Nothing Special   »   [go: up one dir, main page]

THERMOPHOTOVOLTAIC DEVICES AND INFRARED PHOTODETECTORS BASED ON INTERBAND CASCADE STRUCTURES

Download as pdf or txt
Download as pdf or txt
You are on page 1of 232

UNIVERSITY OF OKLAHOMA

GRADUATE COLLEGE

THERMOPHOTOVOLTAIC DEVICES AND INFRARED PHOTODETECTORS

BASED ON INTERBAND CASCADE STRUCTURES

A DISSERTATION

SUBMITTED TO THE GRADUATE FACULTY

in partial fulfillment of the requirements for the

Degree of

DOCTOR OF PHILOSOPHY

By

WENXIANG HUANG
Norman, Oklahoma
2020
THERMOPHOTOVOLTAIC DEVICES AND INFRARED PHOTODETECTORS
BASED ON INTERBAND CASCADE STRUCTURES

A DISSERTATION APPROVED FOR THE


HOMER L. DODGE DEPARTMENT OF PHYSICS AND ASTRONOMY

BY THE COMMITTEE CONSISTING OF

Dr. Rui Yang, Chair

Dr. Michael Santos, Co-chair

Dr. Kieran Mullen

Dr. Bruno Uchoa

Dr. Arne Schwettmann

Dr. Bin Wang


© Copyright by WENXIANG HUANG 2020
All Rights Reserved
To my parents, Caijin Huang and Feng Cheng

iv
1 Acknowledgements
It would not have been possible to write this dissertation without the help and

support of the kind people around me, only some of whom will be given special mention

here.

Above all, I would like to express my sincere gratitude to my advisor Prof. Rui Q.

Yang for his selfless support, for prompt and useful advice on my Ph.D. study and research,

and for sharing his motivation and immense knowledge. He assigned me both theoretical

and experimental topics that were interesting and meaningful, as well as beneficial for

improving my analytical skills and hands-on abilities. It has been wonderful to work with

him and I have had a lot of fun interacting with him all the time. I also want to thank him

for the travel support he provided for me to attend several conferences and share our

research results with community members. Besides my advisor, I would like to thank Prof.

Mike Santos for chairing my committee, and for his support and assistance since the start

of my graduate career. He provided the resources for our group to grow our material,

nominated me for several awards, and helped me improve my presentations.

I am grateful to Dr. Kieran Mullen, Dr. Bruno Uchoa, Dr. Arne Schwettmann and

Dr. Bin Wang for serving on my dissertation committee and squeezing out time to review

my general exam and dissertation.

I would also like to thank the current and former members of the Quantum Device

Laboratory at the University of Oklahoma. During my first year in the lab, I was fortunate

to work closely with Dr. Lin Lei. Many thanks go to him for teaching me the

characterization measurements and data analysis of the devices. Many thanks go to Dr. Lu

Li who performed the fabrication of most of the devices studied in this dissertation. I want

v
to extend sincere thanks to Dr. S.M. Shazzad Rassel. He was helpful in carrying out the

dark current measurements for many laser devices. I also want to thank Dr. Hossein Lotfi

for long discussions on band structure modeling and the device theory. Many thanks go to

Dr. Hao Ye for growing some of the structures presented in this work. I want to thank the

current group member Jeremy Massengale for MBE growth and material characterization

of most of the structures presented in this work. I am grateful to Yuzhe Lin for the

fabrication of some devices in this work.

I am also appreciative of our group’s collaborators. Particularly, I would like to

thank Prof. Matthew Johnson for providing our group with both resources and expertise in

device fabrication and material characterization. I would like to thank Dr. James Gupta of

the National Research Council of Canada who performed the MBE growth of some laser

structures for our research group.

I would also like to acknowledge the sources that have provided me with funding

during my Ph.D. career. I am grateful for funding from the NSF (Award Nos. ECCS-

1202318, DMR-1229678, DMR-1608224, and IIP-1640576) and the AFOSR (Award No.

FA9550-15-1-0067).

Last, but definitely not least, I am greatly indebted to my parents. It is their

unconditional love, care, and tolerance that made the hardship of finishing my Ph.D. career

worthwhile. Without their support, I would not have been able to overcome the difficulties

during these years.

vi
Table of Contents
Acknowledgements ........................................................................................................... v

List of Tables .................................................................................................................. xii

List of Figures ................................................................................................................ xiv

Abstract ........................................................................................................................ xxiii

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

1.1 Infrared radiation ............................................................................................ 1

1.2 Overview of infrared thermophotovoltaic energy conversion ........................ 5

1.2.1 Background ............................................................................................... 5

1.2.2 Active components in a TPV system ........................................................ 8

1.2.3 Thermophotovoltaic cells.......................................................................... 9

1.2.4 Thermodynamic analysis of thermophotovoltaic cell efficiency ............ 12

1.3 Overview of infrared detectors ..................................................................... 15

1.3.1 Background ............................................................................................. 15

1.3.2 Photon detection vs thermal detection .................................................... 17

1.3.3 Examples of photon detectors ................................................................. 19

1.3.4 Performance metrics for infrared detectors............................................. 23

1.4 Dissertation organization .............................................................................. 25

Chapter 2: Sb-based interband cascade devices.............................................................. 27

2.1 6.1 Å Semiconductor family ......................................................................... 27

2.2 Interband cascade lasers-the historic origin .................................................. 28

2.2.1 History and operation principle .............................................................. 28

2.2.2 Current status of ICL performance ......................................................... 31

vii
2.3 Interband cascade thermophotovoltaic devices ............................................ 32

2.3.1 InAs/GaSb type-II superlattices .............................................................. 32

2.3.2 Operation principle of ICTPV cells ........................................................ 34

2.3.3 Enhancement of open-circuit voltage in ICTPV cells ............................ 36

2.3.4 Improvement of carrier collection efficiency in IC structures ................ 37

2.4 Interband cascade infrared photodetectors.................................................... 41

2.4.1 Operation principle of ICIPs ................................................................... 41

2.4.2 Noise reduction in ICIPs ......................................................................... 43

2.4.3 Detectivity improvement in ICIPs .......................................................... 45

2.4.4 Comments on detectivity improvement in ICIP ..................................... 46

2.5 Growth and fabrication of interband cascade devices .................................. 47

Chapter 3: Limiting factors and efficiencies of narrow bandgap thermophotovoltaic cells

............................................................................................................................. 50

3.1 Background and motivation .......................................................................... 50

3.2 Practical limitations on single-absorber TPV devices .................................. 51

3.2.1 Quantum efficiency and dark saturation current density ........................ 51

3.2.2 Open-circuit voltage and fill factor ......................................................... 56

3.2.3 Conversion efficiency ............................................................................. 59

3.3 Efficiency improvement in multistage TPV devices .................................... 64

3.3.1 Enhancement of open-circuit voltage ..................................................... 64

3.3.2 Enhancement of conversion efficiency ................................................... 67

3.4 Performance of TPV devices under variable illumination sources ............... 69

3.4.1 Single-absorber TPV cells ...................................................................... 69

viii
3.4.2 Multistage ICTPV devices ...................................................................... 71

3.5 Summary and concluding remarks................................................................ 73

Chapter 4: Experimental comparison between single-absorber and multistage IC

thermophotovoltaic devices ................................................................................ 75

4.1 Background and motivation .......................................................................... 75

4.2 Device Structure, growth and fabrication ..................................................... 77

4.3 Device characterizations and discussions ..................................................... 79

4.3.1 Quantum efficiency ................................................................................. 79

4.3.2 Particle conversion efficiency ................................................................. 82

4.3.3 Illuminated J-V curve and open-circuit voltage ...................................... 83

4.3.4 Fill Factor and conversion efficiency ..................................................... 87

4.4 Extraction of some important performance related parameters .................... 90

4.4.1 Voltage-dependent collection efficiency ................................................ 90

4.4.2 Thermal generation rate and carrier lifetime .......................................... 93

4.4.3 Series resistance ...................................................................................... 95

4.4.4 Surface leakage ....................................................................................... 96

4.5 Summary and concluding remarks................................................................ 97

Chapter 5: Interband cascade thermophotovoltaic devices with more stages ................. 99

5.1 Background and motivation .......................................................................... 99

5.2 Device structure, growth and fabrication .................................................... 101

5.3 Energy conversion efficiency ..................................................................... 102

5.4 Device characterization and analysis .......................................................... 105

5.4.1 Dark current density and carrier lifetime .............................................. 105

ix
5.4.2 Quantum efficiency and current mismatch ........................................... 107

5.4.3 Collection efficiency of photogenerated carriers .................................. 111

5.5 Quantification of the effects of the performance limiting factors............... 113

5.5.1 Effect of collection efficiency ............................................................... 113

5.5.2 Effect of current mismatch.................................................................... 115

5.5.3 Effect of material quality ...................................................................... 117

5.6 Summary and concluding remarks.............................................................. 119

Chapter 6: Carrier lifetime in mid wavelength interband cascade devices ................... 121

6.1 Introduction ................................................................................................. 121

6.2 Carrier lifetime in mid-wavelength ICIPs................................................... 123

6.2.1 Device structure, growth and fabrication .............................................. 123

6.2.2 Dark current density .............................................................................. 125

6.2.3 Contribution of SRH process to dark current ....................................... 127

6.2.4 Linear fitting of dark current density .................................................... 129

6.2.5 Estimated thermal generation rate and carrier lifetime ......................... 130

6.3 Interband cascade devices vs quantum cascade devices ............................. 133

6.3.1 Device structures ................................................................................... 133

6.3.2 Semi-empirical model for dark current density .................................... 134

6.3.3 Saturation current densities for cascade devices ................................... 135

6.3.4 Effect of J0 on the performances of detectors ....................................... 138

6.3.5 Effect of J0 on the performances of photovoltaic cells ......................... 142

6.4 Summary and concluding remarks.............................................................. 143

Chapter 7: Long wavelength interband cascade infrared photodetectors ..................... 145

x
7.1 Introduction ................................................................................................. 145

7.2 Current matched ICIPs vs noncurrent-matched ICIPs ................................ 147

7.2.1 Device structure, growth and fabrication .............................................. 147

7.2.2 Electrical properties .............................................................................. 148

7.2.3 Responsivity.......................................................................................... 152

7.2.4 Electrical gain ....................................................................................... 155

7.2.5 Johnson-noise limited detectivity ......................................................... 156

7.3 A comprehensive study of electrical gain in ICIPs ..................................... 158

7.3.1 Device structure, growth and fabrication .............................................. 158

7.3.2 Responsivity.......................................................................................... 160

7.3.3 Electrical gain ....................................................................................... 163

7.3.4 Underlying mechanism of electrical gain ............................................. 165

7.3.5 Net effect of electrical gain ................................................................... 167

7.3.6 Electrical characteristics ....................................................................... 171

7.3.7 Johnson-noise limited detectivity ......................................................... 173

7.4 Summary and concluding remarks.............................................................. 176

Chapter 8: Concluding notes and future work .............................................................. 178

8.1 Dissertation summary ................................................................................. 178

8.2 Future works ............................................................................................... 182

References ..................................................................................................................... 185

Appendix A: Publications list ....................................................................................... 205

xi
2 List of Tables
Table 1-1: Summary of some demonstrated TPV system performance. .......................... 7

Table 1-2: Summary of various TPV technologies, classified by absorbing material.... 11

Table 1-3: Summary of various photovoltaic photodetectors, classified by detecting

material. .......................................................................................................................... 23

Table 3-1: Parameters used in calculation for InAs/GaSb superlattice. ......................... 55

Table 3-2: Parameters used in calculation for bulk Ga0.44In0.56As0.5Sb0.5 ....................... 70

Table 4-1: Summary of the PV performance and the related parameters of representative

devices (0.2×0.2 mm2) from the three ICTPV wafers at 300 K. The maximum efficiencies

shown in the table for the 3- and 5-stage devices are obtained at a maximum incident power

density of 36 W/cm2........................................................................................................ 90

Table 5-1: Summary of ICTPV devices that have been reported so far. ...................... 100

Table 5-2: Individual and total absorber thicknesses for the four IC TPV structures. . 102

Table 5-3: Summary of device characteristics and some important performance-related

parameters for the four devices at 300 K. ..................................................................... 113

Table 6-1: Summary of the design and material parameters of the seven wafers. ....... 125

Table 7-1: Summary of material and design parameters for the four devices. ............. 148

Table 7-2: Theoretical calculated and experimental extracted values of R0A ratios at high

temperatures. ................................................................................................................. 152

Table 7-3: Experimentally obtained ratio of responsivity for ICIPs at different

temperatures. ................................................................................................................. 153

Table 7-4: Summary of the design and material parameters of the five wafers ........... 160

Table 7-5: Comparison of electrical parameters of the five ICIPs. .............................. 173

xii
Table 7-6: Comparison of D* at 𝜆=7 m, along with the 100% cutoff wavelengths at 300

K, for the five devices. .................................................................................................. 174

xiii
3 List of Figures
Figure 1-1: (a) An infrared thermography applied for virus screening in airport [2], (b)

Schematic illustration of a TPV system consisting of heat source, radiator, emitter, TPV

cells and cooling system. Figure is from [3]. ...................................................................... 2

Figure 1-2: Spectral radiations for blackbodies at various temperatures. The shaded regions

are of interest for applications such as solar cell, thermophotovoltaic and thermal imaging.

............................................................................................................................................. 3

Figure 1-3: Atmospheric transmittance spectrum of infrared radiation. Figure is from [10].

............................................................................................................................................. 4

Figure 1-4: Schematic illustration of a TPV system. .......................................................... 8

Figure 1-5: The calculated efficiencies based on Equation 1-4 for various blackbody

temperatures. The insert shows the optimal bandgap that maximizes the efficiency. ...... 14

Figure 1-6: The development history of modern infrared detectors and systems............. 15

Figure 1-7: Mid infrared absorption spectra of some molecules and gases. Data were

collected from [69]. ........................................................................................................... 17

Figure 1-8: Block diagram of a thermal detector. Figure from [76]. ................................ 18

Figure 1-9: Images created by uncooled and cooled infrared cameras. Figures are from

[77]. ................................................................................................................................... 19

Figure 1-10: Schematic diagram of (a) a photoconductive photodetector made of a

semiconductor slab, (b) a quantum well infrared photodetector based on bound-to-

continuum transition. ........................................................................................................ 20

Figure 1-11: Schematic diagram of a PV detector made of a single p-n junction ............ 21

xiv
Figure 1-12: Schematics of (a) a nBn barrier detector and (b) a complementary barrier

infrared detector; the biases are applied to improve carrier collection. ............................ 21

Figure 1-13: Schematic diagram of a quantum cascade detector. .................................... 22

Figure 2-1: (a) Bandgap, lattice constant and (b) band alignment of the 6.1 Å

semiconductor materials. .................................................................................................. 27

Figure 2-2: Illustration of the photon emission and cascading effect in an interband cascade

laser. Figure from [115]. ................................................................................................... 29

Figure 2-3: Band diagram of the active core for an interband cascade laser. Figure from

[73]. ................................................................................................................................... 31

Figure 2-4: Room temperature threshold current density for both InAs- and GaSb-based

broad-area ICLs. Figure is from [124]. ............................................................................. 32

Figure 2-5: Band structure, minibands and wavefunctions of electrons and holes for (a)

InAs/GaSb superlattice and (b) M-shape Al(In)Sb/GaSb/InAs/GaSb/Al(In)Sb SL. ........ 33

Figure 2-6: (a) Schematic band diagram of an ICTPV cell, (b) Schematic showing the

operation of an ICTPV cell. .............................................................................................. 36

Figure 2-7: Collection probability of carriers as a function of the distance from the

collection point. The absorber thickness is 3.3 m. The number near the curve indicates

the diffusion length. .......................................................................................................... 38

Figure 2-8: Comparison of collection process in single- and four-stage IC devices for a low

L product (L=0.4). The thickness d of the single-stage device equates the absorption

depth. The individual absorber thickness of four-stage IC device is d/4. ......................... 40

Figure 2-9: (a) schematic diagram of a multistage ICIP and (b) the band profile of one stage

under zero bias. The olive and purples lines in the absorber represent the electron and hole

xv
minibands. The dotted olive wavefunction indicates the electron states in hole barrier while

the dotted purple wavefunction represents the hole states in electron barrier. ................. 42

Figure 2-10: Johnson-noise limited detectivity enhancement for current-matched ICIPs

with two, eleven and thirty stages. Figure is from [141]. ................................................. 46

Figure 2-11: (a) Intevac GEN II MBE system (1993) and (b) Veeco GENxplor MBE

system (2013). ................................................................................................................... 48

Figure 2-12: (a) The schematic of a processed ICTPV or ICIP device and (b) Cross-section

scanning electron microscope image of a wet-etch ICTPV structure, the Figure is from

[157]. ................................................................................................................................. 49

Figure 3-1: Calculated open-circuit voltage (solid) and quantum efficiency (dashed) as a

function of normalized absorber thickness for different values of L. The incident power

density is assumed to be 50 W/cm2................................................................................... 53

Figure 3-2: Calculated dark saturation current density as a function of normalized absorber

thickness for a carrier lifetime of 20 ns, 200 ns and the radiative limit. ........................... 55

Figure 3-3: Simulated J-V curves for different values of L and with incident power density

of 25 and 50 W/cm2. ......................................................................................................... 57

Figure 3-4: Calculated fill factor as a function of normalized absorber thickness for

different values of L. The incident power density is assumed to be 50 W/cm2 except in

the ultimate limit. .............................................................................................................. 59

Figure 3-5: Calculated conversion efficiency vs normalized absorber thickness for different

values of L. The incident power density is assumed to be 50 W/cm2 except for the ultimate

limit. .................................................................................................................................. 62

xvi
Figure 3-6: Calculated (a) conversion efficiency and (b) voltage efficiency and fill factor

for L=0.45, 1.5 and 4.5 and the ultimate efficiency limit as a function of bandgap. The

incident power density is 50 W/cm2 except for the ultimate limit.................................... 63

Figure 3-7: Schematic of a three-stage ICTPV device under forward voltage and

illumination. Optical generation gphm, thermal generation gthm and recombination Rm, along

with the chemical potentials m in each stage are shown, where the index m denotes the

stage ordinal. The flat quasi-Fermi levels (designated with 1, 2, 3 and 4) correspond

to the case where the diffusion length is infinite. ............................................................. 65

Figure 3-8: Calculated open-circuit voltage enhancement Voc(Nc)/Voc(1) as a function of

number of stages. The dashed purple line indicates Voc(Nc)/Voc(1)=Nc. In the calculations,

L was set at 0.45. 1.5 and 4.5. The incident power density is 50 W/cm2. ...................... 66

Figure 3-9: Calculated conversion efficiency for optimized multistage cells as a function

of number of stages. The calculation is done for L=0.45 1.5 and 4.5. The incident power

density is 50 W/cm2. ......................................................................................................... 68

Figure 3-10: Calculated maximum conversion efficiency and conversion efficiency

enhancement as a function of L. The incident power density is 50 W/cm2. .................. 69

Figure 3-11: Calculated conversion efficiency of a GaInAsSb single-absorber device vs

wavelength for various values of diffusion length. The incident power density is 50 W/cm2.

........................................................................................................................................... 71

Figure 3-12: Calculated conversion efficiency for the 5- and 20-stage devices with L=1.5

m (solid curves) and 15 m (dashed curves). The absorbers were adjusted to be

photocurrent matched with an absorption coefficient of 3000 cm-1, corresponding to a

wavelength of 4 m. The calculated maximum efficiencies with optimized multi-stage

xvii
structures at every wavelength are represented by the olive curves. The incident power

density is 50 W/cm2. ......................................................................................................... 73

Figure 4-1: (a) Calculated quantum efficiency and collection efficiency, and (b) open-

circuit voltage factor as a function of normalized absorber thickness for several values of

L. VF initially decreases with increasing d/L due to the nearly linear increase of dark

current when d/L is small. ................................................................................................. 77

Figure 4-2: Schematic layer structures of the three TPV devices with one, three and five

stages. ................................................................................................................................ 79

Figure 4-3: Measured QE spectra of 1-, 3- and 5-stage devices at 300 and 340 K. ......... 80

Figure 4-4: Voltage dependent QE at 4 m for the three devices, where different vertical

scales are used in the top and bottom portions to better show variations. ........................ 81

Figure 4-5: (a) Current-voltage characteristics of the three devices at 300 K under a medium

illumination level where the incident power density was about 19 W/cm2. The solid, dotted

and dashed curves correspond to the measured, Rs corrected and ideal cases, respectively.

(b) Current-voltage characteristics of the three devices at 200 K under the same level of

illumination as in (a). The inset shows the emission spectrum of the ICL. ...................... 85

Figure 4-6: (a) Open-circuit voltage, (b) fill factor, (c) maximum output power density and

(d) conversion efficiency as a function of incident power density for the three devices at

300K.................................................................................................................................. 88

Figure 4-7: (a) Voltage dependence of collection efficiency derived from Equation 4-2

using four different pairs of J-V data at 300 K for the three devices. The numbers in the

legend indicate the incident power densities under different illumination levels. (b)

Average collection efficiency over the four pairs in (a). .................................................. 92

xviii
Figure 4-8: The thermal generation rate and minority carrier lifetime for the 1-, 3- and 5-

stage devices at high temperatures. ................................................................................... 95

Figure 4-9: dV/dI data to obtain series resistance at 300 K, which was found from the

intercept of dV/dI. ............................................................................................................. 96

Figure 4-10: Size dependent R0A for the three devices at 300 K. The sidewall resistivity

was smallest for the one-stage device. .............................................................................. 97

Figure 5-1: Schematic layer structure of the four TPV devices with six, seven, sixteen and

twenty-three stages.......................................................................................................... 102

Figure 5-2: (a) Illuminated current density-voltage characteristics for the representative

200×200 m2 devices from the four wafers at 300 K and at an incident power density of

17 W/cm2. The inset shows the emission spectrum of the IC laser used as the illumination

source, (b) Conversion efficiency as a function of incident power density for the four

devices at 300 K. ............................................................................................................. 104

Figure 5-3: (a) Dark current density for the representative 200×200 m2 devices from the

four wafers at 300 K, (b) Linear fitting (dashed lines) of dark current density at reverse

voltage for the four devices at 300 K. ............................................................................. 106

Figure 5-4: (a) Quantum efficiency spectra of the four devices at 300 K and (b) Bias

dependence of quantum efficiency for the four devices at 300 K and at the wavelength of

4.2 m. ............................................................................................................................ 109

Figure 5-5: (a) Calculated effective quantum efficiency based on Equation 2-5 in each stage

of the four devices, (b) Calculated incident power density vs IC laser current based on

Equation 4-1 for the four devices.................................................................................... 110

xix
Figure 5-6: (a) The measured and the 100% collected J-V curves for the four devices at 300

K and at the incident power density of 17 W/cm2, (b) Extracted collection efficiency at 300

K based on Equation 4-2 using J-V data under incident power densities of 7 and 17 W/cm2

for the four devices. ........................................................................................................ 113

Figure 5-7: Comparison of the measured  and the ideal  in the 100% collected case at

300 K for the 6- and 7-stage devices. ............................................................................. 114

Figure 5-8: Calculated (a) short-circuit current density and (b) conversion efficiency based

on Equation 5-1 as a function of absorption coefficient at incident power density of 17

W/cm2 for the four devices. ............................................................................................ 117

Figure 5-9: Calculated conversion efficiency based on Equation 3-10, along with

measurement for the four devices. For each of the four devices, the carrier lifetime used in

the calculation was 27 and 87 ns..................................................................................... 119

Figure 6-1: Radiative and non-radiative recombination processes in semiconductors... 122

Figure 6-2: Dark current density versus applied voltage for the seven devices at (a) 250 K

and (b) 300 K. ................................................................................................................. 127

Figure 6-3: (a) R0A of the seven devices in the temperature range of 200-340 K. (b)

Temperature dependence of bandgap for M3S-312. The fitting Varshni parameters for the

device are shown. ............................................................................................................ 128

Figure 6-4: Linear fitting (dashed) and experimental measurements (solid) of the dark

current density at reverse bias voltage for the five multistage devices at 300 K. The inset

shows the corresponding results of the two single-stage devices at 300 K. ................... 130

Figure 6-5: The thermal generation rate and minority carrier lifetime for the five multistage

and two single-stage devices at high temperatures. ........................................................ 133

xx
Figure 6-6: The measured and fitted Jd-V curves for an 8-stage ICD and a 50-stage QCD at

300 K. The ICD and QCD were mentioned in [115] (wafer R083) and [96], respectively.

......................................................................................................................................... 136

Figure 6-7: The extracted values of J0 for ICDs and QCDs at 300 K. Some ICDs have been

described previously in [83-84, 96, 222-224], while others are from our unpublished

studies. The QCDs are from [115, 120, 179, 225-226]................................................... 138

Figure 6-8: Measured peak (a) responsivities and (b) detectivities for ICDs, ICD_SLs and

QCDs at 300 K. In addition to some of the ICDs presented in Figure 6-6, two ICDs (devices

A and B) [136] and all ICD_SLs from [137, 151, 199, 229-231] are included. One QWIP

is from [236]. .................................................................................................................. 139

Figure 6-9: Estimated Voc at 300 K for the ICDs, ICD_SLs and QCDs shown in Figure 6-

8....................................................................................................................................... 143

Figure 7-1: Schematic illustration of the multi-stage ICIP with (a) regular and (b) reverse

configurations. The two configurations can be realized by reversing the growth order of

layers in one structure without changing the light illumination direction. ..................... 146

Figure 7-2: Extracted R0A of the four representative devices at various temperatures. . 150

Figure 7-3: The theoretical R0A curves at T=300K. The device dark current was dominated

by the diffusion process at this temperature. .................................................................. 151

Figure 7-4: Zero-bias responsivity spectra for the four devices at different temperatures.

......................................................................................................................................... 153

Figure 7-5: Temperature-dependent responsivity of the four devices at 7 m. .............. 154

Figure 7-6: Absorption coefficient and electrical gain at room temperature. The dips near

4.2 m in the gain curves were due to CO2 absorption in the response spectra. ............ 156

xxi
Figure 7-7: Johnson-noise limited D* spectra of the four devices at various temperatures.

......................................................................................................................................... 158

Figure 7-8: (a) Zero-bias responsivity spectra for the five devices at different temperatures.

(b) Theoretically calculated external quantum efficiency of the five devices vs. absorption

coefficient. ...................................................................................................................... 162

Figure 7-9: Absorption coefficient and electrical gain at room temperature. The dips near

4.2 μm in the gain curves were due to CO2 absorption in the response spectra. ............ 164

Figure 7-10: Theoretically calculated photocurrent based on Equation 7-5 and (b) electric

potential calculated based on Equation 7-7 for each stage of the five devices at room

temperature. .................................................................................................................... 167

Figure 7-11: Theoretically calculated and experimentally measured signal current for the

five devices. .................................................................................................................... 169

Figure 7-12: Theoretical and experimental responsivity spectra for two devices at 250 K

with the IR source and a standard blackbody radiation source at 800 and 1200 K. ....... 171

Figure 7-13: Arrhenius plot of dark current density (measured at -50 mV) and R0A of the

five devices in the temperature range of 200-340 K. ...................................................... 172

Figure 7-14: Johnson-noise limited D* spectra of the five devices at various temperature.

......................................................................................................................................... 174

Figure 7-15: Detectivity derived from Equation 7-10 versus the number of stages with

various ratios of the individual absorber thickness to the diffusion length (d/L), which are

labeled near the curves in the two cases. ........................................................................ 176

xxii
4 Abstract
Mid-infrared (IR) optoelectronic devices form the basis for many practical

applications such as thermophotovoltaic (TPV) energy conversion, gas sensing, thermal

imaging, medical diagnostics, free-space communications, infrared countermeasures and

IR illumination. The mid-IR device family based on interband cascade (IC) structures

includes IC lasers (ICLs), ICTPV cells and IC infrared photodetectors (ICIPs). These are

special types of multistage devices whose operation is made possible by the unique

properties of the 6.1 Å material system: InAs, GaSb and AlSb, and their related alloys. One

of the key properties is the type-II broken-gap alignment between InAs and GaSb.

