Optical Properties and Photoelectric Effects

Similar documents
Lecture 15: Optoelectronic devices: Introduction

Review of Optical Properties of Materials

Chapter 1 Overview of Semiconductor Materials and Physics

3.1 Introduction to Semiconductors. Y. Baghzouz ECE Department UNLV

Optical Properties of Lattice Vibrations

3.1 Absorption and Transparency

The Electromagnetic Properties of Materials

KATIHAL FİZİĞİ MNT-510

Semiconductor Physical Electronics

Luminescence Process

ELECTRONIC DEVICES AND CIRCUITS SUMMARY

Sheng S. Li. Semiconductor Physical Electronics. Second Edition. With 230 Figures. 4) Springer

EE 5344 Introduction to MEMS CHAPTER 5 Radiation Sensors

Optical Properties of Solid from DFT

Chemistry Instrumental Analysis Lecture 8. Chem 4631

ET3034TUx Utilization of band gap energy

EE495/695 Introduction to Semiconductors I. Y. Baghzouz ECE Department UNLV

Introduction to Sources: Radiative Processes and Population Inversion in Atoms, Molecules, and Semiconductors Atoms and Molecules

Optical Properties of Semiconductors. Prof.P. Ravindran, Department of Physics, Central University of Tamil Nadu, India

smal band gap Saturday, April 9, 2011

PH575 Spring Lecture #20 Semiconductors: optical properties: Kittel Ch. 8 pp ; Ch 15 pp

Lecture 20 Optical Characterization 2

Chapter 5. Semiconductor Laser

Density of states for electrons and holes. Distribution function. Conduction and valence bands

PHYSICS nd TERM Outline Notes (continued)

collisions of electrons. In semiconductor, in certain temperature ranges the conductivity increases rapidly by increasing temperature

Chapter 3 Properties of Nanostructures

OPTI510R: Photonics. Khanh Kieu College of Optical Sciences, University of Arizona Meinel building R.626

CME 300 Properties of Materials. ANSWERS: Homework 9 November 26, As atoms approach each other in the solid state the quantized energy states:

Single Photon detectors

Introduction to Engineering Materials ENGR2000. Dr.Coates

Chapter 2 Optical Transitions

EE 446/646 Photovoltaic Devices I. Y. Baghzouz

Session 5: Solid State Physics. Charge Mobility Drift Diffusion Recombination-Generation

1 Name: Student number: DEPARTMENT OF PHYSICS AND PHYSICAL OCEANOGRAPHY MEMORIAL UNIVERSITY OF NEWFOUNDLAND. Fall :00-11:00

n N D n p = n i p N A

LEC E T C U T R U E R E 17 -Photodetectors

EECS130 Integrated Circuit Devices

Course overview. Me: Dr Luke Wilson. The course: Physics and applications of semiconductors. Office: E17 open door policy

sin[( t 2 Home Problem Set #1 Due : September 10 (Wed), 2008

Conductivity and Semi-Conductors

Chapter 12: Semiconductors

Electron Energy, E E = 0. Free electron. 3s Band 2p Band Overlapping energy bands. 3p 3s 2p 2s. 2s Band. Electrons. 1s ATOM SOLID.

PHOTOVOLTAICS Fundamentals

LN 3 IDLE MIND SOLUTIONS

Computer modelling of Hg 1 x Cd x Te photodiode performance

The German University in Cairo. Faculty of Information Engineering & Technology Semiconductors (Elct 503) Electronics Department Fall 2014

Semiconductor Physical Electronics

Uniform excitation: applied field and optical generation. Non-uniform doping/excitation: diffusion, continuity

MTLE-6120: Advanced Electronic Properties of Materials. Intrinsic and extrinsic semiconductors. Reading: Kasap:

Semiconductor device structures are traditionally divided into homojunction devices

Basic cell design. Si cell

SEMICONDUCTOR PHYSICS REVIEW BONDS,

Classification of Solids

8.1 Drift diffusion model

Unit IV Semiconductors Engineering Physics

Lecture 7: Extrinsic semiconductors - Fermi level

Introduction to Optoelectronic Device Simulation by Joachim Piprek

The photovoltaic effect occurs in semiconductors where there are distinct valence and

Chapter 6: Light-Emitting Diodes

Electronic and Optoelectronic Properties of Semiconductor Structures

Engineering 2000 Chapter 8 Semiconductors. ENG2000: R.I. Hornsey Semi: 1

February 1, 2011 The University of Toledo, Department of Physics and Astronomy SSARE, PVIC

Semiconductor Physics Problems 2015

This is the 15th lecture of this course in which we begin a new topic, Excess Carriers. This topic will be covered in two lectures.

nano.tul.cz Inovace a rozvoj studia nanomateriálů na TUL

Physics of Semiconductor Devices. Unit 2: Revision of Semiconductor Band Theory

Basic Semiconductor Physics

ESE 372 / Spring 2013 / Lecture 5 Metal Oxide Semiconductor Field Effect Transistor

SEMICONDUCTOR PHYSICS

Transport Properties of Semiconductors

Chapter 4. Photodetectors

ECE 250 Electronic Devices 1. Electronic Device Modeling

Semiconductor Physics fall 2012 problems

Direct and Indirect Semiconductor

Chapter 7. Solar Cell

THE UNIVERSITY OF NEW SOUTH WALES SCHOOL OF PHYSICS FINAL EXAMINATION JUNE/JULY PHYS3080 Solid State Physics

Solar cells operation

Practical 1P4 Energy Levels and Band Gaps

Carriers Concentration and Current in Semiconductors

Bohr s Model, Energy Bands, Electrons and Holes

Fall 2014 Nobby Kobayashi (Based on the notes by E.D.H Green and E.L Allen, SJSU) 1.0 Learning Objectives

Semiconductor Devices and Circuits Fall Midterm Exam. Instructor: Dr. Dietmar Knipp, Professor of Electrical Engineering. Name: Mat. -Nr.

Misan University College of Engineering Electrical Engineering Department. Exam: Final semester Date: 17/6/2017

Carrier Recombination

ECE 606 Homework Week 7 Mark Lundstrom Purdue University (revised 2/25/13) e E i! E T

UvA-DARE (Digital Academic Repository) Charge carrier dynamics in photovoltaic materials Jensen, S.A. Link to publication

Basics of Infrared Detection

Solar Cell Materials and Device Characterization

FIBER OPTICS. Prof. R.K. Shevgaonkar. Department of Electrical Engineering. Indian Institute of Technology, Bombay. Lecture: 12.

Semiconductors. SEM and EDAX images of an integrated circuit. SEM EDAX: Si EDAX: Al. Institut für Werkstoffe der ElektrotechnikIWE

Practical 1P4 Energy Levels and Band Gaps

Chapter 6. Fiber Optic Thermometer. Ho Suk Ryou

Chalcogenide semiconductor research and applications. Tutorial 2: Thin film characterization. Rafael Jaramillo Massachusetts Institute of Technology

Lecture 12. Semiconductor Detectors - Photodetectors

Electrons are shared in covalent bonds between atoms of Si. A bound electron has the lowest energy state.

