effect of chemical etching on the PCR signal

Some wafers were exposed to variable optical bias with the laser beam incident on the SiO2

and then were etched to hydrophobia in an aqueous solution of 10% vol. HF in water, indicating

that the SiO2 layer was fully removed. The PCR signal amplitudes and phases of a p-Si sample

before and after etching are shown in Fig. 6.8. While the oxidized wafer exhibits complete photo-

saturation at irradiation with approx. 250 mW of continuous laser power, it is clear that the

removal of the oxide also removed the interface-charge layer very efficiently while increasing the

74

non-radiative electron trapping efficiency at the etched surface. Accordingly, the PCR amplitude

dropped significantly and remained independent of the power of the unmodulated laser. The Ar-

ion laser-beam reflectance of the wafer before and after etching was measured to be 0.340 and

0.408, respectively. The reflectance of the primary He-Ne beam was 0.295 and 0.339,

respectively. These differences cannot account for the drastic change in the PCR amplitude after

etching, giving further support to enhanced non-radiative recombination. The PCR phase, Fig.

6.8(b), shows a very reproducible curvature for the oxidized sample. On the other hand, the phase

of the etched sample is independent of PDC and became noisier due to the low signal associated

with this sample. Signal-to-noise ratios (SNR) for oxidized wafers were in the 100 – 200 range

with error bar sizes similar to the symbol size used in the plots, whereas those for the etched

samples were approx. 18 - 25.

0 100 200 3000

20

40

60

80

100

120

140

160

0 100 200 30075

80

85

90

95

100(a)

p-Si Unetched p-Si Etched

PC

R A

mpl

itude

(µV

)

PDC

(mW)

(b)

PC

R p

hase

[deg

.]

PDC

(mW)

75

Figure 6.8 (a) Amplitudes and (b) phases of an oxidized p-Si wafer before and after

etching the SiO2 away. The modulated beam was provided by a mechanically chopped

He-Ne laser. Chopping frequency: 200 Hz.

In the case of oxidized wafer the PCR amplitude exhibits complete photosaturation at

irradiation approximately 250 mW of laser power. Unlike to oxidized wafer, the PCR amplitude

of etched wafer dropped drastically and remained independent of the power of the unmodulated

laser. This significant change in PCR amplitudes was due to etching which enhanced

nonradiative recombination and so attenuated radiative recombination. In figure it is seen that the

PCR phase was reproducible only in case of oxidized wafer signal. The PCR phase of the wafer

after etching depends on the non-modulated laser power and becomes noisier due to the low

signal associated with this sample.

6.5.2 The perturbation effects of the primary modulated laser beam on the PCR signal

The issue of possible perturbation effects of the primary modulated laser beam on the SCL

measurements was investigated by changing the power of the He-Ne laser and repeating the

experiment of Fig. 6.8 using another oxidized p-Si wafer. Typical results are shown in Fig. 6.9.

As expected, the PCR signal amplitude does scale with PAC (or, equivalently, I0), Eq. (6.3),

however, the re-scaled amplitude of the measurement with decreased power at 1.4 mW coincides

with the 3.5-mW amplitude, when normalized to the highest point of the latter curve. Similarly,

the phases coincide with no rescaling within error, with the phase obtained at the lower power

exhibiting higher noise. These results demonstrate that, in the range of the reported

measurements, the modulated laser was not perturbing the electronic properties of the

semiconductors and the observed signal variations with PDC were solely due to the effects of the

dc laser on the sample.

76

0 50 100 150 200 250 3000

20

40

60

80

100

120

140

160 (a)

PAC

= 3.5 mW

PAC

= 1.4 mW

PAC

= 1.4 mW; normalized

PC

R A

mpl

itude

V)

PDC

(mW)

0 50 100 150 200 250 30084

85

86

87

88

89

90(b)

PAC

= 3.5 mW

PAC

= 1.4 mW

PC

R P

hase

(de

g)

PDC

(mW)

Figure 6.9 a) Amplitudes and (b) phases of an oxidized p-Si wafer using two different

power levels of the primary (modulated) laser beam. He-Ne laser chopping frequency:

200 Hz.

6.5.3 The effect of polishing on the PCR signal

Figure 6.10 shows that the effects of surface polishing on the PCR signal from the p-type

wafer of Fig. 6.9 (amplitudes and phases) are minor. The back matte surface was also oxidized

and therefore it was expected that it would exhibit an SCL behavior similar to the front surface.

Based on the fits to the theory as described below, these effects can be accounted for by small

differences in the respective surface recombination velocity, effective lifetime and SCL depth

profile.

77

0 50 100 150 200 250 300-10.0

-9.5

-9.0

-8.5

-8.0

-7.5

-7.0

-6.5

-6.0

-5.5

-5.0

20

40

60

80

100

120

140

160

180

P

CR

Pha

se (

deg)

Polished surface Unpolished surface

PC

R A

mpl

itude

(µV

)

PDC

(mW)

Figure 6.10 Effects of surface polishing on the PCR signal amplitude and phase. He-Ne

laser chopping frequency: 200 Hz.

6.6 Determination of carrier transport properties in SCL and the depth profile

reconstruction

The developed PCR theory [Mandelis, 2005a] resulting in Eq. (6.4) was tested through a

series of PCR measurements with Si-SiO2 interfaces, which were aimed at reconstructing the

depth profile of the SCL from scans of IDC at a fixed frequency. Results on a p-Si sample such as

those shown in Figs. 6.10-6.12 were supplemented by frequency scans launched at several values

of IDC, as shown in Fig. 6.11.

78

101 102 103 104 105

1

10

(a)

Pdc = 0 mW Pdc = 10 mW Pdc = 20 mW Pdc = 30 mW Pdc = 50 mW Pdc = 90 mW Pdc = 200 mW Pdc = 350 mW

PC

R A

mpl

itude

(m

V)

Frequency (kHz)

101 102 103 104 105-100

-80

-60

-40

-20

0 (b)

PDC

= 0 mW

PDC

= 10 mW

PDC

= 20 mW

PDC

= 30 mW

PDC

= 50 mW

PDC

= 90 mW

PDC

= 200 mW

PDC

= 350 mW

Pha

se (

degr

ees)

Frequency (kHz)

Figure 6.11 Amplitude (a) and phase (b) frequency scans of a p-Si – SiO2 interface from

the polished surface of Fig. 6.8, under various dc laser power levels and 830-nm

modulated excitation source. Theoretical fits to Eq. 6.6 are indicated by the continuous

lines. Unique best fits were determined by the set of parameters W0 , τeff , S1, S2, and

effD yielding the minimum variance in Eq. 6.6. For all fits it was found that 0.86 % < Var

< 1.14 %.

The individual frequency scans were fitted to Eq. (6.3) by means of FORTRAN program with ∆W

= W0 since the IAC intensity was too low compared to IDC to affect the band-bending as

demonstrated in Fig. 6.9. The fitted parameters at each value of IDC included τeff, S1, S2,

*neff DD ≈ and W0. During the fitting procedure, the four transport properties τeff , S1, S2, and

effD were set as free parameters. The best-fitting procedure commenced at IDC = IFB , where W0 =

0. Then, the first non-zero value of W0 was incremented for the next lower IDC and all five

parameters were allowed to vary until the absolute minimum of Eq. (6.6) was attained. The

values of the five parameters yielding the absolute minimum in Eq. (6.6) were unique in each

case within the physically expected value ranges and fitting errors were recorded. The procedure

was repeated for all IDC ≥ 0. Figure 6.12 shows the IDC dependence of the front surface

(interface) recombination velocity, S1, in the range used in our experiments.

79

0 50 100 150 200 250 300 350

0

100

200

300

400

500

600

700

S1 (

cm/s

)

PDC

(mW)

Figure 6.12 Recombination velocity for a p-Si – SiO2 interface as a function of the Ar-ion

laser dc power of a p-Si – SiO2 interface. Data obtained from simultaneous best fits to

amplitude and phase frequency scans shown in Fig. 6.11.

The decrease of this parameter essentially down to zero is consistent with the physical

process of optical neutralization of the interface states by photo-excited minority electrons. The

results are also consistent with earlier derived dependencies of the front SRV on the excess

electron density in optically biased photoconductance-decay experiments [Aberle, 1996; J.

Schmidt and A. G. Aberle, 1997], from surface photovoltage measurements using negative

corona charging [Schmidt and Aberle,1997], and from basic Shockley-Reed recombination

theory [A. G. Aberle et al. 1992, M. Schoefthaler et al.,1994]. Using the dependence

11. −×= Sconstsτ [Mandelis, 2005a; Appendix Eq. (A5)] results in an increase of the interface

recombination lifetime with incident dc laser power. Figures 6.13 and 6.14 show the IDC

dependencies of the CDW effective diffusion coefficient, Deff, and effective recombination

lifetime, τeff .

80

0 50 100 150 200 250 300 35010

12

14

16

18

20

22

24

26

28

30

Def

f (cm

2 /s)

PDC

(mW)

Figure 6.13 Effective diffusion coefficient of the carrier-density-wave as a function of the

Ar-ion laser dc power of a p-Si – SiO2 interface. Data obtained from simultaneous best fits to

amplitude and phase frequency scans shown in Fig. 6.11.

0 50 100 150 200 250 300 350400

425

450

475

500

τ eff (µ

s)

PDC

(mW)

Figure 6.14 Effective lifetime of the carrier-density-wave as a function of the Ar-ion laser

dc power of a p-Si – SiO2 interface. Data obtained from simultaneous best fits to

amplitude and phase frequency scans shown in Fig. 6.11.

81

In the high dc-power range, PDC > 100 mW, Fig. 6.12 shows that 01 ≈S . Therefore, in that

range, according to [Mandelis 2005a: Appendix Eq. (A7)]

sBeff τττ111 += ,

Beff ττ ≈ 460≈ µs, a constant value reflecting purely bulk recombination and consistent with

Fig. 6.14. Here τB is the bulk recombination lifetime. Nevertheless, for the same PDC range the

effective diffusion coefficient of the CDW decreases from an essentially electron minority

carrier diffusivity value of 26 cm2/s following a relatively steep increase. The onset of

decrease at high optical bias corresponds to photo-excited carrier densities of 1017 cm-3

(calculated from Eq. (6.4) for excitation at 830 nm) and is consistent with the onset of non-

negligible carrier-carrier scattering reported for Si(111) surfaces [Li et al., 1997]. The increase

of Deff at low PDC bias power is the result of the ambipolar nature of *neff DD ≈ [Mandelis

2005a] as the free-electron CDW increases in the near-interface region with increased optical

interface charge neutralization, thus changing the value of the ambipolar diffusivity from the

majority Dp ( ~ 12 cm2/s) to the minority Dn (~ 30 cm2/s) range. The minimum in the value of

τeff around PDC = 50 mW , Fig. 6.14, is most likely associated with the increasing values of

Deff and τs : The IR emitting photo-carrier density-wave shifts to deeper sub-interface

locations, in agreement with the increased PCR phase lag in Figs. 6.8b, 6.9b and 6.10, and this

results in the disappearance of a number of contributing carriers from the field of view of the

IR detector [Ikari et al.,1999].The computational application of Eq. (6.3), interprets this

relative scarcity of carriers as a decreased recombination lifetime. At higher PDC the flattening

bands bring about an increased free-carrier density-wave in the immediate sub-interface

region (decreased phase lag in Figs. 6.8b, 6.9b and 6.10) which restores the CDW infrared

photon emission within the range of the InGaAs visibility solid-angle. Figure 6.15 is the

reconstructed depth profile of the SCL width from full band-bending to the complete flatband

condition associated with photo-saturation of the PCR signal.