In multistage ICTPV cells and ICIPs, electrons must undergo multiple interband

excitations in order to travel between the electrical contacts. This means that the transport

of a single electron requires multiple photons, which reverses the situation in ICLs where

a single electron can generate multiple photons. Counterintuitively, this transport feature

in ICTPV cells and ICIPs is conducive to improving device performance by enhancing the

open-circuit voltage in ICTPV cells and suppressing the noise in ICIPs. Furthermore, the

collection efficiency of photo-generated carriers in multistage IC devices can be

significantly improved by thinning the absorbers in individual stages. Collectively, these

advantages make IC structures an attractive choice for narrow bandgap optoelectronic

devices, especially for operation at high temperatures. One focus of this dissertation is to

outline and demonstrate the advantages provided by IC structures, both in theory and

experiment. Another focus of this dissertation is to obtain a better understanding of the

physics of IC devices and gain insights into their operation.

Theoretical studies of single-absorber and multistage ICTPV cells are presented.

xxiii
The limitations in efficiency are understood by considering several important practical

factors. These factors are identified to be closely associated with a short carrier lifetime,

high dark saturation current density, small absorption coefficient, and limited diffusion

length. The multistage IC architecture is shown to be able to overcome the diffusion length

limitation that is responsible for the low quantum efficiency (QE) in single-absorber TPV

cells. This ability of the IC architecture offers the opportunity to enhance conversion

efficiency by about 10% for wide ranges of L (product of absorption coefficient and

diffusion length) and bandgaps, resulting in a particle conversion efficiency approaching

100%.

The illustrated theoretical advantage of multistage IC structures is confirmed

experimentally in a comparative study of three fabricated TPV devices, one with a single

absorber and two that are multistage IC structures. The bandgap of the InAs/GaSb type-II

superlattices (T2SLs) in the three devices is close to 0.2 eV at 300 K. The extracted

collection efficiency is considerably higher in multistage IC devices than in the single-

absorber device. To further investigate the prospects of IC TPV cells, detailed

characterization and performance analyses of two sets of four IC devices with similar

bandgaps are performed. The four different configurations enable a comparative study that

shows how device performance is affected by material quality variations, as well as by

current mismatch between stages and collection efficiency.

The carrier lifetime advantage of IC devices over another family of cascade devices,

namely quantum cascade (QC) devices, is manifested in the saturation current density (J0).

The values of J0 extracted using a semi-empirical model, are more than one order of

magnitude lower in IC devices than in QC devices. The significance of J0 on the

xxiv
performances of IR detectors and TPV cells is apparent in a comparison of the measured

detectivity (D*) and the estimated open-circuit voltage (Voc). To extract the carrier lifetime

in IC devices, a simple and effective electrical method is developed. This method is more

generally applicable and considers the parasitic shunt and series resistances found in

practical devices. It provides a simple way to extract the carrier lifetime in InAs/GaSb

T2SLs in a wide range of operating temperatures.

The effect of current mismatch on the performance of ICIPs is investigated using

two sets of devices with current-matched and noncurrent-matched configurations. It is

shown that current matching is necessary to achieve maximum utilization of absorbed

photons for an optimal responsivity. The detectivities of both sets of devices are

comparable largely due to the occurrence of a substantial electrical gain in noncurrent-

matched ICIPs. The electrical gain is shown to be a ubiquitous property for noncurrent-

matched ICIPs through the study of another three devices. To unlock the mechanism

underlying electrical gain, a theory is developed for a quantitative description and the

calculations are in good agreement with the experimental results.

xxv
1 Chapter 1: Introduction

1.1 Infrared radiation

Infrared radiation (IR) is a type of electromagnetic wave with wavelength longer

than for visible light. The wavelength range for IR is between about 700 nm and 1mm,

equivalent to a frequency range of approximate 430 THz to 300 GHz. IR radiation is

commonly divided into several sub-divisions [1]: near-infrared (NIR, 0.7-1.4 m), short

wavelength (SWIR, 1.4-3.0 m), mid wavelength infrared (MWIR, 3-8 m), long

wavelength (LWIR, 8.0-15 m) and far infrared (FIR, 15-1000 m). There are various uses

of infrared radiation in the areas of military, environment, industry, astronomy, climatology

and many more. For example, SWIR is extensively used in fiber-optic communication

wherein pulses of SWIR light are sent though an optical fiber. MWIR is of main interest in

gas sensing areas since many molecules and trace gasses have strong absorption lines in

this band. One of the most useful applications of LWIR is thermal imaging that translates

thermal energy into image in order to analyze an object or scene. A specific example of

thermal imaging is shown in Figure 1-1(a) in which an infrared camera is used to screen

passengers in the airport to prevent virus spread [2]. To implement these applications, one

essential component is the infrared detector. One focus of this dissertation is a special type

of semiconductor infrared detector. The other focus of this dissertation is a

thermophotovoltaic (TPV) cell that is the core element in a TPV system [3]. As shown in

Figure 1-1(b), a complete TPV system includes a heat source, radiator, emitter, set of TPV

cells and cooling system. TPV technology [4-5] has been proposed for applications such

as portable power sources, heat conversion of concentrated solar energy and cogeneration

in remote locations.

1
Figure 1-1: (a) Infrared thermography is applied for virus screening in an airport [2],
(b) Schematic illustration of a TPV system consisting of a heat source, radiator,
emitter, set of TPV cells and cooling system. Figure is from [3].

According to thermodynamic laws, all objects with temperatures higher than

absolute zero emit electromagnetic radiation. Ideally, if the object is a perfect blackbody,

the spectral radiance follows Planck’s law. In this case, the power emitted per unit area,

per unit solid angle and per unit frequency of a blackbody is given by:

2ℎ𝑐 2 1
𝐵𝜆 (𝜆, 𝑇) = ℎ𝑐 , (1-1)
𝜆5 𝑒𝑥𝑝( )−1
𝜆𝑘𝑏 𝑇

where h is the Planck’s constant, c is the speed of light, 𝜆 is wavelength, kb is Boltzmann

constant, and T is temperature. The net power per unit area radiated outward from an ideal

blackbody, considering the temperature difference with the ambient, can be obtained by

integrating Planck’s radiation formula:

𝑃 ∞ 𝑑𝜆 ∞ 𝑑𝜆
= 2ℎ𝑐 2 [∫0 ℎ𝑐 − ∫0 ℎ𝑐
] (1-2)
𝐴 𝜆5 𝑒𝑥𝑝( )−1 𝜆5 𝑒𝑥𝑝( )−1
𝜆𝑘𝑏 𝑇 𝜆𝑘𝑏 𝑇𝑎𝑚𝑏

where A is the surface area and Tamb is the ambient temperature. This integration gives the

final form of Stefan-Boltzmann law that is written as:

𝑃 𝑇
= 𝜎(𝑇 4 − 𝑇𝑎𝑚𝑏 ) (1-3)
𝐴

where  is the Stefan-Boltzmann constant, equal to 5.6704×10-8 W·m-2·K-4.

2
Illustrations of blackbody spectral radiation at various temperatures are shown in

Figure 1-2. The marked regions are linked with several specific technologies: solar cells,

thermophotovoltaics and infrared detectors. The surface temperature of the Sun is around

5800 K; the strongest output of the solar radiation spectrum is in the visible range.

Therefore, the semiconductor materials used in solar cells typically have a wide bandgap

(Eg) such as 1.1 eV for Si, the most common material for commercial solar cells [6-7]. By

comparison, the temperature of the heat source in a TPV system is in a lower temperature

regime, ranging from 1000-2000 K [4-5]. The radiation of the heat source mainly falls in

the NIR and SWIR spectra. On this account, narrower bandgap materials are preferred for

TPV cells. For example, the most prevalent material for TPV cells is GaSb with a 0.7 eV

bandgap [4-5]. Thermal imaging targets usually have a temperature approaching the

ambient; the radiation is mainly distributed over the MWIR and LWIR bands. Hence, the

infrared photodetectors fitted in infrared cameras are typically made of semiconductors

whose bandgaps are lower than 0.4 eV, e.g. InSb with a bandgap of 0.18 eV [8-9].

Photon energy (eV)


8 10 1 0.1
10
Spectral radiance (W.m-2.sr-1.m-1)

106
K
800

TPV
104
T=5

K
000

102
K
T=2

000
T=1

100
0K
30
T=

10-2
MWIR

LWIR

10-4

10-6
0.1 1 10 100
Wavelength (m)

Figure 1-2: Spectral radiation for blackbodies at various temperatures. The shaded
regions are of interest for applications such as solar cell, thermophotovoltaics and
thermal imaging.
3
An important feature of infrared radiation is that it is mostly blocked out by the

atmosphere. The two natural greenhouse gases in Earth’s atmosphere ─ water vapor and

carbon dioxide, absorb most of the infrared light. Only a few infrared wavelength ranges

are likely to travel through the atmospheric window, as shown in Figure 1-3 [10]. Hence,

the better view on the infrared world from ground-based infrared cameras is at infrared

wavelengths with a high atmospheric transmittance. The atmospheric window is also an

important consideration in free space optical communication (FSO) [11]. Because of this,

unlike the earlier mentioned division scheme, a more commonly recognized categorization

framework in the detector community is [10]: NIR (0.7-1 m), SWIR (1-3 m), MWIR (3-

5 m), LWIR (8-14 m), very long wavelength IR (VLWIR, 14-30 m), and far IR (FIR,

30-100 m) bands.

Figure 1-3: Atmospheric transmittance spectrum of infrared radiation. The figure is


from [10].

4
1.2 Overview of infrared thermophotovoltaic energy conversion

1.2.1 Background

In modern society, the overuse of diminishing fossil fuels has driven humanity to

develop alternative non-fossil energy source as well as ways of efficient use of fossil fuels.

TPV is a promising technology that can generate electricity from non-fuel resources such

as radioactive energy and concentrated sunlight. Potentially, it is also a more efficient way

to convert fossil fuel combustions with the ultimate efficiency approaching the Carnot limit

[4-5]. Although the expected high efficiency has not been fulfilled at the current stage, fuel

versatility still motivates further pursuit of this approach.

Early efforts on TPV were dedicated to developing military portable power sources

until the 1970s [12]. After the US Army decided to choose thermoelectrics as the priority

development project, TPV technology experienced a slow pace of development. However,

it still significantly profited from the progress of solar photovoltaics (PVs), particularly

from the rapid development of solar cells. Two examples are GaSb and InGaAs diodes that

are now the two prevalent TPV cells, while they were originally explored as the subcells

in multi-junction solar cells [13-14]. Besides, the experience in controlling the incident

radiation gathered from concentrated solar PV also promotes the development of TPV.

There was a regenerated interest in TPV in the 1990s for space, industry and military

applications. In industry, the use of TPV for waste heat recovery was conceived as a

prospective market niche. Over the same period, the near-field TPV concept started to

emerge, which utilized a sub-micron vacuum gap between the radiator and TPV cells [15-

17]. This method can appreciably improve the heat transfer between the radiator (or

emitter) and TPV cell. Another benefit of this displacement is enhanced incident power

5
density and the resulting higher conversion efficiency.

Until now, TPV is still in a research and development phase, and has not reached

commercial maturity, as it has been impeded by some research barriers. For example, in

the past, the lack of suitable high efficiency TPV cells was the main obstacle. Currently,

the main difficulty is the involvement of various areas of applied science. Unlike solar PVs,

the realization of a TPV system relies on experience in various aspects including optics

with filters, heat transfer over a small scale and materials tolerant of high temperature.

Despite these obstacles, some prototype TPV system demonstrations were reported, as

briefly summarized in Table 1-1.

6
Propane Propane Propane Butane Combined Radioisotope
Heat source JP8 combustion combustor- Sunlight Sunlight
combustion combustion combustion combustion module
emitter
Si/SiO2 tungsten
Emitter planar
SiC SixNy photonic Yb2O3 tungsten on SiC SiC tungsten photonic
material tungsten
crystal crystal

Emitter 1200 ℃ 770 ℃ 700 ℃ 1462 ℃ 1275 ℃ 1309 ℃ 1007 ℃ 1500 ℃ 1200 ℃
temperature
dielectric selective emitter, dielectric surface
Spectral selective selective surface selective
filter on TPV N/A dielectric filter on filter on TPV reflective
control emitter emitter reflective filter emitter
cell TPV cell cell filter

TPV cell GaSb GaSb GaInAsSb Si GaSb InGaAs InGaAs Ge GaSb

Cell

7
0.7 eV 0.7 eV 0.55 eV 1.4 eV 0.7 eV 0.6 eV 0.6 eV 0.66 eV 0.7 eV
bandgap
Cell 75 ℃ 25 ℃ water cooled 14 ℃ 25 ℃ 25 ℃ 50 ℃ 120 ℃ water cooled
temperature

Output
80 W 1 mW 344 mW 48 W 700 W 3.16 W 100 W 415 mW 9.2 W
power

Power 67 6.24
0.4 W/cm2 32 mW/cm2 344 mW/cm2 0.1 W/cm2 1.5 W/cm2 0.79 W/cm2 0.5 W/cm2
density mW/cm2 W/cm2

System emitter/module 20%


2.0% 0.08% 2.5%- 2.4% 16% projected 0.8% 6.2%
Efficiency 23.6% projected

Institution JX Crystal MIT MIT UNSW JX Crystals Inc. Bechtel Bettis Emcore Barcelona UVA
Table 1-1: Summary of some demonstrated TPV system performance.

Inc. Inc. University


Reference [18] [19] [20] [21] [22] [23] [24] [25] [26]
1.2.2 Active components in a TPV system

Solar PV and TPV are similar technologies as they both use PV cells to generate

electricity from high temperature radiation sources. One of the main differences between

the two is the geometry. A TPV system typically consists of a heat source, absorber and

emitter (or radiator), filter and TPV cells. Sometimes a cooling fan is included in the system

to prevent overheating of the TPV cells. The general operating principle of a TPV system

is illustrated in Figure 1-4. The radiation produced from the heat source (either

radioisotope, or fuel combustion or concentrated sunlight) is absorbed by the absorber and

subsequently radiated by the emitter. The filter then converts the broadband radiation

spectrum into a narrowband emission spectrum tuned to the response of the TPV cell.

Afterwards, the radiation is captured by the TPV cell and converted into electricity. In some

cases, the absorber is coupled with a selective emitter with a narrow range of wavelength

emission, thus the filter is no longer needed. Besides the filter, the other approach of

spectral control is to reflect out-of-band photons back to the emitter via reflectors in front

of or behind the TPV cell.

Figure 1-4: Schematic illustration of a TPV system.

The system efficiency is affected by the performance of the individual components

as well as the interaction between them. To build a reliable TPV system, the operations of

8
the components need to be optimized. For example, since the heat source in a TPV system

is generally at 1000-2000 K, the emitter should have high thermal stability. There are

several suitable materials for emitters, classified as ceramics [18-19, 23], metals [22, 24,

26-28], metal oxides [21, 27, 29-30], or other novel materials [31-33]. Conventional metals

and ceramics tend to have broadband emission. In contrast, the pure polished metal oxides

(e.g. rare-earth oxides) can have narrow-band emission. Among these materials, tungsten

is currently the most used, since its emission spectrum is well matched with the bandgap

of GaSb [22, 24, 26-27]. Novel emitters based on artificial structures such as photonic

crystals and metamaterials have the advantage of very narrow emission bands, but at the

expense of more complex structures than conventional emitters [31-33].

1.2.3 Thermophotovoltaic cells

In the early period of development, investigations of TPV cells were mainly

focused on Si [34] and Ge [35]. The low cost and mature production phase of Si made it a

competitive material. However, the bandgap of Si is too wide for efficient conversion of

IR radiation, because most of the photons possess energies lower than its bandgap and are

unable to excite electron-hole pairs. Ge has a narrower bandgap than Si, but its crystal

structure can be easily damaged at high temperatures. Also, the recombination losses in Ge

cells are very high due to the large effective mass and high carrier concentration. Current

generation of TPV cells are mainly made of GaSb [22, 36], InGaAs [23, 37-38], GaInAsSb

[39-40] and InGaSb [41-42]. Among them, GaSb is often regarded as the most suitable

choice for TPV generators. GaSb has a similar bandgap (~ 0.72 eV) with Ge, which allows

it to respond to light with longer wavelengths. Under a perfectly filtered blackbody

(T=1350 K), an efficiency of ~30% was projected for GaSb cells [36].

9
Up to now, without a filter, the best reported efficiencies for TPV cells are 24% for

a 0.6 eV InGaAs cell on InP [23, 37] and 19.7% for a 0.53 eV GaInAsSb cell on GaSb

[39]. These records were measured with a ~1000 °C broadband blackbody radiator and

with a front surface reflector for recovering unabsorbed below-bandgap photons. The

bandgap of a ternary InGaAs diode, exactly lattice matched to InP, is 0.74 eV, but it

underperforms GaSb TPV cells [38]. By changing the ratio of Ga to In, the bandgap of

InGaAs can be tuned from 0.55 to 0.6 eV with some strain from the InP substrate. The

strained InGaAs cells generally outperform GaSb cells [23, 37]. Quaternary GaInAsSb

alloys latticed-matched to GaSb have bandgaps theoretically ranging from 0.25 and 0.75

eV. The fabricated GaInAsSb cells on GaSb substrate have bandgaps from 0.5 to 0.6 eV

[39-40]. The performance of these TPV cells generally falls behind InGaAs TPV diodes.

Also, the manufacture of GaInAsSb cells is expensive and is not commercially available.

Aside from the above-mentioned materials, other TPV cell research interests are

narrow bandgap (0.4 eV) materials such as InAsSbP [43-44], InAs [45-46], InSb [47] and

InAsSb [48]. These narrow bandgap cells have a low open-circuit voltage and fill factor,

as well as a poor efficiency at room temperature as shown in Table 1-2. Even some studies

are only for proof-of-concept demonstrations of potentials. To achieve optimal efficiency,

they were cooled down to overcome some of the downsides [47]. The performance limiting

factors in narrow bandgap TPV cells are identified theoretically and experimentally in

Chapters 3 and 5, respectively. Nevertheless, theoretical calculations following the detailed

balance principle showed that the optimal choice for TPV cell bandgap energy is between

0.2-0.4 eV [49-50]. In the next subsection, a similar bandgap range is calculated from the

thermodynamic perspective. Additionally, up to now, relatively less research work has

10
InAsSbP InGaSb GaInAsSb GaInAsSb InGaAs InGAAs GaSb GaSb Ge Si
MIM

0.35 0.56 0.549 0.53 0.74 0.6 0.73 0.73 0.66 1.1

21 ℃ 27 ℃ 27 ℃ 27 ℃ 30 ℃ 25 ℃ 25 ℃ 25 ℃ 300 K 300 K

1373 K 3250 K 1039 ℃ 1350 K 1200 1100 ℃ 2300 ℃


N/A arc lamp 950 ℃ SiC
blackbody blackbody tungsten blackbody ℃ SiC tungsten radiator
types of single-absorber TPV cells.

front front front perfect front front &


without without without surface surface surface band edge surfac back N/A
filter reflection filter filter e filter surface
reflection

11
0.25 3.0 3.5 2.9 0.288 0.1 3.0 2.83 1.67 9.52

34% 61% 66% 67% 65% 66.2% 75% 73% 67% N/A

120 mV 270 mV 313 mV 306 mV 405 mV 12.5 V 500 mV 477 356 mV N/A
mV

0.024 0.49 0.723 0.58 0.08 0.79 1.26 0.98 0.4 10

0.18% N/A N/A 19.7% 12.4% 24% projected 21% 16% 26%
30%

[44] [42] [40] [39] [38] [23, 37] [36] [22] [35] [34]
Table 1-2: Summary of various TPV technologies, classified by absorbing material.
been done towards narrow bandgap cells; there is still great potential in further

development of them. Table 1-1summary some important device performances of various


Table 1-2 continued

blackbody

without
950 ℃

60 mV
20 ℃
InAs

37%
0.32

0.89

0.02

[45]
3%
blackbody

17.4 mV
without
800 ℃

0.35%
300 K
InAs

1E-3
25%
0.36

0.23

[46]
1248 K IR

without

83 mV
7.2E-3

3.8E-4
source
InSb

77 K

64%
0.23

N/A

[47]
blackbody

projected
InAsSb

without
1500 K
27 ℃
0.286

39.88

162.8

16%
N/A

N/A

[48]
Illuminati

Reference
Efficiency
on source
Material

Spectral

(W/cm2)
(A/cm2)

FF (%)
control
Eg (eV)

Cell T

Pout
Voc
Jsc

1.2.4 Thermodynamic analysis of thermophotovoltaic cell efficiency

In single-absorber TPV cells, without spectral control, the major energy loss arises

from two mechanisms. The first mechanism is that photons with energies lower than the

bandgap energy are not converted. The second mechanism is due to photons with energy

higher than Eg. These photons contribute only Eg and the excess energy is released via hot

carrier heating. Theoretically, both losses can be minimized by means of spectral control,

but this would lead to low, often not acceptable, power densities, and low system

efficiencies. Without spectral control, there is a tradeoff between the intensified below-

bandgap loss and mitigated thermalization loss as the cell bandgap increases, implying an

optimal choice of the bandgap to maximize cell efficiency. Several well-established models

exist to identify the ideal cell bandgap, as well as to predict the upper limits of TPV

efficiency and power density. The efficiencies predicted by different models are compared

12
in [51]. Some models are based on empirical values for the saturation current density [52-

54]. Some models refer specifically to solar TPV conversion [55-57]. The usual assumption

made in these models is full incident spectrum (no spectral control). Here, the ultimate

efficiency and optimal bandgap are calculated by extending Shockley and Queisser’s [58]

limit for solar cells (also known as the detailed balance limit) to the TPV case.

In TPV systems, ideally, there is no radiation lost since the radiator and emitter are

closely arranged. The solid angle subtended by TPV cell can be 4π sr compared to the

6.85×10-5 sr for conventional solar cells. Thanks to this arrangement, from the Stefan-

Boltzmann law (Equation 1-3), the radiation density can reach 16-91 W/cm2 incident on

the TPV cell for a heat temperature at 1000-2000 K, while the average solar radiation on

earth’s surface is only 0.1 W/cm2. To apply detailed balance analysis, several assumptions

need to be made to simplify the scenario. First, there are no non-radiative channels in the

TPV cell; carrier recombination and generation are exclusively radiative. Second, the

bandgap is a sharp demarcation of absorption: photons with above-bandgap energy are

completely absorbed, while below-bandgap photons are hardly absorbed. Third, when a

bias voltage (V) is applied to the TPV cell, it will emit photons as a blackbody with a

chemical potential of eV.

Under these assumptions, the current flowing in a TPV cell under a bias voltage (V)

can be given by:

2𝜋𝑞 ∞ 𝐸2 𝐸2
𝐽 (𝑉 ) = ∫ [ 𝐸 − 𝐸−𝑒𝑉
] 𝑑𝐸 (1-4)
ℎ 3 𝑐 2 𝐸𝑔𝑒𝑥𝑝( )−1 𝑒𝑥𝑝( )−1
𝑘𝑏 𝑇𝑠 𝑘𝑏 𝑇𝑐𝑒𝑙𝑙

where q is electron charge, Ts and Tcell are the temperature of the source and cell,

respectively. The first term in the integral stands for the photocurrent due to light

13
absorption. The second term represents the reverse dark current originated from electron

recombination. Based on Equation 1-4, the calculated efficiencies of TPV cells for various

source temperatures are shown in Figure 1-5. The inset within the figure is the optimal

bandgap that maximizes the efficiency as a function of the source temperature. As can be

seen, the optimal bandgap for a source temperature at 1000-2000 K is in the range of 0.18-

0.37 eV, well less than the bandgap of current mainstream TPV cells made of GaSb,

InGaAs and GaInAsSb. The corresponding maximum efficiency is between 22% and 33%,

remarkably higher than the actual efficiencies of narrow bandgap TPV cells such as

InAsSbP, InAs and InSb (See Table 1-2). This is because the detailed balance limit is a

very idealized and an overestimated limit, as the analysis buries many practical factors. For

example, in real narrow bandgap devices, non-radiative recombination such as Auger and

Shockley-Read-Hall (SRH) tend to prevail over radiative recombination. These non-ideal

factors will seriously limit overall device performance. In Chapter 3, the efficiency limits

of narrow bandgap TPV cells will be re-evaluated by acknowledging some of the practical

factors.

35
Optimum bandgap (eV)

1.2
1.0
30
Conversion efficiency (%)

0.8
0.6
25 0.4
2500 K 0.2
20 0.0
0 2000 4000 6000
2000 K Source temperature (K)
15
1500 K
10

5 1000 K

0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Bandgap (eV)
Figure 1-5: The calculated efficiencies based on Equation 1-4 for various blackbody
temperatures. The inset shows the optimal bandgap that maximizes the efficiency.

14
1.3 Overview of infrared detectors

1.3.1 Background

The historical track record of modern infrared detectors (or systems) is shown in

Figure 1-6. Modern development of infrared detector was possible after the discovery of

lead salt family (PbSe and PbS) [59]. Thereafter, further researches launched the

development of various detecting materials (or structures) including but not limited to: Ge

[60], InSb [61], Si [62], HgCdTe [63], InGaAs [64], quantum well infrared detector

(QWIP) [65], quantum dot infrared detector (QDIP) [66], barrier photodetector [67] and

type-II superlattice (T2SL) [68], as shown in Figure 1-6. Also, there are three generations

of IR detection systems that are generally considered in civil and defense applications. The

first generation is scanning systems with single and linear units. The second generation

includes focal plane array (FPA) technology with monolithic and hybrid detectors.

Combined with the read-out circuit in the FPA, a multiplexing function can be achieved.

The third generation has orders of magnitude more pixel elements than the second

generation FPAs. In addition, a multicolor function and other superior on-chip features are

possible in the third generation.

Figure 1-6: The development history of modern infrared detectors and systems.

15
As mentioned in Section 1.1, MWIR technology finds its application mainly in gas

sensing. Specifically, there are three thriving civil application areas of mid IR gas sensors:

environmental monitoring, industrial process control and medical diagnosis. Many

molecules and gases exhibit strong absorption characteristics in the mid IR band, as shown

in Figure 1-7 [69]. In addition, thanks to the much stronger absorption, gas sensing systems

based on MWIR and LWIR optoelectronics have an inherent advantage over NIR

counterparts in terms of sensitivity (or detection limit). For example, the detection limit for

CH4 at 3.26 m is 1.7 ppb compared to 600 ppb at 1.65 m. Another more contrasting

example is CO2. The detection limit for this greenhouse gas is 0.13 ppb at 4.23 m, while

it is 3000 ppb at 1.55 m. Despite the real advantages, MWIR and LWIR optoelectronics

had received considerably much less research attention than NIR optoelectronics. The main

reason for this difference is the revolution of communication systems with the advent of

optical fiber systems, which directly lead to the rapid development of NIR optoelectronics.

Nevertheless, the impressive accomplishments in MWIR and LWIR lasers such as

quantum cascade lasers (QCLs) [70-71] and interband cascade lasers (ICLs) [72-73] will

significantly promote the research and development of MWIR detectors.

16
10

HCOOH
CH3OH
T=300 K

C2H2
C2H4

H2O
CH4

CO2

NH3

NO2
H2S
P=1atm

CO

NO

O2
O3
Absorptance (Ln(I/Io))
1

0.1

0.01
2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Wavelength (m)
Figure 1-7: Mid infrared absorption spectra of some molecules and gases. Data were
collected from [69].