- Outline. Chapter 4 Optical Source. 4.1 Semiconductor physics

UNIVERSITY OF CALIFORNIA College of Engineering Department of Electrical Engineering and Computer Sciences. EECS 130 Professor Ali Javey Fall 2006

UNIT I: Electronic Materials.

Luminescence basics. Slide # 1

Transcription:

9 Optical Properties and Photoelectric Effects 9.1. Introduction This chapter presents the fundamental optical properties and bulk photoelectric effects in a semiconductor. The optical properties associated with the fundamental and free-carrier absorption processes and the internal photoelectric effects such as photoconductive (PC), photovoltaic (PV), and photomagnetoelectric (PME) effects in a semiconductor are depicted. Important fundamental physical and electronic properties such as the energy band structures, excess carrier phenomena, and recombination mechanisms can be understood by studying the optical absorption processes and photoelectric effects in a semiconductor. Many practical applications have been developed using the internal photoelectric effects such as PV and PC effects in semiconductors. Future trends are moving toward more developments of various optoelectronic devices for a wide variety of applications in PV devices (solar cells), light-emitting diodes (LEDs) and laser diodes (LDs), and the optoelectronic integrated circuits (OEICs) for use in optical computing, optical communications, signal processing, and data transmission. In fact, many photonic devices such as solar cells, LEDs, LDs, and photodetectors using silicon, II-VI, and III-V compound semiconductors have been developed for commercial use in power generation, light sources and signal displays, optical communications and data transmission, and signal processing systems. Depending on the energy of incident photons, there are two types of optical absorption processes that may occur in a semiconductor. The first type of optical absorption process involves the absorption of photons, which have energies equal to or greater than the band gap energy of a semiconductor. This type of optical absorption is called the fundamental or interband absorption process. The fundamental absorption process is usually accompanied by an electronic transition across the forbidden gap, and as a result excess electron hole pairs are generated in the semiconductor. The absorption coefficient due to the interband transition is usually very large. For example, in the ultraviolet (UV) to visible spectral range, typical values of the absorption coefficient for most semiconductors are in the range from 10 6 cm 1 near the UV wavelength to around 1 cm 1 near the cutoff wavelength of the semiconductor. However, the absorption coefficient becomes 246

9.2. Optical Constants of a Solid 247 very small (e.g., less than 1 cm 1 ) when the photon energies fall below the band gap energy of the semiconductor. In this case, another type of optical absorption process takes place in the semiconductor. This type of optical absorption results only in electronic transitions within the allowed energy band, and is called the free-carrier absorption process. The fundamental absorption process, which leads to an interband transition, must be treated quantum mechanically, while the freecarrier absorption process can be described by the classical electromagnetic (EM) wave theory. Finally, absorption of photons with energies below the band gap energy of the semiconductor may also lead to electronic transitions from localized impurity states to the conduction or valence band states. For example, the extrinsic photoconductivity observed at low temperatures is due to the photoexcitation of free carriers from the shallow-impurity states to conduction or valence band states. Since the energy band gap varies between 0.1 and 6.0 ev for most semiconductors, the fundamental optical absorption could occur in the UV, visible, and infrared (IR) spectral regimes (i.e., 0.3 to 10 µm). Therefore, most semiconductors are opaque from the UV to the IR spectral range, and become transparent in the IR spectral regime for λ>10 µm. In order to better understand the optical absorption processes in a semiconductor, it is necessary to first consider two optical constants, namely, the index of refraction and the extinction coefficient. These two optical constants may be derived by solving the Maxwell wave equations for the EM waves in a solid, as will be depicted in Section 9.2. The free-carrier absorption process is depicted in Secton 9.3. Section 9.4 deals with the fundamental absorption process in a semiconductor. The internal photoelectric effects such as PC, PV, and PME effects in a semiconductor are described in Sections 9.5, 9.6, and 9.7, respectively. Section 9.5 presents both the intrinsic and extrinsic PC effects in a semiconductor. The internal PV effect also known as the Dember effect is depicted in Section 9.6. The PME effect in a semiconductor is presented in Section 9.7. 9.2. Optical Constants of a Solid The optical constants such as the index of refraction and extinction coefficient may be derived from solving the Maxwell equations for EM waves propagating in a solid. It is well known that some solids are transparent while others are opaque, that some solid surfaces are strongly reflective while others tend to absorb optical radiation that falls on them. The optical absorption depends on the wavelength of the incident optical radiation. For example, most semiconductors show strong absorption from UV (λ <0.4 µm), visible to near-ir (0.4 <λ<2 µm), mid-wavelength IR (3 5 µm), and long-wavelength IR (8 12 µm), and become transparent in the far-ir (λ >14 µm) spectral regime. Therefore, in order to obtain a better understanding of the optical absorption process in a semiconductor, it is important to derive the expressions of the two basic optical constants (i.e., index of refraction and extinction coefficient) in the UV to IR spectral range.

248 9. Optical Properties and Photoelectric Effects The propagation of EM waves in a solid can be described by the Maxwell wave equations, which are given by ε = 0, (9.1) ε = B t, (9.2) B = 0, (9.3) ε H = σε+ ε 0 ε s t. (9.4) In free space, the EM wave equation can be obtained from solving Eqs. (9.1) through (9.4) by setting B = µ 0 H,σ = 0, and ε s = 1, which yields 2 2 ( ε 1 ε = µ 0 ε 0 t = 2 c 2 ) 2 ε t 2, (9.5) where c = 1/ ε 0 µ 0 is the speed of light in free space. Inside the solid, the wave equation can also be obtained by solving Eqs. (9.1) through (9.4), and the result yields 2 2 ε ε = µ 0 ε 0 ε s t + µ 0σ ε 2 t. (9.6) A comparison of Eqs. (9.5) and (9.6) shows that the difference between the waves propagating in free space and in a solid is due to the difference in the dielectric constant and electrical conductivity in both media. It is clear that Eq. (9.6) will reduce to Eq. (9.5) if the dielectric constant ε s equals 1 and the electrical conductivity σ is zero. The first term on the right-hand side of Eq. (9.6) is the displacement current density, while the second term represents the conduction current density. An EM wave with frequency ω propagating in the z-direction and polarizing in the x-direction can be expressed by [ ( z )] ε x = ε 0 exp iω v t = ε 0 exp [ i ( k z ωt )], (9.7) where k is the complex wave vector and v is the velocity of the EM waves inside the solid which could also be a complex number. It is noted that k and v are related by k = ω v (9.8) Now by substituting Eq. (9.7) into Eq. (9.6), one obtains or where k 2 = ω2 v 2 = µ 0ε 0 ε s ω 2 + iµ 0 σω (9.9) k = ω ( ω ) ( v = ε s + iσ ) 1/2 = c ωε 0 n = ( ω c ) n (9.10) ( ε s + iσ ) 1/2 = εs 1/2 (9.11) ωε 0