82

0 50 100 150 200 250 300 350

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

SC

L w

idth

(µm

)

PDC

(mW)

Figure 6.15 SCL width of a p-Si – SiO2 interface as a function of the Ar-ion laser dc

optical bias. The reconstruction was obtained from simultaneous best fits to

amplitude and phase frequency scans shown in Fig. 6.11.

6.7 Summary

The interface modulated-charge-density theory developed by Mandelis [Mandelis,

2005a] was used for physical simulations involving PCR signals in order to study the effects

of the various SCL optoelectronic transport parameters on the PCR amplitude and phase

signal channels. Furthermore, an experimental configuration was used involving n- and p-

doped Si – SiO2 interfaces and a low-intensity (non-perturbing) modulated laser source as

well as a co-incident dc laser of variable intensity acting as interface-state neutralizing optical

bias. The application of the theory to the experiments yielded the various transport parameters

of the samples as well as the depth profile of the SCL. It was shown that PCR can monitor the

complete flattening of the energy bands at the interface of p-Si – SiO2 with dc optical powers

up to a 300 mW. The uncompensated charge density at the interface was also calculated from

the theory. The SCL profile for the n-type Si wafers was also investigated. Typical results are

shown in Fig. 6.16.

83

0 50 100 150 200 250 30010

15

20

25

30

35

40

75

80

85

90

95

100

105

110

PC

R P

hase

(de

g)

unetched etched

PC

R A

mpl

itude

(µV

)

PDC

(mW)

unetched etched

Figure 6.16 Amplitudes and phases of an oxidized n-Si wafer before and after etching

the SiO2 away. The modulated beam was provided by a mechanically chopped He-

Ne laser chopping frequency: 200 Hz.

Both lasers were incident on the exposed SiO2 layer. In all cases, large differences were

observed in the PCR amplitudes between intact and etched samples. Nevertheless, there was

no indication of flatband saturation within the range of IDC intensities available and phases

were always essentially independent of the dc laser power. A comparison among the

amplitude shapes and absolute signal levels in Figs. 6.8 and 6.16 shows that it is much harder

for the optical bias to induce complete flattening of the bands in n-Si than in p-Si. This

observation indicates the relatively low efficiency of this methodology for driving n-Si into

the flatband state. The independence of PCR phases from IDC is an additional indicator that

the optical bias does not affect the transport properties of the SiO2 - n-Si interface. This

behavior can be explained by the much lower (ca. 100 times [Aberle, 1992]) minority carrier

capture cross-section of the n-type interface. This fact leads to much shorter interface

lifetimes, τs [Mandelis, 2005a], making it harder for an optical source to build up the

neutralized interface charge coverage observed with p-Si – SiO2 interfaces, Fig. 6.15. The

same effect makes the interface recombination velocity essentially independent of excess

carrier density at the n-Si – SiO2 interface [Aberle, 1992] which results in the independence of

the PCR phase from PDC, Fig. 6.16.

84

Chapter 7: Non-linear dependence of

Photocarrier Radiometry signals from p-Si

wafers on optical excitation intensity and its

effect on charge carrier transport properties

In Chapter 4 an introduction to the PCR technique was presented. It was shown that the

PCR signal is sensitive to the carrier transport properties and its linear function of the optical

excitation intensity. In this chapter the nonlinear dependence of PCR signals from p-Si wafers

on optical excitation intensity and its effect on charge carrier transport properties is presented.

Part of the following results were published by J. Tolev, A. Mandelis and M. Pawlak in J.

Electrochem. Soc. J. Electrochem. (2007) and Eur. Phys. J. Special Topics (2008)

7.1 Introduction

In chapter 6 the PCR signal nonlinearities associated with the existence of a space

charge layer in the silicon wafer and the effects of optical biasing and conditioning of this

layer in the flatband configuration under low-intensity conditions were presented. If the space

charge layer disappears then the PCR signal was considered as linear function on optical

excitation intensity [Mandelis et al., 2003] in sense that the amplitude of the signal depends

linearly on laser power at low power ranges. This was attributed to the linear relation between

the PCR signal and the free carrier absorption coefficient. With this assumption the

expression of the PCR signal is given by (4.5). Furthermore, it was also assumed that the free

85

carrier absorption coefficient is a linear function of the carrier concentration ((4.3)). This

coefficient can be written in the frame of the Drude model [Smith, 1978]

2 2

2 3( )

8 * ( )fS

NqN

nc m N

λαπ τ

= . (7.1)

where N is carrier concentration τ is the recombination lifetime and other quantities are

defined in Abbreviations. From above formula it is seen that the free carrier IR absorption

coefficient is a function of the carrier recombination lifetime. Since the carrier recombination

lifetime is independent on the carrier concentration in a semiconductor the ambipolar

diffusion equation can be described by Eq.3.5. As a consequence the PCR signal is a linear

function of the optical excitation intensity (carrier concentration). The CDW field is expressed

by the solution of the ambipolar diffusion equation which is given by (3.12). This is true at

low carrier concentrations where the recombination lifetime is dominated by the SRH

recombination as indicated in Figure 3.5. At high carrier concentrations the recombination

lifetime can be affected by other recombination processes such those discussed in Section 3.5.

This implies an ambipolar diffusion equation that is described by equation (3.5) instead of

Equ. (3.7). The former equation is a nonlinear differential equation that is difficult to solve.

The concept of the nonlinear coefficient β (defined later) is introduced to take into account

possible effects of the carrier concentration on the recombination lifetimes (and subsequently

on the PCR signal).

Then the PCR is sensitive only on the emitting photons. Photons (in the spectral detection

range of the PCR measurements) are produced by the radiative recombination of an electron

to the defects state (monopolar limit) or by an electron-hole pair (bipolar limit). If the

recombination lifetime of the carriers is dominated by the radiative recombination, then the

recombination lifetime for p-type silicon wafer in high injection level ( 0pn >>∆ ) can be

described by the SRH recombination (monopolar limit) [Schroder ,1998]

pnSRH 00 τττ += ,

The bipolar recombination can be expressed by

nBbandbandrad ∆=−

1,τ ,

86

The used quantities are defined in Section 3.5. In general, the relation between the effective

radiative recombination lifetime and photo-injected carrier concentration can be written as:

γτ −∆= nCrad (7.2)

where 0<γ<1 and C are constants. At near-degenerate densities of the order of 1018 cm-3 and

higher, it has been found that γ = 1/2 for Ge [Seidel, 1961]. In this range of densities

electronic transport is limited by impurity or carrier-carrier scattering.

In summary, the nonlinear coefficient is related with the PCR amplitude. Phenomenological,

the amplitude of the PCR signal SPCR(ω)is a nominally linear function of the incident laser

power P0, or intensity I0, but, in view of the N-dependence of the Drude equation (7.1), it can

be generalized as:

0 0( )PCRS aP bIβ βω = = (7.3)

where the experimental value of β is given by the slope

( )

( )0log

log

P

SPCR

∆∆

β (7.4)

Here a and b are constants, and β is the non-linearity coefficient/exponent. Therefore,

β = 1 + γ and αf(N) ∝ I01+γ.

The concept of the non-linear coefficient was introduced by Guidotti et al. [Guidotti et

al, 1988; Guidotti et 1989]. The authors investigated theoretically and experimentally the

laser power dependence of the modulated photoluminescence in 10 – 15 Ω cm p-Si

corresponding to equilibrium hole density p0 = 1x1015 cm-3 under room temperature

conditions, Figure 7.1.

87

Figure 7.1 Power dependence of Y(ω) and Y(2ω) [Guidotti et al., 1988]

They found a sharp linear-to-quadratic transition using a 647-nm Krypton-Argon laser

with the quadratic threshold at incident power ~ 5 mW corresponding to an excess electron-

hole density ~ 3x1016 cm-3. The linear range they attributed to recombination of photoexcited

carriers via donor (or acceptor) density of states present primarily due to semiconductor

doping. Recombination in the presence of impurities and dopants in Si has been known to be a

source of room-temperature photoluminescence (PL) since the 1950s [Haynes et al.,1956].

Consequently the quadratic behavior was explained as due to bipolar recombination via

photogenerated electron and hole densities of states. A superposition of linear and quadratic

dependence on laser intensity of spectrally integrated dc PL emission in layered

semiconductors (InP/InGaAs/InP) has been reported by Nuban et al. [Krawczyk and Nuban,

1994; Nuban and Krawczyk, 1997].

88

7.2 Experimental methodology and materials

During the study of nonlinear dependence of PCR signal on the optical excitation

intensity two PCR systems were used. The super-bandgap laser intensity and frequency

dependence of PCR signals were studied using a set-up with 532 nm wavelength and 18 µm

spotsize and another set-up with 830 nm wavelength and 24, 387 and 830 µm spotsizes.

7.2.1 Low resolution PCR system.

The PCR system with the diode laser with λ=830 nm operating wavelength is shown in

Figure 7.2.

LOCK-IN AMPLIFIER

FUNCTION GENERATOR

COMPUTER

XY

ZInGaAs

DETECTOR

LONG PASSFILTER

SAMPLE

MOTORIZEDSTAGES

OFF-AXISPARABOLOIDAL

MIRRORS

SUPER-BAND GAPLASER 830 nm

LENS GROUP AND SPOTSIZE FILTER GROUP

MIRROR

Pos. 1 – NO FILTER

Pos. 2 – DEFOCUSINGLENS f’ = 1 m

Pos. 3 – DEFOCUSINGLENS f’ = 0.5 m

Pos. 1 – NO FILTER

Pos. 2 – GROUP NEUTRALDENSITY FILTERS

Pos. 3 – VARIABLE NEUTRALDENSITY FILTER

GRADIUMLENS

24 µm

387 µm

830 µm

LOCK-IN AMPLIFIER

FUNCTION GENERATOR

COMPUTER

XY

Z

XY

Z

XY

ZInGaAs

DETECTOR

LONG PASSFILTER

SAMPLE

MOTORIZEDSTAGES

OFF-AXISPARABOLOIDAL

MIRRORS

SUPER-BAND GAPLASER 830 nm

LENS GROUP AND SPOTSIZE FILTER GROUP

MIRROR

Pos. 1 – NO FILTER

Pos. 2 – DEFOCUSINGLENS f’ = 1 m

Pos. 3 – DEFOCUSINGLENS f’ = 0.5 m

Pos. 1 – NO FILTER

Pos. 2 – GROUP NEUTRALDENSITY FILTERS

Pos. 3 – VARIABLE NEUTRALDENSITY FILTER

Pos. 1 – NO FILTER

Pos. 2 – GROUP NEUTRALDENSITY FILTERS

Pos. 3 – VARIABLE NEUTRALDENSITY FILTER

GRADIUMLENS

24 µm

387 µm

830 µm

Figure 7.2 Experimental system for PCR measurements with super-bandgap laser at

the operating wavelength 830 nm.

A 150 mW diode laser (Melles Griot model 561CS115/HS) with a operating wavelength

of 450 MHz through voltage input from a function generator (Stanford Research Systems

model DS335). Output power modulation is achieved through current stealing so that when a

positive voltage is applied some or the entire laser current is shunted through the modulation

89

circuitry. The laser beam is directed to the surface and focused onto the sample surface using

a gradium lens. The position of the laser beam is coincident with the focal point of an off-axis

paraboloidal mirror that collects a portion of any diffuse back-scattered photons. The

collected IR is then focused onto detector. The system is designed such that the specular

reflection of the excitation beam is not collected by the paraboloidal mirrors and thus is not

focused onto the detector (Chapter 4). The angle of incidence of the excitation beam is ~28°.