1.3.2 Photon detection vs thermal detection

Most infrared detectors can be classified into two categories [9, 74-76]: photon

detectors and thermal detectors. Photon detection occurs when incident photons, absorbed

by the detecting material, excite free electron-hole pairs. In most instances, the material is

a bulk semiconductor or a low dimensional material such as those mentioned in Subsection

1.3.1 (refer to Figure 1-6). The electrical signal arises from the change of electron

distribution inside the detector. Thermal detection is defined as the mechanism that change

some measurable property of the detecting material due to the temperature increase of that

material resulting from the absorption of radiation, as illustrated in Figure 1-8 [76]. Among

the various thermal mechanisms, the most important three are the thermoelectric effect, the

resistive bolometric effect, the pyroelectric effect and its modification known as the

ferroelectric bolometer [9, 74-76]. Although many other mechanisms were proposed, only

these three have been shown to be practical to date. The electrical output in a resistive

bolometer (typically made of VOx) arises from the change of its electrical resistance as the

17
temperature rises. The pyroelectric effect is demonstrated with certain materials which

could generate electrical polarization that can be measured as an electrical charge on the

opposite face. The thermoelectric effect, i.e. Seeback effect, is the buildup of an electrical

potential across a temperature gradient; the resulting voltage is proportional to the

temperature difference between the hot and cold ends.

Figure 1-8: Block diagram of a thermal detector. Figure from [76].

In general, thermal detectors do not require cryogenic cooling, while the photon

detectors in MWIR and LWIR regions are cooled to suppress thermal generation of

carriers. The coolers are normally costly devices, making the detection system (e.g. an

infrared camera) more expensive than uncooled systems. Also, the coolers make the

systems bulky, and more steps are needed in manufacturing, therefore reducing the yields.

In addition, photon detectors are selective in wavelength, while thermal detectors have no

wavelength dependence. Nevertheless, cooled systems based on photon detectors are

incredibly more sensitive than uncooled thermal systems, as illustrated in Figure 1-9. As

can be seen, the image captured by a cooled infrared camera has a quality much better than

that created by an uncooled camera. In addition, the imaging speeds of cooled systems are

much higher than uncooled systems. The high-speed thermal imaging of cooled systems

18
allows capturing frame rates as high as 62000 fps.

Figure 1-9: Images created by uncooled and cooled infrared cameras. The figures are
from [77].

1.3.3 Examples of photon detectors

Photon detectors can be further subdivided into photoconductive (PC) and

photovoltaic detectors according to satisfaction or violation of symmetry in the

configuration. The simplest form of a symmetric PC photodetector consists of a slab of

semiconductor, as shown in Figure 1-10(a). By contrast, the structure of a PV photodetector

is asymmetric, in most cases, it is made of a p-n junction [See Figure 1-11]. Such an

asymmetric structure enables the separation of photo-generated carriers without applying

external bias. The resulting difference between PC and PV photodetectors is the operation

bias: PV detectors can operate at zero bias, while PC detectors require an external bias to

initiate the operation. In addition to the simplest semiconductor slab, a comparably more

complex PC photodetector structure is a QWIP [65, 78], as illustrated in Figure 1-10(b).

As can be seen, the basic elements of a QWIP are quantum wells (QWs) separated by wide

barriers. The incident light is absorbed via intersubband transitions of electrons within the

QWs. Once the electrons are optically excited into the continuous upper states, they will

be measured as a signal current. However, to collect these electrons, an external bias needs

19
to be applied and the signal current responds in an almost linear fashion to the applied bias.

Among various types of QWIPs, technology based on GaAs/AlGaAs multiple QWs is most

mature [65, 78]. QWIP detectors have relatively low quantum efficiencies, generally lower

than 10%, partially resulting from the selection rule of intersubband transitions in

conduction band [79].

Figure 1-10: Schematic diagram of (a) a photoconductive photodetector made of a


semiconductor slab, (b) a quantum well infrared photodetector based on bound-to-
continuum transitions.

The most common configuration for PV detectors is a single p-n junction as shown

in Figure 1-11. The optically excited electrons and holes are separated by the built-in

electric field in the depletion region and then contribute to the signal current. One route to

increase light absorption in a p-n junction is to sandwich a thick intrinsic layer between the

p- and n- doped layers, forming the so-called p-i-n structure. Some p-i-n detectors can use

avalanche multiplication but they must be reverse-biased [80].

20
Figure 1-11: Schematic diagram of a PV detector made of a single p-n junction

Another simple but refined PV detector technology is the barrier photodetector [67,

81]. Among various types of barrier photodetectors, the most popular one is nBn detector

as shown in Figure 1-12(a). Such a configuration is conducive to reducing majority-

electron dark current, while the signal current from minority holes is unaffected. The

barrier also takes a role to reduce the surface current, a benefit equivalent to self-

passivation. In addition, the absence of a depletion region eliminates the excess dark

current associated with the SRH process and trap-assistant tunneling. A special

modification of the nBn detector is the complementary barrier infrared detector (CBIRD)

[82] with an additional hole barrier introduced in the valance band, as shown in Figure 1-

12(b). The electron and hole barriers complement one another to impede the flow of dark

current. As with nBn detectors, the benefit of reduced dark current from elimination of a

depletion region also extends to CBIRD detectors.

Figure 1-12: Schematics of (a) an nBn barrier detector and (b) a complementary
barrier infrared detector; the biases are applied to improve carrier collection.

21
Apart from p-n junction and barrier structures, there is another more complex

photodetector operating in PV mode: the quantum cascade detector (QCD) [83-84], as

shown in Figure 1-13. As an intersubband detector, the QCD is a special variation from the

standard QWIP structure. The QCD is configured to operate in PV mode to reduce the dark

current present in a QWIP. However, despite this improvement, the dark current in QCDs

is still relatively high due to the short carrier lifetime (~ ps at 300 K) in intersubband

transitions. This fundamental problem severely undermines the ability to achieve a high

detectivity for QCDs especially at high temperatures, which will be discussed in detail in

Chapter 6. A brief summary of various photovoltaic photodetectors is presented in Table

1-3.

Figure 1-13: Schematic diagram of a quantum cascade detector.

22
Table 1-3: Summary of various photovoltaic photodetectors, classified by detecting
material.

Material or cutoff or peak Ri D*


T (K) R0A or Jd Ref
Structure (m) (A/W) (Jones)
Ge 1.55 300 10 A/cm2 0.1 N/A [60]

InGaAs 1.55 295 10 nA/cm2 0.62 1.0×1012 [85]

PbS 3.0 300 0.1 ·cm2 1.22 3.0×109 [86]

PbSe 4.0 300 6.5 ·cm2 1.6 1.0×109 [87]

InAs 3.4 295 8.1 ·cm2 0.54 1.2×1010 [88]

InSb 4.0 77 6.3E6 ·cm2 N/A 1.0×1011 [89]

InAsSb 4.0 300 0.19 ·cm2 0.76 2.6×109 [90]

HgCdTe 3.0 300 1.0 ·cm2 0.5 6.5×109 [91]

HgCdTe 5.0 300 0.01 ·cm2 1.0 2.0×109 [91]

HgCdTe 8.0 230 2E-4 ·cm2 0.8 4.0×108 [91]

HgCdTe 10.6 230 1E-4 ·cm2 0.4 2.0×108 [91]

InAs/GaSb SL 2.2 300 15 ·cm2 0.57 1.7×1010 [92]

InAs/GaSb SL 4.2 150 5.1E3 ·cm2 1.9 1.1×1012 [93]

InAs/GaSb SL 9.9 77 1.4E4 ·cm2 1.5 1.1×1011 [82]

InAs/InAsSb SL 10.0 77 119 ·cm2 4.47 2.8×1011 [94]

InAs/InAsSb SL 14.6 77 0.84 ·cm2 4.8 1.4×1010 [95]

QCD 5.4 300  ·cm2 7E-3 2.5×1010 [96]

QCD 8 300 0.028 ·cm2 1.7E-2 1.4×107 [97]

1.3.4 Performance metrics for infrared detectors

The most important performance coefficient for infrared detectors is the specific

detectivity D* that describes the smallest detectable signal. It equates to the reciprocal of

noise-equivalent power (NEP, in unit of W) that is normalized per square root of frequency

bandwidth and detector area. That is, the expression of D* is given by:

√∆𝑓𝐴
𝐷∗ = (1-5)
𝑁𝐸𝑃

23
where Δf is the bandwidth and A is the detector area. The unit of D* is cm·Hz1/2/W or more

frequently it is expressed as Jones. The noise equivalent power NEP is the incident flux

required to generate an output signal current/voltage equivalent to the noise

current/voltage. For most photon detectors, the noise current is used to define NEP:

NEP = 𝐼𝑛 /𝑅𝑖 (1-6)

where In is the noise current, and Ri is current responsivity that is equal to 1.24·QE/𝜆 (QE

is quantum efficiency). The noise sources in a photodetector include low-frequency noise,

Johnson noise, shot noise and generation-recombination (G-R) noise. In some instances,

the dominant noises are Johnson and shot noises. They occur as results of thermal

fluctuation during carrier motion (Johnson noise) and statistical fluctuation of thermal

generation of carriers (shot noise). Since the two noises are not coupled, the total mean

square noise current is the sum of both noise currents:

4𝑘𝑏 𝑇
𝑖𝑛2 = ∆𝑓 + 2𝑒𝐽𝐴∆𝑓 (1-7)
𝑅0

where R0 is zero-bias resistance and J is the dark current density. The first term in this

equation describes Johnson noise and the second term corresponds to shot noise.

Substituting Equations (1-6) and (1-7) into Equation (1-5), one can obtain the expression

of Johnson- and shot-noise limited detectivity:

𝑅𝑖
𝐷∗ = (1-8)
√4𝑘𝑏 𝑇⁄𝑅0 𝐴+2𝑒𝐽

From this equation, the D* can be improved either by reducing the noise or by increasing

the QE. The most effective way to maximize D* in conventional single-absorber detectors

is to increase the QE. In contrast, the D* can be effectively improved in multistage detectors

via noise reduction, as will be described in Chapter 2.

24
1.4 Dissertation organization

Chapter 2 concentrates on the fundamentals of the interband cascade (IC) device

family including IC lasers (ICLs), IC infrared photodetectors (ICIPs) and IC

thermophotovoltaic (ICTPV) cells. The main purpose of this chapter is to explain the

historic development, constituent materials, operation principles and basic theories of these

quantum engineered devices. It commences with the introduction of the 6.1 Å material

system: InAs, GaSb and AlSb and their unique properties. Subsequently, it presents the

attractive features of IC structures when functioning as lasers, PV cells and detectors.

Chapter 3 presents the theoretical comparison between single-absorber and

multistage ICTPV cells. The efficiency limits are calculated considering some practical

factors that apparently violate the assumptions made in the idealized thermodynamic

analysis in Subsection 1.2.4. This is in keeping with the relatively low efficiencies

demonstrated for current narrow bandgap TPV technologies. Several limiting factors are

identified, which turn out to be closely associated with short carrier lifetime, small

absorption coefficient and high dark saturation current density.

After the theoretical comparison, experimental details and comparisons between

single-absorber and multistage ICTPV cells are given in Chapter 4. A set of three TPV

cells with single-absorber and multistage architectures are characterized and analyzed in

detail. The experimental data confirmed the advantages of the multistage IC architecture

for TPV cells. It is shown that a multistage IC structure can be successful in resolving the

diffusion length limitation in single-absorber cells, and to achieve a collection efficiency

approaching 100% for photogenerated carriers.

Speculatively, the performance should be better for ICTPV cells with more stages,

25
as will be shown in Chapters 3 and 4. The initial goal of the fabricated four ICTPV devices

in Chapter 5 is to examine this speculation. However, the experimental study reaches the

opposite conclusion that significantly increasing the number of stages may penalize device

performance. Detailed device characterization and analysis are developed to explain this

contradiction, as well as to identify and quantify three factors: current mismatch, material

quality and collection efficiency.

Chapter 6 and 7 are mostly focused on the deep knowledge and strategies of IC

infrared photodetectors. Chapter 6 first describes an effective and simple approach to

extract carrier lifetime in the InAs/GaSb SLs. The developed method is applied to some

ICIP devices to extract the carrier lifetime at high temperatures. This chapter then

introduces a unified figure of merit for interband and intersubband devices, i.e. the

saturation current density J0. The significance of J0 on the performances of detectors and

PV cells is illustrated with measured D* and calculated Voc, respectively.

Chapter 7 first provides a comparative study of two sets of four ICIP devices with

current-matched and noncurrent-matched configurations. This study demonstrated the

necessity of current matching in ICIPs to maximize the utilization of absorbed photons for

an optimal responsivity. Following this study, the universally observed electrical gain in

noncurrent-matched ICIPs is explained with a unique mechanism. Furthermore, a theory is

developed to quantitatively describe the electrical gain, and the calculations agree well with

experimental data. Finally, Chapter 8 gives some prospective points for the future work

arising from these studies.

26
2 Chapter 2: Sb-based interband cascade devices

2.1 6.1 Å Semiconductor family

Interband cascade (IC) optoelectronic device is an umbrella term that refers to IC

lasers (ICLs) [72-73, 98], IC infrared photodetectors (ICIPs) [99] and ICTPV cells [100].

The materials that make up these devices are the 6.1 Å material system including InAs,

GaSb, AlSb and their related alloys. The crystal structures of the three compounds are all

zin blende. The main advantages of the three materials are small lattice constant mismatch

and similar growth windows. Specifically, the lattice constants are respectively 6.0584,

6.0959 and 6.1355 Å for InAs, GaSb and AlSb. Thus, these binary materials can be

incorperated together to the same heterostrucutre with low densities of defets and

dislocations. The bandgaps of them and the related alloys are between 0.41 eV (for InAs)

and 1.70 eV (for AlSb) as shwon in Figure 2-1(b). This bandgap range is of great interst

for the design of optoeelctronic devices in the SWIR and MWIR spectral regimes.

Figure 2-1: (a) Bandgap, lattice constant and (b) band alignment of the 6.1 Å
semiconductor materials.

The operations of IC devices are possible due to the unique properties of the 6.1 Å

materials. One of the key properties is the type-II broken-gap alignment between InAs and

27
GaSb. As shown in Figure 2-1(b), the conduction band edge of InAs is about 150 meV

lower than the valence band edge of GaSb. The benefits of this type of misaligned structure

are twofold. It enables smooth transition of electrons from valence band in GaSb layer to

conduction band in InAs layer without energy loss [101-102]. Also, due to this alignment,

the InAs/GaSb type-II SLs (T2SLs) have very flexible engineering capability [103-106]

and can cover a wide range of infrared spectra from SWIR to VLWIR. On the other hand,

the InAs/AlSb interface forms a type-II staggered alignment where the conduction band

edge of InAs is slightly above the valence band edge of AlSb. This staggered alignment,

tougher with the wide bandgap of AlSb, results in an extremely large conduction bandgap

offset of nearly 1.45 eV. This enables the realizations of very deep quantum wells and very

large tunneling barriers. Because of this feature, InAs/AlSb heterostructure has been

frequently used in resonant interband tunneling diodes (RITDs) [107-108] and short-

wavelength QCLs [109-110].

2.2 Interband cascade lasers-the historic origin

2.2.1 History and operation principle

Both ICIPs and ICTPV cells spring from ICLs, so for better understanding of their

evolutions and operations, first a brief review of ICL is given before moving on to ICIPs

and ICTPV cells. The concept of ICL was originally proposed in 1994 [98]. The main

innovation behind the concept is the capability to manipulate electron transport to form an

interband cascade scheme, whereby a single electron can generate multiple photons based

on interband transitions, as shown in Figure 2-2. Prior to the proposal of ICL, another

cascade laser, i.e. QCL, based on intersubband transitions was demonstrated in the same

28
year [70]. Both ICL and QCL consist of multiple cascade stages connected in series, and

each cascade stage ideally acts as an individual photon generator. However, unlike QCLs

in which the photons are generated via intersubband transition, ICLs use interband

transitions for active generation of photons. The injected carriers in ICLs relax to the lower

energy level at a rate much slower than in QCLs, so the threshold condition can be much

easier to establish in ICL. This is because the interband transitions in ICLs are characterized

by radiative, Auger and SRH processes, in which carrier lifetimes are on the order of

nanosecond. In contrast, the intersubband relaxation in QCLs is accompanied with

longitudinal phonon emission and has a picosecond time scale. The use of interband

transition in ICLs makes the threshold current and input power much lower than that in

QCLs. Even compared with other types of mid IR lasers such as Sb-based type-I QW diode

lasers [18-19] and II-VI lead salt lasers [20-21], the threshold current and input power of

ICLs are considerably lower. This makes them the preferred option for applications where

low power consumption is strongly prioritized.

Figure 2-2: Illustration of the photon emission and cascading effect in an interband
cascade laser. Figure from [115].

Compared to the conventional diode lasers, the cascade design requires a higher

voltage to reach threshold. This is because each cascade stage needs to consume a voltage

29
to invert the population. Nevertheless, the current required to trigger the lasing action is

significantly reduced, as multiple photons are generated for each injected electron. This

tradeoff between voltage and current is in favor of reducing Ohmic losses from the series

resistance, especially for high-power semiconductor lasers operating with high currents. In

this regard, IC structures can be beneficial to improving the overall power efficiency by

lowering the operating current.

The active core of an ICL is schematically shown in Figure 2-3. In each stage of an

ICL, the active region is sandwiched between the electron and hole injectors. The active

region, the electron injector and the hole injector are typically made of GaInSb-InAs “W”

QW, multiple InAs/AlSb QWs and multiple GaSb/AlSb QWs, respectively. Under a

forward bias, the electrons are injected from the injector into the conduction band of the

active region. The injected electrons are confined in the active region by the AlSb barriers

and transit to the valence band via photon emission. The transited electrons subsequently

enter the electron injector in the next stage via interband tunneling through the broken gap

between InAs and GaSb. This process is orders of magnitude faster than the interband

transition (~1 ns) in the active region. Therefore, the electrons relaxed to valence band in

the active region are efficiently swept out and population inversion can be readily achieved.

30
Figure 2-3: Band diagram of the active core for an interband cascade laser. Figure
from [73].

2.2.2 Current status of ICL performance

Since its first demonstration, the performance/capability of ICL has been

transformed a lot. In CW operation, ICLs can cover a broad range of wavelengths

extending from 2.8 m to 6.0 m at room temperature (RT) or above [116-120]. Further

preparation for high temperature operation with a longer wavelength is in progress [115,

121-124]. Typically, the epitaxy growth of ICL is done on either a GaSb [72-73, 116-118,

125-126] or InAs [115, 119-124] substrate. As wavelength increases, the InAs/AlSb SL

cladding layers in GaSb-based ICLs need to be thick, in order to provide strong optical

confinement. This is problematic for heat dissipation, as InAs/AlSb SLs have very low

thermal conductivity (~2.7 W/m·K). Also, thick InAs/AlSb SLs are challenging in MBE

growth due to many shutter movements. These issues can be readily resolved in InAs-based

ICLs wherein the SL cladding layers are replaced with highly doped InAs layers [115, 119-

124]. Besides, this approach offers another benefit: the low refractive index for highly

31
doped InAs layers increases the optical confinement. Figure 2-4 shows the room

temperature threshold current densities for both InAs- and GaSb-based ICLs in the

wavelength range of 2.7-7.2 m. Most of the data are collected in pulsed modes at 300 K.

As can be seen, the technology maturity for GaSb-based ICLs is well demonstrated in the

3-4 wavelength region. By comparison, InAs-based ICLs aim to cover wavelengths longer

than 4 m. In the 4-5 m wavelength region, the two types of ICLs have comparable

performances. However, as the wavelength goes beyond 6 m, InAs-based ICLs

outperform GaSb-based ICLs in terms of threshold current density.

Figure 2-4: Room temperature threshold current density for both InAs- and GaSb-
based broad-area ICLs. Figure is from [124].

2.3 Interband cascade thermophotovoltaic devices

2.3.1 InAs/GaSb type-II superlattices

The ideal of InAs/GaSb T2SLs was first introduced in 1977 [127]. Ten years later,

it was proposed for detector application [68]. Since then, it has been recognized as a

promising material for mid IR detectors due to the predicted reduction of Auger

recombination rates [128-130]. Measurements of the Auger recombination coefficient by

32
pump-probe transmission likewise showed suppressed Auger rates compared to bulk

materials [131]. Factors considered to contribute to this suppression include strain induced

splitting in valence band, quantum confinement and off-resonance positions of the spin-

orbit split-off band. On the other hand, as shown in Figure 2-5(a), the electrons and holes

are confined separately in InAs and GaSb layers, which reduces the light absorption. The

bandgap of InAs/GaSb SLs is the difference between the minibands for electrons and holes.

The miniband for holes is very narrow since the effective mass of holes is large. Moreover,

the energy level of hole is almost quasi-constant with GaSb well thickness. Hence, the

bandgap of InAs/GaSb SLs is mainly controlled by conduction band level, via the change

of InAs and GaSb layer thicknesses.

1.2 2.0 (b) Al70In30Sb


(a)
1.0
1.5
0.8
Energy (eV)
Energy (eV)

0.6 Eg 1.0

0.4 Eg
0.5
0.2

0.0 0.0
InAs GaSb InAs GaSb
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Distance (nm) Distance (nm)

Figure 2-5: Band structure, minibands and wavefunctions of electrons and holes for
(a) InAs/GaSb superlattice and (b) M-shape Al(In)Sb/GaSb/InAs/GaSb/Al(In)Sb SL.

When the bandgap is wide, the binary InAs/GaSb SL is not the preferred option.

This is because a wide bandgap necessitates thin InAs layers, which can make the bandgap

very sensitive to layer variations during growth. Also, it can cause interface

mixing/roughness, as lower material and interface quality were reported in literature [132-

133]. A solution to these issues is inserting thin Al(In)Sb layer in the middle of GaSb layers,

forming the so-called M structure [104, 134-135]. The letter “M” stands for the shape of

33
the band alignment of the Al(In)Sb/GaSb/InAs/GaSb/Al(In)Sb layers, as shown in Figure

2-5(b). There are several potential advantages of the M-shape SL. First, the AlSb blocking

barrier can reduce the dark current and improve the R0A product of devices made from this

structure [104]. Second, the AlSb layer can compensate the tensile strain induced by InAs

layers. Third, it reduces the wavefunction penetration into barrier layers, thereby narrowing

the minibands and allowing a sharp increase of absorption coefficient near bandgap. In

addition to M-shape SL, there are other modifications of the normal InAs/GaSb SL,

namely, the W- [105] and N-shape SLs [106]. These various modifications manifest the

flexible heterostructure design of T2SLs based on InAs/GaSb/AlSb material system.

2.3.2 Operation principle of ICTPV cells

The photovoltaic operation of IC structures was first demonstrated with devices

that were fabricated from ICL wafers [136]. The light absorption region was simply

composed of a single pair of coupled quantum wells; small absorption was revealed by the

measured low responsivity of the fabricated PV detectors. To address this problem, it is

necessary to make some modifications to the structure. One prominent alteration is

replacing the quantum well absorber with much thicker InAs/GaSb T2SLs [99-100]. This

structural change was shown to be very effective to improve light absorption characteristics

and overall device performance [99]. Further refinement of the structure was made on the

hole injection region: additional QWs are added to better block intraband tunneling of

electrons, thus reducing the dark current density [137].

Overall, the structure of an ICTPV cell is roughly similar with that of an ICL. Each

stage of an ICTPV cell consists of an electron barrier (eB), a hole barrier (hB), and a T2SL

absorber sandwiched between the two barriers, as shown in Figure 2-6. The electron and

34
hole barriers correspond to the hole and electron injectors in an ICL structure, respectively.

They are assigned different names in ICLs and ICTPV cells to distinguish between their

functions in the two structures. In ICTPV cells, the unipolar barrier plays a function as

blocking the namesake carrier while allowing smooth transport of the otherwise carrier, as

shown in Figure 2-6. The unipolar barriers work as intended because of the proper energy

alignment at the interfaces. For example, the first electron miniband energy level of the

T2SLs lies within the bandgap of GaSb layer in the electron barrier, therefore the photo-

generated electrons can only move to the hole barrier. This provides a novel way for

constructing PV devices with perfect current rectification without appealing to p-n

junctions.

The basic operation principle of an ICTPV cell is illustrated in Figure 2-6(a). If the

concept of hole is disregarded, the electron and hole barriers serve as the tunneling and

relaxation regions for electrons, respectively. As shown, electrons optically excited in the

absorber first travel to the hole barrier by diffusion. Following the diffusion process, the

electrons then relax to the bottom state in the digitally grated QWs of the hole barrier. The

transition in this energy ladder times on the order of picosecond, much faster than the

interband excitation in the absorber region. As such, the photo-generated electrons can be

transferred to the bottom of the energy ladder with very high efficiency. This mechanism

allows efficient and quick removal of electrons in the absorber region. Finally, the electrons

return to the valence band state in the adjacent absorber through interband tunneling

facilitated by the broken gap alignment between InAs and GaSb.

35
Figure 2-6: (a) Schematic band diagram of an ICTPV cell, (b) Schematic showing the
operation of an ICTPV cell.

2.3.3 Enhancement of open-circuit voltage in ICTPV cells

The advantage of IC structure for light emission is apparent: one electron can be

reused to generate multiple photons. In the reversing situation as in light-to-electron

converting devices, generation of a single electron requires multiple photons. Given this

situation, the achievable maximum quantum efficiency (or photocurrent) for ICTPV cells

is reduced by a factor of 1/Nc, where Nc is number of stages. This seems to make it

counterintuitive to explore this type of TPV cells. To resolve this problem, one needs to

really understand the benefits provided by multistage design. One of the key benefits is

enhanced collection efficiency of photo-generated carriers. This benefit will be described

in detail in next subsection, both physically and mathematically. Another important benefit

36
is the enhanced open-circuit voltage Voc, as it is equal to the sum of the photovoltages

created in every stage. As shown in Figure 2-6(b), the unipolar barriers repeat their roles

to separate photo-generated electrons and holes in all stages. This yields an effective

photovoltage in each individual stage. The recycling of electron across the device make

them add up to the total open-circuit voltage of the device. This behavior is analogous to

that seen in extensive study of multijunction solar cells [13-14]. As will be shown in

Chapter 3, at high incident power densities, the Voc of an ICTPV cell approximately scales

with the number of stages.

Because of enhanced Voc, the conversion efficiency of ICTPV cells can be higher

than conversional single-absorber cells even though the photocurrent is lower, which will

be shown in Chapter 3 and 4 in both theory and practice. From another perspective, like

ICLs, the reduction of photocurrent can be beneficial for mitigating the Ohmic power loss

in series resistances. In practice, TPV cells may experience significant Ohmic loss in cases

such as power delivery in free space [138-139] and near-filed TPVs [15-17]. In these

instances, the TPV cell often encounter an intensive illumination condition and generate a

high photocurrent, consequently suffering a heavy Ohmic loss.

2.3.4 Improvement of carrier collection efficiency in IC structures

The QE of a TPV cell depends on both the absorption of incident photons and the

collection of photo-generated carriers. The carrier collection probability fc (x) can be found

using Green’s function solution to the diffusion equation, as described in [140-141]. Its

expression at distance x from the collection point (x=0) is given by:

cosh[(𝑑−𝑥)/𝐿]
𝑓𝑐 (𝑥) = (2-1)
cosh(𝑑/𝐿)

37
where d is the absorber thickness and L is the diffusion length. Here, the light is assumed

to be incident from the collection point and travels through the absorber in a direction

opposite to the flow of minority carriers. In the other case where light is incident opposite

the collection point, most electrons are generated far from the collection point, therefore

the QE is likely reduced [142]. In the subsequent discussion, only the regular illumination

pattern will be treated. The calculated fc (x) based on Equation 2-1 in a 3.3 m absorber

for various diffusion lengths is plotted in Figure 2-7. As shown, the fc (x) is a strong

function of diffusion length. Also, it decreases dramatically with x if the diffusion length

is shorter than the absorber thickness. For example, given L=1 m, fc (x) is even lower than

0.4 when x is longer than L. Evidently, for a single-absorber device, increasing the absorber

thickness enhances the absorption, but may fail to improve QE, especially when the

diffusion length is short.