9.2. Optical Constants of a Solid 249 is a complex refractive index of the solid and εs is the complex dielectric constant; complex refractive index n can be expressed by n = n + ik e (9.12) where n is the index of refraction of the media and k e is the extinction coefficient, which is a constant relating the attenuation of the incident EM wave inside the solid to its penetration depth. For example, if an incident EM wave propagates into a solid at a distance equal to one wavelength in free space (i.e., λ 0 = 2πc/ω), then its amplitude is decreased by a factor of e 2πk e where k e is the extinction coefficient of the solid. It is noted that the fundamental optical absorption coefficient α is related to k e by α = 4πk e /λ, aswill be shown later. Solving Eqs. (9.11) and (9.12), one obtains the real and imaginary parts of the complex refractive index, which are given respectively by n 2 ke 2 = ε s (9.13) 2nk e = σ (9.14) ωε 0 Thus, the optical properties of a solid, as observed macroscopically, may be described in terms of the complex refractive index n.now substituting Eqs. (9.10) and (9.12) into Eq. (9.7), the solution for the EM waves inside the solid becomes ( ) ke ωz [ ( nz )] ε x = ε 0 exp exp iω c c t (9.15) which shows that the speed of incident electric wave in a solid is reduced by a factor of n (n is the refractive index), and its amplitude decreases exponentially with distance. The attenuation of incident electric waves is associated with the absorption of EM energy by the dissipating medium. However, the optical constant commonly measured in a solid is not the extinction coefficient k e,but the absorption coefficient α. The optical absorption coefficient is related to the Poynting vector of the EM wave energy flow by S (z) = S 0 e αz (9.16) where S(z)isthe Poynting vector which is proportional to the amplitude square of the electric waves (i.e., εx 2 )given by Eq. (9.15). Thus, from Eq. (9.15) to Eq. (9.16), one obtains the optical absorption coefficient α = 2k eω c = 4πk e λ 0 (9.17) where λ 0 is the wavelength of the EM waves in free space. Thus, the extinction coefficient k e can be determined from the optical absorption coefficient α of the semiconductor. It is noted that both the real (n 2 k 2 e ) and imaginary (2nk e) parts of the complex refractive index n are the quantities measured in a solid. In practice, (n 2 k 2 e ) and 2nk e can be obtained by measuring the reflection and transmission coefficients of a solid.

250 9. Optical Properties and Photoelectric Effects Figure 9.1. An eletromagnetic wave propagating into a solid under normal incidence. To derive the reflection coefficient in a solid let us consider the case of normal incidence, as shown in Figure 9.1. If ε x (H y ) and ε x (H y ) denote the incident and reflected electric (magnetic) waves, and ε x (H y )isthe transmitted electric (magnetic) wave into the solid in the z-direction, then the transmitted wave for z > 0 can be expressed by [ ( n ε x )] = ε z 0 exp iω c t (9.18) For z < 0 (i.e., in free space), the electric waves are composed of the incident and reflected waves, which can be expressed by [ ( z )] [ ( z )] ε x = ε 1 exp iω c t + ε 2 exp iω c + t (9.19) The magnetic wave components polarized in the y-direction (i.e., H y ) may be related to the electric wave components in the x-direction by the characteristic impedance of the medium, which are given by ε x µ0 ε x µ = = Z 0, = = Z ε x µ0, = = Z H y ε 0 H y ε 0 ε s H y 0 (9.20) ε 0 Equation (9.20) relate the incident, transmitted, and reflected EM waves to the characteristic impedance Z 0 and Z in the free space and in the solid. The boundary conditions at z = 0 plane requires that the tangential components of both ε x and H y be continuous. Thus, one can write ε x = ε x + ε x and H y = H y + H y (9.21) Now solving Eqs. (9.20) and (9.21) yields ε x = Z Z 0 ε0 /µ 0 ε 0 ε s /µ = ε x Z + Z 0 ε0 /µ 0 + (9.22) ε 0 ε s /µ From Eq. (9.20) one obtains H y H y = ε x ε x = Z Z 0 Z + Z 0 (9.23)

9.2. Optical Constants of a Solid 251 Since the Poynting vector is equal to the product of the electric and magnetic field strengths (i.e., S =ε x H y, S =ε x H y ), the reflection coefficient R can be obtained from Eqs. (9.22) and (9.23) using the definition R = S /S, and the result yields R = S ( ε )( S = x H ) ( y Z ) Z 2 0 = ε x H y Z + Z 0 [ ε0 /µ 0 ] 2 ( ε 0 ε s /µ = 0 n ) ε0 /µ 0 + n 2 0 = (9.24) ε 0 ε s /µ 0 n + n 0 For nonmagnetic materials, µ = µ 0,ε = ε 0 ε s, and n = n + ik e ; for free space, n 0 = 1. Thus, the absolute value of the reflection coefficient for normal incidence can be written as R = (n 1) 2 + k 2 e (n + 1) 2 + ke 2 (9.25) where n and k e are the index of refraction and the extinction coefficient of the solid, respectively. The transmission coefficient T, defined as the ratio of the transmission power and the incident power, can be derived in a similar way as that of the reflection coefficient depicted above or by using the relation that T = 1 R, which yields ε T = x H y ε x H = 4Z 0Z y (Z + Z 0 ) 2 = 4n 0n (9.26) (n 0 + n ) 2 Since n 0 = 1 for free space, the absolute value of the transmission coefficient can be obtained from Eq. (9.26), which yields 4n T = (9.27) (n + 1) 2 + ke 2 For normal incidence, it is seen from Eqs. (9.25) and (9.27) that by measuring T and R one can determine both n and k e.however, for incident angles other than normal incidence, the reflection coefficient will, in general, depend on the polarization, and from observation of different angles of incidence, both n and k e values can be determined if k e is not too small. If both n and k e values are large, then R will approach unity. An inspection of Eq. (9.13) reveals that the dielectric constant ε s can also be determined directly from the refractive index n, provided that k e is much smaller than unity. Values of n can be found directly from measurements of the reflection coefficient if k e is very small. There is considerable practical interest in measuring the transmission and reflection coefficients in free space under normal incidence by using a thin plane-parallel sheet of crystal with refractive index n and thickness d.ifi 0, I t, and I r denote the incident, transmitted, and reflected wave intensities through the thin specimen, then the normalized transmitted and reflected wave intensities can be expressed,