The signal from the detector is demodulated by a lock in amplifier (EG&G model 5210). All

instruments, data acquisition, and data storage are controlled by a computer running MATLAB

program with a graphical user interface and real-time display of experimental data.

7.2.2 High resolution PCR system.

The PCR system with the λ=532 nm operating wavelength is shown in Figure 7.3.

LOCK-IN AMPLIFIER

FUNCTION GENERATOR

COMPUTERInGaAs

DETECTOR

LONG PASSFILTER

SAMPLEMOTORIZED

STAGES

SUPER-BAND GAPLASER 532 nm

MIRROR

GRADIUMLENS

FILTER GROUP

Pos. 1 – FREE SPACE

Pos. 2 – GROUP NEUTRALDENSITY FILTERS

Pos. 3 – VARIABLE NEUTRALDENSITY FILTER

AOM DRIVER

XY

Z

MIRROR

REFLECTINGOBJECTIVES

ACOUSTO_OPTICMODULATOR

DIAPHRAGM

45°MICRO-MIRROR

LOCK-IN AMPLIFIER

FUNCTION GENERATOR

COMPUTERInGaAs

DETECTOR

LONG PASSFILTER

SAMPLEMOTORIZED

STAGES

SUPER-BAND GAPLASER 532 nm

MIRROR

GRADIUMLENS

FILTER GROUP

Pos. 1 – FREE SPACE

Pos. 2 – GROUP NEUTRALDENSITY FILTERS

Pos. 3 – VARIABLE NEUTRALDENSITY FILTER

AOM DRIVER

XY

Z

XY

Z

MIRROR

REFLECTINGOBJECTIVES

ACOUSTO_OPTICMODULATOR

DIAPHRAGM

45°MICRO-MIRROR

Figure 7.3 Experimental system for PCR measurements with super-bandgap laser

wavelength 532 nm.

90

Radiation from a Coherent Model Verdi V10 diode-pumped laser was harmonically

modulated by an acousto-optic modulator (AOM) (ISOMET Model 1205C-2). The modulated

beam was directly focused by a Gradium lens (focal length 60 mm) to the polished surface of

the wafers following a 90-degree reflection from a 45° micro-mirror attached to the collecting

reflecting objective (Ealing/Coherent Model 25-0522). The beam spotsize was 18 µm. Under

the high-intensity conditions for this experimental configuration, a signal stabilization period

was necessary, during which the PCR amplitude reached a constant level while the phase

remained constant, before any measurements were recorded. In order to avoid unnecessary

exposure of the sample to laser irradiation leading to PL fatigue, the beam was interrupted

between successive measurements. The PCR NIR emissions from the samples were collected

and collimated by the silver coated reflecting objective and focused onto an InGaAs Detector

( Thorlabs Model DET410,Chapter 3) with a second reflecting objective (Ealing/Coherent

Model 25-0506). The output signal was fed to a lock-in amplifier (Model SRS 850 DSP) and

processed by a personal computer.

7.2.3 Materials

Experiments were performed with two (100) p-type (boron-doped) Si wafers. One wafer

with relatively long lifetime, 6// diameter, thickness 675 ± 20 µm, and resistivity 14 - 24 Ωcm,

was labeled W1. The other wafer, with relatively short lifetime, 4// diameter, thickness 525 ±

20 µm, resistivity 20 - 25 Ωcm, was labeled W2. Both wafers had a 500 Å thermal oxide layer

and were placed on a mirror sample holder to amplify the PCR signals [Mandelis et al.,2003].

In the previous chapter the influence of the SCL on the PCR signal was found to be important

during estimation of the transport parameters at the Si-Si02. It was found that the influence of

the SCL is very important for the long lifetime silicon wafer. Since the W1 and the W2 wafers

have relatively small recombination lifetimes and the thermal oxide layer is considerably

thinner to that in Chapter 6, the effect of the SCL is neglected.

91

7.3. Numerical simulations of the PCR signal as a function of the non-linear coefficient ββββ

and photo-injected carriers.

If there is no the carrier concentrations depend on the recombination lifetime τ, the one

dimensional PCR signal can be expressed by the formula (3.5):

( ) ( ) ( ) ( ) ( ) ( )∫ =∆=L

DDD MEFdxxNFS0

1121211 ,,, ωωλλωλλω . (7.5)

Otherwise the non-linear coefficient β (defined in (7.4)) is introduced to the one-dimensional

PCR signal expression and then it is given by

( ) ( ) ( ) ( ) ( ) ( )∫ =∆=L

DDD MEFdxxNFS0

1121211 ,,,,,,, βωβωλλωλλβω β (7.6)

where ( ) ( )( )βσαν

ηαβω22*

01

2

1,

e

DDh

RIE

−= , (7.6a)

and ( ) ( ) ( ) ( )β

ασ

ασασ

ασγγβω

−−

Γ−Γ+Γ−+Γ= −

−+−−Le

L

LLL

D ee

eeeM

e

ee

1,2

12

)(1221

1 , (7.6b)

where existing quantities were defined in Section 3.4.

The numerical simulations of formula (7.4) using MATLAB program were performed. The

results of these simulations as a non-linear coefficient β set a parameter are shown in Figure

7.4.

92

10-3 10-2 10-1 100 101

-1.4

-1.2

-1.0

-0.8

-0.6

-0.4

10-3 10-2 10-1 100 10110-16

10-11

10-6

10-1

Intensity I0 [W/cm-2]

PC

R p

hase

[deg

.]

β=1.0 β=1.2 β=1.4 β=1.6 β=2

PC

R a

mpl

itude

[a.u

.]

Intensity I0 [W/cm-2]

Figure 7.4 The PCR amplitude (a) and phase (b) of p-type silicon wafer as a function

of the intensity with the different values of nonlinear coefficient. Other parameters

were set: τ=50µs ,S1=100cm/s, S2=105 cm/s, Dn* = 30 cm2/s, L = 500 µm,α(λ=514nm)

= 7.76×103 cm-1,f=5 kHz.

It is clearly seen from Figure 7.4 that with increasing non-linear coefficient the PCR

amplitude also increases. Additionally, slopes of the PCR amplitudes increase with increasing

non-linear coefficient. This can be explained by the fact that with increasing non-linear

coefficient the probability of the radiative recombination processes also increase. As a result

also the PCR amplitude increases. Moreover small changes in PCR phase are also observed. It

is seen that with increasing non-linear coefficient increase the PCR phase lag. Regardless of

the fact that CDW field expressed by Equ. 7.6 doesn’t describe correctly dependence of the

carrier concentration on the PCR, some effects of the nonlinearity can be seen in numerical

simulations of the linear PCR signal expression with the recombination lifetime, described by

formula (3.17). The numerical simulations results are shown in Figure 7.5 and it was

performed by means of the MATLAB program.

93

10-3 10-2 10-1 100 101109

1014

1019

1024 β=1.0 β=1.2 β=1.6 β=1.8 β=2.0

PC

R a

mlit

ude[

a.u.

]

Intensity [W/cm2]

1E-3 0,01 0,1 1 10-1,5

-1,4

-1,3

-1,2

-1,1

-1,0

-0,9

-0,8

-0,7

-0,6

-0,5 β=1.0 β=1.2 β=1.6 β=1.8 β=2.0

Intensity [W/cm2]

PC

R p

hase

[deg

.]

Figure 7.5 The PCR amplitude (a) and phase (b) of p-type silicon wafer as a function

of the intensity with the different values of nonlinear coefficient. Other parameters

were set: τ=50µs ,S1=100cm/s, S2=105 cm/s, Dn* = 30 cm2/s, L = 500 µm,α(λ=514nm)

= 7.76×103 cm-1,f=5 kHz and in Chapter 3.

The three dimensional expression of PCR signal can be similarly generalized to the one

dimensional ones in terms of the non-linear coefficient and can be written as

( ) ( ) ( )∫∞

=0

2132

33, ,,

~, λλβωλ

πβω daJN

a

CS D

DDPCR , (7.7)

94

where

( ) ( ) ( )βωλβωλβωλ ,,,,,,~

333 DDD MEN = (7.8a)

( ) ( )( )β

λ

ξαναηβωλ

22*

40

32

1,,

22

e

W

DDh

eRIE

−=−

(7.8b)

and ( ) ( ) ( ) ( )[ ] ( ) βαξ

ξ

αωλωλ

ξβωλ

−−+−=

−−

− LL

e

L

D

eeCC

eM e

e 1,,

1,, 213 (7.8c)

7.4 Experimental results and discussion

7.4.1 Laser power dependencies

In this section the laser power dependence, P0, of the amplitude and phase of the PCR

signal is reported. In the frame of this chapter experimental results are presented as a function

of power or laser spotsize, instead of optical excitation intensity. This is due to dimensionality

of signals which affects the theoretical interpretation and determination of the transport

parameters (Chapter 3). Figure 7.6 shows results from wafer W1 made on the Low Resolution

PCR system at 10 kHz.

95

0.2 0.4 0.6 0.8 1.0 1.2

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

0.2 0.4 0.6 0.8 1.0 1.2

-60

-50

-40

-30

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log(Laser Power [mW])

Figure 7.6 Log – Log plot of PCR amplitude (a) and Lin – Log plot of phase (b) vs.

laser power at 830 nm and 10 kHz from wafer W1 and 24-µm (), 387-µm () and 830-

µm () spotsizes of the focused laser beam. Slopes of the corresponding best fits (—)

for amplitudes: β = 1.90 (24 µm); 1.82 (387 µm); 1.72 (830 µm).

96

1.1 1.2 1.3 1.4 1.5 1.6

-2.0

-1.5

-1.0

-0.5

0.0

1.1 1.2 1.3 1.4 1.5 1.6

-2

-1

0

1

2

3

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log(Laser power [mW])

Figure 7.7 Log – Log plot of PCR amplitude (a) and Lin – Log plot of phase (b) vs. laser

power at 830 nm and 10 kHz from wafer W2 and 24-µm (), 387-µm () and 830-µm ()

spotsizes of the focused laser beam. Slopes of the corresponding best fits (—) for

amplitudes: β = 1.76 (24 µm); 1.76 (387 µm); 1.60 (830µm)

Laser power dependencies of the PCR signal were obtained and slopes in log-log plots of

amplitudes (Fig. 7.6a) vs. power were calculated. The laser power was varied in the range 1.8

mW – 15.6 mW. Slopes β were calculated from Eq.(7.6) and were found to be 1.90, 1.82, and

1.72 for spotsizes 24 µm, 387 µm and 830 µm, respectively. It is clear that the values of β

decrease with increasing spotsize, that is, with decreasing laser intensity. Except for a rigid

phase shift associated with the change in the diffusing carrier-wave dimensionality with

changing spotsize shown in Fig. 7.6b, there is also a dependence of the PCR phase on laser

power, indicative of the non-linear origin of the PCR signal (simluation). For fixed spotsize,

the decrease in phase lag is consistent with enhanced carrier-carrier scattering with increasing

laser power leading to accelerated recombination and a decrease in the distance below the

97

wafer surface of the location of the carrier density wave centroid, physically a diffusion-

limited free carrier mean free path [Mandelis, 2001]. Decreasing the spotsize affects the

dimensionality of the diffuse carrier density wave, enhancing the radial diffusion degrees of

freedom and shifting the carrier-wave centroid phase lag farther away from surface. The

experimental results of the optical power dependence for the W2 are shown in Fig. 7.7.The

laser power was varied in the range 13.6 mW – 37.5 mW and, again, the slopes β were found

to decrease with decreasing intensities. Furthermore, they were consistently lower than those

measured with the long-lifetime wafer W1. There was no measurable phase shift with laser

spotsize or power changes in Fig. 7.7(b). This behavior at the considered frequency

(f=10kHz) may be the result of a weaker non-linearity mechanism sample W2 than in W1, as

expected from the shorter lifetime and smaller photogenerated root mean square rms free-

carrier-density in the former Si wafer.