1.0

L=5 m
0.8 L=4 m
Collection probability

L=3 m
0.6
L=2 m

0.4 L=1 m

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Position (m)

Figure 2-7: Collection probability of carriers as a function of the distance from the
collection point. The absorber thickness is 3.3 m. The number near the curve
indicates the diffusion length.

The mechanism that affects collection probability also affect QE. Considering

carrier collection probability, the QE of a single-absorber device is given by:

38
𝑑
𝑄𝐸 = −1 𝑞 ∫0 𝑓𝑐 (𝑥)𝑔𝑝ℎ (𝑥)𝑑𝑥 (2-2)

where  is the incident photon flux density per unit area, and gph (x) is the photon

generation rate per unit volume. Here, the top surface reflection is left aside, which is

practically possible by adding a front anti-reflection layer. Note that Equation 2-2 can be

used to calculate the effective QE in each stage of a multistage IC device as well. The gph

(x) in Equation 2-2 exponentially deceases with position following the rule:

𝑔𝑝ℎ (𝑥) = 𝛼𝑒 −𝛼𝑥 (2-3)

where  is absorption coefficient. When applied to multistage IC device, Equation 2-3

needs to be modified to reconcile the light absorption in the optically upper stages. Based

on Equations 2-2 and 2-3, the QE of a single-absorber TPV device is given as:

𝛼𝐿 𝛼𝐿𝑒 −𝛼𝑑
𝑄𝐸 = 1−(𝛼𝐿)2 × [tanh(𝑑 ⁄𝐿) + cosh(𝑑⁄𝐿) − 𝛼𝐿] (2-4)

Likewise, the effective QE in the Nth stage of a multistage IC device is given by:

𝑁−1 𝛼𝐿 𝛼𝐿𝑒 −𝛼𝑑𝑁


𝑄𝐸𝑁 = 𝑒 −𝛼 ∑𝑚=1 𝑑𝑚 1−(𝛼𝐿)2 × [tanh(𝑑𝑁 ⁄𝐿) + cosh(𝑑 − 𝛼𝐿] (2-5)
𝑁 ⁄𝐿 )

𝑁−1
where dm is the absorber thickness in the mth stage, and the term “𝑒 −𝛼 ∑𝑚=1 𝑑𝑚 ” represents

light absorption in all the upper stages.

Based on the above equations, a numerical example is provided to illustrate the

improvement of carrier collection in multistage architecture, as shown in Figure 2-8. It

presents the calculated fc (x)·gph (x) in single-absorber and four-stage IC devices for

L=0.4. The total absorber thicknesses (d) for both structures were set to be the absorption

depth. Therefore, if no absorption occurs in the barrier regions (indicated by thick grey

lines in Figure 2-8), the total absorption in the absorbers is equal in the two cases. The four-

stage IC device has identical absorber thickness in each stage, meaning that the individual

39
absorber thickness is equal to d/4. According to Equation 2-2, the QEs of the two cells are

marked by the shaded regions in Figure 2-8. As can be seen, the total effective QE of the

four-stage IC device is appreciably higher than the single-absorber device. This result can

be considered the equivalent of much higher total collection efficiency (c) in the four-

stage IC device. Here, the c is defined as the ratio of the total effective QE in any of the

stages to the total absorption of incident photons (1-e-d). The calculated c for the single-

absorber cell is only 46% due to the low collection probability at positions far from the

collection point. At the right edge of the absorber, the carrier generated over there has a

collection probability of only 16%. In contrast, since the absorbers are made thin, the

collection probability is much higher in the four-stage IC device. For example, the

collection probability is enhanced to 83% at the right edge of each individual absorber.

This enables it to achieve a total c as high as 89%.

1.0
Barriers
0.8
g
ph (x)
f
g (x)f (x) (a.u.)

(x) g gph (x)


0.6 (x)
f
ph
(x)
gp (x
h )f
0.4 (x)
gp (x
g ( h )f (x)
ph x
)f
0.2 (x)

0.0
d/4 d/4 d/4 d/4

Figure 2-8: Comparison of collection process in single- and four-stage IC devices for
a low L product (L=0.4). The thickness d of the single-stage device equates the
absorption depth. The individual absorber thickness of four-stage IC device is d/4.

It should be commented that the above analysis ignored a couple of unfavorable

factors that may affect collection efficiency. For example, TPV cells generally operate at

40
forward bias for performing power output. The applied external field may impede the

collection of photo-generated carriers. To calculate c with accounting the external bias,

it’s going to be more complex and challenging whatever method one chooses. The

experimental investigation of this subject is presented in Chapter 4, while the theoretical

aspect will continuous to be one of future research focuses. Another neglected factor is the

recombination at the absorber-electron barrier interface, characterized by surface

recombination velocity. The complete calculation with consideration of interface

recombination is described in [141].

2.4 Interband cascade infrared photodetectors

2.4.1 Operation principle of ICIPs

The configuration and operation principle of an ICIP are quite analogous to those

of an ICTPV cell. In fact, there is no essential difference between them except operating

bias voltage and light intensity encountered by them. As shown in Figure 2-9, like an

ICTPV cell, each stage of an ICIP consist of an electron barrier, a hole barrier and a T2SL

SL absorber. The constituent layers of the three components in each stage are same in the

two different types of devices. Also, electrons almost undergo the same transport path in

them. The only notable difference is the operating voltage as illustrated in Figure 2-6 and

2-9. To better differentiate them from an ICL, the detailed band profile of one stage of an

ICIP is shown in Figure 2-9(b), which differs markedly in the absorber region from an ICL.

41
Hole Barrier Electron Barrier
2.5
(b)
2.0
Energy (eV)

1.5
Absorber Absorber

1.0
 Tuneling

0.5
Relaxing
0.0
0 20 40 60 80 100 120 140
Distance (nm)

Figure 2-9: (a) schematic diagram of a multistage ICIP and (b) the band profile of
one stage under zero bias. The olive and purples lines in the absorber represent the
electron and hole minibands. The dotted olive wavefunction indicates the electron
states in hole barrier while the dotted purple wavefunction represents the hole states
in electron barrier.

Compared to ICTPV cells, ICIPs have relatively looser design requirement. For

example, the individual absorber thickness in an ICTPV cell is better adjusted to keep

current match between stages, as done elsewhere in multijunction tandem solar cells [13-

14], otherwise the photocurrent will be largely reduced. However, such a requirement does

not need to be fulfilled for ICIPs due to significant electrical gain, as will be descried in

Chapter 7. Besides, ICIPs can operate at zero bias, so the device design is not concerned

with the effect of external bias.

42
2.4.2 Noise reduction in ICIPs

As with an ICTPV cell, an ICIP also benefits from the high collection efficiency,

but suffers from the relatively low achievable maximum QE. The other profit offered by

multistage architecture in detectors is the reduced noise level. As shown in Figure 2-9(a),

a single electron must undergo Nc interband excitations in an ICIP to travel across the

contacts. This fact means that the noise is naturally reduced in ICIPs due to the averaging

process. A similar example is quantum well infrared detector (QWIP) [65, 78]. The noise

in these intersubband detectors is reduced by a factor of 1/Nw (number of quantum wells),

provided that the emission and capture of electrons are uncorrelated in each QW. Another

easy-to-understand example is the reduction of random error by increasing the sample size

and averaging over all the samples.

There are various sources of noise that can affect a photodetector’s detectivity. The

dominant noise changes with the environment and the temperature of the detector. For

example, when the signal is strong or the detector temperature is low, the dominant noise

is from either the fluctuation of signal current or the fluctuation of current induced by

background radiation. Conversely, when the signal is weak or the detector temperature is

high, the detectivity is generally regulated by shot or Johnson noise. In realistic

applications, the operation of detector is neither shot- nor Johnson-noise limited, since the

performance is poor and does not satisfy application requirement. However, for most

LWIR detectors such as MCT and T2SL detectors, the detectivity in this regime represents

an ultimate limit for the detector operating at room temperature [130]. The focus here will

only involve this situation. In addition, as mentioned before, the unipolar barriers allow

ICIPs to operate in unbiased mode. This means that shot noise can be neglected in an ICIP,

43
and the detectivity will be exclusively limited by Johnson noise. The mean square Johnson

noise current is inversely proportional to zero-bias resistance R0, as seen from Equation 1-

7. Hence, in order to proceed, the expressions of dark current and R0 needs to be derived

first.

Analogous to QE, the dark collection current (which has the same direction with

photocurrent) in the mth stage in an ICIP can be calculated as:


𝑑
𝐽0𝑚 = 𝑞 ∫0 𝑓𝑐 (𝑥)𝑔𝑡ℎ 𝑑𝑥 = 𝑞𝑔𝑡ℎ 𝐿tanh(𝑑𝑚 /𝐿) (2-6)

where gth is thermal generation rate per unit volume. Unlike the optical generation, the

thermal generation can be uniform across the device if the bandgaps of absorbers are made

equal in each stage.

In addition, there is another contribution of dark current: the injection current. It

has opposite direction with photocurrent and has a magnitude of 𝑒 𝑉𝑚 ⁄𝑘𝑏𝑇 𝐽0𝑚 (Vm is the

voltage that falls across the mth stage). Collectively, considering the two current

components, the total dark current of an ICIP can be written as:

𝑐𝑁
𝐽𝑑 (𝑉) = 𝑞𝑔𝑡ℎ 𝐿 ∑𝑚=1 tanh(𝑑𝑚 /𝐿)[𝑒𝑥𝑝(𝑞𝑉𝑚 ⁄𝑘𝑏 𝑇) − 1] (2-7)

Based on Equation 2-7, the R0A of an ICIP can be extracted and expressed as:

𝑘𝑏 𝑇 1
𝑅0 𝐴 = ∑𝑁𝑐
𝑚=1 (2-8)
𝑞2 𝑔𝑡ℎ 𝐿 tanh(𝑑𝑚 ⁄𝐿)

For an ICIP with identical stages, the expression of R0A of can be simplified to:

𝑘𝑏 𝑇 𝑁𝑐
𝑅0 𝐴 = 2
(2-9)
𝑞 𝑔𝑡ℎ 𝐿 tanh(𝑑𝑚 ⁄𝐿)

Evidently, from Equation 2-8 and 2-9, the R0A is larger for detectors with more stages and

thinner absorbers. In other words, according to Equation 1-7, the Johnson noise is

effectively reduced in ICIPs compared to single-absorber detectors.

44
2.4.3 Detectivity improvement in ICIPs

In principle, the device QE of an identical-stage ICIP is decided by the stage with

minimum effective QE. This will be the last stage due to most significant light attenuation.

However, to maintain current continuity, there is additional injection current induced to

offset the higher photocurrent in other stages. This undermines some of the benefits

provided by multistage architecture. Another design option to eliminate this downside is to

make current-matched absorbers. In this revised design, the individual absorber thicknesses

are increased from first stage to last stage to achieve equal photocurrent in each stage. In

practice, perfect current match is hard to accomplish unless the diffusion length and

absorption coefficient are accurately grasped. Nevertheless, even with inexact match in

photocurrent, the device QE in principle can still be improved. Here, only current-matched

ICIPs will be considered while ICIPs with identical absorbers will be detailed in Chapter

7.

The detectivity enhancement in ICIPs has been covered in [141], a brief review of

the calculation results is provided here. Substituting Equation 2-8 into Equation 1-8, one

can obtained the expression of Johnson-noise limited detectivity for an ICIP:

√ 𝑁𝑐
∗ λ 𝑄𝐸 ∑𝑚=11⁄𝑡𝑎𝑛ℎ(𝑑𝑚 ⁄𝐿)
𝐷 = (2-11)
ℎ𝑐 √4𝑔𝑡ℎ 𝐿

The current match condition in the ICIP is first obtained using an iterative process by

varying the thickness of each stage so that the contribution of QE is equal. The absorber

thicknesses are then determined by selecting the optimal photocurrent-matched absorber

sequence that maximizes detectivity. In this way, the calculated detectivity enhancement

as a function of L for ICIPs with two, eleven and thirty stages are shown in Figure 2-10

45
[141]. The detectivity enhancement is defined as the D* (Nc) of the optimized multistage

ICIP normalized to the value D* (1) of the optimized single-absorber detector. As can be

seen, the detectivity enhancement is pronounced when L1 for different designs. Also,

the detectivity is raised as the number of stages increases since the noise is further

suppressed, although the signal current is slightly reduced. At larger L, multistage ICIPs

do not make obvious advantage, but there is still a small advantage can be gained. For

example, for optimized ICIPs with many stages, the upper limit improvement is about 1.1

times higher than single-absorber detectors [141]. This conclusion can be derived from

Figure 2-10 where the platform value of detectivity enhancement is slightly higher than

unity at large L.

Figure 2-10: Johnson-noise limited detectivity enhancement for current-matched


ICIPs with two, eleven and thirty stages. Figure is from [141].

2.4.4 Comments on detectivity improvement in ICIP

The above calculations clearly quantify the possible detectivity enhancement when

L1 for current-matched ICIPs. In realistic, for InAs/GaSb T2SLs, the absorption

coefficient near bandgap is about 3000 and 2000 cm-1 in MWIR [143-145] and LWIR [145-

46
147] regimes, respectively. The diffusion length is shorter than 1.5 m at RT, as estimated

from the temperature or bias dependence of responsivity for the detectors made of

InAs/GaSb T2SLs [149-151]. Taken together, the product L can be smaller than unity at

high temperatures ( 200 K), especially for LWIR T2SL detectors. Hence, the prospect of

detectivity enhancement in ICIPs is real at high temperatures. At lower temperatures, the

diffusion length is appreciably increased as carrier lifetime is extended. For example, the

diffusion length can be far longer than 6 m at 77 K, as evaluated in [152]. The increased

diffusion length is very likely to make L larger than unity, therefore it will be bootless to

use ICIP structure at low temperatures. However, in applications where the response speed

is prioritized over sensitivity, ICIP is still the better option. For single-absorber detectors,

high response speed requires a thin absorber, which compromises light absorption and thus

sacrifices the detectivity. However, for ICIPs, they have been demonstrated with high

frequency operation (higher than 1.3 GHz) as well as decent detectivity [153-154].

2.5 Growth and fabrication of interband cascade devices

The IC devices are relatively complex structures; some devices even have

thousands of layers. This complexity rules out the possible growth by conventional growth

techniques as well as some epitaxy growth techniques such as chemical vapor deposition

(CVD), physical vapor deposition (PVD) and liquid phase epitaxy (LPE). The only reliable

and feasible growth method is molecular beam epitaxy (MBE) [75-76]. Up to present,

almost without exception the reported IC devices were grown by MBE systems. Compared

to other epitaxy growth techniques, MBE is better able to grow sophisticated structures

with high degree of success. This is due to its nature of utilizing atomic layer-by-layer

47
growth, which is accomplished through a good monitor of molecular or atomic beams onto

a heated substrate in ultrahigh environments. In this dissertation, all the devices involved

were grown by the two MBE systems in the University of Oklahoma as shown in Figure

2-11. The first one is an Intevac Gen II that has been operational since 1994. The system

is equipped with two Sb and As crackers, three In, Ga and Al effusion cells, as well as two

Si and Be doping cells. The second one is a new Veeco Genxplor MBE system launched

in 2015, which has many new and improved features. For example, all the group-III cells

are comprised of dual-filament heaters to generate more stable flux. This new MBE system

has ten cells including two In and two Ga cracked cells, two Al Sumo cells, one cracked

As cell, one cracked Sb cell and three Si, Te and Be doping cells.

Figure 2-11: (a) Intevac GEN II MBE system (1993) and (b) Veeco GENxplor MBE
system (2013).

Manufacture of IC devices involves various fabrication processes. The general

processing flow of IC devices (e.g. ICTPVs and ICIPs) include: (1) standard cleaning, (2)

mesa etching, (3) insulating layer deposition, (4) contact opening, (5) top contact

deposition, (6) lapping, (7) bottom contact deposition, and (8) mounting and wire bonding.

Specifically, after cleaning and standard contact photolithography, wet chemical etching is

used to define a mesa structure by etching deep down below the active region. Then, a ~

48
200 nm thick silicon nitride followed by ~ 200 nm silicon dioxide is sputter deposited as

an insulating layer. This step is followed by reactive ion etching (RIE) to open a window

on top of mesa. This window is opened to deposit 30/300 nm of Ti/Au layer by sputtering

technique as top metal contact. The schematic of a typical fabricated 3-stage ICTPV device

is shown in Figure 2-12(a). The cross-sectional scanning electron microscope image of the

3-stage ICTPV device is presented in Figure 2-11(b) [157].

InAs

InAs/GaSb SL

InAs/Al(In)Sb

InAs/GaSb SL

InAs/Al(In)Sb

Figure 2-12: (a) The schematic of a processed ICTPV or ICIP device and (b) Cross-
section scanning electron microscope image of a wet-etch ICTPV structure, the
Figure is from [157].

49
3 Chapter 3: Limiting factors and efficiencies of narrow bandgap
thermophotovoltaic cells
3.1 Background and motivation

In Chapter 1, the efficiency limits of TPV cells were calculated based on detailed

balance theory, which however tends to be overestimated due to some unrealistic

expectations. There are many theoretical works attempting to predict the efficiency limit

of TPV cells. For example, in [49, 51], a prospective efficiency exceeding 30% was pointed

out when the heat source is at 1000-2000 K, even without spectral control. For solar TPV,

even a maximum efficiency of 85% was projected with full concentration of incident

sunlight [55]. Realization of this extremely high efficiency requires that the incident light

spectrum is perfectly tailored to the cell absorption spectrum and non-absorbed is recycled

back to the heat source. At current stage, the highest reported TPV cell efficiencies at 300

K are 24% for a 0.6 eV InGaAs diode on InP [37] and 19.7% for a 0.53 eV GaInAsSb

diode on GaSb [39], which were measured using a 950 °C broadband radiator with spectral

control filters mounted on the front surface of the TPV cells. As for narrow bandgap TPV

cells (Eg 0.4 eV), the demonstrated efficiencies at 300 K are far below 10% (See table

1.2). Evidently, there is a large gap between the efficiencies of existing TPV cells and

theoretical predictions, and little work has been dedicated to narrow bandgap TPV cells to

clarify their efficiency limits. It is therefore necessary to have ongoing work to bridge the

efficiency gap and to determine the practical efficiency limits as well.

Most of previous theoretical works assumed very ideal behaviors of carrier

recombination and collection. Specifically, they assumed purely radiative recombination

and an infinite diffusion length. However, in real devices, non-radiative recombination is

often involved and even prevails, and carrier collection can be limited by a short diffusion

50
length. In this chapter, practical factors such a finite diffusion length (L) and absorption

coefficient () are considered and their effects on conversion efficiency () are inspected.

As examples, calculations are carried out for narrow bandgap InAs/GaSb T2SLs and

quaternary GaInAsSb materials in several different scenarios under monochromatic light

illumination. This narrow bandwidth light illumination can be accomplished through the

use of spectral filters or selective emitters that can be made based on nanostructured

materials and metamaterials [31-33]. The calculations start from single-absorber TPV cells

and then are performed for multistage IC architecture to show how it can be used to

improve the performance of narrow bandgap TPV cells.

3.2 Practical limitations on single-absorber TPV devices

3.2.1 Quantum efficiency and dark saturation current density

The conversion efficiency of a TPV device is intimately related to its output current

and voltage. These two quantities are characterized by quantum efficiency QE and voltage

efficiency V (defined as the ratio of open-circuit voltage eVoc to the bandgap). Both QE

and Voc are largely ruled by dark saturation current density J0, as well as minority carrier

transport and lifetime . Therefore, QE and Voc will be severely limited if the carrier

lifetime and diffusion length are short and the J0 is significant. As an example of such

limitation, InAs/GaSb SL absorber with a bandgap of 0.29 eV will be first used for

illustration purpose. At 300 K, the diffusion length and carrier lifetime are estimated to be

1.5 m and 20 ns based on the experimental results of type-II InAs/GaSb infrared detectors

[149-151, 158]. The conversion efficiency of a TPV device under monochromatic

illumination is given by:

51
𝑞𝑉𝑜𝑐
𝜂 = 𝐹𝐹 ⋅ 𝑄𝐸 ⋅ (3-1)
ℎ𝜐

where FF is the fill factor and  is the frequency of incident photons. Hence, FF, QE and

Voc are the three main performance metrics that controls the desired conversion efficiency.

Below, their respective behaviors are studied in narrow bandgap TPV devices. The

frequency of incident photons also plays a role in affecting conversion efficiency, but is

less significant than the above-mentioned three quantities, which will be described in

Subsection 3.4.

The expression of QE for a single-absorber TPV device is given by Equation 2-4.

Here the light is assumed to travel through the absorber in a direction opposite to the flow

of minority carriers. Based on this equation, the calculated QE as a function of normalized

absorber thickness (d/L) for different values of L is shown in Figure 3-1. As can be seen,

the QE peaks a at a certain value of d/L and falls off with further increasing the absorber

thickness, irrespective of the value of L. This common tendency of QE was identified due

to the reduction of collection efficiency as the absorber thickness increases [159].

Particularly, for L=0.45, the maximum QE is only 32%, which would significantly limit

the conversion efficiency as will be shown later.

52
0.30 90
0.28 80
0.26 L=4.5

Open-circuit votlage (V)

Quantum efficiency (%)


70
0.24 solid: Voc
60
0.22 dashed: QE
50
0.20 L=1.5
40
0.18
0.16 30
0.14 L=0.45 20
0.12 10
0.10 0
0.01 0.1 1 10
Normalzied absorber thickness (d/L)

Figure 3-1: Calculated open-circuit voltage (solid) and quantum efficiency (dashed)
as a function of normalized absorber thickness for different values of L. The incident
power density is assumed to be 50 W/cm2.

The dark saturation current density is the pre-factor in standard diode equation and

measures the recombination loss in PV devices. Normally, the open-circuit voltage is a

logarithmic function of the ratio between photocurrent density and J0. In solar cells, the

thermal current density is sometimes ignored because it is low when the bandgap is

relatively wide. In contrast, J0 is orders of magnitude higher in TPV devices and therefore

cannot be neglected. The value of J0 can be calculated based on Equation 2-6 for the simple

single-stage case. The thermal generation rate gth in this equation for p-type doped

absorbers can be written as: gth=n0/, where n0 is the electron concentration at thermal

equilibrium. By replacing n0 with ni2/p0, thermal generation rate can be further written as:

ni2/Na, where ni and Na are the intrinsic carrier concentration and doping concentration,

respectively. Hence, a short carrier lifetime (e.g. 20 ns) will manifest itself as a high J0,

thus severely limiting the open-circuit voltage. An increase of carrier lifetime will naturally

reconcile this issue and enhance QE as well since the diffusion length is increased with

raised carrier lifetime. For example, if carrier lifetime is extended to 200 ns, on a

53
conservative estimate, the diffusion length will be increased is 5 m, assuming the electron

mobility (43 cm2·V-1·s-1) remains the same. In this scenario, the J0 will be an order of

magnitude lower and the QE will be appreciably improved.

Nevertheless, the J0 is still much higher than the radiative limit set by the detailed

balance theory [58]. In this fundamental limit, the dark saturation current density is given

by:

2𝜋𝑞 ∞ 𝑛2 (1−𝑒 −𝛼𝑑 )𝐸 2


𝐽0 = ℎ3𝑐 2 ∫𝐸 𝑑𝐸 (3-2)
𝑔 𝑒 𝐸/𝑘𝑏 𝑇 −1

where n is refractive index. Here, several assumptions were made: the surface reflections

and photon recycling effect [160-161] are ignored, and the radiative photons are assumed

to have a single path and a solid angle of . The term (1-e-d) in Equation 3-2 describes

incomplete absorption of photons due to the finite absorber thickness, compared to the full

absorption for a blackbody. With ignoring recycling factor, the calculated radiative carrier

lifetime is about 2.3 s [See table 3-1]. As a result, the diffusion length is around 15 m,

assuming a constant electron mobility of 43 cm2·V-1·s-1. Based on Equation 2-6 and

Equation 3-2, the calculated dark saturation density is shown in Figure 3-2 for three

different carrier lifetimes. As can be seen, for  =20 ns, J0 is on the order of 0.1 A/cm2, in

agreement with the measurements for ICIPs [158]. This substantially high J0 poses a

difficulty in realizing a high open-circuit voltage. By comparison, in the radiative limit, J0

is approximately two orders of magnitude lower. This implies that there is a still plenty

room for improvement of performance for existing TPV devices based on InAs/GaSb SLs.

54
Table 3-1: Parameters used in calculation for InAs/GaSb superlattice.

Temperature and bandgap Tdevice=300 K, Eg=0.29 eV


Effective mass 𝑚𝑒 = 0.03𝑚0 , 𝑚ℎ = 0.4𝑚0
Effective density of states Nc=1.3×1017 cm-3, Nv=6.3×1018 cm-3
−𝐸𝑔
Intrinsic carrier concentration 𝑛𝑖 = √𝑁c 𝑁v exp ( )= 3.4×1015 cm-3
2𝑘𝑏 𝑇
p-type doping concentration
in the absorber Na=2.4×1016 cm-3

Refractive index n=3.5


Absorption coefficient: =3000 cm-1
𝜃=𝜋 𝜑=2𝜋 𝐸=∞
𝑛2 2𝐸 2 𝐸
Radiative 𝐵= 2 ∫ ∫ ∫ 𝛼(𝐸) 3 2
[𝑒𝑥𝑝 ( ) − 1]−1 𝑠𝑖𝑛𝜃𝑑𝜃𝑑𝜑𝑑𝐸
𝑛𝑖 ℎ 𝑐 𝑘𝑏 𝑇
recombination 𝜃=0 𝜑=0 𝐸=𝐸𝑔
coefficient [162]
B=1.75×10-11 cm-3·s-1

Actual lifetime = ns, L=1.5 m, L=0.45


Electron mobility e=43 cm2·V-1·s-1, calculated from L=1.5 m and =20 ns.

Medium lifetime = ns, L=5 m, L=1.5


Radiative lifetime 𝜏𝑟𝑎𝑑 = 1⁄𝐵𝑁a = 2.3 μs (no photon recycling), L=15 m,
L=4.5
Ultimate limit L=15 m, L=4.5, Voc=0.29 V

100
Dark saturation current density (A/cm2)

10-1
ns

10-2 ns
200
=
it
10-3 e lim
iativ
Rad

10-4
0.01 0.1 1 10
Normalized absorber thickness (d/L)

Figure 3-2: Calculated dark saturation current density as a function of normalized


absorber thickness for a carrier lifetime of 20 ns, 200 ns and the radiative limit.