252 9. Optical Properties and Photoelectric Effects respectively, by I t = (1 ( R)2 e αd 1 + k 2 e /n 2) (9.28) I 0 1 R 2 e 2αd I r = R ( ) 1 e 2αd (9.29) I 0 1 R 2 e 2αd Equations (9.28) and (9.29) show that both n and k e can be found by measuring I t and I r.for most transmission experiments, it is valid to assume that ke 2 n2.if the sample thickness d is chosen such that R 2 e 2αd 1, then Eq. (9.28) becomes I t = (1 R) 2 e αd (9.30) I 0 From Eq. (9.30), it is noted that the optical absorption coefficient α of a semiconductor near the bandedge can be determined by measuring the transmission coefficient as a function of wavelength on two thin samples of different thickness without the knowledge of reflectance. This is valid as long as both samples have the same reflection coefficients at the front surface of the sample. For elemental semiconductors such as Si and Ge, the main contribution to the dielectric constant arises from electronic polarization. However, in compound semiconductors (such as III-V and II-VI compounds), both electronic and ionic polarizations can contribute to the dielectric constant. The increase in the degree of ionicity in these compounds relative to the group IV elements will lead to a significant difference between the static and optical (high-frequency) dielectric constants. The high-frequency dielectric constant εs is equal to n 2. The static dielectric constant ε s can be calculated using the relation ( ) 2 ε s = εs ωl (9.31) where ω l and ω t are the longitudinal- and transverse-mode optical phonon frequencies, respectively. Table 9.1 lists values of dielectric constants and refractive indices for Si, Ge, and some III-V and II-VI compound semiconductors. ω t 9.3. Free-Carrier Absorption Process When the energy of incident EM radiation is smaller than the band gap energy (i.e., hν E g )ofthe semiconductor, excitation of electrons from the valence band into the conduction band will not occur. Instead, the absorption of incident EM radiation will result in the excitation of lattice phonons and the acceleration of free electrons inside the conduction band. In the conduction band, free-carrier absorption is proportional to the density of conduction electrons. Since free-carrier absorption involves only electronic transitions within the conduction band, one can apply the classical equation of motion to deal with the interaction between the EM waves and the conduction electrons. The equation of motion for an electron due to

9.3. Free-Carrier Absorption Process 253 Table 9.1. Refractive indices and dielectric constants for Si, Ge, and some III-V and II-VI semiconductors. Materials n ε s ε s Si 3.44 11.8 11.6 Ge 4.00 16 15.8 InSb 3.96 17 15.9 InAs 3.42 14.5 11.7 GaAs 3.30 12.5 10.9 GaP 2.91 10 8.4 CdS 2.30 8.6 5.2 CdSe 2.55 9.2 6.4 CdTe 2.67 9.7 7.1 ZnS 2.26 8.1 5.1 ZnSe 2.43 8.7 5.9 a time-varying electric wave (i.e., ε 0 e iωt )offrequency ω and propagating in the z-direction is given by ( m 2 z m ) z t + 2 τ t = qε 0 e iωt (9.32) where τ is the relaxation time and m is the effective mass of electrons in the conduction band. The solution of Eq. (9.32) is given by z = (qε 0/m ) e iωt ( ) (9.33) iω/τ ω 2 If the electron density in the conduction band is equal to N 0, then the total polarization P, which is equal to the product of displacement z and electron density N 0, can be expressed by P = qn 0 z (9.34) The polarizability p, which is defined as the polarization per unit electric field, can be written as p = qn 0z (9.35) ε 0 The complex dielectric constant εs given by Eq. (9.11) is related to the polarizability p by ( N0 q 2 /m ε 0 ) εs = ε s iε s = n 2 = ε s + p = ε s + ( ε ) (9.36) 0 iω/τ ω 2 Note that the second term on the right-hand side of Eq. (9.36) is due to the contribution of free-carrier absorption. Thus, from Eq. (9.36), the real and imaginary parts of the complex dielectric constant can be written as ε s = n2 ke 2 = ε τσ 0 s ( ε 0 1 + ω2 τ 2) (9.37)

254 9. Optical Properties and Photoelectric Effects and ε s = 2nk σ 0 e = ( ωε 0 1 + ω2 τ 2) (9.38) where σ 0 = N 0 q 2 τ/m is the dc electrical conductivity. In Eqs. (9.37) and (9.38), it is assumed that τ is a constant and is independent of energy. Solving Eqs. (9.17) and (9.38) one obtains the optical absorption coefficient as α = 4πk e σ 0 = ( λ 0 ncε 0 1 + ω2 τ 2) (9.39) which shows the frequency dependence of the optical absorption coefficient. Two limiting cases for ωτ 1 and ωτ 1, are discussed next. (i) Long-wavelength limit (ωτ 1). In this case, the absorption coefficient given by Eq. (9.39) becomes α = σ 0 (9.40) ncε 0 The real part of the dielectric constant in Eq. (9.37) is reduced to ε s = ε s τσ 0 (9.41) ε 0 Equation (9.40) shows that the absorption coefficient is independent of frequency, but depends on temperature through σ 0.For example, for an n-type germanium sample with ωτ = 10 12 s this corresponds to a wavelength of about 2 mm. For alightly doped semiconductor with large dielectric constant, the contribution of τσ 0 /ε 0 in Eq. (9.41) to the real part of the dielectric constant ε s is quite small, and hence ε s is equal to the dielectric constant of the semiconductor, ε s. For heavily doped semiconductors with large σ 0,value of τσ 0 /ε 0 becomes much larger than ε s and hence ε s becomes negative. This corresponds to the metallic case. If one assumes ε s τσ 0 /ε 0, then solving Eqs. (9.41) and (9.37) one obtains the real part of the dielectric constant as ε s = n2 ke 2 = τσ 0 (9.42) ε 0 Similarly, from Eq. (9.38) one obtains the imaginary part of the dielectric constant as ε s = 2nk e = σ 0 (9.43) ωε 0 Now, solving Eqs. (9.42) and (9.43) yields ( ) n 2 ke 2 ωτ = (9.44) 2nk e

9.3. Free-Carrier Absorption Process 255 From Eq. (9.44) for ωτ 1, k e n. Thus, by setting n = k e in Eq. (9.43), the refractive index is given by ( ) 1/2 σ0 n = (9.45) 2ωε 0 The absorption coefficient can be deduced from Eqs. (9.40) and (9.45), and the result yields ( ) 2σ0 ω 1/2 α = (9.46) ε 0 c 2 which shows that α is proportional to the square root of the frequency. Therefore, in this case, the material exhibits metallic behavior. This corresponds to the well-known skin effect in which the penetration depth (δ) ofthe incident EM wave is inversely proportional to the square root of frequency and electrical conductivity. (ii) Short-wavelength limit (ωτ 1). This usually occurs in the wavelength regime extending upward from far-ir toward the fundamental absorption edge of the semiconductor. The short wavelength free-carrier absorption becomes negligible when the photon energy exceeds the band gap energy of the semiconductor. To understand the free-carrier absorption process in the short-wavelength limit, one can solve Eqs. (9.37) to (9.39), and the results yield ε s = ε s σ 0 ε 0 ω 2 τ = n2 ke 2 (9.47) ε x = 2nk e = σ 0 (9.48) ε 0 ω 3 τ 2 σ 0 α = ncε 0 ω 2 τ = N 0 q 3 λ 2 0 (9.49) 2 4π 2 c 3 m 2 µnε 0 Equation (9.49) shows that, for ωτ 1, the absorption coefficient is directly proportional to the square of the wavelength, which was observed in a number of semiconductors. Figure 9.2 shows the absorption coefficient versus square of the wavelength for two n-type InSb specimens with different dopant concentrations. (1) The results are in good agreement with the prediction given by Eq. (9.49). It is seen from Eq. (9.47) that ε s changes sign from positive to negative value as ω decreases. The condition for which ε s = 0 corresponds to the total internal reflection, and the frequency for this to occur is called the plasma resonance frequency, ω p. Solving Eq. (9.47) one obtains ( ) 1/2 ( σ0 N0 q 2 ) 1/2 ω p = = (9.50) ε 0 ε s τ m ε 0 ε s where ω p is the frequency in which the classical undamped plasma of free electrons exhibits its normal mode of oscillation. For a germanium sample with N 0 = 10 16 cm 3, m = 0.12 m 0, and ε s = 16, one finds that ω p is equal to 2 10 2 GHz, which falls into the rather difficult millimeter-wavelength regime. In order