Another set of measurements was made at the High Resolution PCR System using the

highly focused set-up shown in Fig. 7.3 at 10 kHz. Figure 7.8(a) shows the power dependence

of PCR amplitude for wafer W1 in the range 6 mW – 25 mW. The slope was found to be 1.89.

The phase lag exhibited a slight decrease with increasing laser power, Fig. 7.8(b), as expected

from enhanced carrier scattering processes. The same measurement was repeated with wafer

W2 and the results are shown in Fig. 7.9

98

0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

-4.0

-3.8

-3.6

-3.4

-3.2

-3.0

-2.8

-2.6

0.8 0.9 1.0 1.1 1.2 1.3 1.4

-23.0

-22.5

-22.0

-21.5

-21.0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log(Laser power [mW])

Figure 7.8 Log – Log plot of PCR amplitude (a) and Lin – Log plot of phase (b) vs.

laser power () at 532 nm and 10 kHz from wafer M 28 and 18-µm spotsize of the

focused laser beam. Slope of the corresponding best fit (—) for amplitude: β = 1.89

(18 µm).

99

0.09 0.10 0.11 0.12 0.13 0.14

-4.00

-3.75

-3.50

-3.25

0.09 0.10 0.11 0.12 0.13 0.14

-4.9

-4.8

-4.7

-4.6

-4.5

-4.4 b

Log(

PC

R A

mpl

itude

[mV

])

a

PC

R P

hase

[deg

rees

]

Log(Laser power [mW])

Figure 7.9 Log – Log plot of PCR amplitude (a) and Lin – Log plot of phase (b) vs. laser

power () at 532 nm and 10 kHz from wafer W2 and 18-µm spotsize of the focused laser

beam. Slope of the corresponding best fit (—) for amplitude: β = 1.94 (18 µm).

Laser power was varied between 8.7 mW and 22.4 mW and the amplitude slope was 1.94. No

discernible PCR phase dependence on laser power was found. This result at 532 nm is

consistent with the 830-nm results obtained with this wafer. They can both be understood in

terms of the short recombination lifetime, which renders the subsurface position of the

diffusive carrier-wave density centroid [Mandelis, 2001] essentially independent of lifetime at

the frequency of these experiments. The higher slopes than those obtained at 830 nm are due

to the increased laser-induced carrier-wave densities as discussed below.

100

7.4.2. Modulation frequency dependence at the low resolution system.

Figure 7.10 shows frequency scans obtained from wafer W1 and three laser powers at a

spotsize of 830 nm. The corresponding best fits of the 3-D theory, Eq. (7.10) with β = 1 to the

data are also shown. The phases, Fig. 7.10(b) nearly overlap and thus they appear not to

depend strongly on laser power. Figure 7.6(b) shows a change of ~ 10 degrees between 1.8 –

15.6 mW at 10 kHz. Small differences on the order of 2° exist among phases across the

narrower power range in Fig. 7.10(b), however, they are nearly imperceptible over the full 75°

phase range. The frequency dependence of the PCR signals using the same laser powers and

24 µm and 387 µm spotsizes exhibited similar behavior.

0.1

1

10-2 10-1 100 101 102

-70

-60

-50

-40

-30

-20

-10

0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

101

Figure 7.10 Experimental frequency scans (10 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W1 for various 830-nm laser powers focused to an 830-µm

spotsize, and the corresponding multiparameter theoretical 3-D best fits (—) using

Eq. 7.10 with β = 1. Laser power: 20 mW (), 15.2 mW () and 9.6 mW ().

Figure 7.11 shows the frequency dependent PCR signals from wafer W1 using 20 mW and

various spotsizes, as well as the corresponding best fits to the 3-D theory, Eq. (7.10),

with β = 1. Figure 7.11(b) shows significant phase changes with spotsize at high frequencies

due to changes in the dimensionality of the PCR signal [Rodriguez et al., 2000].

1E-1

1

1E-2 1E-1 1 10 100-80

-70

-60

-50

-40

-30

-20

-10

0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

102

Figure 7.11 Experimental frequency scans (10 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W1 for 830-nm and 20 mW laser beam focused to various

spotsizes: 24 µm (), 387 µm () and 830 µm (). Best theoretical 3-D fits (—) using

Eq. 7.4 with β = 1 are also shown.

The results of best-fitting the 3-D theory to the data and extracting transport parameters are

shown in Table 7.1 The frequency dependencies of the PCR signals generated with laser

powers 15.2 mW and 9.6 mW and focused on the same spotsizes exhibited similar behaviors.

Table 7.1. Best-fitted computational results for sample W1. Experimental data at 830

nm 3-D theory and integer values of the non-linearity coefficient β

3-D Model, Eq. (7.10)

β = 1 β = 2

Power

P0 [mW]

Transport and Beam Parameters 24 µm 387 µm 830 µm 24 µm I0 [W/m2] 21.2×106 8.2×104 1.8×104 21.2×106

∆N[1/cm3] 8. ×1018 9.51016 3.9×1016 6.6×1017 τ [µs] 39.3 49.3 84.9 35.9

D [cm2/s] 0.31 4.2 9.9 11.1 S1 [cm/s] 0.07 980 960 4.3×103

9.6

S2 [cm/s] 2.3×104 2.4×104 3.7×103 1.2×104 I0 [W/m2] 33.6×106 1.3105 2.8×104 33.6×106

∆N[1/cm3] 1.5×1019 1.31017 4.3×1016 9.6×1017 τ [µs] 43.5 42.6 55.1 28.8

D [cm2/s] 0.26 4.9 5.5 9.9 S1 [cm/s] 62 890 787 4×103

15.2

S2 [cm/s] 4.4×103 4×103 6.8×103 1.2×104 I0 [W/m2] 44.2×106 1.7×105 3.7×104 44.2×106

∆N[1/cm3] 1.7×1019 1.7×1017 5.4×1016 1.2×1018 τ [µs] 37.2 43.8 51.6 26.7

D [cm2/s] 0.26 5.3 4.7 10 S1 [cm/s] 60 890 740 4×103

20

S2 [cm/s] 1.7×104 6.5×103 3.6×104 1.2×104

The theoretical best fits in Figs. 7.10 and 7.11 were made using a computational program

which minimizes the variance of the combined amplitude and phase fits [Li et al., 2005].

Furthermore, the PCR signals were fitted using a non-linear extension of the foregoing 3-D

computation algorithm with β = 2. It is important to recall that fractional powers 1 < β < 2

103

could not be efficiently accommodated in the 3-D PCR theory, Eq. (7.10). A 1-D non-linear

best-fit algorithm was generated based on Eq. (7.9) and capable of accommodating arbitrary

values 1 ≤ β ≤ 2. The computational algorithm was implemented with the corresponding

experimental values for β (Fig. 7.6 for sample W1) and numerical quantities are presented in

Table 7.2.

Table 7.2. Best-fitted computational results for sample W1. Experimental data at 830

nm 1-D theory and experimental values of non-linearity coefficient β.

1-D Model, Eq. (7.9)

β = 1.82 β = 1.72

Power

P0 [mW]

Transport and Beam Parameters

387 µm 830 µm I0 [W/m2] 8.2×104 1.8×104

∆N[1/cm3] 4×1016 3.2×1016 τ [µs] 31.1 81.4

D [cm2/s] 25 28.7 S1 [cm/s] 1×104 7×103

9.6

S2 [cm/s] 420 9.1×103 I0 [W/m2] 1.3105 2.8×104

∆N[1/cm3] 7.4×1016 4.1×1016 τ [µs] 27.9 65.8

D [cm2/s] 9.2 27.5 S1 [cm/s] 4.8×104 6.9×103

15.2

S2 [cm/s] 0.3 9.7×103 I0 [W/m2] 1.7×105 3.7×104

∆N[1/cm3] 8.7×1016 5×1016 τ [µs] 28 59.8

D [cm2/s] 15.3 24.7 S1 [cm/s] 1.1×104 6.5×103

20

S2 [cm/s] 0.2 9.9×103

In all tables the photogenerated excess carrier-wave densities (m-3) shown are mean values

calculated using the expression:

104

∆Ν =( )

019 2

0

1

21.6 10 ( ) [ ( ) ] 1/D P

P

W L f

τω π α−

× + h

, (7.11a)

where P0 is the incident power [W]; 0ωh is the incident photon energy expressed in eV

,ω0 = hc/λP is the photon angular frequency at the wavelength of the laser; and αP is the

optical absorption coefficient at the excitation wavelength. LD(f) is the magnitude of the

complex carrier diffusion length ( )1 2

D

DL f

i f

τπ τ

=+

. (7.11b)

Values of LD and τ used in Eq. (7.11) were calculated from the multiparameter best-fit values

obtained from Tables 7.1-7.4. From Table 7.1 the values of lifetimes obtained with the linear

3-D theory, β = 1, and 24-µm spotsize, are reasonable for p-Si [Rodriguez et al., 2000],

however, the corresponding diffusion coefficients are outside the typical range for p-type

wafers [Rodriguez et al., 2000]. Table 7.1 also shows the results of the same experimental

data, Figs. 7.8 and 7.9, fitted with the 3-D theory and β = 2, Eq. (7.10). The experimental

slope for 24 µm laser spotsize (β = 1.9) is very close to 2 and validates this fitting procedure.