55
3.2.2 Open-circuit voltage and fill factor

The illuminated J-V characteristic needs to be known prior to calculating the open-

circuit voltage and fill factor. The net current density flowing out from a TPV device under

illumination is simply the superposition of the dark current density (Jd) and the

photocurrent density (Jph):


𝑞𝑉
𝐽 = 𝐽𝑝ℎ − 𝐽0 (𝑒 𝑘𝑏𝑇 − 1) (3-3)

where Jph equates eQE0 and J0 is given by Eq. (4) or Eq. (5) (for the radiative limit). The

term J0eqV/kbT stands for the injection current density under a forward bias, which is in the

opposite direction of Jph and thus can strongly affect the fill factor. Figure 3-3 shows the

simulated J-V curves for different values of L (0.45, 1.5 and 4.5). The diffusion lengths

and carrier lifetimes are different but the absorption coefficient (~ 3000 cm-1) is the same

in the three scenarios as shown in Table 3-1. The incident power density Pinc was assumed

to be 25 and 50 W/cm2. In each case, the absorber thickness is the optimal value that

maximizes the conversion efficiency. As shown in Figure 3-3, the simulated J-V curve is

more square-like for larger L, suggesting the increase of fill factor with L. As can be

seen in Figure 3-4, the FF decreases with d/L and is lower than 55% for L=0.45, which is

significantly lower than the 85% reported for high-quality crystalline Si and thin film GaAs

solar devices. Likewise, the Voc exhibit similar trends with d/L and is low when L is small

due to relatively high J0. These two quantities (Voc and FF) both increase with L due to

the decrease of J0 as well as the increase of the QE. Raising the incident power density

from 25 W/cm2 to 50 W/cm2 led to the insignificant enhancements of FF and conversion

efficiency for each value of L. Specifically, the FF () increases from 50% (6%), 58%

(19%) and 54% (69%) to 53% (7%), 61% (23%) and 71% (59%) for L equal to 0.45, 1.5

56
and 4.5, respectively. In the following analysis, the incident power density is set at a fixed

value of 50 W/cm2. Nevertheless, the fundamental insights gained in the analysis are

expected to be applicable to lower incident power densities.

In practice, the goal of a 50 W/cm2 incident power density is difficult to achieve for

conventional TPV configurations, but is still feasible under some circumstances. For

instance, adding a concentrator in a TPV system, analogous to concentration solar cells,

can significantly enhance the incident power density. Another example is the PV device

used in power beaming as the light is sent from a high-power laser source. In this case, the

incident power density is likely to exceed 50 W/cm2 for adequate power delivery. In

addition, in the near filed transfer technology where the TPV device is placed in extreme

proximity (typically < 100 nm) to the heat source (or radiator) [15-17], the incident power

density of the device can be very high as well. On the other hand, the high incident power

density can incur the high injection effect, as observed in a GaSb p-n junction near filed

TPV cell [163]. Narrow bandgap TPV devices with low doping level may be also subject

to this effect, but this is beyond the scope of this chapter.

160
140 solid: 50 W/cm2
Current density (A/cm2)

120 dashed: 25 W/cm2


L=4.5
100
L=1.5
80
60
40 L=0.45
20
0
0.00 0.05
0.10 0.15 0.20 0.25 0.30
Voltage (V)
Figure 3-3: Simulated J-V curves for different values of L and with incident power
density of 25 and 50 W/cm2.

57
Based on Equation 2-6 and Equation 3-3, the expression of open-circuit voltage can

be written as:

𝑘𝑏 𝑇 𝛷0 𝑄𝐸
𝑉𝑜𝑐 = 𝑙𝑛( + 1) (3-4)
𝑞 𝑔𝑡ℎ 𝐿𝑡𝑎𝑛ℎ(𝑑/𝐿)

Under highly intensive illumination, the second term in the natural logarithm can be

neglected, then the open-circuit voltage can be expressed as:


𝐸𝑔
𝑘𝑏 𝑇 𝛷0 𝑄𝐸𝜏𝑝0 𝐸𝑔 𝑘𝑏 𝑇 𝛷0 𝑄𝐸𝜏𝑝0
𝑉𝑜𝑐 = 𝑙𝑛 ( 𝑒 𝑘𝑏𝑇 ) = + 𝑙𝑛 ( ) (3-5)
𝑞 𝑁𝑐 𝑁𝑣 𝐿𝑡𝑎𝑛ℎ(𝑑/𝐿) 𝑞 𝑞 𝑁𝑐 𝑁𝑣 𝐿𝑡𝑎𝑛ℎ(𝑑/𝐿)

Here, the well-known relationship for non-degenerate absorbers was used:

𝑛𝑖2 = 𝑁𝑐 𝑁𝑣 𝑒 −𝐸𝑔/𝑘𝑏𝑇 (3-6)

where Nc (Nv) is the effective density of state for the conduction (valence) band of the

absorbers (See Table 3-1). Based on Equation 3-5, the calculated open-circuit voltage for

different values of L (0.45, 1.5 and 4.5) is presented in Figure 3-1. As shown, the Voc

gradually decreases with the absorber thickness due to the sharper increase of Jo (as shown

in Figure 3-2) than QE. As an example, the Voc decreases from 0.128, 0.187 and 0.287 V

to 0.118, 0.163 and 0.266 V while the normalized absorber thickness increases from 0.01

to 10 for L of 0.45, 1.5 and 4.5, respectively. Hence, in practical device with L=0.45,

the Voc seldom exceeds 0.13 V even at high incident power density, which sets a boundary

(<0.45%) of the voltage efficiency. As the carrier lifetime increases via improvement of

material quality, the Voc can be increased substantially as shown in Figure 3-1 with a higher

L. The Voc in the radiative limit is quite close to bandgap voltage, but never allowed to

exceed it. This is because the amplified stimulated emission will be triggered when the

separation of quasi-fermi levels for electrons and holes exceeds the bandgap. Such a

process will further reduce the carrier lifetime thus increase the saturation dark current

58
density. In [164], unexpectedly, the value of Voc was evaluated to be higher than bandgap

voltage for a solar device under monochromatic light illumination. However, the

calculation did not account the reduction in carrier lifetime. Below, based on Equation 3-

5, a convincing argument is provided to support why Voc< Eg/q.

Under steady state condition, the sweep-out of photo-generated electrons needs to

be compensated by the absorption of photons. This signifies that the density of escaping

photogenerated electrons equates 0QE/v, where v=Ltanh(d/L)/ represents the average

escaping speed of photogenerated electrons. The upper limit of photo-generated electron

density is the available density of states NcNv/p0. Hence, based on Equation 3-5, Voc is

always lower than Eg/q. This implies that the carrier lifetime  reduces with increasing 0

in order to keep consistent with the upper limit.

74
72 ultimate limit
70
L=4.5
68
Fill factor (%)

66
64
62
60 L=1.5
58
56
54 L=0.45
52
0.01 0.1 1 10
Normalzied absorber thickness (d/L)
Figure 3-4: Calculated fill factor as a function of normalized absorber thickness for
different values of L. The incident power density is assumed to be 50 W/cm2 except
in the ultimate limit.

3.2.3 Conversion efficiency

The fact that Voc is always lower than Eg/q offers an effective approach to evaluate

59
the ultimate limit of conversion efficiency for single-absorber devices. To do this, one

needs to first define the ratio of the photon flux to the thermal flux as:

𝛷0
=𝑔 (3-7)
𝑡ℎ𝐿

According to Equation 3-4, the maximum value of  is exp(-Eg/kbT)·tanh(d/L)/QE to keep

Voc lower than Eg/q. When  reaches this value at sufficiently high incident power density,

the conversion efficiency will be stretched to its ultimate limit. This means the ultimate

efficiency limit of a single-absorber device can be obtained by maximizing following

equation with an optimal voltage:

𝜂 = [𝑒 𝐸𝑔/𝑘𝑏𝑇 − (𝑒 𝑞𝑉/𝑘𝑏𝑇 − 1)] ∙ 𝑉/(𝑒 𝐸𝑔 /𝑘𝑏𝑇 𝐸𝑔 ) ∙ 𝑄𝐸 (3-8)

According to this equation, the ultimate efficiency equates the quantum efficiency

multiplied by a factor of 0.71 for a 0.29 eV bandgap. The diffusion length in the ultimate

limit will be assumed to be 15 m, identical to the value in the radiative limit (See Table

3-1). In the ultimate limit, the fill factor remains constant with of d/L as shown in Figure

3-4. It should be emphasized that there are two approximations were made to derive

Equation 3-8. First, the incident photon energy is precisely matched with the bandgap.

Second, the illumination source has an ideal monochromatic spectrum with a shape of delta

function. In practice, the incident photons should possess an energy higher than bandgap

to excite electron-hole pairs. Thus, the first assumption would somewhat overestimate the

conversion efficiency. The second assumption significantly simplifies the illumination

source, which in fact does not make too much difference in conversion efficiency. For

example, provided that the incident photon has Gaussian distribution with the central

energy being 0.34 eV (50 meV higher than the bandgap) and a FWHM of 26 meV (equal

to kbT), the calculated conversion efficiency is 5.4% for L=0.45 at the power density of

60
50 W/cm2. This value of efficiency is slightly lower than the 5.6% calculated for the case

with a perfect monochromatic light (delta function) at 0.34 eV.

Figure 3-5 shows the calculated conversion efficiency of single-absorber devices

in different cases. As shown, the efficiencies in the radiative and ultimate limit are quite

close to each other, especially at smaller d/L. The peak efficiencies are 59% and 63% in

the radiative and ultimate limit, respectively. The vast gap between the radiative limit and

the practical efficiency (L=0.45) reveals a huge potential for improvement. To bridge this

gap, the material quality needs to be greatly improved. For L=0.45, the actual achievable

efficiency is less than 7% as a directly result of low Voc (Figure 3-1) and FF (Figure 3-4)

that spring from a high J0 with a short carrier lifetime (~ 20 ns). If, however, the carrier

lifetime increases by an order of magnitude, the efficiency is possible to reach up to 23%.

These results explicitly show that carrier lifetime is the key issue in narrow bandgap TPV

devices. Besides, another important issue is the relatively low QE ( 32%) due to a small

product of L. The main tendency of conversion efficiency with d/L is resembles that of

QE with d/L (Figure 3-1). That is, the conversion efficiency peaks at a certain absorber

thickness, then slowly drops, and finally reaches a plateau value with further increasing

absorber thickness. The maximum value of  occurs at an optimal d/L equal to 1.8, 1.1 and

0.7 for L value of 0.45, 1.5 and 4.5, respectively, consistent with the order of the optimal

d/L for maximum QE. Compared to the optimal d/L for maximum QE, the optimal d/L for

maximum  is slightly lower due the decrease of Voc and FF with increasing d/L.

61
70
ultimate limit
60

Conversion efficiency (%)


50 L=4.5

40

30
L=1.5
20

10 L=0.45

0
0.01 0.1 1 10
Normalized absorber thickness (d/L)
Figure 3-5: Calculated conversion efficiency vs normalized absorber thickness for
different values of L. The incident power density is assumed to be 50 W/cm2 except
for the ultimate limit.

Figure 3-6 shows the calculated conversion efficiency as a function of bandgap

within 0.2-0.4 eV for single-absorber devices. In principle, the variation of bandgap should

result in systematic changes in carrier lifetime. However, because of little relevant

experimental data and uncertainties in carrier lifetime for InAs/GaSb SLs with different

bandgaps, the carier lifetime, absorption coefficient and diffusion length are remained same

for different bandgaps as given in Table 3-1. This assumption, together with same doping

concentration, implies that the thermal generation is proportioanl to e-Eg/kbT. Hence, a

modest increase in bandgap will result in a large redcution in J0 and significant increases

in FF and votlage efficiency, eventually raising conversion efficiency sustantially as shwon

in Figure 3-6. For example, for L=0.45, the conversion efficiency is raised from 3% to

12% while the bandgap is increased from 0.2 eV to 0.4 eV.

62
0.20 0.25 0.30 0.35 0.40
70
68

Conversion efficiency (%)


(a)
66 mate
64 Ulti
62
60 L=4.5
58
30
25 L=1.5
20
15
10 L=0.45
5
0 80
100
75
90 Ultimate
Voltage efficiency (%)

L=4.5 70
80
65

Fill factor (%)


70 60
L=1.5
60 55
50 50
(b)
40 45
solid: voltage efficiency
30 40
L=0.45 dahsed: fill factor
20 35
0.20 0.25 0.30 0.35 0.40
Bandgap (eV)
Figure 3-6: Calculated (a) conversion efficiency and (b) voltage efficiency and fill
factor for L=0.45, 1.5 and 4.5 and the ultimate efficiency limit as a function of
bandgap. The incident power density is 50 W/cm2 except for the ultimate limit.

In the radiative limit, the J0 detemined by Equation 3-2 decreases with increasing

bandgap, but at a slower rate than that for for L=0.45 or 1.5. Consequently, the FF and

Voc increase gradually with bandgap, while the voltage efficiency decreases with bandgap

as shown in Figure 3-6(b). Hence, in the radiative limit, the increase of conversion

efficiency with bandgap is insignificant as it ranges between 58%-60%, as shown in Figure

3-6(a). Note that, in Equation 3-2, the absorption spectrum was assumed to have same

shape but different take-off points for different bandgaps. The diffusion length in the

radiative limit was still taken to be 1.5 m for different bandgaps. In addition, the QE at

the optimal absorber thickness is almost identical for different bandgaps. This means that

63
the maximum QE is purely decided by the value of L and insensitive to the change of

bandgap. Also, the small value of L (0.45) in narrow bandgap materials serves as an

obstacle to achieving a high conversion efficiency (15% as shown in Figure 3-6) in single-

absorber TPV devices.

3.3 Efficiency improvement in multistage TPV devices

3.3.1 Enhancement of open-circuit voltage

The structure and operation principle of ICTPV devices are described in Chapter 2.

Figure 3-7 shows the chemical potentials (designated by the flat lines) across individual

stages for an ICTPV device under illumination, which adds up to generate a high open-

circuit voltage. Each stage in a multistage ICTPV device operates in the same manner as a

single-absorber device. The equations in the preceding section can be directly applied to

the individual stages in a multistage device. The net current flowing in the mth stage is

given by:
𝑞𝑉𝑚
𝐽𝑚 = 𝛷𝑚 𝑄𝐸𝑚 − 𝑔𝑡ℎ 𝐿𝑡𝑎𝑛ℎ(𝑑𝑚 /𝐿)(𝑒 𝑘𝑏𝑇 − 1) (3-9)

where m is the incident flux on the mth stage, QEm is the effective quantum efficiency

given by Equation 2-5, Vm is the voltage across the mth stage, and dm is the absorber

thickness. The optimized multistage device is designed to have an equal photocurrent in

each stage. This current matching condition is realized with an iterative process by varying

the thickness of each stage so that the contribution of photocurrent from each is equal. The

optimal absorber thicknesses are then found by selecting the photocurrent-matched

absorber sequence that maximizes the output power.

64
Figure 3-7: Schematic of a three-stage ICTPV device under forward voltage and
illumination. Optical generation gphm, thermal generation gthm and recombination Rm,
along with the chemical potentials m in each stage are shown, where the index m
denotes the stage ordinal. The flat quasi-Fermi levels (designated with 1, 2, 3
and 4) correspond to the case where the diffusion length is infinite.

Based on Equation 3-9, the J-V characteristic of a multistage TPV device is

obtained by adding together the voltage across each stage:

𝑁
𝑐 𝑘𝑏 𝑇 𝛷𝑚 𝑄𝐸𝑚 −𝐽
𝑉 = ∑𝑚=1 𝑙𝑛[ + 1] (3-10)
𝑞 𝑔𝑡ℎ 𝐿𝑡𝑎𝑛ℎ(𝑑𝑚 /𝐿)

Then the open-circuit voltage of a multistage device can be derived by setting J=0 in

Equation 3-10. After correcting m with absorption in the upper stages, the expression of

Voc can be written as:

𝑘𝑏 𝑇 𝑁 𝑄𝐸𝑚 𝑁
𝑉𝑜𝑐 = [𝑁𝑐 𝑙𝑛() + ∑𝑚=1
𝑐
𝑙𝑛 ( 𝑐
) − ∑𝑚=2 ∑𝑚−1
𝑖=1 𝛼𝑑𝑖 ] (3-11)
𝑞 tanh(𝑑𝑚 /𝐿)

where  is the ratio of photon flux to thermal flux, defined by Equation 3-7. The third term

on the right side of Equation 3-11 represents light attenuation. According to Equation 3-

11, when  is substantially high, the Voc of a multistage device is dominated by the first

term in Equation 3-11 since the last two terms are negligible. This implies that the Voc of a

multistage device nearly scales with number stages when the photon flux to thermal flux

ratio is very high. This speculation is confirmed by the calculations for the incident power

65
density of 50 W/cm2 (corresponding to a minimum  of 305) as shown in Figure 3-8. The

open-circuit voltage enhancement in this figure is the Voc(Nc) of the optimized multistage

device normalized to the Voc(1) of the optimized single-absorber device. Note that, the

parameters used in the calculation are same as those for single-absorber devices, as

presented in Table 3-1. In different scenarios, the normalized open-circuit voltage almost

scales with Nc. The slopes are only slightly lower than unity (indicated by the dashed purple

line in Figure 3-8, i.e. Voc(Nc)/Voc(1)=Nc) due to light attenuation in the optically deeper

stages. For example, the slope is about 0.9 for L=0.45 and 0.95 for L=1.5 and 4.5. This

good consistent linear proportionality for a wide range of L will lead to a universal

enhancement of conversion efficiency in multistage ICTPV devices compared to single-

absorber devices.
Open-circuit voltage enhancement

30
Voc(Nc)/Voc(1)=Nc
25 L=0.45
L=1.5
20 L=4.5

15

10

0
0 5 10 15 20 25 30
Number of stages
Figure 3-8: Calculated open-circuit voltage enhancement Voc(Nc)/Voc(1) as a function
of number of stages. The dashed purple line indicates Voc(Nc)/Voc(1)=Nc. In the
calculations, L was set at 0.45. 1.5 and 4.5. The incident power density is 50 W/cm2.

66
3.3.2 Enhancement of conversion efficiency

Figure 3-9 shows the calculated conversion efficiency as a function of number of

stages for different values of L. The  is increased from 7%, 23% and 59% in a single-

absorber structure to 17%, 33% and 68% in a multistage IC architecture for L equal to

0.45, 1.5 and 4.5, respectively. Therefore, the  has a universal absolute increase of 9-10%

regardless of the value of L. In terms of relative change, it is more pronounced for small

value of L. For example, for L=0.45, the  of multistage devices is more than twice that

for single-stage cells. This can be explained by resorting to the preceding analysis. When

 is high enough, the following equations hold: Voc(Nc)/Voc(1)Nc (Figure 3-8) and

FF(Nc)/FF(1)1 according to Equation 3-9 and 3-11. Then efficiency enhancement in a

multi-stage structure (Nc)/(1) is approximated as:

𝜂(𝑁𝑐 ) 𝑄𝐸(𝑁𝑐 ) 𝑉𝑜𝑐 (𝑁𝑐 ) 𝐹𝐹(𝑁𝑐 ) 𝑁𝑐 ∙𝑄𝐸(𝑁𝑐 )


= ∙ ∙ ≈ (3-12)
𝜂(1) 𝑄𝐸(1) 𝑉𝑜𝑐 (1) 𝐹𝐹(1) 𝑄𝐸(1)

According to this equation, it is evident that the efficiency enhancement of a multistage

device is essentially due to its increased particle conversion efficiency 𝜂part that is defined

as Nc·QE(Nc) for current-matched IC structures. Increasing the number of stages increases

𝜂part, although it shortens the absorber thicknesses and reduces QE(Nc). This explains why

the  increases with the number of stages as shown in Figure 3-9. From Figure 3-1, QE(1)

is low for small L, hence the (Nc)/(1) can be substantial. For large L, QE(1) is

relatively high, so the (Nc)/(1) is less significant, but still exceeds unity. This manifestly

shows how the multistage structures enhance through an increased particle efficiency with

shortened individual absorbers for high collection of photo-generated carriers, which could

otherwise be lost to recombination with a long single absorber.

67
70
L=4.5

Conversion efficiency (%)


60

50

40
L=1.5
30

20 L=0.45
10

0 5 10 15 20
Number of stages
Figure 3-9: Calculated conversion efficiency for optimized multistage cells as a
function of number of stages. The calculation is done for L=0.45 1.5 and 4.5. The
incident power density is 50 W/cm2.

According to Equation 3-12, if Voc(Nc)/Voc(1)=Nc and Nc·QE(Nc)=1 (with

sufficiently large number stages), the efficiency enhancement is equal to 1/QE(1). This

maximum efficiency enhancement is calculated as a function of L and is indicated by the

solid purple curve in Figure 3-10. However, it is higher than the real efficiency

enhancement as represented by the dashed olive curve in Figure 3-10, since Voc(Nc)/Voc(1)

is slightly lower than Nc in practical case (See Figure 3-8). Also displayed in Figure 3-10

are the calculated maximum efficiencies that can be achieved by single-absorber and

multistage devices with two different bandgaps (i.e. 0.29 and 0.4 eV). The number of stages

of the multistage devices is twenty, which is large enough to reach the plateau value of

achievable efficiency (as shown in Figure 3-9). For the two different bandgaps, almost the

same improvement is observed. When the bandgap is increased to 0.4 eV, it’s possible to

achieve a conversion efficiency of 30% with a multistage IC architecture even for a small

L value of 0.45. This elucidates that multistage IC architecture is an effective strategy to

universally improve the device performance in a wide infrared spectral range.

68
3.3 50
red curve: Eg= 0.29 eV
3.0 blue curve: Eg=0.4 eV

Conversion efficiency (%)


Efficiency enhancement
40
ge
i-sta
2.5 mult
30
ber
bsor
sin gle-a
2.0
20
Eq. (
15 )
1.5 actual
10

0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0


L
Figure 3-10: Calculated maximum conversion efficiency and conversion efficiency
enhancement as a function of L. The incident power density is 50 W/cm2.

3.4 Performance of TPV devices under variable illumination sources

3.4.1 Single-absorber TPV cells

In the preceding two sections, the incident photon energy is set equal to bandgap,

which inevitably overestimates the conversion efficiency to some degree. This is because,

to excite electron-hole pairs, the photon energy needs to be higher than the bandgap. On

the other hand, a high photon energy will escalate thermalization loss. Here, to evluate the

dependence of device performance on incident photon energy, an energy-dependent

absorption coefficient (E)=1.9×(hv-Eg)1/2 m-1 is used, which matches well with that for

quaternary Ga0.44In0.56As0.5Sb0.5 with a bandgap of 0.29 eV [165]. This bulk material has

been employed as the absorbers in IC detectors and the pre-factor 1.9 m-1 is assumed as

the best fit to the experiment data [165]. Table 3-2 shows all the parameters used in the

calculations for this Ga0.44In0.56As0.5Sb0.5 material, some of which are same with those for

InAs/GaSb SL in Table 3-1.

69
Table 3-2: Parameters used in calculation for bulk Ga0.44In0.56As0.5Sb0.5.

Temperature and bandgap Tdevice=300 K, Eg=0.29 eV


Effective mass 𝑚𝑒 = 0.028𝑚0 , 𝑚ℎ = 0.51𝑚0
Density of states Nc=1.2×1017 cm-3, Nv=9.2×1018 cm-3
−𝐸𝑔
Intrinsic carrier concentration 𝑛𝑖 = √𝑁c 𝑁v exp ( )= 3.8×1015 cm-3
2𝑘𝑏 𝑇
p-type doping concentration
in the absorber Na=2.8×1016 cm-3

Refractive index n=3.5


Absorption coefficient: α = 1.9 × √𝐸 − 𝐸𝑔 (m-1)
𝜃=𝜋 𝜑=2𝜋 𝐸=∞
𝑛2 2𝐸 2 𝐸
𝐵= 2 ∫ ∫ ∫ 𝛼(𝐸) [𝑒𝑥𝑝 ( ) − 1]−1 𝑠𝑖𝑛𝜃𝑑𝜃𝑑𝜑𝑑𝐸
Radiative recombination 𝑛𝑖 ℎ3 𝑐 2 𝑘𝑏 𝑇
coefficient [29] 𝜃=0 𝜑=0 𝐸=𝐸𝑔

B=2.81×10-11 cm-3·s-1
Actual lifetime = ns, L=1.5 m
Electron mobility e=43 cm2·V-1·s-1, calculated with L =1.5-m and =20-ns.
Radiative lifetime 𝜏𝑟𝑎𝑑 = 1⁄𝐵𝑁a = 1.3 μs (no photon recycling), L=12 m

Figure 3-11 shows the calculated conversion efficiency of a single-absorber device

as a function of incident wavelength for different diffusion lengths and under illumination

at 50 W/cm2. Note that the  at each wavelength and diffusion length is the maximum

value with the optimal absorber thickness. Given a constant electron mobility of 43 cm2·V-

1
·s-1, the diffusion length increases from 1.5 m to 15 m while the carrier lifetime is

enhanced from 20 ns to 2.0 s. For a long diffusion length, the QE is high (Figure 3-1) and

the reaches the maximum value at an incident photon energy that is closely matched with

the bandgap. Conversely, when the diffusion length is short, the QE is low at a photon

energy close to the bandgap, thus resulting in a low . By increasing the energy of incident

photons, the  can be improved since the absorption coefficient and QE is enhanced. This

leads to a blue shift of the peak value of  as indicated by the black arrow in Figure 3-11.

This result for a short diffusion length goes against the conventional view that the incident

70
photon energy should be very close to the bandgap for best conversion efficiency.

However, based on above analyses, it is comprehensible from the perspective of the QE

for narrow-bandgap TPV cells with small L.

40

Conversion efficiency (%)


35 m
= 15
30 L
2 m
25 L=1
m
L=8
20
L=5 
m
15
L=3 m
10
L=1.5 m
5
0
2.0 2.5 3.0 3.5 4.0
Wavelength (m)
Figure 3-11: Calculated conversion efficiency of a GaInAsSb single-absorber device
vs wavelength for various values of diffusion length. The incident power density is 50
W/cm2.

3.4.2 Multistage ICTPV devices

As for multistage TPV cells, there is more flexibility to maximize the conversion

efficiency under different incident photon energies because of the multiple adjustable

parameters. Despite more complicated than the single-absorber structure, the multistage IC

architecture offers an effective way of dealing with the diffusion length limitation and thus

to maximize the  at a photon energy close to the bandgap. One important consideration

in the design of a multistage TPV device is the photocurrent match between stages. If the

current is mismatched, the QE decreases with incident photon energy, which can be

partially caused by the light attenuation. Figure 3-12 shows the calculated  for four

multistage structures. These structures have the optimal current matched absorbers that

71
were designed based on an absorption coefficient of 0=3000 cm-1 (at a wavelength 0 of

4 m, close to the cutoff wavelength of 4.3 m) to maximize conversion efficiency. With

the same number of stages, the individual absorber thicknesses differ considerably for

different diffusion lengths. For example, for the 5-stage devices, the optimal absorber

thickness sequence (nm) is 476/562/693/923/1567 and 603/733/941/1318/2225 for a

diffusion length equal to 1.5 and 15 m, respectively. The light attenuation is significant

in the thick absorbers for the case of L=15 m, therefore, there are dramatic reductions in

the QE and  at short wavelengths. This explains why  of the 20-srage device with L=15

m is even lower than that for L=1.5 m at wavelengths shorter than 3 m. These results

illustrate the importance of retaining current match when selecting the illumination source

for multistage devices.

In addition, the  of four devices peaks at a wavelength slightly shorter than 0,

where the QE reaches its maximum with current match. This is because the open-circuit

voltage and fill factor are both somewhat higher at a wavelength slightly shorter than 0.

Alternatively, one can optimize the multistage structure based on the measured absorption

coefficient at every given wavelength such that the  at each wavelength reaches the

maximum value that is achievable by a multistage architecture. This is illustrated by the

olive curves in Figure 3-12 for two diffusion lengths. The total absorber thickness of each

optimized structure is about 8 m for wavelength near the bandgap, and the number of

stages for each structure is twenty. For example, for =4 m and L=1.5 m (represented

by the solid olive curve in Figure 3-12), the optimal absorber thickness (nm) sequence is

148/156/165/174/183/194/206/220/236/254/276/301/332/370/419/483/573/711/960/175

with a total absorber thickness of about 8.1 m. In contrast to single-absorber devices, the

72
maximum  of the optimized multistage devices always occurs at an incident photon

energy very close to the bandgap, regardless of the magnitude of the diffusion length. This

further validates the advantages and flexibility of the multistage architecture.