256 9. Optical Properties and Photoelectric Effects Figure 9.2. Optical absorption coefficient versus the square of the wavelength for two InSb specimen with different doping densities over the wavelength range in which free-carrier absorption is dominant. After Moss, (1) by permission. to observe plasma resonance in the microwave frequency range, one should use an ultrapure semiconductor specimen for the experiment. Otherwise the experiment must be performed at extremely low temperature. For a germanium crystal, a carrier concentration of 10 13 cm 3 or less is required for the plasma resonance to be observed in the microwave-frequency range. For metals, since the electron concentration is very high (i.e., 10 22 cm 3 ), the plasma resonance frequency usually falls in the solar blind UV spectral range, which corresponds to photon energies of 10 20 ev. Free-carrier absorption has been used extensively in determining the relaxation time constant and the conductivity effective mass of electrons in a semiconductor. 9.4. Fundamental Absorption Process The fundamental absorption process takes place when photons with energies greater than the band gap energy of the semiconductor (i.e., hν E g ) are absorbed in a semiconductor. This process usually results in the generation of electron hole pairs in a semiconductor. For most semiconductors, the fundamental absorption process may occur in the UV, visible, and IR wavelength regimes. It is the most important optical absorption process because important photoelectric effects are based on such absorption processes to generate excess electron hole pairs in a semiconductor. There are two types of optical transitions associated with the fundamental absorption process, namely, direct and indirect band-to-band transitions, as shown in Figures 9.3a and b. In a direct transition only one photon is involved, while in an

9.4. Fundamental Absorption Process 257 Figure 9.3. Direct and indirect transitions associated with the fundamental absorption processing in a semiconductor. indirect transition additional energy is supplied or released in the form of phonons. The absorption coefficients associated with these two transition processes depend on the probability per unit time that an electron makes a transition from the valence band into the conduction band when an incident photon is absorbed. The transition probability P i can be calculated by using the first-order time-dependent perturbation theory, which is given by P i = ( 2π h ) M if 2 g n (E) (9.51) where P i is the transition probability per unit time from the initial state k i in the valence band to the final state k f in the conduction band, M if denotes the matrix element due to perturbation which connects the initial states k i and the final states k f of the system, and g n (E) isthe density of final states in the conduction band. In the present case, the perturbation is due to incident EM radiation, and the matrix element corresponding to the electric dipole transition is given by M if = ψi rψ f d 3 r = u v (k i, r)e iki r r u c (k f, r)e ikf r d 3 r (9.52) = u v (k i, r) r u c (k f, r)e i(k f k i ) r d 3 r + ik f u v (k f, r)u c (k f, r)e i(k f k i ) r d 3 r where ψ i and ψ f denote the electron wave functions in the valence and conduction bands, respectively; u v (k i, r) and u c (k f, r) are the Bloch functions for the valence and conduction bands, respectively. Both terms on the right-hand side of Eq. (9.52)

258 9. Optical Properties and Photoelectric Effects contain the factor e i(k f k i ) r, which oscillates rapidly. Thus, the integrand of Eq. (9.52) will vanish unless k f = k i (9.53) which is the condition of momentum conservation for such a transition. In fact, the contribution of photon momentum in Eq. (9.53) is negligible because it is very small compared to the crystal momentum. It is noted that the first term on the right-hand side of Eq. (9.52) is known as the allowed transition; its value is real and independent of the wave vector k f of the final state. The second term on the right-hand side of Eq. (9.52) is imaginary, and it depends on the wave vector of the final state k f.transitions associated with the second term are called the forbidden transitions. Since the absorption coefficient is directly related to the rate of transition probability P if, Eq. (9.52) can be employed to derive the absorption coefficient for the direct and indirect interband transitions taking place between the valence band and the conduction band of a semiconductor. Author: should this be g n (E) as given in Eq. (9.51)? 9.4.1. Direct Transition Process The direct (or vertical) transition shown in Figure 9.3a is the dominant absorption process taking place in a direct band gap semiconductor when the conduction band minimum and the valence band maximum are located at the same k-value in the reciprocal space (i.e., typically at the Ɣ-point of the Brillouin zone center). In order to derive an expression of the absorption coefficient near the conduction band minimum, it is necessary to find the density of final states g(e n )ineq. (9.51). It is noted that electron energy in the conduction band can be expressed by E n = E c + h2 k 2 2m n (9.54) and in the valence band by E p = E v h2 k 2 2m p (9.55) The photon energy corresponding to such a verical transition can be written as hν = E n E p = E g + h2 k 2 2m r (9.56) where E g is the band gap energy of the semiconductor and m r = m n m p /(m n + m p ) is the reduced electron effective mass. Equations (9.54) through (9.56) enables one to express the density of final states for the conduction band with a parabolic band structure as ( ) 4π g(e n ) = (2m r )3/2 (hν E g ) 1/2 (9.57) h 3

9.4. Fundamental Absorption Process 259 Figure 9.4. Direct transition in a p-type GaAs specimen with N A = 10 17 cm 3 and an absorption coefficient greater than 9 10 3 cm 1. The threshold energy is (1.39 ± 0.02) ev. After Kudman and Seidel, (2) by permission. From Eqs. (9.52) and (9.57), the absorption coefficient for a direct allowed transition can be expressed by α a d = K a d (hν E g) 1/2 (9.58) In the direct allowed transitions, the square of the matrix element (i.e., M if 2 )is independent of the wave vector, and hence Kd a is a constant and is independent of electron energy. Equation (9.58) shows that for an allowed direct optical transition, the optical absorption coefficient αd a varies as (hν E g) 1/2. Therefore, a plot of αd 2 versus hν near the fundamental absorption edge enables one to determine the band gap energy of a semiconductor. This is illustrated in Figure 9.4 for a p-type GaAs (2) ; the intercept of this plot with the horizontal axis yields the band gap energy of GaAs. In the forbidden direct transitions, as given by the second term of Eq. (9.52), the matrix element M if is proportional to k f, and hence the optical absorption coefficient in this case is given by α f d = K f d (hν E g) 3/2 (9.59) The energy dependence ( E 3/2 )ofα f d given by Eq. (9.59) is due to the fact that the transition probability for the direct forbidden transitions varies with the product of kf 2 ( E = (hν E g)) and the density-of-states function ( E 1/2 ). 9.4.2. Indirect Transition Process For an indirect bangap semiconductor, the conduction band minimum and the valence band maximum are not located at the same k-value in the reciprocal space. Therefore, the indirect optical transition induced by photon absorption is