Fitting error was very low, 0.6% – 0.9%. The measured transport parameters for all laser

intensities and 24 µm spotsize can be compared for β = 1 and β = 2 from Table 7.1. It is seen

that recombination lifetimes consistently decrease in the non-linear fit. This decrease is larger

for higher intensities reaching up to 43%. The absolute value of the lifetime monotonically

decreases with increasing intensity for β = 2, as expected from the increased carrier-wave

densities and the proportionately increased scattering probabilities which limit the diffusion

length (or mean free path). In the non-linear fit, carrier diffusivity values increase

dramatically from unrealistically low levels at β = 1: They rise to values approx. 10 cm2/s,

which is normal medium to high injection conditions (∆N ~ 1018 – 1019 cm-3). Front surface

recombination velocities also increase considerably in the non-linear fit, and remain

essentially unchanged with increasing intensity, as expected, since the defect / recombination

center density on the surface is not affected by the incident photon flux. Back surface

recombination velocities do not change as dramatically under the non-linear fit. They are

higher than S1, as expected from the matte nature of the back surfaces of our wafers compared

to the polished front surfaces. No attempt for non-linear fits to the data corresponding to 387

µm and 830 µm spotsizes was made, as the slopes (β = 1.82 and 1.72) were far from β = 2 to

guarantee the validity of the 3-D computation. As shown in Fig. 3.8, the best fits of PCR

signals generated with the relatively large spotsizes, 2W ~ 830 µm , yield very similar sets of

105

transport parameters using either 1-D or 3-D linear theories (β = 1) only when τ < 20 µs. This

is not the case with sample W1 as seen in Table 7.2. However, for comparison with Table 7.2,

the β = 1 parameters were calculated using the 3-D theory and are shown in Table 7.1. Table

7.2 shows results from PCR signals with 2W = 387 µm and 830 µm fitted to the non-linear 1-

D theory, Eq. (12), with β = 1.82 and 1.72, from Fig. 7.6(a). In view of Fig. 3.7 and the best-

fitted values of recombination lifetime under all excitation intensities, the calculated transport

parameters with the 1-D model can only be considered to be approximate. The excess

photoinjected carrier densities are somewhat lower than the 2W = 24-µm case. Recombination

lifetimes under all incident intensities do not change much from the β = 1 values but they

decrease monotonically with increased intensity, as expected for enhanced carrier-carrier

scattering [Li et al.,1997]. It is noted that the lifetime obtained under 9.6 mW excitation and

2W = 830 µm is 81.4 µs, close to the value 84.9 µs obtained from fitting the same data with

the 3-D theory and β = 1. These differences increase with increasing incident intensity. As

observed with 2W = 24 µm spotsize, for non-linear fits with β = 1.82 (2W = 387 µm) and β =

1.72 (2W = 830 µm) the values of D increased manifold over the β = 1 values, attaining the

normal range of p-Si [Rodriguez et al., 2000]. However, the values of D are larger than those

obtained with 2W = 24 µm as expected from the much lower intensities accompanying the

larger spotsizes. Only slight decreases are observed with increasing intensity [Li et al.,1997].

Front and back surface recombination velocities do not undergo significant changes for all

non-linear fits, β = 2 (2W = 24 µm), β = 1.82 (2W = 387 µm) and β = 1.72 (2W = 830 µm).

This is as expected from the intrinsic defect structure of the wafer surfaces. S2 is consistently

higher than S1 for all incident intensities. The comparison between Tables 7.1 and 7.2 show

that the difference between the β = 1 values and the non-linear results is large and due to the

non-linearity exponents β = 1.72 and 1.82. This comparison demonstrates the need for precise

knowledge of the degree of non-linearity of the PCR signal in order to obtain physically

reasonable and reliable (i.e. self-consistent) measurements of semiconductor transport

properties. Figure 7.12 shows two sets of modulation frequency scans obtained from wafer

W2 including the corresponding best fits to 3-D theory with β = 1. The phase plots at fixed

spotsize nearly overlap, Fig. 7.12b, and are essentially independent of laser power, consistent

with Fig. 7.7b. Figure 7.13 shows frequency scans with 38-mW laser power and various

spotsizes. The corresponding 3-D theoretical best fits to Eq. (7.10) and β = 1 are also shown.

106

9E-31E-2

2E-2

3E-2

4E-2

5E-2

1 10 100

-25

-20

-15

-10

-5

0

5

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

Figure 7.12 Experimental frequency scans (500 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W2 for various 830-nm laser powers focused to an 830-µm

spotsize and the corresponding theoretical 3-D best fits (—) using Eq. 7.4 with β = 1.

Laser power: 38 mW () and 12 mW ().

107

1E-1

1

1 10 100

-25

-20

-15

-10

-5

0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

Figure 7.13 Experimental frequency scans (500 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W2 for 830-nm, 38 mW laser beam focused to various

spotsizes: 24 µm (), 387 µm () and 830 µm (). Best theoretical 3-D fits (—) using

Eq. 7.4 with β = 1 are also shown.

The phase shifts at the high-frequency end of Fig. 7.13b are due to changing signal

dimensionality from 3-D (2W = 24 µm) to 1-D (2W = 830 µm). The frequency dependence of

the PCR signals with excitation power 12 mW and the same spotsizes exhibited similar

behavior. The PCR signals of Fig. 7.12 were fitted to 1-D theory using the actual

experimental non-linear values for β, Fig. 7.7a. Calculated transport parameters are presented

in Table 7.3.

108

Table 7.3. Best -fitted computational results for sample W2. Experimental data at 830

nm with 1-D and 3-D theories, linear and various non-linearity coefficients

3-D Fitting Model 1-D Fitting Model

β = 1 β = 1.76 β = 1.60

Power P0

[mW]

Transport and Beam Parameters 387 µm 830 µm 387 µm 830 µm I0 [W/m2] 1×105 2.2×104 1×105 2.2×104

∆N[1/cm3] 3.6×1015 6.2×1014 1.6×1015 3.9×1014 τ [µs] 1.3 0.9 0.54 0.55

D [cm2/s] 24.6 34.5 23.2 23.4 S1 [cm/s] 9.3×103 370 390 390

12

S2 [cm/s] 3.9×104 2.2×104 1.6×104 1.6×104 I0 [W/m2] 3.2×105 7×104 3.2×105 7×104

∆N[1/cm3] 1×1016 2.2×1015 4.8×1015 1.5×1015 τ [µs] 1.3 1.1 0.51 0.65

D [cm2/s] 32 32.2 23.4 19.5 S1 [cm/s] 8×103 430 407 920

38

S2 [cm/s] 1.5×104 2.4×104 1.5×104 1.6×104

Unlike sample W1, the short-lifetime (τ ∼ 1 µs) wafer W2 did yield PCR frequency scans

which could be considered fully one-dimensional under the expanded laser beam spotsizes of

387 and 830 µm, as the numerical simulations of PCR signal presented on Fig. 4.9. At 2W =

24 µm the supra-linear slope β = 1.76, Fig. 7.7a, did not allow using the 3-D Eq. (7.10) to fit

the data and yield representative values of the transport parameters since we could only use

integer values of β. Therefore, only fits for the largest two spotsizes are shown in Table 7.3.

Here the lifetimes are very short for both linear and non-linear fits and they do not change

perceptively with increased laser intensity. This fact may be traced to the relatively low

photoexcited carrier-wave densities in the 1014 – 1015 cm-3 range, using Eq. (7.11). These are

up to four orders of magnitude lower than those estimated for the sample W1. The best-fitted

diffusivities are in the range of acceptable values for p-Si [Rodriguez et al., 2000] and vary

little with increased intensity; again, owing to the relatively low ∆Ν. Front-surface

recombination velocities are lower than back-surface velocities, as expected. There are large

differences between the linear and non-linear fits for the same W which lead to the same

conclusion as in the case of wafer W1 in terms of the importance of using the actual non-

linear exponent obtained experimentally in order to calculate the transport parameters. In

summary, there were neither large differences in values of the transport parameters between

109

excitation with 12 and 38 mW, nor were clear trends in those parameters established. This

would be expected under low-injection conditions, however, in these experiments the ∆Ν

range corresponded to intermediate conditions [Bullis and Huff, 1996].

7.4.3. Modulation frequency dependence at 532 nm

Figures 7.14 and 7.15 show the frequency dependent amplitude and phase of the PCR

signal from wafers W1 and W2, respectively, using several laser powers and 18 µm spotsize.

The corresponding best fits to the 3-D theory, Eq. (7.10), are also shown.

1E-4

1E-3

1E-2 1E-1 1 10 100

-40

-30

-20

-10

0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

Figure 7.14 Experimental frequency scans (10 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W1 for various 532-nm laser powers focused to an 18-µm

spotsize and the corresponding theoretical 3-D best fits (—) using Eq. 7.4 with β = 2.

Laser power: 16 mW (), 13.2 mW () and 9.6 mW ().

110

6E-5

7E-58E-59E-51E-4

2E-4

1E-2 1E-1 1 10 100-12

-10

-8

-6

-4

-2

0

a

Log(

PC

R A

mpl

itude

[mV

])

b

PC

R P

hase

[deg

rees

]

Log (Frequency [kHz])

Figure 7.15 Experimental frequency scans (10 Hz – 100 kHz) of PCR amplitudes (a)

and phases (b) from wafer W2 for various 532-nm laser powers focused to an 18-µm

spotsize and the corresponding theoretical 3-D best fits (—) using Eq. 7.4 with β = 2.

Laser power: 16 mW (), 13.2 mW () and 9.6 mW ().

111

While the amplitudes scale non-linearly with laser power with slopes β = 1.89 (W1, Fig. 7.8)

and β = 1.94 (W2, Fig.7.7), the phases at the various laser powers nearly overlap, consistently

with Figs. 7.8b and 7.7b. The best fits to Eq. (7.10) presented in Figs. 7.14a and 7.15a were

made using β = 2. This value of β is very close to the experimentally measured slopes,

therefore it is expected that transport parameters obtained from these fits are close to those

that would have been obtained using the exact experimental supra-linear slopes. The

multiparameter best-fit results to the data are presented in Table 7.4 for three laser powers.

Table 7.4. Best-fitted computational results for samples W1 and W2. Experimental

data at 532 nm with 3-D theory.

3-D Fitting Model for Wafer W1

3-D Fitting Model for Wafer W2

Power P0

[mW]

Transport and Beam Parameters Β = 1 β = 2 β = 1 β = 2 I0 [W/m2] 3.8×107 3.8×107 3.8×107 3.8×107 ∆N[1/cm3] 3.6×1020 3.1×1019 3×1018 8.3×1017 τ [µs] 238 70.5 4.2 0.34

D [cm2/s] 0.46 2.7 5.7 8.9 S1 [cm/s] 123 5.6×104 6.7×103 3.7×103

9.6

S2 [cm/s] 2×104 9.4×104 4.2×104 9.8×103 I0 [W/m2] 5.2×107 5.2×107 5.2×107 5.2×107 ∆N[1/cm3] 6.1×1020 3.4×1019 4.4×1018 1.2×1018 τ [µs] 259 56.7 3.5 0.34

D [cm2/s] 0.37 2.9 4.7 8.9 S1 [cm/s] 184 4.3×104 6.3×103 3.7×103

13.2

S2 [cm/s] 4.3×104 8.6×104 3.7×104 9.9×103 I0 [W/m2] 6.3×107 6.3×107 6.3×107 6.3×107 ∆N[1/cm3] 5.9×1020 3×1019 5.5×1018 1.6×1018 τ [µs] 200 44.5 4.1 0.33

D [cm2/s] 0.35 3.2 4.8 6.2 S1 [cm/s] 241 4×104 7.4×103 3.6×103

16

S2 [cm/s] 3.9×104 6.8×104 6.9×104 9.6×103

For sample W1 the values of τ decrease with increasing laser intensity, in a manner similar to

that observed with 830-nm irradiation, Table 7.1. The relative values of τ for excitation with

beam spotsizes 18 and 24 µm at the two distinct wavelengths, Tables 7.1 and 7.4, are quite

close, within a factor of two. On the other hand, in Table 7.4 the values of τ obtained with β =

2 for the three incident intensities I0, are more than three times smaller than those obtained

112

with β = 1. Once again, this shows the dramatic difference the non-linearity coefficient value

can make in PCR measurements. The diffusion coefficients obtained using β = 1 are far from

the range of acceptable values for this parameter for p-type wafers [Rodriguez et al., 2000]

and those obtained with β = 2 are marginally within the range of typical values. This is

approx. one order of magnitude smaller than the D values obtained in Table 7.1 and is

probably due to the much stronger carrier-carrier scattering at high-injection carrier-wave

densities in the 1019 – 1020 cm-3 range [Li et al., 1997]. The best-fitted values of S1 obtained

with β = 1 are more than 100 times lower than those obtained with β = 2. The latter values are

essentially independent of laser intensity (also the case in Table 7.1 under 830-nm excitation)

but are clearly higher than their β = 2 counterparts in Table 7.1. This, along with differences

in τ (532 nm) and τ (830 nm), is expected, because the optical absorption depth at 523 nm is

much shorter (0.96 µm; αP = 10,340 cm-1) than that at 830 nm (15.2 µm; αP = 658.9 cm-1)

with the consequence that optical injection samples the defect densities in the near-surface

region must more strongly at 523 nm than at 830 nm, thus exhibiting enhanced sensitivity to

surface recombination and yielding a larger effective surface recombination velocity S1. The

back-surface recombination velocities are all in the range of 104 cm/s for both wavelengths,