50
5-stage L=15 m
20-stage

Conversion efficiency (%)


40
optimized

30

20

10
L=1.5 m

0
2.0 2.5 3.0 3.5 4.0
Wavelength (m)
Figure 3-12: Calculated conversion efficiency for the 5- and 20-stage devices with
L=1.5 m (solid curves) and 15 m (dashed curves). The absorbers were adjusted to
be photocurrent matched with an absorption coefficient of 3000 cm -1, corresponding
to a wavelength of 4 m. The calculated maximum efficiencies with optimized multi-
stage structures at every wavelength are represented by the olive curves. The incident
power density is 50 W/cm2.

3.5 Summary and concluding remarks

To recapitulate, in this chapter, the efficiency limiting factors in narrow bandgap

TPV cells are identified and how they affect the device performance is discussed. These

factors are highly correlated with high dark saturation current density, short carrier lifetime,

relatively small absorption coefficient and finite diffusion length. As an example, narrow-

bandgap InAs/GaSb SLs are used to illustrate the specific impact of these factors on

conversion efficiency and how the device performance can be improved by adjusting

material parameters such as the product L. One way to increase L is to employ Ga-free

73
InAs/InAsSb SLs for absorbers with a relatively long carrier lifetime [166-167].

Furthermore, it is shown that multistage IC structure is capable to overcome the diffusion

length limitation and achieve a particle conversion efficiency approaching 100%, therefore

increasing the conversion efficiency by about 10% in a wide range of L values and

bandgaps. The enhancement of conversion efficiency with multistage IC structure is

especially impressive for small values of L, for which the conversion efficiency is more

than double that in the single-absorber TPV devices. In addition, the entire structure’s

flexibility and other advantages of multistage structure offer the possibility to achieve

maximum conversion efficiency with the incident photon energy close to the bandgap.

Nevertheless, as with single-absorber TPV devices, the issues of relatively low fill factor

and voltage efficiency (=qVoc/(NcEg) for IC structures) remain. These issues are directly

related to the high dark saturation current density in narrow bandgap materials. To resolve

them, an approach that can significantly increase the photocurrent without requiring a

higher incident power density needs to be implemented, which should be one of future

research focuses.

74
4 Chapter 4: Experimental comparison between single-absorber and
multistage IC thermophotovoltaic devices
4.1 Background and motivation

In chapter 3, theoretical evidence and illustrated scenarios are presented to prove

the advantage of multistage ICTPV devices over single-absorber TPV devices.

Specifically, IC structures are shown to be capable to enhance the conversion efficiency by

promoting the collection of photo-generated carriers. In this chapter, broad experimental

proof is furnished to support the advantage of IC devices, especially in concerns of

enhanced collection efficiency. Aside from InAs/GaSb T2SLs that are treated in Chapter

3, the advantage of IC structures is also true for other narrow bandgap materials since their

diffusion length and absorption coefficient are limited as well. For example, the bulk InAs

and InSb (either intrinsic or lightly doped) typically have  in the range of 1000-3000 cm-
1
near bandgap. Their L can be several microns at room temperature but may be shortened

significantly under strong illumination due to the high concentration of excess carriers.

The effects of small  and short L on single-absorber TPV performance are

illustrated in Figure 4-1, where the calculated QE and collection efficiency (c) are plotted

as functions of normalized absorber thickness (d/L). The calculation of QE is carried out

based on Equation 2-4 without considering the surface reflection of light. The collection

efficiency is defined as the ratio of collected carriers to absorbed photons and is equal to

QE/[1-exp(-d)]. For single-absorber devices, adequate absorption of incident light

necessities a thick absorber, especially with a small . However, if the diffusion length is

short, QE will not increase further with absorber thickness after d≈L as shown in Figure 4-

1(a). This is because some photogenerated carriers recombine before being collected and

75
the collection efficiency is reduced with absorber thickness. The reduction of collection

efficiency with increasing d is more significant when L<1, as shown in Figure 4-1(a).

Also, for L1, the QE peaks at a certain finite absorber thickness, because the collection

probability (defined by Equation 2-1) of photogenerated carriers is reduced with absorber

thickness. A high collection efficiency (>90%) can be obtained only when the absorber is

thinner than the diffusion length (or thinner than 0.6L for L<1) as shown in Figure 4-1(a).

In addition, the open-circuit voltage, defined by Equation 3-4, is reduced with a limited

collection efficiency. This is illustrated by the open-circuit voltage factor VF=

ln[QE/tanh(d/L)] in Figure 4-1(b), where the dotted curves are calculated assuming

complete collection of carriers while solid curves are based on the calculated QE in Figure

4-1(a) with a limited collection efficiency. As can be seen, comparatively, VF is decreased

considerably with a limited collection efficiency especially when L<1 and d>L. For

instance, for L=0.35 and d=3L, VF is decreased by 0.91, resulting in a reduction of Voc by

24 mV at 300 K. Hence, the considerably reduced VF coupled with the limited QE due to

the finite diffusion length will result in a poor conversion efficiency when L is less than

unity.

In this chapter, a comparative study of three TPV devices is presented to

experimentally confirm the advantage of multistage architecture, as well as to examine how

different configurations affect device performance. One of the three devices has single-

absorber structure while the others are three- and five-stage IC devices. The bandgap of the

InAs/GaSb T2SLs in these devices is about 0.2 eV at 300 K, which is the narrowest

bandgap ever reported so far in TPV cells.

76
Normalized absorber thickness (d/L)
0.1 0.5 1 10
1.0 1.0
0.9 L=5.1 0.9
(a)
0.8 0.8

Quantum efficiency
Collection efficiency
0.7 0.7
0.6 L=1.1 0.6
0.5 0.5
0.4 0.4
0.3 L=0.35 0.3
0.2 0.2
0.1 0.1
0.0 0.0
Solid: L-limited collection
1.0 L=5.1 Dotted: 100% collected
Normalized Voc factor

(b)
0.5
L=1.1
0.0

-0.5

-1.0 L=0.35

-1.5
0.1 0.5 1 10
Normalized absorber thickness (d/L)
Figure 4-1: (a) Calculated quantum efficiency and collection efficiency, and (b) open-
circuit voltage factor as a function of normalized absorber thickness for several values
of L. VF initially decreases with increasing d/L due to the nearly linear increase of
dark current when d/L is small.

4.2 Device structure, growth and fabrication

The three TPV structures are grown by GENxplor MBE system (Figure 2-11) on

nominally undoped p-type GaSb (001) substrates. In the three structures, each period of

the SL absorber is composed of four layers: InSb (1.2 Å), InAs (20.5 Å), InSb (1.2 Å) and

GaSb (25.1 Å). The two thin InSb layers were inserted to balance the tensile strain of the

InAs layer [168]. The absorbers in the three structures are p-type doped to 2.6×1016 cm-3.

In the two multistage structures, the individual absorber thickness was increased in the

optically deeper stages to achieve current match between stages by compensating for light

77
attenuation. The current-matched absorbers were deigned based on the absorption

coefficient of 3000 cm-1 for a monochromatic light source and the assumption of full

collection of photo-generated carriers. The absorber thickness for the 1-stage device is 2.31

m. The 3-stage device has a total absorber thickness equal to that of the 1-stage device

with the discrete individual thicknesses of 624, 749 and 936 nm from surface to the

substrate. The individual absorber thicknesses in the 5-stage device are 360, 408, 480, 576

and 696 nm, and the total absorber thickness is 2.52 m, slightly longer than the 1- and 3-

stage devices. The electron barriers in the three devices were made of four digitally

GaSb/AlSb QWs with GaSb well thicknesses of 33/43/58/73 Å. The hole barriers consist

of eight digitally graded InAs/AlSb QWs with the InAs well thicknesses (in Å) of

32/34/36/40/45/52/60/71. The schematic layer structures of the three devices are shown

Figure 4-2. After MBE growth, square mesa devices with edge lengths ranging from 50 to

1000 m are processed by using conventional contact lithography and wet etching. For

passivation, two layers composed of Si3N4 followed by SiO2 are used for improving overall

stress management and minimizing pin holes. Finally, Ti/Au contacts are deposited by

sputtering, and then the devices are mounted on heat sinks and wire bonded for

characterization.

78
Figure 4-2: Schematic layer structures of the three TPV devices with one, three and
five stages.

4.3 Device characterizations and discussions

4.3.1 Quantum efficiency

The QEs of the three devices were measured using a FTIR spectrometer and a

calibrated blackbody radiation source with a temperature of 800 K and a 2 field of view

(FOV). The blackbody source had an aperture of 0.76 cm and was placed at 30 cm from

the device. Figure 4-3 shows the calibrated QE spectra at 300 and 340 K for the

representative 0.2×0.2 mm2 devices processed from the three wafers. Because of current

continuity in multistage IC structure, the device QE is decided by the stage with weakest

response, therefore the measured QE reflects the actual device performance and is more

meaningful than the effective QE for any individual stages. As can be seen in Figure 4-3,

at 300 K, the 1- and 3-stage devices have a 100% cutoff wavelength of 5.5 m, which

corresponds to a bandgap of 225 meV. By comparison, the 5-stage device has a slightly

79
longer 100% cutoff wavelength of 5.8 m with the SL absorber bandgap estimated to be

214 meV. Since the QE is roughly proportional to the individual absorber thickness, the 5-

stage device with thinnest individual absorbers has the lowest QE, while the 1-stage device

with a 2.31-m absorber has the highest QE among the three devices. For example, at 𝜆=4

m and T=300 K, the QEs are 29.5%, 12.0%, and 8.8% for the 1-, 3- and 5-stage devices,

respectively. As the temperature in increased to 340 K, the QEs of the 1- and 3-stage

devices were decreased, while the QE of the 5-stage device was nearly unchanged. Also,

the decline of QE with temperature for the 1-stage device is more pronounced than the 3-

stage device. For example, at 𝜆=4 m, the QE was reduced to 23.6% for the 1-stage device,

compared to a small reduction to 11.3% for the 3-stage device at 340 K. The QEs were

reduced because the diffusion length was shorter at a higher temperature, leading to a

smaller collection efficiency as illustrated in Figure 4-1. This speculation is further proved

by the bias dependence of the QE at =4 m for the three devices as shown in Figure 4-4.

60

50
solid: 300 K
dashed: 340 K
40
QE (%)

30 1-stage

3-stage
20

10 5-stage

0
2 3 4 5 6
Wavelength (m)
Figure 4-3: Measured QE spectra of 1-, 3- and 5-stage devices at 300 and 340 K.

80
As can be seen in Figure 4-4, for the 1-stage and 3-stage devices, a reverse bias is

required to achieve the saturation (or maximum) value of QE with complete collection of

photo-generated carriers at 300 K. This is because the diffusion length is either shorter than

or comparable to the absorber thicknesses in the 1- and 3-stage devices. Hence, at zero

bias, some of the photo-generated carrier recombine during transport paths and do not

contribute to photocurrent. At higher temperature (e.g. 340 K), the diffusion length is even

shorter, consequently, a larger reverse bias is required to saturate the QE for the 1- and 3-

stage devices. By comparison, the diffusion length has much less impact on the 5-stage

device since its individual absorbers are much thinner. Also, the saturation values of QE

for all the devices are higher at 340 K since the absorption coefficient is enhanced due to

the bandgap narrowing with rising temperature. Thanks to the thickest absorber, the 1-stage

device has the highest QE among the three devices. However, this highest QE does not

necessarily result in the best performance among the three devices when they operate at a

forward bias voltage.

45
40
1-stage
35
solid: 300 K
30 open: 340 K
= m
25
QE (%)

3-stage
14

12

10 5-stage

8
0 -100 -200 -300 -400 -500 -600 -700
Bias (mV)

Figure 4-4: Voltage dependent QE at 4 m for the three devices, where different
vertical scales are used in the top and bottom portions to better show variations.

81
4.3.2 Particle conversion efficiency

As pointed out in Chapter 3, instead of QE, a more proper figure of merit for

multistage TPV device is the particle conversion efficiency PCE [169-170]. It is defined

as the sum of effective QEs in individual absorbers and is equal to Nc×QE for a current-

matched configuration. At 𝜆=4 m and T=300 K, current match condition is nearly

fulfilled based on the measured absorption coefficient (3159 cm-1 for 1- and 3-stage devices

and 3470 cm-1 for the 5-stage device) from the transmission measurement. Hence, the PCE

at zero bias is 29.5%, 36.0%, and 44% for the 1-, 3- and 3-stage devices at 300 K,

respectively. The highest PCE for the 5-stage device among them agrees with the projected

high collection efficiency due to thin individual absorbers. In principle, the value of PCE

can be increased up to maximum 69% (estimated by subtracting the 31% reflection loss

from the top surface) by adding more stages to fully absorb the incident photons. Also,

adding an anti-reflection coating onto the surface can raise the PCE beyond 69%.

In theory, the effective QE in the Nth stage of a multistage ICTPV device can be

calculated based on Equation 2-5. Based on Equation 2-5, together with the measured

absorption coefficient and QE, the diffusion length was extracted to be about 1.5 m at 300

K for the three devices. Evidently, at 𝜆=4 m, the product of absorption coefficient and

diffusion length (L) is smaller than unity in the three devices. Consequently, according to

Figure 4-1, the individual absorber thicknesses need to be shorter than 0.6L in order to

achieve a collection efficiency higher than 90%. The 1-stage device has an absorber

thickness that is about 1.5 times of the diffusion length and thus it has the lowest collection

efficiency at zero bias (~60% as illustrated in Figure 4-1). In comparison, the individual

absorbers in the 5-stage device are thinner than 0.6L, thus resulting in a collection

82
efficiency over 90% and the highest PCE at zero bias as discussed above.

4.3.3 Illuminated J-V curve and open-circuit voltage

In the lighted current density-voltage (J-V) measurement, a type-II IC laser (ICL)

was employed to illuminate the three devices. The narrow emission spectrum of the ICL

reproduced the characteristics of a selective emitter (or a narrow-band filter), which is

usually included in a TPV system to minimize the thermalization and below-bandgap

losses. Both experimental and theoretical efforts were devoted to nanostructured materials

for efficient narrowband emissivity near 4 m or longer wavelengths [32-33]. These

studies reinforce the feasibility and applicability of narrow bandgap TPV devices. During

the lighted J-V measurement, the IC laser was cooled down to ~80 K and continuously

delivered high output power at an emission wavelength near 4.2 μm (See inset in Figure 4-

5(b)). This emission wavelength corresponds to a photon energy of 295 meV that is 70-80

meV higher than the bandgap of the three TPV devices at 300 K. Hence, there is some

thermalization loss (20-27%) from above-bandgap photons. Nevertheless, at laser emission

wavelength, current match was almost satisfied in the 3- and 5-stage devices. The PV

characteristics of the three devices were studied at different incident power densities simply

by adjusting the injection current of the laser. The measured J-V curves at 300 K under a

medium level of illumination from the ICL are shown in Figure 4-5(a). The incident power

density Pinc was about 19 W/cm2, which was assessed through the connection between QE

and Jsc as expressed by the following equation:

1.24𝐽𝑠𝑐
𝑃𝑖𝑛𝑐 = (4-1)
𝜆𝑙𝑎𝑠𝑒𝑟 𝑄𝐸

where laser is the laser emission wavelength. This simple and effective method to estimate

83
incident power density allows to circumvent the difficulties associated with the nonuniform

and divergent beam of the edge emitting ICL laser.

Also displayed in Figure 4-5(a) are the series resistance (Rs) corrected J-V curves

and the ideal curves that were plotted in the same manner with [171]. Or rather, the ideal

J-V curve is the superposition of dark current density and the maximum photocurrent

density (Jphmax), where the photo-generated carriers are completely collected. The

magnitude of Jphmax is the difference between the saturated current densities at a reverse

bias under dark and illuminated conditions. For example, at T=300 K and Pinc=19 W/cm2,

the saturation value of current density under illuminated (dark) condition was 25.3 (2.9),

9.1 (1.1) and 5.9 (0.9) A/cm2 for the 1-, 3- and 5-stage devices, respectively. Therefore, the

corresponding Jphmax is 22.4 (1-stage), 8.0 (3-stage) and 5.0 A/cm2 (5-stage), proportional

to their individual absorber thicknesses. At the same incident power density, the Jsc values

are 9.2 A/cm2, 6.7 A/cm2, and 4.9 A/cm2 for the 1-, 3- and 5-stage devices, respectively.

These values of Jsc are higher than Jphmax values for the three devices, primarily due to

incomplete collection of photo-generated carriers particularly in the 1-stage device. Even

though the Jsc is highest in the 1-stage device, its PCE and collection efficiency are lowest,

which results in the lowest conversion efficiency described in next subsection. The high

current in the 1-stage device also results in a significant Ohmic loss in series resistance, as

reflected by the notable shift between the Rs-corrected and measured J-V curves. Instead,

the Rs-corrected J-V curves for the 3- and 5-stage devices almost coincide with the

measured J-V curves due to the relatively lower currents.

84
25
1-stage 1-stage ICL
emmision
20 (a) 15 spectrum

Current density (A/cm2)

Current density (A/cm2)


T=300 K (b)
solid: measured
15 dotted: Rs corrected
T=200 K
3.5 4.0 4.5 5.0
10 Wavelength (m)
dashed: 100% collected
10
3-stage 3-stage
5
5 5-stage 5-stage

0 0
-0.4 -0.2 0.0 0.2 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8
Voltage (V) Voltage (V)

Figure 4-5: (a) Current-voltage characteristics of the three devices at 300 K under a
medium illumination level where the incident power density was about 19 W/cm2. The
solid, dotted and dashed curves correspond to the measured, Rs corrected and ideal
cases, respectively. (b) Current-voltage characteristics of the three devices at 200 K
under the same level of illumination as in (a). The inset shows the emission spectrum
of the ICL.

The primary reason for the low collection efficiency in the 1-stage device at 300K

was because the diffusion length was shorter than absorber thickness. This can be further

confirmed by examining the behaviors at a low temperature where the diffusion length

should be longer. Figure 4-5(b) shows the measured J-V curves of the three devices at 200

K under the same illumination level as in Figure 4-5(a) from the ICL. As shown, for the 3-

and 5-stage devices, the onset of current saturation occurs at a certain forward voltage

rather than a reverse voltage. This suggests that complete collection of photogenerated

carriers was achieved under a forward voltage and the diffusion length was increased

significantly beyond the absorber thicknesses in the 3-and 5-stage devices. The increased

diffusion length also improved the collection efficiency (~72% at zero bias) in the 1-stage

device, although it was still below 100% since the diffusion length was shorter than the

absorber thickness (2.31 m). Also, because of the reduced dark saturation current (orders

of magnitude lower than the photocurrent), the Voc was appreciably higher for the three

85
devices at 200 K. On the other hand, at 200 K, the Jphmax under this illumination level

dropped to 18.5, 6.9 and 4.4 A/cm2 for the 1-, 3- and 5-stage devices, respectively. This is

because the absorption coefficient decreased due to bandgap widening at lower

temperatures.

Aside from a higher collection efficiency compared to the single-stage device, the

multistage IC structure can also create a Voc far exceeding the individual absorber bandgap.

For example, at T=200 K and Pinc=19 W/cm2, the measured Voc was 170 (1-stage), 513 (3-

stage) and 745 meV (5-stage), corresponding to a voltage efficiency of 67%, 68% and 63%,

respectively. As the temperature increased to 300 K, the Voc at the same illumination level

dropped to 72 (1-stage), 223 (3-stage) and 287 meV (5-stage) with a corresponding voltage

efficiency of 32%, 33% and 27%, respectively. Presumably, the slightly lower voltage

efficiency in the 5-stage device was due to the narrower bandgap and poorer material

quality, which collectively resulted in a much higher thermal generation rate (about two

times higher as estimated in Subsection 4.4.2) than in the 3-stage devices at 300 K.

Specifically, the Voc could be reduced by ~90 mV (amplified by about 5 times with five

cascade stages [169-170]) due to the doubling of the thermal generation rate. On the same

account, the Voc of the 5-stage was lower than the 3-stage device in the ideal case as well.

In addition, the Voc and voltage efficiency increased when the incident power density was

enhanced. For example, at T=300 K and Pinc=36 W/cm2 (highest illumination level

available from the ICL), the measured Voc was 85, 271 and 371 mV for the 1-, 3- and 5-

stage devices, respectively.

86
4.3.4 Fill Factor and conversion efficiency

Figure 4-6 shows the measured Voc, FF, maximum output power density (Pmax), and

conversion efficiency () as functions of incident power density at 300 K for the three

devices. At the maximum incident power density (36 W/cm2), the FF was 25%, 28% and

38% for the 1-, 3- and 5-stage devices, respectively. Throughout the whole range of

incident power density, the 1-stage device had the lowest FF due to the lowest collection

efficiency and a greater series resistance loss, while the 5-stage device had the highest FF

because of the highest collection efficiency. Under the highest illumination level, the

maximum output power was harvested at a voltage of 43, 136 and 226 meV for the 1-, 3-

and 5-stage devices, respectively. At this voltage, the extracted collection efficiencies (See

Figure 4-7) were about 29% (1-stage), 53% (3-stage) and 87% (5-stage). If, however, the

photogenerated carriers were fully collected as in the ideal case, the FF would increase to

32%, 36% and 39% for the 1-, 3- and 5-stage devices, respectively. From this point of

view, the 5-stage device with thin absorbers is nearest to the ideal case for maximum output

power. The FFs of the 1- and 3-stage devices were also observed to peak at a certain

incident power density and then fall off with further increasing the incident optical power.

This behavior was possibly related to the larger current and the resulting higher Ohmic

losses in series resistances. In contrast, the FF of the 5-stage device exhibited a monotonic

rise with increasing incident power. The FFs of the three devices were considerably lower

than the typical values (~60-70%) of TPV cells with bandgaps of 0.5-0.6 eV [5], but they

are reasonable for narrow bandgap (~0.2 eV) TPV cells with un-optimized structures.

87
38
350 one-stage one-stage
three-stage three-stage 36
300 five-stage five-stage
34
250
32
Voc (mV)

FF (%)
200
150 (a) (b) 30

100 28
50 26
0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Pinc (W/cm2) Pinc (W/cm2)

one-stage one-stage 3.5


1.2
three-stage three-stage
3.0
1.0 five-stage five-stage
2.5
Pmax (W/cm2)

0.8
2.0

 ()
0.6
(c) (d) 1.5
0.4
1.0
0.2 0.5
0.0 0.0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Pinc (W/cm2) Pinc (W/cm2)

Figure 4-6: (a) Open-circuit voltage, (b) fill factor, (c) maximum output power density
and (d) conversion efficiency as a function of incident power density for the three
devices at 300K.

The maximum conversion efficiencies at 300 K are 0.9% (1-stage), 2.5% (3-stage)

and 3.6% (5-stage) as shown in Figure 4-6(d). The 5-stage device attained the highest

power efficiency thanks largely to the efficient collection of photogenerated carriers. This

unambiguously verifies the advantage of multistage IC structures with thin individual

absorbers for narrow bandgap TPV cells. The main reason for the relatively low conversion

efficiency in the three devices was because the dark current was significant in such a

narrow band gap (~0.2 eV) structure. Other factors include the contact resistances, some

thermalization loss (20-27%), surface reflection (31%), as well as incomplete absorption

88
(~50%) due to insufficient thick total absorber (≤2.52 m). At lower illumination levels

(Pinc<5 W/cm2), the conversion efficiency of the 5-stage device was slightly lower than the

3-stage device due to the narrower bandgap and the higher thermal generation rate, as will

be given in Subsection 4.4.2. For example, at Pinc=3.5 W/cm2, the conversion efficiency

was respectively 0.94% and 0.88% for the 3- and 5-stage devices, although the Voc of the

5-stage device was somewhat higher than the 3-stage device (103 vs. 95 mV). In fact, the

conversion efficiencies of the two multistage devices can be further enhanced by increasing

the incident power, as the conversion efficiencies have not yet saturated even at 36 W/cm2.

This can be accomplished with built-in lenses on the device surface and by exploring the

photonic structure or metamaterial. In contrast, the conversion efficiency of the 1-stage

device dropped by about 16% after saturation, which is consistent with the trend of FF

with incident power. In addition to FF, the increased Ohmic losses at higher incident power

in the 1-stage device provided another mechanism for reducing the conversion efficiency

after saturation. In contrast, for the 3-stage device, the rapid increase of Voc overcame the

decrease of FF with increasing incident power, and the Ohmic loss in the 3-stage device

was lower than in the 1-stage device. Consequently, similar trends of conversion efficiency

and Voc were observed for the 3-stage device. Table 4-1 summarizes the PV performance

characteristics and related parameters for the representative devices from the three wafers.

These data collectively show the capabilities and advantages of multistage ICTPV devices,

and the limitation of the single-stage TPV devices.

89
Table 4-1: Summary of the PV performance and the related parameters of
representative devices (0.2×0.2 mm2) from the three ICTPV wafers at 300 K. The
maximum efficiencies shown in the table for the 3- and 5-stage devices are obtained
at a maximum incident power density of 36 W/cm2.

c (0) Jsc Jphmax Voc FF Pmax Maximum Rs


Device
(%) (A/cm2) (A/cm2) (mV) (%) (W/cm2)  (%) ()

1-stage 40 12.9 40.8 85 25 0.27 0.9 4.9

3-stage 76 12.2 15.7 271 28 0.91 2.5 4.6

5-stage 95 9.3 9.7 371 38 1.29 3.6 4.7

4.4 Extraction of some important performance related parameters

4.4.1 Voltage-dependent collection efficiency

In Figure 4-5(a), there is a common characteristic for the three devices, namely a

shift between the measured and ideal J-V curves. This shift is particularly striking for the

1-stage device, significantly reduces for the 3-stage device, and almost disappears for the

5-stage device. The implies that the collection efficiencies and the photocurrents in the

three devices are voltage-dependent, and the illuminated J-V curves do not comply with

the usual superposition principle [7]. This voltage-dependent characteristic has been

reported for solar cells made of Silicon [172-174], CdS/CdTe [171, 175-176], CdS/CdInSe2

[177-178] and GaAs [173]. In thses solar cells, the voltage-dependent characteristic mainly

arises from the variation of the electrical field in the depletion region when the applied

external voltage is changed. By comparison, the diffusion process plays a more important

role in ICTPV structures.

The voltage-dependent collection efficiency c (V) can be obtained through the

approach described in [171, 174, 178]. This approach relies on two assumptions: First, the

photocurrent density can be written as the Jphmax times c (V): Jph (V)= Jphmax·c (V).

90
Second, the dark current density is assumed to remain unchanged at different incident

power densities [171, 174, 178]. Applying this approach to the current three devices, the

c (V) can be expressed as:

𝐽2 (𝑉)−𝐽1 (𝑉)
𝜂𝑐 (𝑉) = (4-2)
𝐽2𝑝ℎ𝑚𝑎𝑥−𝐽1𝑝ℎ𝑚𝑎𝑥

where J1(V) and J2(V) are the current densities at two different incident power densities,

and J1phmax and J2phmax are the corresponding maximum photocurrent densities. For each

device at 300 K, four J-V curves were selected at incident power densities of 19, 13, 7

W/cm2 and the dark condition to extract c (V) as shown in Figure 4-7(a). As can be seen,

the extracted c (V) from different pairs of J-V data does not exactly overlap for the 1- and

3-stasge devices. This suggests that the dark current might change with the incident power

density, which can be partially explained by large number of photo-generated excess

carriers shortening the carrier lifetime. Another possibility was the small variation of

device temperature (<1 K according to the estimated thermal resistance for IC structures

[179] and incident power), which may affect the dark injection current contribution,

especially at high incident power densities. For this reason, the J-V pairs at relatively low

incident power densities were used to extract c (V) as shown in Figure 4-7(a). However,

this effect somehow becomes insignificant when the individual absorbers are thin, as

evidenced by the almost overlapped c (V) profiles with different pairs of incident power

densities for the 5-stage device. Another factor is the surface leakage due to imperfect

passivation and active surface stages on the etched sidewalls, which will be discussed in

Subsection 4.4.4. Note that the possible variations of the diffusion length due to the small

change of temperature (<1 K) under different incident power densities should be negligible,

since the QE would only differ by at most 0.15% with a 1 K deviation at 4.25 m as shown

91
Figure 4-3. The temperature variation for a larger size device might be larger under

intensive illumination, but still can be addressed with effective thermal dissipation through

a heat sink. For example, based on the previously extracted data for IC structures [179],

the specific thermal resistance (Rsth) for a device with side dimension of 1 mm is lower

than 100 Kcm2/kW. An incident power density of 36 W/cm2 would increase the device’s

temperature by at most 3.6 K (with effective heat conduction through the substrate to a heat

sink) compared to its temperature in the dark.