260 9. Optical Properties and Photoelectric Effects usually accompanied by the simultaneous absorption or emission of a phonon. As illustrated in Figure 9.3b, conservation of momentum in this case is given by k f = k i ± q (9.60) Where k f and k i denote the wave vectors of the final and initial states of electrons, respectively, and q is the phonon wave vector. The plus sign in Eq. (9.60) corresponds to the phonon emission and the minus sign is for the phonon absorption. The conservation of energy for the indirect optical transitions requires that and hν = E n E p ± hω q = E g + h2 (k n k c ) 2 2m n + h2 k 2 p 2m p E n = E c + h2 (k n k c ) 2 2m n ± hω q (9.61) (9.62) E p = E v h2 kp 2 (9.63) 2m p where E n is the electron energy in the conduction band, E p is the electron energy in the valence band, and (k n k c ) k c. It is noted from Eq. (9.61) that in an indirect optical transition the conservation of energy is accompanied by the emission or absorption of a lattice phonon. The plus sign in Eq. (9.61) is for the phonon emission and the minus sign is for the phonon absorption; k p denotes the initial state in the valence band, and k c is the state at the conduction band minimum, k n is the final state in the conduction band. Now consider the case in which transition from the k p state in the valence band is induced by a photon with energy hν. The density of states in the valence band can be described by g ν (E p ) = A ν Ep 1/2, (9.64) where E p = E ν E, E is a small energy interval in the valence band in which transitions can take place, and A ν = (4π/h 3 )(2m p )3/2. The density of final conduction band states involving phonon absorption is given by g c (E n ) = A c (E n E c ) 1/2 = A c (hν E g E p + hω q ) 1/2 = A c ( E E p ) 1/2, (9.65) where A c = (4π/h 3 )(2m n )3/2. Equation (9.65) is obtained by solving Eqs. (9.61) through (9.64). In Eq. (9.65), the relation hν = (E g ± hω q + E) isused in the derivation. Therefore, the total effective density of states for transitions involving absorption and emission of a phonon can be expressed by E E g(hν) = g c (E n )g v (E p )de p = A c A v ( E E p ) 1/2 Ep 1/2 de p 0 = K a i E 2 = K a i (E v E p ) 2 = K a i (hν E g ± hω q ) 2, (9.66) 0

9.4. Fundamental Absorption Process 261 where the plus sign denotes phonon absorption and the minus sign is for phonon emission. Note that the integral on the right-hand side of Eq. (9.66) is carried out by letting u = Ep 1/2 so that E 0 = 2 = 2 ( E E p ) 1/2 Ep 1/2 de p E 1/2 0 u( E u 2 ) 1/2 du { u 4 ( E u2 ) 3/2 + E 8 [ u( E u 2 ) 1/2 + E sin 1 ( u )]} E 1/2 E 1/2 = π E2 = π(hν E g + hω q ) 2. (9.67) 8 8 The probability of phonon absorption and phonon emission is directly proportional to the average phonon density, which is given by n q = ( e hω q/k BT 1 ) 1. (9.68) Combining Eqs. (9.66) and (9.68) one obtains the optical absorption coefficient due to the indirect transitions with phonon absorption as ( ) 2 hν Eg + hω q α ia = n q g(hν) = K ia ( e hω q /k BT 1 ) (9.69) Similarly, for transitions involving the phonon emission, the optical absorption coefficient can be expressed by ( ) 2 hν Eg hω q α ie = n q + 1 g(hν) = K ie ( ). (9.70) 1 e hω q /k B T Now combining Eqs. (9.69) and (9.70), the optical absorption coefficients for the indirect allowed transitions involving both the emission and absorption of a phonon can be written as 0 α i = α ie + α ia { (hν Eg hω q ) 2 = K i ( ) + (hν E } g + hω q ) 2 ( 1 e hω q /k B T e hω q /k BT 1 ) (9.71) The first term in Eq. (9.71) is due to phonon emission, while the second term is attributed to phonon absorption. It is clearly shown in Figure 9.5 that the optical absorption coefficient curve for the indirect allowed transitions involving phonon absorption will extend to longer wavelengths than those associated with phonon emission. The optical absorption coefficient for the indirect allowed transitions varies with the square of photon energy. A plot of α 1/2 i versus hv at different temperatures should yield a straight line, and its intercept with the horizontal axis allows one to determine the phonon energy and the energy band gap of a semiconductor.

262 9. Optical Properties and Photoelectric Effects Figure 9.5. α 1/2 i versus photon energy hν for indirect optical transitions with temperature as a parameter. Note that T 1 > T 2 > T 3 > T 4. Figure 9.5 shows the plot of α 1/2 i versus hν involving the emission and absorption of a phonon for four different temperatures. According to Eq. (9.70), two straight-line segments can be observed in the α 1/2 i versus hν plot. For small photon energy, only α ia (i.e., associated with phonon absorption) contributes, and the α 1/2 ia versus hν plot intersects the axis at hν = E g hω q.forhν >E g + hω q,α ie becomes dominant at lower temperatures. Since the intersection of α 1/2 ie versus hν occurs at hν = E g + hω q, one can determine both the energy band gap and the phonon energy from this plot. gure 9.6 shows the square root of the absorption coefficient versus photon energy near the fundamental absorption edge of a germanium crystal with temperature as a parameter. (3) The results show that the phonon emission process becomes dominant for T < 20 K. Several effects could influence the accuracy of determining the band gap energy from the optical absorption measurements in a semiconductor. The first effect is due to the Burstein shift in a degenerate semiconductor. In a heavily doped n-type semiconductor, the Fermi level lies inside the conduction band. Therefore, in order for photon-generated electrons to make transitions from the valence band into the conduction band, the photon energy must be greater than the band gap energy of the semiconductor so that electrons can be excited into the empty states above the Fermi level in the conduction band. This shifts the optical absorption edge to a higher energy with increased doping concentration. This problem is particularly severe for small-band gap semiconductors such as InSb and InAs, since the electron effective masses and densities of states in the conduction band are small for these materials. In calculating the Burstein shift, the effect of energy-band nonparabolicity should also be considered. The second effect is related to the formation of impurity band-tail states (or impurity bands) arising from high concentration of shallow