Tables 7.1 and 7.4. Physically, they should be the same, however our simulations show that

the shallower optical penetration depth at 523 nm renders the PCR signal less sensitive to S2

than that at 830 nm. The best fits to Eq. (7.10) resulting from the W2-sample experimental

data with β = 1 and 2 are also shown in Table 7.4. The values of τ obtained with β = 2 which

is very close to the experimental value β = 1.94 for all laser powers are in the sub-µs range,

consistently with those obtained at 830 nm in Table 7.2. In fact, they are approx. two times

smaller under 523-nm excitation as expected from a significantly higher intensity range

(about three orders of magnitude) and the corresponding high-injection ramifications. The τ

values obtained with the linear fit are very unrealistic in comparison with other data in Table

7.4, more than ten times larger than those obtained with β = 2. The carrier diffusivity values in

Table 7.4 are considerably smaller than those in Table 7.2 and decrease with increased power

16 mW. These trends are consistent with the shallower penetration depth, higher photoinjected

carrier-wave densities and higher laser intensities under 523 nm excitation [Li et al., 1997]. In

a manner similar to the behavior of sample W1, the S1 values under 523-nm excitation are

higher than under 830-nm, whereas values of S2 are very close. Overall, values of the

transport parameters in Table 7.4 fitted with β = 1 show some significant deviations from the

non-linear fits, especially for τ and S1.

113

7.5 Summary

It is interesting to note for both samples that the ca has three-orders-of-magnitude

higher intensities and shorter optical absorption depth at 523 nm which are associated with

increased non-linearity exponents which tend to values closer to 2, a limiting value for

electron-hole band-to-band bipolar recombination [Guidotti et al, 1988-1989; Nuban and

Krawczyk, 1997; Nuban and Krawczyk, 1997]. In that limit, band-to-defect recombination

may be dominated by the N×P product of electron-hole density recombination. A different

mechanism for quadratic PL dependence on photo-carrier density may have to be considered

such as proposed by Guidotti et al. [Guidotti et al, 1988-1989] and Nuban et al. [Nuban and

Krawczyk, 1997; Nuban and Krawczyk, 1997] and which have been discussed in the context

of Eq. (3.1).

114

Chapter 8: Conclusion and outlook

This final chapter summarizes the main conclusions of this dissertation and outlines the

directions for the future research. Following the organization of the thesis the last chapter is

also devided into two parts. The first section is concerned with the thermal properties of

Cd1-xMgxSe single crystals, the second one is devoted to the new developed technique of

photocarrier radiometry (PCR).

Thermal properties of Cd1-xMgxSe single crystals

Pyroelectric experiments have proven to be a valuable tool to characterize the thermal

parameters of Cd1-xMgxSe single crystals in a limited temperature range around room

temperature. With a home-made apparatus the thermal diffusivity of Cd1-xMgxSe single

crystals could be determined experimentally. It was shown on the glassy carbon (reference

sample) that the thermal diffusivity can be determined directly from measured PPE

amplitudes and phases when sample and detector are thermally thick and optically opaque.

From these data the thermal conductivity of these compounds was estimated. It was found that

thermal conductivity of Cd1-xMgxSe single crystals decreases by a factor of three as the Mg

concentration is increased from zero to about 50 %. This strong variation could be explained

by structural effects of the mixed crystal.

Future work should concentrate on development of the photopyroelectric cell able to

measure in broader temperature range. From the theoretical point of view it is very interesting

to develop theoretical model for mixed crystals to elucidate the relative role of the electronic

and phonon contributions to thermal conduction and the mass-scattering effects as well.

115

Photocarrier radiometry signal

It was shown that in some cases the existence of a space charge layer or bipolar

recombination mechanism can be important when the electrical transport properties of a

silicon wafers are measured by photocarrier radiometry (PCR).

In chapter 6 the interface modulated-charge-density theory developed by Mandelis

[Mandelis, 2005a] was experimentally proved. The interface modulated-charge-density theory

developed by Mandelis [Mandelis, 2005a] was used for physical simulations involving PCR

signals in order to study the effects of the various SCL optoelectronic transport parameters on

the PCR amplitude and phase signal channels. Furthermore, an experimental configuration

was used involving n- and p-doped Si – SiO2 interfaces and a low-intensity (non-perturbing)

modulated laser source as well as a co-incident dc laser of variable intensity acting as

interface-state neutralizing optical bias. The application of the theory to the experiments

yielded the various transport parameters of the samples as well as the depth profile of the

SCL. It was shown that PCR can monitor the complete flattening of the energy bands at the

interface of p-Si – SiO2 with dc optical powers up to a 300 mW. The uncompensated charge

density at the interface was also calculated from the theory. It was found that this

methodology is the relatively low efficient for driving n-Si into the flatband state. The

independence of PCR phases from IDC is an additional indicator that the optical bias does not

affect the transport properties of the SiO2 - n-Si interface. This behavior was explained by the

much lower (ca. 100 times [Aberle, 1992]) minority carrier capture cross-section of the n-type

interface. This fact leads to much shorter interface lifetimes, τs [Mandelis, 2005a], making it

harder for an optical source to build up the neutralized interface charge coverage observed

with p-Si – SiO2 interfaces, Fig. 6.15. The same effect makes the interface recombination

velocity essentially independent of excess carrier density at the n-Si – SiO2 interface [Aberle,

1992] which results in the independence of the PCR phase from PDC, Fig. 6.16.

In chapter 7 it was shown that the introduced nonlinear coefficient, which could be

determined from intensity-scans, can considerably improve estimated values of the carrier

transport parameters. It is interesting to note for both samples that the ca. has three-orders-

of-magnitude higher intensities and shorter optical absorption depth at 523 nm which are

associated with increased non-linearity exponents which tend to values closer to 2, a limiting

value for electron-hole band-to-band bipolar recombination [Guidotti et al, 1988-1989; Nuban

and Krawczyk, 1997; Nuban and Krawczyk, 1997]. In that limit, band-to-defect

116

recombination may be dominated by the N×P product of electron-hole density recombination.

A different mechanism for quadratic PL dependence on photo-carrier density may have to be

considered such as proposed by Guidotti et al. [Guidotti et al, 1988-1989] and Nuban et al.

[Nuban and Krawczyk, 1997; Nuban and Krawczyk, 1997] and which have been discussed in

the context of Eq. (3.1).

Altogether it is demonstrated that the photocarrier radiometry technique is a powerful

tool for the investigation of electronic properties of silicon wafers. The results of this work

show that a careful search has to take into account the influence of different experimental

conditions. Future work should be concentrate on developing 3 D photocarrier radiometry

signal theory extended for discussed phenomena in this work. Additional for ion implanted

wafers (such as presented in Section 4.1.7) the 3 D photocarrier radiometry signal theory in

layered sample developed by Li eat al. [Li, 2004] should be extended to take into account

phenomena related with a space charge layer and bipolar recombination mechanism.

117

Bibliography

[Abeles, 1963] B. Abeles, Phys. Rev. 131, 1906 (1963)

[Aberle et al.,1992] A. G. Aberle, S. Glunz and W. Warta, J. Appl. Phys. 71, 4422 (1992).

[Aberle et al., 1996] A. G. Aberle, J. Schmidt and R. Brendel, J. Appl. Phys. 79, 1491 (1996).

[Adachi, 1983] S. Adachi, J. Appl. Phys. 54, 1844 (1983)

[Almond and Patel, 1996] Almond, D. P. and Patel, P. M. (1996). Photothermal Science and

Techniques. Chapman & Hall, London.

[Angstrm, 1861] Angstrm, A. J. (1861). Ann. Physik. Lpz., 114:513.

[Augustine et al., 1992] G.Augustine, A. Rohatsi, N.M.Jokerst, Based doping optimization for

radiation-hard Si,GaAs and InP solar cells.,IEEE Trans. Electron Dev., 39:2395-2400, 1992

ASTM Standard F723-88, 1996 Annual Book of ASTM Standards, Am. Soc. Test. Mat., West

Conshohocken, PA (1996).

[Balageas et al. 1986] D. Balageas, J.C. Krapez, P. Cielo, J.Appl. Phys. 59,348-357

[Bandiera et al., 1982] Bandiera I.N., Closs, H. and Ghizoni, C.C. J. Photoacoust. 1982, 1,

275

[Bell, 1880] Bell, A. G. (1880), Upon the production of sound by radiant energy. Am.J. Sci.,

20, 305:324.

[Bhandari and Rowe, 1998] C. M. Bhandari, D. M. Rowe, Thermal Conduction in

Semiconductors, 1st edn. (Wiley Eastern Limited, New Delhi, 1998), pp. 59-77

[Boccara et al., 1980] Boccara, A. C., Fournier, D., and Badoz, J. (1980). Thermooptical

spectroscopy - detection by the 'mirage effect'. Appl. Phys. Lett., 36, 130:132.

[Bullis and H.R.Huff, 1996] W.M. Bullis and H.R.Huff, Interpretation of carrier

recombination lifetime and diffusion measurements in silicon, J.Electrochem Soc.,vol.143,

No.4, April 1996

[Carslaw and Jaeger, 1959] Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of heat

in solids. Oxford University Press, 2nd edition

[Chirtoc and Mihailescu, 1989] M. Chirtoc, G. Mihailescu, Phys. Rev. B, 40, 9606 (1989)

[Delenclos et al, 2001] S. Delenclos, M. Chirtoc, A. Hadj Sahraoui, C. Kolinsky, J. M.

Buisine, Anal. Sci. 17, 161 (2001)

[Dietzel, 2001] Dietzel, D. (2001). Photothermische Mikroskopie an Ionenstrahl-

strukturierten Halbleitern und Halbleiterbauelementen. PhD thesis, Ruhr University Bochum.

118

[Dietzel et al., 2003a] Dietzel, D., Chotikaprakhan, S., Filip, X., Chirtoc, M., Bangert,H.,

Eisenmenger-Sittner, C., Neubauer, E., Pelzl, J., and Bein, B. K. (2003a). Cu-coatings on

carbon-Effects of plasma processing analysed by means of thermoreflectance microscopy. In

SFV, France, editor, Proceedings of 14th International Colloquium on Plasma Processes,

CIP'2003, pages 264-66, Antibes-Juan-les-Prins, France.

[Dziewoir and Schmid, 1977] J.Dziewor and W. Schmid, Auger coefficients for highly doped

and highly excited silicon, Appl. Phys. Lett. 31: 346-348, 1977

[Flashir and Cahen, 1986] H.Flashir and D.Cahen, IEEE Trans. UFFC, 1986, 13, 622

[Firszt et al.,1995] F.Firszt,H.Męczyńska, B.Sekulska, J.Szatkowski, W.Paszkowicz,

J.Kachniarz, Semicond. Sci. Technol. 10 (1995) 197-200.

[Firszt et al., 1996]. F. Firszt, S. Legowski, H. Męczyńska, H. Oczkowski, W. Osinska, J.