1.0 1.0
(a) (b)
0.8 0.8
5-stage 5-stage
0.6 0.6
c

c

0.4 3-stage 0.4 3-stage


19-7
19-dark
13-7 0.2
0.2 1-stage 1-stage
13-dark

0.0 0.0
-0.5 0.0 0.5 1.0 -0.5 0.0 0.5 1.0
Voltage (V) Voltage (V)
Figure 4-7: (a) Voltage dependence of collection efficiency derived from Equation 4-
2 using four different pairs of J-V data at 300 K for the three devices. The numbers
in the legend indicate the incident power densities under different illumination levels.
(b) Average collection efficiency over the four pairs in (a).

For ease of comparison, the average of the four c (V) curves in Figure 4-7(a) is

plotted in Figure 4-7(b). As shown, the 5-stage device had the highest average c (V), while

the 1-stage device had the lowest average c (V) among the three devices. At zero bias, the

c (0) was 40%, 76% and 95% for the 1-, 3- and 5-stage devices, respectively. For the 1-

stage device, at least 80% of the photo-generated carriers were not collected at forward

bias (>0.1 V), as reflected by the small c (V) (<0.2). This small c (V) severely penalized

the fill factor and conversion efficiency as discussed in Subsection 4.3.4. The extracted c

92
(0) was substantially smaller than the theoretical projection (~60%) shown in Figure 4-

1(a), especially for the 1-stage device. This was likely caused by the shutting of surface

leakage as mentioned earlier. As shown in Figure 4-10, there is significant surface leakage

in dark condition especially in the 1-stage device. Likewise, under illuminated condition,

large number of photogenerated carriers could leak through the rough sidewalls, thus

reducing the collection efficiency. At this moment, why the 1-stage device had most

notable surface leakage is not fully understood, and it is worth exploring in the further

research.

4.4.2 Thermal generation rate and carrier lifetime

The relatively low conversion efficiencies (5%) in the three devices were

primarily due to the high dark current density associated with the high thermal generation

rate (gth) and a relatively short carrier lifetime () in narrow bandgap InAs/GaSb T2SL

absorbers. As will be described in Chapter 6, there is a simple and effective method to

extract thermal generation rate and carrier lifetime in IC structures. This method is

particularly suitable for multistage IC devices since their dark current densities usually

exhibit clear and large linear regions at reverse bias [158]. In this method, the gth is first

found from the intercept of the linear fitting of dark current at large reverse bias [158]. The

carrier lifetime then can be calculated from the gth based on the equation:

𝑛𝑖2
𝑔𝑡ℎ = (4-3)
𝑁𝑎 𝜏

where ni is the intrinsic carrier concentration and Na is the p-type doping concentration.

The intrinsic carrier concentration is given by:

93
2𝜋𝑘𝑏 𝑇 1.5
𝑛𝑖 = 2 ( ) (𝑚𝑒 𝑚ℎ )0.75 𝑇 1.5 𝑒 −𝐸𝑔⁄2𝑘𝑏𝑇 (4-4)
ℎ2

where me and mh are the electron and hole effective masses, taken to be 0.03m0 (m0 is

electron mass) and 0.4m0, respectively. Based on Equation 4-4, the calculated intrinsic

carrier concentrations at 300 K were 1.15×1016 (1- and 3-stage) and1.44×1016 (5-stage) cm-
3
. From the linear fitting of dark current density, the thermal generation rate at 300 K was

found to be 3.81×1022, 4.55×1022 and 8.35×1022 cm-3·s-1 for the 1-, 3- and 5-stage devices,

respectively. Based on Equation 4-3, the carrier lifetime at 300 K was calculated to be 134

(1-stage), 113 (3-stage) and 89 (5-stage) ns. Compared to the 1- and 3-stage devices, the

shorter carrier lifetime and higher thermal generation rate in the 5-stage device are ascribed

to its narrower bandgap (214 meV vs. 225 meV) and poorer material quality (with

somewhat more defects and larger perpendicular lattice mismatch). In addition, both gth

and  are very strong functions of temperature in the three devices, as shown in Figure 4-

8. The sharp decrease of carrier lifetime with temperature is likely due to the growing

prevalence of Auger processes linked with bandgap narrowing of the SL absorber at high

temperatures. The thermal generation rate in the three devices is many orders of magnitude

higher than those in solar cells. For example, for a crystalline Si solar cell, the Na and  at

300 K are normally in the ranges of 1015-1016 cm-3 [7] and 0.1-1 ms [180-181], respectively.

Therefore, the gth is estimated to be 2.25×107-2.25×109 cm-3·s-1, about 13-15 orders of

magnitude lower than that in the current three ICTPV devices. Evidently, reducing the gth

either by increasing carrier lifetime or cooling down the device, even by one order of

magnitude, will boost the conversion efficiency of ICTPV devices.

94
Temperature (T)
340320300 280 260 240 220 200
700 1024
550
400 1023

Carrier lifetime  (ns)

gth (cm-3s-1)
250 1022

1021

1-stage 1020
85
70 2-stage
55 3-stage 1019
3 4 5
1000/T (K-1)
Figure 4-8: The thermal generation rate and minority carrier lifetime for the 1-, 3-
and 5-stage devices at high temperatures.

4.4.3 Series resistance

As identified in Subsection 4.4.1, the photocurrents in the three devices were

voltage dependent. This voltage-dependent characteristic creates significant complexities

when extracting the series resistance Rs using illuminated J-V curves. Even with a relatively

weak voltage dependence of photocurrent, the series resistance extracted based on a

generalized Suns-Voc method could be somewhat overestimated [182]. Hence, to avoid the

complexity caused by the voltage-dependent photocurrent, the series resistance of the three

devices were extracted from the dark condition based on the following equation [174, 178]:

𝑅𝑠 = 𝑙𝑖𝑚𝑖𝑡(𝑑𝑉 ⁄𝑑𝐼 ) (4-5)


1⁄𝐼 →0

Figure 4-9 shows the plots of dV/dI under dark conditions, as well as the extracted series

resistances for the three devices. The Rs was acquired by finding the intercept of dV/dI vs.

1/I. The extracted series resistances were respectively 4.9, 4.6 and 4.7  for the 1-, 3- and

5-stag devices, which were close to each other. This implies that the series resistances in

95
the three devices were mainly from the contacts and wires, while the resistances between

cascade stages can be ignored due to the smooth carrier transport in the type-II broken-gap

heterostructure.

12
one-stage, Rs=4.9 
11
three-stage, Rs=4.6 
10 five-stage, Rs=4.7 
dV/dI () 9

4
0 5 10 15 20 25 30 35 40
I -1 (A-1)

Figure 4-9: dV/dI data to obtain series resistance at 300 K, which was found from the
intercept of dV/dI.

4.4.4 Surface leakage

Surface leakage has been a long-standing issue for III-V based, especially T2SL

based, infrared devices [183]. Various passivation techniques were developed for T2SL

detectors with varying degree of reliability and effectiveness [184]. In principle, under dark

condition, the effect of surface leakage can be quantified through the linear fitting between

P/A and 1/R0A [185]:

1 1 1 𝑃
=( ) + ( ) (4-6)
𝑅0 𝐴 𝑅0 𝐴 𝑏𝑢𝑙𝑘 𝜌𝑠𝑤 𝐴

where sw is the device sidewall resistivity, and P and A are the device area and perimeter.

Figure 4-10 shows the size dependence of R0A, along with the sw obtained through above

fitting for the three devices at 300 K. For the 200×200 m2 devices, the R0A values were

96
0.02 (1-stage), 0.11 (3-stage) and 0.18 .cm2 (5-stage). Hence, surface leakage contributed

to 74%, 62% and 48% of the total dark current for the 1-, 3- and 5-stage devices,

respectively. For devices with larger sizes, the surface leakage affects the dark current to a

lesser degree. However, the larger size device has a relatively low R0 (e.g. only 26  for

the 0.5×0.5 mm2 device from the 1-stage wafer at 300 K), which makes it difficult to

accurately extract the device QE. Hence, to optimize the tradeoff, the 0.2×0.2 mm2 devices

with comparatively high R0 in the three wafers were selected for device analysis.

14
1/R0A (−.cm-2)

70 12
10
60 8
1/R0A (−cm-2)

6
50 4
100 200 300
40 P/A (cm-1)
1-stage sw=5.1 cm
30
3-stage sw=34.5 cm
20 5-stage sw=76.9 cm

10
0
50 100 150 200 250 300 350
P/A (cm-1)
Figure 4-10: Size dependent R0A for the three devices at 300 K. The sidewall
resistivity was smallest for the one-stage device.

4.5 Summary and concluding remarks

In this chapter, rigorous experimental justifications of the advantage of multistage

ICTPV devices over conventional single-absorber devices are presented. This is done by a

comparative study of three narrow bandgap (~0.2 eV) TPV devices with a single-absorber

and multistage IC structures. It is shown that the performance of a single-absorber TPV

cell with T2SL absorbers is mainly limited by the small collection efficiency associated

with a relatively short diffusion length (1.5 m at 300 K). Instead, multistage IC structure

97
is proven to be capable of overcoming the diffusion length limitation and achieving a

collection efficiency of about 100% for photogenerated carriers. Consequently, the open-

circuit voltage, fill factor and conversion efficiency are greatly improved compared to the

single-absorber TPV structure. At current stage, although the demonstrated room-

temperature conversion efficiency (3.6%) is relatively low, there is still great room for

further improvement. Possible ways to improve the efficiency include increasing the total

absorber thickness, adding an anti-reflection coating onto the surface, attaching a back

reflector, as well as reducing the contact resistance. The fundamental limitation of a high

dark current in narrow bandgap absorbers can be overcome by applying an even stronger

optical illumination. This will increase the conversion efficiency since the  in multistage

ICTPV devices has not yet saturated as shown in Figure 4-6(d). Alternatively, these narrow

bandgap TPV devices can be cooled down to lower temperatures with substantially reduced

dark current density and increased power efficiency for applications such as in space (e.g.

Jupiter and Saturn missions) where the environment temperature is well below 300 K.

98
5 Chapter 5: Interband cascade thermophotovoltaic devices with more
stages
5.1 Background and motivation

Unlike the Esaki tunnel junctions routinely used in multijunction solar cells [13,

14], type-II broken-gap heterostructures are used to connect adjacent cascade stages in IC

structures so that the interband tunneling is smooth and the electrical resistances between

stages are negligible. As often implemented in ICLs, many stages (>20) can be

concatenated together without impacting carrier transport. Hence, for ICTPV cells, many

IC stages are desirable to maximize the absorption of incident light and produce a high

open-circuit voltage for optimizing power efficiency. However, in contrast to ICLs where

the light is generated inside the active cascade stages, each stage in an ICTPV cell sees a

different intensity due to the absorption in preceding stages. Consequently, to satisfy the

current match condition between cascade stages for optimized device operation, the

absorber thickness in the optically deeper stages is increased based on the absorption

coefficient. In practice, if there are many stages in an ICTPV cell, the deviation of exact

current match condition due to the variation of material parameters can be significant. Also,

ICTPV cells are relatively complex structures that are very vulnerable to the instable

growth conditions, thus the material quality may differ vastly from structure to structure.

In this chapter, the effects of current mismatch and material quality will be identified and

quantified in four ICTPV devices with different number of stages and absorber thickness.

In addition, in Chapter 4, the better device performance in the 5-stage compared to

the 1- and 3-stage devices implies that IC structure with more stages should be preferred.

This inference is also in accordance with other experimental data of ICTPV cells [157, 182,

186-188] and the theoretical projection in Chapter 3. Hence, another purpose of this chapter

99
is to examine this inference with ICTPV devices with many stages. Note that these ICTPV

cells were designed for achieving a better understanding of the underlying physics rather

than reaching optimized device performance. At current stage, the conversion efficiencies

of ICTPV cells do not reach respected levels, and they are not comparable with those

achieved from the TPV cells with relatively wide bandgaps [36-42], as shown in Table 5-

1. This is because the conversion efficiency of an ICTPV device is primarily limited by a

significantly high dark saturation current density J0 associated with the narrower bandgap

and a short carrier lifetime. In Chapter 4, it has been shown that although the IC structure

is able to overcome the limitations of a short diffusion length and low absorption

coefficient in conventional single-stage TPV cells, the issues of low fill factor and voltage

efficiency that result from the high J0, remain in narrow bandgap ICTPV cells even under

monochromatic illumination with high incident power density, as shown in Table 5-1.

Table 5-1: Summary of ICTPV devices that have been reported so far.

Temperature Illumination Pinc


Nc Eg (eV) Voc (V)  (%) Ref.
(K) Source (W/cm2)
Blackbody @
7 80 0.24 1.11 0.67 NA 100
1323 K
Blackbody @
7 80 0.31 1.68 0.23 4.3 189
1323 K
ICL emitting
7 300 0.24 0.65 19 2.1 190
@ 4.3 m
ICL emitting
3 300 0.23 0.18 7 NA 157
@ 4.3 m
ICL emitting
3 300 0.39 0.80 130 9.6 182
@ 2.81 m
ICL emitting
5 300 0.23 0.37 36 3.6 187
@ 4.2 m
ICL emitting
6 300 0.23 0.52 21 4.1 191
@ 4.2 m

100
5.2 Device structure, growth and fabrication

The four structures were grown using GENxplor MBE system on nominally

undoped p-type GaSb (001) substrates. The first two structures have six and seven stages

and were grown earlier. The other two structures have substantially increased stages

(sixteen and twenty-three) and were grown a year later after the system maintenance.

Hence, the growth conditions and material qualities can be somewhat different between the

two sets of structures. The absorbers in the four structures were made of InAs/GaSb T2SLs

and each period of the SL consist of four layers: InSb (1.2 Å), InAs (20.5 Å), InSb (1.2 Å)

and GaSb (25.1 Å). The purpose of including the two InSb layers is to balance the tensile

strain of the InAs layer [168]. The absorbers in the four structures were p-type doped to

2.6×1016 cm-3. The schematic layer diagram of the four structures are shown in Figure 5-1,

and the individual absorber thicknesses are presented in Table 5-2. As can be seen, the

individual absorbers in the 16- and 23-stage structures are much thinner than in the 6- and

7-stage devices. Conversely, the total absorber thicknesses in the 16- and 23-stage

structures are thicker compared to those in the 6- and 7-stage ones. The electron and hole

barriers in the four structures were identical to those in the three devices described in

Chapter 4. After the MBE growth, the wafers are processed into square mesa devices with

dimensions ranging from 50 to 1000 m by using standard contact UV photolithography

and wet-chemical etching. A RF-sputter deposited two-layer passivation (Si3N4 then SiO2)

is used for minimizing pin holes and improving overall stress management, and then the

Ti/Au layers are sputter deposited for top and bottom contacts. Finally, the devices were

mounted on heat sinks and wire bonded for characterization.

101
Figure 5-1: Schematic layer structure of the four TPV devices with six, seven, sixteen
and twenty-three stages.

Table 5-2: Individual and total absorber thicknesses for the four IC TPV structures.

Device Individual absorber thickness (nm) dtotal (m)

6-stage 360/403.2/456/523.2/619.2/758.4 3.12


7-stage 307.2/336/374.4/417.6/480/556.8/662.4 3.13
144/153.6/158.4/168/177.6/187.2/196.8/211.2/225.6/240/259.2/283.2
16-stage 3.90
/312/345.6/388.8/446.4
96/100.8/105.6/110.4/115.2/120/124.8/129.6/134.4/139.2/144/148.8/
23-stage 3.87
158.4/168/172.8/182.4/196.8/206.4/220.8/240/259.2/283.2/312

5.3 Energy conversion efficiency

The energy conversion efficiency  of the four TPV structures was investigated

under the illumination from an IC laser. The narrow emission spectrum of the IC laser is

analogous to a selective emitter that would be included in a TPV system to reduce the

thermalization and below-bandgap losses. During the experiment, the laser was cooled to

102
80 K and continuously emitted at a wavelength near 4.2 m (photon energy is 295 meV)

as shown in the inset within Figure 5-2(a). The output power of the laser can be controlled

by adjusting the injection current, thereby the performance of the four devices was

investigated under different incident power densities. Figure 5-2(a) shows the measured

illuminated J-V characteristics at 300 K for representative 200×200 m2 devices from the

four wafers. The incident power density of 17 W/cm2 was assessed through the connection

between quantum efficiency and short-circuit current density Jsc, as expressed by Equation

4-1. As can be seen in Figure 5-2(a), the short-circuit current density decreases with

number of stages Nc, primarily due to reduced optical absorption in individual stages with

thinner absorbers. Conversely, the open-circuit voltage increases with the number of

stages, since it is proportional to Nc when the individual stages are connected in series, as

stated by Equation 3-11. For example, at T=300 K and Pinc=17 W/cm2, the Jsc was 4.4, 3.2,

1.3 and 1.0 A/cm2, while the Voc was 350, 518, 910 and 1461 meV for the 6-, 7-, 16- and

23-stage devices, respectively. The trade-off of Jsc for Voc with increasing the number of

stages can in principle be beneficial for improving the conversion efficiency in many cases,

according to the previous experimental results [157, 182, 186-188]. However, such benefit

may not always be demonstrated, as will be discussed in the analysis of the characteristics

of the current four devices.

103
4 6-stage
4

Conversion efficiency (%)


6-stage ICL
emission 7-stage

Current density (A/cm2)


spectrum 3 16-stage
3 7-stage 23-stage
3.5 4.0 4.5 5.0
2 Wavelength (m)
2
16-stage (b)
(a) 1
1
23-stage

0 0
-0.5 0.0 0.5 1.0 1.5 0 5 10 15 20
Voltage (V) Pinc (W/cm2)

Figure 5-2: (a) Illuminated current density-voltage characteristics for the


representative 200×200 m2 devices from the four wafers at 300 K and at an incident
power density of 17 W/cm2. The inset shows the emission spectrum of the IC laser
used as the illumination source, (b) Conversion efficiency as a function of incident
power density for the four devices at 300 K.

Figure 5-2(b) shows the energy conversion efficiencies as a function of incident

power density for the four devices at 300 K. As shown, the four devices can be arranged

as 7-, 6-, 23- and 16-stage devices according to their  values, from best to worst. For

example, at the maximum incident power density (~21 W/cm2) available from the

illumination of the IC laser, the  is 3.5%, 4.1%, 2.7% and 3.3% for the 6-, 7-, 16- and 23-

stage devices, respectively. As theoretically illustrated in Chapter 3, the efficiency of

ICTPV cells should monotonically increase with the number of stages. This is because the

particle conversion efficiency (part), a more appropriate figure of merit for ICTPV devices,

is enhanced as the number of stages increases [169-170]. However, the results of the

current four devices indicate that the device performance in terms of  is better with fewer

cascade stages (6 and 7) than with more stages (16 and 23. This goes counter with the

theoretical forecasting in Chapter 3 and the previous experimental results [157, 182, 186-

188]. Nevertheless, the  was higher with more stages for devices grown in the same

104
campaign. For example, the device performance is better for the 7-stage compared to the

6-stage, and for the 23-stage compared to the 16-stage. Give that the four devices have

nominally identical absorber and barrier structures, what causes the different device

performances between the two sets? One possible factor is that the current mismatch is

more significant in the 16- and 23-stage devices compared to the 6- and 7-stage devices.

Another possible cause is that the 16- and 23-stage devices have poorer material quality

than the 6- and 7-stage devices, since the two sets of structures were grown in different

campaigns. In the following sections, the two possible factors will be inspected and

quantified through the analysis of detailed device characteristics such as dark current

density, carrier lifetime and quantum efficiency.

5.4 Device characterization and analysis

5.4.1 Dark current density and carrier lifetime

The dark current density-voltage (Jd-V) characteristics of the four devices were

measured using a Keithley 2636A source meter. During the measurement, the device was

put in a cryostat for temperature control between 78 to 340 K, and a top copper shield was

used to block background radiation from the environment. The measured dark current

densities at 300 K for the representative 200×200 m2 devices from the four wafers are

shown in Figure 5-3(a). As shown, the Jd decreases with number of stages due to the

reduced thermal generated carriers in thinner individual absorbers [141]. Also, the Jd in the

four devices is orders of magnitude higher than in conversional solar cells made of Si and

GaAs, which severely limits the device performance of these TPV cells. This is mainly due

to their narrow bandgaps that are 0.22-0.25 eV at 300 K as estimated from the 100% cutoff

105
wavelength of the quantum efficiency spectra [see Figure 5-4(a)]. From the measured dark

current, the carrier lifetime (), an important indicator of material quality, can be extracted.

As will be described in Chapter 6, a simple and effective method to extract carrier lifetime

is to apply a linear fit of the dark current density at large reverse bias and first obtain the

thermal generation rate gth. This approach is particularly useful for multistage IC devices

since their dark current densities usually have a large linear region under reverse bias, as

shown in Figure 5-3(b) for the four TPV devices. There is an explicit linear relationship

between current density and voltage at reverse bias starting from -2 V. The linear fittings

of current density with good accuracy from -4 to -2 V are indicated by the dashed lines in

Figure 5-3(b). Based on Equation 6-4, the thermal generation rate at 300 K acquired from

the intercept of the linear fitting is 6.6×1022, 4.3×1022, 9.6×1022 and 7.9×1022 cm-3s-1 for the

6-, 7-, 16- and 23-stage devices, respectively. The shunt resistance obtained from the slope

of the linear fitting is 10547 (6-stage), 21258 (7-stage), 29294 (16-stage) and 63459  (23-

stage).

0.0
102 6-stage
-0.2 23-stage
Dark current density (A/cm2)
Dark current density (A/cm2)

7-stage
101 16-stage -0.4 e
23-stage 16-stag
-0.6
100 ge
7-sta
-0.8 (b)
T=300 K
10-1 -1.0
e
(a) -1.2 tag
10-2 6-s solid: measurement
dashed: linear fitting
-1.4
10-3
-4 -3 -2 -1 0 1 2 -4 -3 -2 -1 0
Voltage (V) Votlage (V)

Figure 5-3: (a) Dark current density for the representative 200×200 m2 devices from
the four wafers at 300 K, (b) Linear fitting (dashed lines) of dark current density at
reverse voltage for the four devices at 300 K.

106
With the extracted gth, the minority carrier lifetime can be calculated according to

Equation 4-3. The carrier lifetime is determined by the comprehensive effect of the

radiative, Auger and SRH processes. Usually, based on experimental results in literature

[158, 192-193], Auger and SRH processes are dominant in InAs/GaSb T2SLs. The intrinsic

carrier concentration at 300 K is calculated to be 1.21×1016 (6-stage), 9.9×1015 (7-stage),

8.3×1015 (16-stage) and 7.6×1015 (23-stage) cm-3, according to Equation 4-4. Finally, based

on Equation 4-3 and the obtained gth values, the extracted carrier lifetime at 300 K is 86,

88, 28, 28 ns for the 6-, 7-, 16- and 23-stage devices, respectively. Compared to the 16-

and 23-stage devices, the longer carrier lifetime in the 6- and 7-stage devices suggests their

better material quality. This agrees with higher activation energies Ea (213 and 217 eV

between 200 and 340 K) for the 6- and 7-stage devices than that (204 and 207 eV) for the

16- and 23-stage devices. The activation energies were extracted from the temperature

dependence of the zero-bias resistance. These values of Ea are 50%-100% of the zero-

temperature bandgap values (~ 275-302 meV), which implies a non-negligible contribution

of the SRH process to the dark current. The variations of material quality and the

corresponding contributions to the SRH process among the four TPV wafers result in

different carrier lifetimes, which ultimately affects the TPV device performance that will

be quantified in Section 5-5.

5.4.2 Quantum efficiency and current mismatch

In the quantum efficiency measurement, a fourier transform infrared spectrometer

(FTIR) was used to measure the relative spectra response. The calibrated QE spectrum was

obtained by measuring the device’s photocurrent, while it was illuminated by chopped

radiation from a standard blackbody source (800 K). The measured QE spectra of the four

107
devices at 300 K are shown in Figure 5-4(a). The 100% cutoff wavelength where the QE

fast turns on is 5.6, 5.3, 5.1 and 5.0 m, corresponding to a bandgap of 221, 234, 243 and

248 meV at 300 K for the 6-, 7-, 16- and 23-stage devices, respectively. The bandgap

difference results from variations in MBE growth conditions, although the SL absorbers in

each device were designed to have identical compositions and period. The difference is

more outstanding between devices grown in different campaigns. As shown in Figure 5-

4(a), the QE decreases with number of stages due to the reduced optical absorption in in

thinner individual absorbers. For example, at T=300 K and =4.2 m, the QE is 6.46%,

5.41%, 2.31% and 1.57% for the 6-, 7-, 16- and 23-stage devices, respectively. Figure 5-

4(b) shows the measured bias dependence of the QE at 300 K and at the same wavelength.

As shown, the devices with more stages and thinner absorbers tend to have weaker bias

dependences of QE. The QEs of the 6-, 7- and 16-stage devices slightly increase with

reverse bias, while the QE of the 23-stage device is nearly a constant value. Specifically,

the QE changes from 6.46%, 5.41%, 2.31% and 1.57% to 7.03%, 5.66%, 2.37% and 1.58%,

while the reverse bias is increased from 0 mV to -700 meV for the 6-, 7-, 16- and 23-stage

devices, respectively. The moderate degree of bias dependence for the QEs is due to the

relatively thin individual absorbers compared to the conventional TPV structure with a

single thick absorber. This leads to the unique advantage of high collection efficiency of

photogenerated carriers, as described in Chapter 4.

108
7.5
12 6-stage 7.0
7-stage 6-stage

Quantum efficiency (%)


10 6.5 T=300 K

Quantum efficiency (%)


16-stage
6.0 =4.2 m
23-stage
8 5.5
(a) T=300 K 7-stage
zero bias 5.0
6 2.4
2.2 16-stage
4
2.0 (b)
2 1.8
23-stage
1.6
0 1.4
2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 0 100 200 300 400 500 600 700
Wavelength (m) Bias (mV)

Figure 5-4: (a) Quantum efficiency spectra of the four devices at 300 K and (b) Bias
dependence of quantum efficiency for the four devices at 300 K and at the wavelength
of 4.2 m.

In theory, provided that the absorption coefficient () and diffusion length (L) are

known, the effective QE in each stage of an IC device can be calculated from Equation 2-

5 in the diffusion limited case. At 300 K, the measured absorption coefficient at 4.2 m is

2984, 2643, 2334 and 2200 cm-1 for the 6-, 7-, 16- and 23-stage devices, respectively.