9.5. Fundamental Absorption Process 263 Figure 9.6. Square root of the absorption coefficient versus photon energy for a germanium specimen with temperature as a parameter. The inserts show the spectral resolution. After Macfaclane et al., (3) by permission. impurities or defects, which can merge into the conduction band (for n-type) or the valence band (for p-type). This effect will result in an exponential absorption edge in the semiconductor. The third effect is associated with exciton formation in the semiconductor. An exciton is an electron hole pair bound together by Coulombic interaction. Excitons may be free, bound, or constrained to a surface, or associated with a defect complex. The binding energies for excitons are slightly below the conduction band edge, and hence exciton features are sharp peaks just below the absorption edge. Excitons are usually observed at low temperatures and become dissociated into free carriers at room temperature. It is seen that the absorption coefficient increases rapidly above the fundamental absorption edge (i.e., hν E g ). In the visible spectral range, values of the absorption coefficient for most semiconductors may vary from 10 3 to 10 6 cm 1. In general, the magnitude of the absorption coefficient represents the degree of interaction between the semiconductor and the incident photons. The internal photoelectric effects in a semiconductor are closely related to the optical absorption coefficient. Experimental results of absorption coefficient versus photon energy for some elemental and compound semiconductors (i.e., Si, Ge, GaAs, GaP, and InSb) are shown in Figures 9.7 through 9.10. (4 7) Information concerning the optical absorption coefficient versus photon energy is essential for analyzing the photoelectric effects in a semiconductor.

264 9. Optical Properties and Photoelectric Effects Figure 9.7. Absorption coefficient versus photon energy for a GaP sample at room temperature. After Spitzer et al., (4) by permission. 9.5. The Photoconductivity Effect Author: Citations of fig. 8to10are missing. In this section, the photoconductivity effect in a semiconductor is depicted. In the absence of illumination, the dark conductivity of a semiconductor is given by σ 0 = q(n 0 µ n + p 0 µ p ), (9.72) Figure 9.8. Absorption coefficient versus photon energy for a GaAs sample at room temperature. After Moss and Hawkins, (5) by permission.

9.5. The Photoconductivity Effect 265 Figure 9.9. Absorption coefficient versus photon energy for silicon and germanium crystals measured at 300 K. After Dash and Newman, (6) by permission. where n 0 and p 0 denote the densities of electrons and holes in thermal equilibrium, while µ n and µ p are the electron and hole mobilities, respectively. When photons with energies equal or greater than the band gap energy (hν E g ) of a semiconductor are absorbed in a semiconductor, intrinsic photoconductivity results. The absorbed photons create excess electron hole pairs (i.e., n and p), and as a result the densities of electrons and holes (i.e., n and p) increase above their equilibrium values of n 0 and p 0 (i.e., n = n 0 + n, p = p 0 + p). Figure 9.10. Absorption coefficient versus photon energy for a pure InSb sample measured at three different temperatures. After Johnson, (7) by permission.

266 9. Optical Properties and Photoelectric Effects The photoconductivity is defined as the net change in electrical conductivity under illumination and can be expressed by σ = σ σ 0 = q ( nµ n + pµ p ), (9.73) where n and p are the excess electron and hole densities, respectively. In a degenerate semiconductor, p and n are generally much smaller than p 0 and n 0, and the effect of incident photons can be considered as a small perturbation. However, in an insulator or a nondegenerate semiconductor, values of n and p can become comparable or larger than their equilibrium carrier densities. If the effect of electron or hole trapping by the defect levels is negligible and the semiconductor remains neutral under illumination, then n = p holds throughout the specimen. Depending on the incident photon energies, there are two types of photoconduction processes that are commonly observed in a semiconductor. One type of photoconduction process is known as the intrinsic photoconductivity (PC), in which the excess electron hole pairs are generated across the energy band gap by the absorption of photons with energies greater than the band gap energy of the semiconductor (i.e., hv E g ). This type of photoconduction process is illustrated in Figure 9.11a. The other type of photoconduction process is known as the extrinsic photoconductivity, in which electrons (or holes) are excited from the localized donor (or acceptor) states into the conduction (or valence) band states by the absorption of photons with energy equal to or greater than the activation energy of the donor (or acceptor) levels, but is less than the band gap energy of the semiconductor (i.e., E D hv E g for n-type conduction, and E A hv E g for p-type conduction). This is illustrated in Figure 9.11b. In intrinsic photoconduction, both the photogenerated electrons and holes are participating in the photoconduction process, and the photoconductivity is described by Eq. (9.73). However, for the extrinsic photoconductivity, the Figure 9.11. (a) Intrinsic and (b) extrinsic photoconductivity in a semiconductor.

9.5. The Photoconductivity Effect 267 photoconduction process usually involves only one type of charge carriers (i.e., either electrons or holes), and the expressions for the extrinsic photoconductivity are given by σ n = qn D µ n for n-type, (9.74) σ p = qp A µ p for p-type, (9.75) where n D and p A are the photogenerated excess electron and hole densities from the donor and acceptor centers, respectively. An extrinsic photoconductor usually operates at cryogenic temperatures because at very low temperatures freeze-out occurs for electrons in the conduction band states or for holes in the valence band states. The transition of electrons from the conduction band states to the shallow-donor states or of holes from the valence band states to the shallow-acceptor states because of the freeze-out effect is the basis for the extrinsic photoconductivity. At very low temperatures, the electrical conductivity under the dark condition and the background noise are generally very low. When photons with energies of E D hv E g are impinging on an n-type specimen, the electrical conductivity of the sample will increase dramatically by the absorption of these incident photons. These photons excite the electrons in the shallow-donor impurity states into the conduction band states, resulting in the increase of electrical conductivity in the sample. The sensitivity of extrinsic photoconductivity depends greatly on the density of sensitizing shallow-impurity centers and the thickness of the specimen. Extrinsic photoconductivity has been widely used in long-wavelength IR detection. For example, a Cu-doped germanium extrinsic photoconductor operating at 4.2 K can be used to detect photons with wavelengths ranging from 2.5 to 30 µm, while a Hg-doped germanium photodetector operating at 28 K can be used for 10.6 µm wavelength detection. Figure 9.12 shows the schematic diagram of an intrinsic photoconductor under illumination and bias conditions. In the intrinsic photoconduction process, electron hole pairs are generated in a semiconductor when photons with energies exceeding the band gap energy of the semiconductor are absorbed. The rate of generation of electron hole pairs per unit volume per unit time can be written as g E = αφ 0 (1 R) for αd 1 (9.76) g E = αφ 0 (1 R) e αy for αd 1 (9.77) where R is the reflection coefficient of the semiconductor defined by Eq. (9.25), α is the absorption coefficient, φ 0 is the photon flux density (i.e., φ 0 = I 0 /hv), and I 0 is the incident light intensity per unit area (W/cm 2 ). Equation (9.76) is valid for a very thin photoconductor (i.e., αd 1) in which photons are uniformly absorbed throughout the sample, while Eq. (9.77) is applicable for a thick specimen (i.e., αd 1) in which the photogeneration rate decays exponentially with penetration distance. These are discussed next.