Szatkowski, W. Paszkowicz, Z. Spolnik, J. Cryst. Growth 159, 167 (1996)

[Firszt et al., 1998] F. Firszt, H. Męczyńska, B. Sekulska, J. Szatkowski, W. Paszkowicz, J.

Cryst. Growth 184/185, 1053 (1998)

[Fournier, 1992] Fournier, D. (1992). Competition between thermal and plasma waves for the

determination of electronic parameters in semiconductor samples. In Bicanic, D. and

Mandelis, A., editors, Photoacoustic and Photothermal Phenomena III, Springer Ser. Opt.

Sci., pages 339-348.

[Fotsing et al., 2003] Fotsing, J. L. N., Hoffmeyer, M., Chotikaprakhan, S., Dietzel, D., Pelzl,

J., and Bein, B. K. (2003). Laser modulated optical reflectance of thin semiconductor films on

glass. Rev. Sci. Instru., 74(1):873-876.

[Gerlach et al., 1972 ] W. Gerlach, H. Schlangenotto and H.Maeder, On the radaitive

recombination in silicon, Phys. Stat.Sol. A., 13:277-283, 1972

[Gruss et al., 1997] Gruss, C., Bein, B. K., and Pelzl, J. (1997). Measurement of high

temperatures by the probe beam reflection technique. High Temp. High Press., 29.

[Gou et al., 1997] Y. Gou, G. Aizin, Y. C. Chen, L. Zeng, A. Cavus, M. C. Tamargo, Appl.

Phys. Lett. 70, 1351 (1997)

[Guidotti et al, 1988] D. Guidotti, J. S. Batchelder, A. Finkel, and J. A. Van Vechten, Phys.

Rev. B, 38,1569 _1988;

[Guidotti et al., 1989] D. Guidotti, J. S. Batchelder, A. Finkel, P. D. Gerber, and J. A. Van

Vachten, J. Appl. Phys., 66, 2542 1989;

[Hall, 1952] R.N.Hall, Electron-hole recombination in germanium, Phys.Rev., 87:387, 1952

[Harren et al., 2000] Harren, F. J. M., Cotti, G., Oomens, J., and Hekkert, S. L. (2000).

119

Encyclopedia of Analytical Chemistry, chapter Photoacoustic spectroscopy in trace gas

monitoring, pages 2203-2226. John Wiley & Sons Ltd, Chichester.

[Haynes et al., 1956] J. R. Haynes and C. W. Westphal, Phys. Rev., 86, 647 _1956

[Holeman, 1972] B. R. Holeman, Infrared Phys. 12, 125 (1972)

[Ikari et al., 1999] T. Ikari, A. Salnick and A. Mandelis, J. Appl. Phys. 85, 7392 (1999).

[Quimby and Yen, 1980] R.S.Quimby, W.M.Yen, J.Appl.Phys, 1980, 51, 4985

[Ikari et al., 1999] T.Ikari, A.Salnick and A. Mandelis, Theoretical and experimental aspects

of three-dimensional infrared radiometry of semiconductors,J.Appl.Phys.85: 7292-7297, 1999

[Johnson, 1958] E. O. Johnson, Phys. Rev. 111, 153 (1958).[Krawczyk and Nuban, 1994]S.

K. Krawczyk and M. F. Nuban, Mater. Sci. Eng., B, 28, 452 1994.

[Kiepert et al., 1999] Kiepert, W., Obramski, H. J., Bolte, J., Brandt, K., Dietzel, D.,

Niebisch, F., and Bein, B. K. (1999). Plasma etching and ion implantation on silicon,

characterized by laser-modulated optical reflectance. Surf. Coat. Tech., 116:119:410-418.

[King and Hall, 1994] King O and Hall D G 1994-I Phys. Rev. B 50 10661

[Kitgawara, 1992] Y.Kitgawara, R.Hoshi, T.Takenada, J.Electrochem.Soc.,139:2277-

2281,1992

[Li et al.,1997] C. M. Li, T. Sjodin, and H. L. Dai, Phys. Rev. B, 56, 15252 _1997

[Li et al., 2004] D.Li, D.Shaughnessy, A.Mandelis and J.Batista, J.of Appl. Physics, Vol.95,

12: 7832-7840, 2004

[Li et al., 2005] B. Li, D. Shaughnessy, and A. Mandelis, J. Appl. Phys., 97, 023701 _2005

[Mandelis, 1998] A.Mandelis, Solid State Electronics, Vol.42, No.1., pp.1-15, 1998

[Mandelis and Zver, 1985] A.Mandelis and M.Zver, J.Appl.Phys. 57 (9), 1985, 4421-4430

[Mandelis, 1996] A.Mandelis, R.E.Wagner, Jpn. J.Appl.Phys., Part 1, 35, 1786 (1996)

[Mandelis et al, 1999] A.Mandelis, A.Salnick, L.Chen, J.Opsal, A.Rosencwaig, J.Appl. Phys.

85, 1811 (1999)

[Mandelis, 2000] A.Mandelis, Diffusion Waves and their uses, Physics Today, August 2000,

pages 29-34

[Mandelis, 2001]A. Mandelis, Diffusion-Wave Fields: Mathematical Methods and Green

Functions, Springer, New York _2001.

[Mandelis et al., 2003] A. Mandelis, J. Batista and D. Shaughnessy, "Infrared photo-carrier

radiometry of semiconductors: Physical principles, quantitative depth profilometry and

scanning imaging of deep sub-surface electronic defects", Phys. Rev. B 67, 205208-1-18

(May 2003)

120

[Mandelis, 2005a] A. Mandelis, “Theory of Space Charge Layer Dynamics at Oxide-

Semiconductor Interfaces under Optical Modulation and Detection by Laser Photocarrier

Radiometry”, J. Appl. Phys., 97, 083508-1 - 11 (April, 2005)

[Mandelis, 2005b] A. Mandelis, J. Batista, J. Gibkes, M. Pawlak and J. Pelzl, “Non-

Contacting Laser Photocarrier Radiometric Depth Profilometry of Harmonically Modulated

Band-Bending in the Space Charge Layer at Doped SiO2 – Si Interfaces”, J. Appl. Phys., 97,

083507-1 - 11 (April 2005)

[Marinelli 1992] M.Marinelli, U.Zammit, F.Mercuri, R.Pizzoferrato, J.Appl.Phys. 72 (3), 1

August 1992

[Meyer and Sigrist, 1990] Meyer, P. L. and Sigrist, M. W. (1990). Atmospheric pollution

monitoring using CO2-laser photoacoustic spectroscopy and other techniques.

Rev. Sci. Instru., 61(7):1779-1807

[Murphy and Aamont, 1977] Murphy, J. C. and Aamodt, L. (1977). Photoacoustic

spectroscopy of luminescent solids: Ruby. J. Appl. Phys., 48(8), 3502-3509.

[Nuban and Krawczyk, 1997] M. F. Nuban, S. K. Krawczyk, and N. Bejar, Mater. Sci. Eng.,

B, 44, 125 1997.

[Nordal and Kanstad, 1979] Nordal, E. and Kanstad, S. O. (1979) Photothermal radiometry.

Phys. Scripta, 20:659.

[Pelzl, 1984] J.Pelzl, In: Photoacoustic Effect, Braunschweig, pp.52-79

[Pelzl and Bein, 1990] J.Pelzl, and B.Bein, Photothermal chracterisation of plasma surface

modifications, Pure & Appl. Chem., 64 (5): 729-744, 1990

[Pelzl and Bein,1992] Pelzl, J. and Bein, B. K. (1992). Photothermal characterization of

plasma surface modications. Pure & Appl. Chem., 64(5):739-744.

[Perzynska et al., 2000] K. Perzynska, F. Firszt, S. Legowski, H. Meczynska, J. Szatkowski,

M. Biernacka, A. Gajlewicz, S. Tarasenko, P. Zaleski, J. Crys. Growth 214/215, 904 (2000)

[Philips et al., 1992] M. C. Philips, M. W. Wang, J. F. Swenberg, J. O. McCaldin, T. C. Mc

Gill, Appl. Phys. Lett. 61, 1962 (1992)

[Rombouts et al., 2005] M. Rombouts, L. Froyen, A. V. Gusarov, E. H. Bentfour, C.

Glorieux, J. App. Phys. 97, 24905 (2005)

[Rodriguez et al., 2000] M. E. Rodriguez, A. Mandelis, G. Pan, L. Nicolaides, J. A. Garcia,

and Y. Riopel, J. Electrochem. Soc., 147, 687 _2000.

[Rosencwaig and Gersho, 1976] Rosencwaig, A. and Gersho, A. (1976). Theory of the

photoacoustic effect in solids. J. Appl. Phys., 47:64-69.

121

[Rosencwaig et al., 1985] Rosencwaig, A., Opsal, J., Smith, W. L., andWillenborg, D. L.

(1985). Detection of thermal waves through optical re°ectance. Appl. Phys.Lett., 46(11):1013-

1015.

[Salazar, 2003] A. Salazar, Rev. of Sci. Instrum.,74, 825 (2003)

[Schaub, 2001] Schaub, E. (2001). New calibration procedure for absolute temperature

measurement on electronic devices by means of thermoreflectance technique. Anal. Sci., 17.

[ Schmidt and Aberle, 1997] J. Schmidt and A. G. Aberle, J. Appl. Phys. 81, 6186 1997.

[Schoefthaler et al., 1994] M. Schoefthaler, R. Brendel, G. Langguth and J. H. Werner, First

WCPEC (1994), p. 1509.

[Schroder, 1997] D. K. Schroder, IEEE Trans. Electron Dev. 44, 160 (1997).

[Schroder, 1998] D.K.Schroder, Semiconductor Material and Device Characterization,

chapter 7:439-448, Wiley, New York, 2nd edition, 1998

[Seidel, 1961] T. Seidel, Metallurgy of Elemental and Compound Semiconductors (Ed. R. O.

Grubel, Interscience, New York 1961).

[Shaughnessy, 2005] D.Shaughnessy, PhD thesis, University of Toronto, 2005

[Shockley and Read, 1952] W.Shockley and Jr.W.T.Read, Statistic of the recombination of

holes and electrons, Phys. Rev., 87:832-842, 1952

[Slack, 1972] G. Slack, Phys. Rev. B, 6, 3791 (1972)

[Smith, 1978] R.A. Smith, Semiconductors, Cambridge University Press, New York, 2nd

edition, 1978

[Sze, 1981] S. M. Sze, Physics of Semiconductor Devices, 2nd ed. Wiley, New York, 198l,

Chap. 7.2.

[Tritt, 2004] T.Tritt, Thermal conductivity: Theory, Properties and Applications, Kluwer

Academic, 2004[Tsen et al., 1997] K. T. Tsen, D. K. Ferry, A. Botchkarev, B. Sverdlov, A.

Salvador, H. Morkoc, Appl. Phys. Lett. 71, 1852 (1997)[Thielemann and Rheinlaender, 1985]

W. Thielemann and B.Rheinlaender, Solid State Electron, 1985, 28, 1111

[Wagner et al, 1996] R.E.Wagner, A.Mandelis, Semicond.Sci.Technol. 11, 300 (1996)

[Varshni, 1967] Y.P.Varshni, Band-to-band radiative recombination in groups IV,VI and III-

VI semiconductors (I), Phys. Stat.Sol., 19:459-514, 1967

[Zakrzewski, 2003] J.Zakrzewski, PhD thesis, Torun, Poland, 2003

[Tolev, 2007] J. Tolev, A. Mandelis and M. Pawlak, J. Electrochem. Soc. J. Electrochem.

Soc., Volume 154, Issue 11, pp. H983-H994 (2007)

122

[Tolev et al., 2008] On the non-linear dependence of photocarrier radiometry signals from Si

wafers on the intensity of the laser beam, J. Tolev, A. Mandelis and M. Pawlak, Eur. Phys. J.

Special Topics 153, 317-320 (2008)

123

Curriculum Vitae

PERSONAL DETAILS

Name Michał Janusz Pawlak

Nationality Polish

Date and place of birth 30.01.1978, Bydgoszcz, Poland

Address Moniuszki 5/3

87-100 Toruń

Poland

Contact [email protected]

tel.: +48 608 349 125

EMPLOYMENT HISTORY

Date Position

Since April 2009 IT specialist at Kujawsko-Pomorskie Voivodeship, Torun,

Poland

Nov 2008 – March 2009 Assistant at Ruhr University Bochum, Germany

Sept 2007 – April 2008 Postdoctoral Fellow at University of Toronto, Canada

May 2007 – August 2007 Assistant at Ruhr University Bochum, Germany

April 2004-March 2007 Scholarship holder at Ruhr University Bochum, Germany

Oct 2002 – March 2004 PhD student at Nicolaus Copernicus University, Torun,

Poland

Oct 2001 – June 2002 Assistant at Nicolaus Copernicus University, Torun, Poland

124

EDUCATION

Date Title Institution

April 2004-March 2007 Scholarship holder Ruhr University Bochum,

Germany

October 2002 – March 2004 PhD student Nicolaus Copernicus

University, Torun, Poland

Since April 2004 Electrical Eng. Warsaw University of

Technology, Poland

October 2000 – June 2002 Master of science

(Physics)

Nicolaus Copernicus

University, Torun, Poland

October 1997 – June 2000 Bachelor (Physics) Nicolaus Copernicus

University, Torun, Poland

PERSONAL QUALITIES

Programming in Java, Matlab, Labview; digital image processing, biomedical eng. Languages: Polish - native, English - good, German – basic.

125

List of publications and conference contributions

Publications

1. Photoacoustic Study of Zn1-x-yBexMnySe Mixed Crystals J. Zakrzewski, F. Firszt, S.

Łęgowski, H. Męczyńska, M. Pawlak, A. Marasek, abstract book 31st Winter School on

Molecular and Quantum Acoustics and 7th workshop on Photoacoustics and Photothermics,

page. 81, 2002.

2. Thermal Diffusivity Measurements of Zn1-x-yBexMnySe Mixed Crystals by Photoacoustic

Method J. Zakrzewski, F. Firszt, S. Łęgowski, H. Męczyńska, M. Pawlak, abstract book

28. Deutsche Jahrestagung fur Akustik, page 114, Bochum, 2002

3. Piezoelectric and pyroelectric study of Zn1-x-yBexMnySe mixed crystal, J. Zakrzewski, F.

Firszt, S. Legowski, H. Meczynska, M. Pawlak, and A. Marasek, Rev. of Scienfic

Instruments, vol. 74 nr 1, p. 566-568 (January 2003)

4. Photoacoustic investigation of Cd1-XMnXTe mixed crystals J. Zakrzewski, F. Firszt, S.

Łęgowski, H. Męczyńska, A. Marasek, and M. Pawlak, Rev. of Scienfic Instruments, vol.

74 nr 1, 572-574 (January 2003)

5. Photoacoustic Study of Cd1-x-yBexMnyTe Mixed Crystals, J. Zakrzewski, F. Firszt, S.

Łęgowski, H. Męczyńska, A. Marasek, M. Pawlak, Journal de Physic IV 109 (2003), 123-

126

6. Growth and characterization of selected wide – gap II – VI ternary solid solutions, F.

Firszt, S. Legowski, A.Marasek, H. Meczynska, M.Pawlak and J.Zakrzewski, abstract

book, International Symposium on 50th Anniversary of the Death of Prof. Dr. Jan

Chochralski, Torun, Poland, p. 4

126

7. Photoelectric and Photothermal Properties of Selected II – VI Compounds Mixed Crystals,

F. Firszt, S. Łęgowski, A. Marasek, H. Męczyńska, M. Pawlak and J. Zakrzewski, Opto-

electronics Rev. 12 (1), p. 161-164 (2004)

8. Study of optical properties of Zn1-xBexTe mixed crystals by means of combined modulated

IR radiometry and photoacoustics, M. Pawlak, J. Gibkes, J. L. Fotsing, J. Zakrzewski, M.

Mali ński, B. K. Bein, J. Pelzl, F. Firszt and A. Marasek,

J. Phys. IV France 117, p. 47-56 (2004)

9. Carrier - density - wave transport and local internal electric field measurements in biased

metal-oxide-semiconductor n-Si devices using contactless laser photo-carrier

radiometry,A.Mandelis, M.Pawlak, D.Shauhnessy, Semicond.Sci.Technol. 19, p. 1240-

1249 (2004)

10. Growth and Optical Characterization of Cd1-xBe xSe and Cd1-xMgxSe crystals,

F. Firszt, A. Wronkowska, A. Wronkowski, S. Łęgowski, A. Marasek,

H. Męczyńska, M. Pawlak, W. Paszkowicz, K. Strzałkowski, and J. Zakrzewski,

Cryst. Res. Technol. 40, No. 4/5, p. 386-394 (2005)

11. Piezoelectric photothermal study of Cd1-x-yBexZnySe crystals,J. Zakrzewski, F. Firszt,

S. Łęgowski, H. Męczyńska, A. Marasek, M. Pawlak, K. Strzałkowski, M. Maliński, and

L. Bychto, Journal de Physique IV 125, p. 473-476 (2005)

12. Photoacoustic Study of Cd1-XBeXSe Mixed Crystals, J.Zakrzewski, M.Pawlak, F. Firszt,

S. Łęgowski, A. Marasek, H. Męczyńska,M.Malinski , Int. Journal of Thermophysics 26

(1), p. 285 (2005)

13. Non – contacting Laser Photocarrier Radiometric Depth Profilometry Of Harmonically

Modulated Band – Bending In The Space Charge Layer At Doped SiO2 – Si Interfaces,

A.Mandelis, J. Batista, J.Gibkes ,M.Pawlak, J.Pelzl, Journal Of Applied Physics 97

083507 (2005).

127

14. Space charge layer dynamics at oxide-semiconductor interfaces under optical

modulation: Theory and experimentalstudies by non-contacr photocarrier radiometry, A.

Mandelis, J.Batista, J.Gibkes, J.Pelzl, J.Phys.IV France 125, 565-567 (2005)

15. Investigations of AII-BVI mixed crystals with the piezoelectric photothermal method,

M. Mali ński, L. Bychto, J. Zakrzewski, F. Firszt, M. Pawlak, P. Binnebesel,

Journal de Physique IV 125, p. 379-382 (2005)

16. Time-domain and lock-in rate-window photocarrier radiometric measurements of

recombination processes in silicon,A. Mandelis, M. Pawlak, Ch. Wang, I. Delgadillo-

Holtfort, J. Pelzl, Journal of Applied Physics 98, p.123518 (2005)

17. Photoacoustic and photothermal radiometry spectra of implanted Si wafers,

D.M.Todorovic, M. Pawlak, I. Delgadillo-Holtfort, J . Pelzl, European Physical Journal -

Special Topics 153,259-262 (2006).

18. Non-linear Dependence of Photocarrier Radiometry Signals from p-Si Wafers on Optical

Excitation Intensity, J. Tolev, A. Mandelis and M. Pawlak, J. Electrochem. Soc. J.

Electrochem. Soc., Volume 154, Issue 11, pp. H983-H994 (2007)

19. On the non-linear dependence of photocarrier radiometry signals from Si wafers on the

intensity of the laser beam, J. Tolev, A. Mandelis and M. Pawlak, Eur. Phys. J. Special

Topics 153, 317-320 (2008)

20. Laser Photothermal Radiometric Instrument For Industrial Steel Hardness Inspection, X.

Guo, K. Sivagurunathan, M. Pawlak, J. Garcia, A. Mandelis, S. Giunta, S. Milletari and

S. Bawa, conference material for 15th International Conference Of Photoacoustic And

Photothermal Phenomena in Leuven, Belgium, 2009 (July 2009)

21. Thermal Transport Properties of Cd1-xMgxSe Mixed Crystals Measured by Means of the

Photopyroelectric Method, M. Pawlak, F. Firszt, S. Łęgowski, H. Męczyńska, J. Gibkes

and J. Pelzl , Intern. Journal of Thermophysics 31, 1, 187-198 (2010)

128

Conference contributions

32nd Winter School on Molecular and Quantum Acoustics and 8th workshop on

Photoacoustics and Photothermics) w Szczyrku, Poland, (oral), 2003

Photoacoustic Study of Cd1-x-yBexMnyTe Mixed Crystals

International Symposium on 50th Anniversary of the Death of Prof. Dr. Jan Chochralski,

2003, Torun, Poland (2 posters), 2003

Growth and characterization of selected wide – gap II – VI

ternary solid solutions

XXXII International School on the Physics of Semiconducting Compounds Jaszowiec,

Poland, 2003.

XVII School of Optoelectronics, Kazimierz Dolny, Poland, 2003 (poster)

Photoelectric and Photothermal Properties of Selected II – VI

Compounds Mixed Crystals,

33rd Winter School on Molecular and Quantum Acoustics and 9th workshop on Photoacoustics

and Photothermics) w Szczyrku, Poland, (oral), 2004,

Study of optical properties of Zn1-xBexTe mixed crystals by

means of combined modulated IR radiometry and

photoacoustics

13th International Conference Of Photoacoustic And Photothermal Phenomena in Rio de

Janeiro, Brazil, 2004

Investigation Of AII BVI Mixed Crystals With The Piezoelectric

Photothermal Method, oral

Piezoelectric and Microphone Photoacoustic Study Of

BeZnCdSe Mixed Crystals, poster

Space Charge Layer Dynamics At Oxide – Semiconductor

Interface Under Optical Modulation: Theory and Experimental

Studies By Non – Concact Photocarrier Radiometry, poster

129

34th Winter School on Molecular and Quantum Acoustics and 10th workshop on

Photoacoustics and Photothermics) in Szczyrku, Poland, 2005:

Determination of the Thermophysical Parameters of Zn1-

XBeXSe Mixed Crystals by means of Standard

Photopyroelectric and Piezoelectric Photothermal Techniques,

Gordon Research Conference on "Photoacoustic and Photothermal Phenomena"

Trieste, Italy, 2005, poster:

Photocarrier Radiometric Imaging of H+ Ions Implanted in Si

wafers,

14th International Conference Of Photoacoustic And Photothermal Phenomena in Cairo,

Egypt, 2007, poster:

The Non-linear Dependence of Photocarrier Radiometry

Signals from Si Wafers on the Intensity of the Laser Beam,

Michal Pawlak, Jordan Tolev and Andreas Mandelis.

14th Winter Workshop on Photoacoustics and Thermal Waves Methods w Korbielów, Poland,

(oral), 2009.

130

Appendix A

Current controller

In order to drive a Peltier element from PPE experimental set up presented in Fig.

(4.13) the dedicated current controller had been designed. The electrical schema of this

controller is presented in Figure A.1.

131

Figure A.1. The electrical schema of the current controller.