268 9. Optical Properties and Photoelectric Effects Figure 9.12. Photoconductivity process in a semiconductor specimen. First consider the case of a thin specimen with αd 1. Here, the excess electron and hole densities are related to the generation rate g E by n = g E τ n (9.78) p = g E τ p (9.79) where τ n and τ p denote the electron and hole lifetimes, respectively. As shown in Figure 9.12, the change of electrical conductance as a result of the incident photons can be expressed by ( ) A G = σ = q ( ) ( ) Wd nµ n + pµ p l l ( ) ( ) Wd = qg E τn µ n + τ p µ p (9.80) l and ( ) τn µ n + τ p µ p G = qg E, (9.81) l 2 where G E = g E (Wdl) = g E V 0 is the total volume generation rate (i.e., total number of carriers generated per second), and A = Wd is the cross-section area. If V is the applied voltage, the photocurrent I ph can be expressed as ( ) τn µ n + τ p µ p I ph = V G = qvg E l 2 = qvg E S, (9.82)

9.5. The Photoconductivity Effect 269 where S = (τ nµ n + τ p µ p ) = µτ (9.83) l 2 l 2 is the photosensitivity factor. It is seen that the value of S is directly proportional to the product µτ. This means that in order to obtain a high photosensitivity factor, the lifetimes and mobilities of the excess carriers must be as large as possible and the sample length l between two electrodes should be as small as possible. As an example, consider a silicon photoconductor. If the wavelength of the incident photon is λ = 0.5 µm, the absorption coefficient α = 10 4 cm 1, τ n = 100 µs, the reflection coefficient R = 0.3, and the photon flux density φ 0 = 10 14 cm 2 s 1, then the excess electron density can be calculated by using the following formula: n = αφ 0 (1 R) τ n = 7 10 13 cm 3, (9.84) which shows that a relatively large density of excess electrons can be generated even with a relatively small incident light intensity. Another parameter that has often been used to assess the performance of a photoconductor is the photoconductivity gain G p. This figure of merit (G p )isdefined as the ratio of the excess carrier lifetime τ to the carrier transit time t r across the specimen, which can be defined by G p = τ t r = SV, (9.85) where t r = l/v d = l 2 /µv is the transit time for the excess carriers to drift across the photoconductor specimen, ν d is the drift velocity, and µ is the carrier mobility. A photoconductivity gain of 10 4 can be readily obtained for a CdS photoconductor. In the above formulation, loss due to surface recombination was neglected. For a thin-film photoconductor, the effect of surface recombination can be incorporated into an effective excess carrier lifetime as 1 = 1 + 1, (9.86) τ τ B τ s where τ is the effective excess carrier lifetime, τ B is the bulk carrier lifetime, and τ s is the surface recombination lifetime given by τ s = d 2s, (9.87) where s is the surface recombination velocity and d is the sample thickness. For example, for a chemically polished silicon specimen, with s = 500 cm/s and thickness d = 2 µm, the surface recombination lifetime τ s,isfound equal to 2 10 7 s. Therefore, for a thin-film photoconductor, if τ s is less than τ B, then the surface recombination lifetime rather than the bulk lifetime may control the effective excess carrier lifetime.

270 9. Optical Properties and Photoelectric Effects In general, the photocurrent for a thin-film photoconductor can be derived from Eqs. (9.76) and (9.82), and the result yields ( ) I ph = GV a = q (1 R) αφ 0 τ Va (1 + b) µ p (Wd), (9.88) l where τ is given by Eq. (9.86) and b = µ n /µ p is the electron-to-hole mobility ratio. From Eq. (9.88) it is seen that, for a constant τ, the photocurrent I ph is directly proportional to the light intensity I 0 (= φ 0 /hν)orphoton flux density φ 0. This is generally true under low- and high-injection conditions (i.e., for n n 0 or n n 0 ). However, for the intermediate-injection range (i.e., n n 0 ),τ may become a function of the injected excess carrier density, n, and hence the photocurrent is no longer a linear function of the light intensity. Depending on the relationship between the excess carrier lifetime and the injected excess carrier density, a superlinear or sublinear region may exist in the intermediate-injection regime. Next consider the case of a thick photoconductor with αd 1. Here, the generation rate is given by Eq. (9.77). Because of nonuniform absorption the diffusion of excess carriers along the direction of incident photons plays an important role in this case. As shown in Figure 9.12, the excess carrier densities as a function of distance along the y-direction can be obtained by solving the following continuity equation: D n 2 n y 2 n τ n = g E = αφ 0 (1 R) e αy (9.89) If the electron diffusion length L n (= (D n τ n ) 1/2 )ismuch smaller than the sample thickness d, then Eq. (9.89) has the solution n = αi [( 0 (1 R) τ n αl 2 ) ] hv ( n + sτ n α 2 L 2 n 1) e y/l n e αy, (9.90) L n + sτ n where I 0 = φ 0 /hv is the incident light intensity. Note that Eq. (9.90) was obtained by using the boundary condition that n D n y = s n. (9.91) y=0 y=0 The photocurrent can be obtained by integrating Eq. (9.90) with respect to y from y = 0toy =, and the result yields ( ) W I ph = V a qµ p (1 + b) l 0 n dy = qi 0WL n µ p (1 + b)τ n (1 R)V a l(l n + sτ n )hv [ ] sτ n 1 +, (9.92) L n (1 + αl n ) where l is the sample length. In Eq. (9.92), the upper limit of the integral y = d is replaced by y = in the integration. This is valid as long as the sample thickness is much larger than the diffusion length of electrons.

9.5. The Photoconductivity Effect 271 Figure 9.13. Relative photoresponse versus wavelength for different surface recombination velocities, s i. Figure 9.13 shows the photocurrent versus wavelength of incident photons for different surface recombination velocities. For sτ n L n, the photocurrent I ph reaches a maximum for α 1/L n.however, if the surface recombination velocity is small and sτ n L n, then the photocurrent will increase monotonically with decreasing wavelength. This is shown in Figure 9.13 for the case of s = 0. The sharp decrease in photocurrent is usually observed near the absorption edge (i.e., hv E g )inwhich the absorption coefficient decreases exponentially with increasing wavelength. However, in very short-wavelength regime (i.e., near the UV regime), the absorption coefficient is usually very large (i.e., α 10 5 cm 1 ) and αl n 1. In this regime, the excess carriers are generated near the surface of the photoconductor where the excess carrier lifetime is controlled by the surface recombination. Thus, the photocurrent is expected to decrease rapidly with increasing surface recombination velocity in the short-wavelength regime. In order to improve the short-wavelength photoresponse, careful preparation of the sample surface is necessary so that the surface recombination velocity of a photoconductor can be kept low. 9.5.1. Kinetics of Photoconduction Since the photocurrent is directly related to the excess carrier densities generated by the incident photons, a study of photocurrent as a function of light intensity usually yields useful information concerning the recombination mechanisms of the excess carriers in a semiconductor. As an example, consider an n-type direct band gap semiconductor. If the band-to-band radiative recombination dominates the excess carrier lifetimes, then the kinetic equation for the photoconduction process can be expressed by dn dt = g E U, (9.93)