Based on Equation 2-5, the calculated effective QE at 4.2 m in each stage of the four

devices is shown in Figure 5-5(a). In the calculation, the diffusion length was assumed to

be 1.5 m for the 6- and 7-stage devices, while it was taken to be 0.7 m for the 16- and

23-stage devices. These values of L were adopted to achieve close agreement with the

experimental results. As can be seen, the calculated effective QEs of the 7-stage device are

nearly equal in each stage and are quite close to the measured device QE. Contrarily, the

calculated effective QEs of the 6-, 16- and 23-stage devices are mismatched between

stages. In this scenario, as will be described in Chapter 7, an electrical gain will be delivered

across the device to ensure current continuity and will enhance the device’s QE to the

average value over all stages [194]. On average, the effective QE is 6.65%, 2.42% and

1.63% for the 6-, 16- and 23-stage devices, respectively. These values are well matched

109
with the measured device QEs with an error less than 5%. This also indirectly verifies the

appropriateness of the values used for the diffusion lengths for these devices. Compared to

the 16- and 23-stage devices, the longer diffusion length for the 6- and 7-stage devices

agrees with their longer carrier lifetime. In addition, the mismatch of effective QE is less

significant in the 6-stage device than in the 16- and 23-stage devices. For example, the

minimum (maximum) of the effective QEs is 6.39% (6.90%), 2.20% (2.69%), 1.38%

(1.95%) in the 6-, 16- and 23-stage devices, corresponding to a mismatch of 8% (6-stage),

22% (16-stage), 41% (23-stage) in their QEs.

7.0

Incident power density (W/cm2)


6-stage 6-stage
6.5 20
7-stage 7-stage
6.0 (a) 16-stage
16-stage
Effective QE (%)

23-stage 15
5.5 23-stage
=4.2 m
5.0
10
2.5
(b)
2.0 5
1.5
1.0 0
0 5 10 15 20 0 50 100 150 200 250 300 350 400
Stage number IC laser current (mA)

Figure 5-5: (a) Calculated effective quantum efficiency based on Equation 2-5 in each
stage of the four devices, (b) Calculated incident power density vs IC laser current
based on Equation 4-1 for the four devices.

The direct result of current mismatch in these multistage devices is the reduction of

their photocurrents, which are decided by the stage with the minimum effective QE. From

Figure 5-5(a), the photocurrent was determined by the last stage in the 6-stage device, while

it was decided by the first stage in the 16- and 23-stage devices. This statement is tenable

when the photocurrent is dominant in the device under intense illumination from the IC

laser, which can be validated through the assessed incident power densities on the four

110
devices. A simple and effective method to assess Pinc is based on the relationship between

Jsc and QE (at laser emission wavelength), as expressed by Equation 4-1. Note that, the QE

in Equation 4-1 should be the minimum effective QE in the individual stages. According

to this equation, the calculated Pinc as a function of the IC laser current is shown in Figure

5-5(b). As can be seen, the calculated values of Pinc onto the four devices are close to each

other. This is anticipated since they were illuminated by the same IC laser, even with some

possible experimental uncertainties due to alignment. The good consistency of the Pinc also

validates the above-mentioned statement that there was no electrical gain in the four

devices when they were illuminated by the IC laser. Based on this commonality, the effect

of current (or effective quantum efficiency) mismatch between stages on device

performance will be quantified in Subsection 5.5.2.

5.4.3 Collection efficiency of photogenerated carriers

As described in Chapter 4, a special feature of some ICTPV devices is that the

photocurrent is voltage dependent. But this feature is likely to be less notable for devices

with thinner individual absorbers and more stages. It would be interesting to examine this

feature in the 16- and 23-stage devices which have even more stages and thinner individual

absorbers. This can be done by comparison between the 100% collected and the measured

J-V curves. At T=300 K and Pinc=17 W/cm2, the 100% collected and the measured J-V

curves for the four devices are shown in Figure 5-6(a). As mentioned in Chapter 4, the

100% collected J-V curve refers to the ideal case where the photogenerated carriers are

completely collected. It can be plotted in the same manner as in [171] and is the

superposition of the dark current density and the maximum photocurrent density Jphmax.

The magnitude of Jphmax is the difference between the saturated current densities under dark

111
and illuminated conditions. As can be seen in Figure 5-6(a), there are noticeable shifts

between the ideal and the measured J-V curves for the 6- and 7-stage devices. This means

that the photocurrents (or the collection efficiencies) in the two devices are voltage

dependent. In contrast, for the 16- and 23-stage devices, the ideal and the measured J-V

curves almost overlap with each other. This indicates that the collection efficiencies in the

two devices are close to unity as well as being voltage independent.

The collection efficiency c in the four devices can be extracted based on Equation

4-2 whose validity relies on two assumptions, as mentioned in Subsection 4.4.1. For each

of the four devices, the J-V data at four different illumination levels were chosen for

subtraction to make a fair comparison. It was found that, although not presented here, the

extracted c using different J-V data pairs overlap each other. This verifies the assumption

that the dark current density and collection efficiency remain almost unchanged at under

different illumination levels. In particular, the extracted c based on Equation 4-2 using J-

V data at incident power densities of 7 and 17 W/cm2 is shown in Figure 5-6(b). As can be

seen, the c in the 6- and 7-stage devices decreases dramatically with forward voltage. In

contrast, the c is always close to unity in the 16- and 23-stage devices throughout the

forward voltage range of interest. At this moment, this difference of c between ths two

sets of devices is not fully understood. Presumably, one factor is that the photocurrent in

the 16- and 23-stage devices is determined by the first stage [Figure 5-5(a)] with an

absorber that is much thinner than the one in the last stage of the 6- and 7-stage devices.

This factor along with more stages (to share forward voltage) could contribute to the nearly

100% collection efficiency in the 16- and 23-stage devices. This phenomenon may need

further investigation in the future.

112
1.0
4 6-stage
solid: measured

Current density (A/cm2)


0.8

Collection efficiency
dahsed: 100% collected
3 7-stage
0.6 6-stage
7-stage
2 T=300 K 16-stage
(a) 0.4
16-stage 23-stage
1 0.2 (b)
23-stage

0 0.0
-0.5 0.0 0.5 1.0 1.5 -0.5 0.0 0.5 1.0
Voltage (V) Voltage (V)

Figure 5-6: (a) The measured and the 100% collected J-V curves for the four devices
at 300 K and at the incident power density of 17 W/cm2, (b) Extracted collection
efficiency at 300 K based on Equation 4-2 using J-V data under incident power
densities of 7 and 17 W/cm2 for the four devices.

5.5 Quantification of the effects of the performance limiting factors

5.5.1 Effect of collection efficiency

In the preceding section, the 16- and 23-stage devices are identified to have poor

material quality and more severe current mismatch that that the 6- and 7-stage devices. On

the other hand, the collection efficiency was higher in the 16- and 23-stage devices

compared to the 6- and 7-stage devices. Table 5-3 summaries the three factors and

characteristics, and some important performance-related parameters at 300 K. In this

section, the effects of the three performance limiting factors will be quantified.

Table 5-3: Summary of device characteristics and some important performance-


related parameters for the four devices at 300 K.
Current c-Voltage
Eg (meV) ni (cm-3) gth (cm-3s-1) Rshunt ()  (ns) L (m) 
mismatch dependence
6-stage 221 1.21×1016 6.6×1022 10547 86 1.5 3.5% mild substantial

7-stage 234 9.9×1015 4.3×1022 21258 88 1.5 4.1% none substantial

16-stage 243 8.3×1015 9.6×1023 29294 28 0.7 2.7% severe insensitive

23-stage 248 7.6×1015 7.9×1023 63459 28 0.7 3.3% severe insensitive

113
Among the three factors, the effect of voltage dependent collection efficiency is

simplest to quantify. This can be done through a comparison between the measured  and

the ideally collected case, as shown in Figure 5-7. As shown, at the maximum incident

power density, the  was 4.4% and 4.6% in the ideal case for the 6- and 7-stage devices,

respectively. This corresponds to a 0.9% (6-stage) and 0.5% (7-stage) increase relative to

the actual measured values. The more significant increase for the 6-stage device is due to

the lower collection efficiency than in the 7-stage device. Also, the increase of  was less

appreciable at the lower incident power density. This occurs because the operating voltage

at the maximum output power was smaller at the lower incident power density. From

Figure 5-6(b), the collection efficiency at the operating voltage is higher for the lower

incident power density. For example, at Pinc=17 W/cm2, the  was increased from the

measured 3.1% and 3.7% to the ideal 3.8% and 4.1%, corresponding to a 0.7% and 0.4%

increase for the 6- and 7-stage devices, respectively.

5
6-stage
4 7-stage

3
 (%)

1 solid symbol: measured


open symbol: 100% collected

0
0 5 10 15 20
Pinc (W/cm2)

Figure 5-7: Comparison of the measured  and the ideal  in the 100% collected case
at 300 K for the 6- and 7-stage devices.

114
5.5.2 Effect of current mismatch

As for the current mismatch, it is commonly recognized as a serious problem in PV

arrays. For example, it can even cause localized heating of the cell and possible cell

damage, which is known as hot-spot heating [195-196]. By comparison, although current

mismatch between stages is significant in the 16- and 23-stage devices, but far from being

able to cause any substantial damage or heating issues in a single stage when under intense

illumination. The direct negative impact of current mismatch in ICTPV devices is the

reduction of photocurrent. According to Equation 2-5, current mismatch in an IC structure

can result from the deviation of either the absorption coefficient or diffusion length from

the original reference values that were used to design current-matched absorbers.

Comparatively, the deviation of  is more prone to occur in practice and has a greater

impact on the calculated effective QE. Hence, here only the deviation of  will be

considered. In addition, the voltage dependence of collection efficiency in the 6- and 7-

stage devices should not be ignored. In this regard, the effect of current mismatch can be

quantified by decoupling the photocurrent and dark current densities. Proceeding in this

way, the illuminated J-V relation can be expressed as:

𝐽 (𝑉) = 𝐽𝑠𝑐 𝜂𝑐 ⁄𝜂𝑐 (0) − 𝐽𝑑 (𝑉) = 𝑃𝑖𝑛𝑐 𝑄𝐸 𝜂𝑐 ⁄𝜂𝑐 (0) − 𝐽𝑑 (𝑉) (5-1)

where c is shown in Figure 5-6(b) and c (0) is the collection efficiency at zero voltage.

As previously emphasized, the QE in Equation 5-1 should be the minimum effective QE

in individual stages and can be calculated from Equation 2-5. For direct connection to

actual devices, the Jd (V) in Equation 5-1 was replaced by the experimental data for the

four devices. With these specifications, the effect of the deviation of  and consequential

current mismatch will only be embodied in Jsc and QE in Equation 5-1.

115
The calculated Jsc and  as functions of  based on Equation 5-1 are shown in

Figure 5-8. In the calculation, the incident power density was taken to be 17 W/cm2, and

the diffusion length was assumed to be 1.5 m for the 6- and 7-stage devices and 0.7 m

for the 16- and 23-stage devices. The kinks in the calculated Jsc and  curves correspond

to the condition where the effective QE is perfectly matched between stages. This occurs

at an  of 2687 (6-stage), 2721 (7-stage), 2849 (16-stage) and 3061 cm-1 (23-stage). As can

be seen in Figure 5-8, the Jsc and  of the 16- and 23-stage devices peak at the current-

matched condition, while the Jsc and  in the 6- and 7-stage devices slightly increase when

 passes the current-matched condition with further increases. This is because the total

absorbers in the 6- and 7-stage devices are relatively thin so that the higher absorption

coefficient will increase absorption of photons and enhance the photocurrent. In contrast,

the total absorbers of the 16-and 23-stage devices are much thicker than the 6- and 7-stage

devices, so the light attenuation (and thus the current-mismatch) is more dominant in the

optically deeper stages. The circles in Figure 5-8 represent the calculated Jsc and  with the

measured . As can be seen, the 16- and 23-stage devices depart far more from the current-

matched condition than the 6- and 7-stage devices. At the current-matched condition, the

calculated  is 3.0%, 3.9%, 3.4% and 4.5% for the 6-, 7-, 16- and 23-stage devices,

respectively. This corresponds to a difference of 0.1% (6-stage), 0.2% (7-stage), 1.0% (16-

stage) and 1.7% (23-stage) compared to the actual measured . The impact of current

mismatch is comparable at different incident power densities. For example, at Pinc=21

W/cm2, the calculated  at the current-matched condition is 3.43%, 4.29%, 3.85%, 5.08%,

corresponding to a difference of 0.11%, 0.21%, 1.1%, 1.84% compared to the actual

obtained  for the 6-, 7-, 16- and 23-stage devices, respectively.

116
 (cm-1)
2000 2500 3000 3500 4000

6-stage

Jsc (A/cm2)
3
7-stage
16-stage
23-stage
1.5

1.0 (a)
4.5
Pinc=17 W/cm2
4.0

3.5
 (%)

3.0
6-stage
2.5 7-stage
16-stage
2.0 23-stage (b)

2000 2500 3000 3500 4000


 (cm-1)
Figure 5-8: Calculated (a) short-circuit current density and (b) conversion efficiency
based on Equation 5-1 as a function of absorption coefficient at incident power density
of 17 W/cm2 for the four devices.

5.5.3 Effect of material quality

Lastly, regarding the effect of material quality, it can be quantified through the variation of

carrier lifetime , an important parameter for material quality. The variation of  brings

corresponding variations of thermal generation and dark saturation current density [197],

which can significantly affect the fill factor and open-circuit voltage [197], consequently

making a substantial impact on conversion efficiency. The effect of material quality can be

evaluated based on a diffusion limited model as described in detail in Chapter 3. In this

model, the J-V characteristic of the device is given by Equation 3-10. Based on Equation

3-10, the calculated conversion efficiency as well as the measurement are shown in Figure

117
5-9. In the calculation, the carrier lifetime was assumed to be 28 and 87 ns, close to the

extracted values shown in Table 5-3. As can be seen, the calculated  using the extracted

carrier lifetime was higher than the measured value for all the four devices. This is mainly

because the extracted lifetime was somewhat overestimated due to the occurrence of the

SRH process. For the 6- and 7-stage devices, this is also due to the voltage dependence of

collection efficiency that was instead ignored in the calculation. Nevertheless, the

calculations based on Equation 2-5 evidently indicate the considerable impact of carrier

lifetime on device performance. As shown in Figure 5-9, there is a distinct gap between

the calculated conversion efficiencies with different values of carrier lifetime. For example,

for =28 ns and Pinc=17 W/cm2, the calculated  was 2.2% (6-stage), 2.6% (7-stage), 3.1%

(16-stage) and 3.3% (23-stage). However, as the carrier lifetime increased to 87 ns, the

calculated  at the same Pinc was 4.1%, 4.6%, 5.3% and 5.4% for the 6-, 7-, 16- and 23-

stage devices, respectively. This corresponds to an efficiency increase of 1.9% (6-stage),

2.0% (7-stage), 2.2% (16-stage) and 2.1% (23-stage). Clearly, this increase is much more

significant than those due to the eliminations of voltage-dependent collection efficiency

and current mismatch. Therefore, the material quality plays the most important role among

the three factors. If carrier lifetime is kept the same, the  is higher in the 16- and 23-stage

devices than in the 6- and 7-stage devices, even though the current mismatch is more

significant in the 16- and 23-stage devices. In this respect, given comparable material

quality, ICTPV devices with more stages and thinner absorbers are advantageous,

consistent with previous experimental results [157, 182, 186-188]. When the current

mismatch is minimized, ICTPV devices will have further conversion efficiency with more

stages.

118
6 6
measured measured
5 calculated-28 ns calculated-28 ns 5
calculated-87 ns calculated-87 ns
4 4

 (%)
 (%)
3 3

2 2

1 6-stage 7-stage 1

6 6
measured measured
calculated-28 ns calculated-28 ns
5 5
calculated-87 ns calculated-87 ns
4 4
 (%)

 (%)
3 3

2 2

1 16-stage 23-stage 1

0 0
0 5 10 15 20 0 5 10 15 20
2 2
Pinc (W/cm ) Pinc (W/cm )

Figure 5-9: Calculated conversion efficiency based on Equation 3-10, along with
measurement for the four devices. For each of the four devices, the carrier lifetime
used in the calculation was 27 and 87 ns.

5.6 Summary and concluding remarks

This chapter deals with detailed characterization and performance analysis of two

sets of four narrow bandgap (~0.22-0.25 eV at 300 K) ICTPV devices. The four ICTPV

devices have increased number of stages compared to the three devices in Chapter 4. With

different numbers of stages and individual absorber thicknesses, it was shown that current

mismatch between stages could be significant with more stages due to the variation of

absorption coefficient. On the other hand, the collection efficiency of photogenerated

carriers can be much improved with thinner individual absorbers and more stages. Also,

the carrier lifetime was extracted from dark current density to evaluate the material quality.

119
The extracted shorter carrier lifetime, together with substantial current mismatch, explains

the lower conversion efficiencies in the 16- and 23-stage devices compared to that in the

6- and 7-stage devices. Furthermore, the effects of material quality, current mismatch and

collection efficiency on device performance are quantified. The quantitative analysis shows

that the material quality has the most significant impact on the device performance among

the three factors. This indicates the importance of good material quality and its consistency

for realizing efficient IC TPV devices. This conclusion also challenges the inference put

forward in Section 5.1 as more cascade stages may not succeed to improve device

performance if the material quality is poor.

120
6 Chapter 6: Carrier lifetime in mid wavelength interband cascade
devices
6.1 Introduction

Starting from this chapter, experimental studies of IC structure for infrared detector

will be presented. The operation principle and theoretical background of IC infrared

photodetectors (ICIPs) are reviewed in Chapter 2. Specifically, the noise reduction and

detectivity enhancement in multistage detectors compared to single-absorber detectors are

detailed in Chapter 2. These advantages enable ICIPs to operate at high temperatures with

decent detectivity, as has been manifested in experiment [99, 137, 151, 199]. Nevertheless,

at the current stage, ICIPs does not outperform the state-of-art HgCdTe detectors in the

MWIR regime. For example, at 300 K, the detectivity of an ICIP with a cutoff wavelength

of 4.3 m is close to 1×109 Jones [151], slightly lower than the claimed ≥ 3.0×109 Jones

for an uncooled photovoltaic HgCdTe detector with similar cutoff wavelength (~4 m)

[91]. This is partially because the carrier lifetime in InAs/GaSb T2SLs is lower than in the

HgCdTe materials, although the Auger reduction is theoretically projected to be suppressed

in T2SLs [128-130]. For example, the reported lifetimes are 30-100 ns in MWIR T2SLs

[200-203], and 10-55 ns for LWIR T2SLs [192-193, 201, 204], which are mainly limited

by SRH recombination. Speculatively, the origin of the recombination centers is ascribed

to the presence of gallium, as the gallium-free InAs/InAsSb SLs possess much longer

radiative-dominated lifetimes (e.g. >400 ns or 9 s at 77 K) [166-167]. Because of the

shorter lifetime, the dark current densities in InAs/Ga(In)Sb T2SLs detectors are generally

higher than the benchmark known as “Rule 07” [205] for MCT materials.

In this chapter, a simple and effective electrical method is developed to the extract

121
carrier lifetime in InAs/GaSb T2SLs. This method differs from the frequently used optical

methods based on time- or frequency-domain photoluminescence (PL) measurements [192,

200-201, 204]. These optical methods are mainly focused on low temperatures (<200 K),

while the developed method can extract lifetime in a wide range of temperature, especially

at high temperatures (e.g. 300 K and above). There have been a few studies on carrier

lifetime using different approaches, such as measuring photoconductive response and

modeling dark current characteristics of T2SL detectors [206-209]. However, as with the

optical methods, these approaches fail to work at high temperatures. Sometimes, a more

meaningful carrier lifetime that is different from the recombination lifetime needs to be

realized and extracted. For example, for a photodiode that is operated under a reverse bias,

the generation lifetime is more relevant to the device performance and could be far longer

than the recombination lifetime, depending on the defect energy level as discussed for Si-

based devices [210]. In practical devices, carrier lifetime is often a mixture of various

mechanisms (See Figure 6-1), which are challenging to separate.

Figure 6-1: Radiative and non-radiative recombination processes in semiconductors.

The carrier lifetime in IC devices (QCDs) is lower than MCT materials. However,

compared to the other cascade device family ─ quantum cascade (QC) devices, it can be

much longer. QC devices (QCDs) operate based on intersubband transitions within the

122
same band (e.g. the conduction band). This contrasts to IC devices (ICDs) that are based

on the interband transitions between the conduction and valence band. This fundamental

difference in carrier transport results in distinct carrier lifetimes and device performances,

especially at high temperatures. For example, the lifetime in QCDs is in the picosecond

range due to fast phonon scattering, while ICDs have a nanosecond lifetime scale due to

Auger and SRH recombination. Like IC devices, QC family include QC lasers and QC

detectors. Although QC structures were also proposed and simulated theoretically for PV

cells [211-212], none have been reported experimentally. The two families of devices are

both based on quantum-engineered layer structures, and they nearly went through a parallel

rapid evolution, especially in lasers [70-73]. However, they were often discussed and

presented separately but seldom compared with their counterparts. There is particularly no

evaluation or comparison based on a unified figure of merit to fairly describe their

characteristics with different device functionalities. In this chapter, the saturation current

density J0 is identified as the common figure of merit. A semi-empirical model is employed

to extract the J0 from many QCDs and ICDs published in literature and some of

unpublished ICDs.

6.2 Carrier lifetime in mid-wavelength ICIPs

6.2.1 Device structure, growth and fabrication

The seven devices presented in this section have ICIP structures with different numbers

of stages (Nc) and absorber thicknesses. They were grown using a GENxplor MBE system

on nominally undoped p-type GaSb (001) substrates. Table 6-1 presents the individual

absorber thicknesses of the seven ICIPs in order from the surface to substrate. For

123
convenience, they are denoted as 1S-1040, 1S-2340, M3S-312, M6S-312, N8S-312,

M12S-156 and N16S-156, where “M” and “N” stand for current-matched and noncurrent-

matched configurations, respectively. As mentioned in Chapter 2, the individual absorbers

in current-matched ICIPs are designed thicker in the optically deeper stages to ensure equal

photocurrent in each stage. In contrast, the individual absorber thicknesses in a non-

current-matched ICIP are made identical. For the current-matched ICIPs studied in the

section, the individual absorber thicknesses were designed based on the absorption

coefficient of 3000 cm-1 and the assumption of complete collection of photogenerated

carriers. 1S-1040, 1S-2340, M3S-312, M6S-312 and N8S-312 were grown earlier as

descried in [151], while M12S-156 and N16S-156 were grown in a later growth campaign

(just after system maintenance) with possibly varied conditions and material qualities. The

seven detectors have identical electron and hole barriers as described in [151]. The

absorbers in the seven detectors consist of InAs/GaSb/Al(In)Sb/GaSb M-shape SLs [104,

134-135] with layer thicknesses of 27/15/815 Å, respectively. The GaSb layers in the SLs

were p-type doped to 5.1×1016 cm-3 for all the seven detector structures. The average

doping concentration in the SLs is estimated to be 2.4×1016 cm-3 according to the ratio of

the GaSb thickness over the SL period. Upon this doping level, the carrier transport in the

absorbers is expected to be determined by the dynamics of minority electrons. The bandgap

of the absorbers was designed with a cutoff wavelength (c) near 4.3 µm at 300 K, which

closely matched the observed 100% cutoff wavelengths for devices made from the seven

wafers, implying good control of layer thicknesses and alloy compositions during MBE

growth.

124
The important design and material parameters such as surface defect density and

perpendicular (⊥) lattice mismatch of the seven wafers are summarized in Table 6-1. After

the MBE growth, the wafers were processed into square mesa devices with dimensions

from 50 to 1000 m using standard contact UV photolithography followed by wet-

chemical etching. A two-layer passivation (Si3N4 then SiO2) was RF sputter deposited to

improve overall stress management and minimize pin holes. Sputter deposited Ti/Au layers

provided top and bottom contacts. Finally, the devices were mounted on heat sinks and

wire bonded for characterization.

Table 6-1: Summary of the design and material parameters of the seven wafers.

Absorber thickness # of dtotal Doping c Defect Lattice


Wafer
(nm) stage (m) (cm3) (m) (cm-2) mismatch

1S-1040 1040 1 1.04 2.4×1016 4.4 6.0×103 -0.09%

1S-2340 2340 1 2.34 2.4×1016 4.4 5.0×104 ~0

M3S-312 312/344.5/383.5 3 1.04 2.4×1016 4.3 2.0×104 -0.027%


312/344.5/383.5/435
M6S-312 6 2.59 2.4×1016 4.3 2.0×104 -0.10%
.5/507/604.5
N8S-312 312×8 8 2.50 2.4×1016 4.3 5.0×104 -0.08%
156/169/182/195/20
M12S-156 8/227.5/247/273/299 12 3.08 2.4×1016 4.3 4.4×104 0.051%
/331.5/370.5/422.5
N16S-156 156×16 16 2.49 2.4×1016 4.3 4.3×104 0.065%

6.2.2 Dark current density

The dark current density-voltage (Jd-V) characteristics of the seven ICIPs were

measured at various temperatures. Figure 6-2 (a) and (b) shows the measured Jd at 250 and

300 K for the representative 400×400 (1S-1040 and 1S-2340) and 500×500 (other five

wafers) m2 devices made from the seven wafers. As shown, at reverse voltage, the seven

125
devices in the ascending order of Jd are N16S-156, M12S-156, N8S-312, M6S-312, M3S-

312, 1S-1040 and 1S-2340. This sequence is precisely in the descending order of number

of stages or increasing order of absorber thickness. This is because ICIPs with more stages

and thinner individual absorbers are better able to suppress the dark current. Given carrier

transport is diffusion limited, according to Equation 2-6 and 2-7, the dark current density

in the mth stage of an ICIP can be written as:

𝐽𝑑,𝑚 = 𝑒𝑔𝑡ℎ 𝐿𝑡𝑎𝑛ℎ(𝑑𝑚 ⁄𝐿)(𝑒 𝑒𝑉𝑚 ⁄𝑘𝑏𝑇 − 1) (6-1)

where Vm is the applied voltage across the mth stage and dm is the individual absorber

thickness of the mth stage. Here, the parasitic series resistance Rs and shunt resistance Rshunt

are ignored. The voltage drop across each stage equates V/Nc in a noncurrent-matched ICIP

with identical absorbers. However, in a current-matched ICIP, to ensure dark current

continuity, the Vm will be smaller in an optically deeper stage with a thicker individual

absorber. Based on Equation 6-1, given a similar cutoff wavelength and minority carrier

lifetime, the dark current density at the same voltage will be lower in ICIPs with more

stages and thinner individual absorbers. This essentially agrees with the measured Jd-V

characteristics of the seven ICIPs as shown in Figure 6-2 (a) and (b).

However, at large reverse voltage where all the carriers are swept out from the

absorbers [213-214], a more appropriate equation for dark current density is given by:

𝑉−𝐽𝑑 𝑅𝑠 𝐴
𝐽𝑑 = 𝑒𝑔𝑡ℎ 𝑑1 + (6-2)
𝑅𝑠ℎ𝑢𝑛𝑡𝐴

Because the cascade stages are connected in series, the dark current density is decided by

the stage with the thinnest individual absorber (i.e. the first stage). The second term on the

right side of Equation 6-2 represents the average leakage current density with a constant

shunt resistance. Hence, from Equation 6-2, there is a liner relationship between current

126
density and voltage at large reverse voltage, which forms the important basis to extract

thermal generation rate and carrier lifetime.

101 101
M3S-312

Dark current density (A/cm2)


Dark current density (A/cm2)

M6S-312 1S-2340
0 N8S-312 0
1S-1040
10 10
1S-1 1S- M12S-156
040 234
0 N16S-156
10-1 10-1
M3S-312
-2 -2 M6S-312
10 10 N8S-312
T=250 K T=300 K M12S-156
-3 (a) -3 (b) N16S-156
10 10
-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5
Voltage (V) Voltage (V)

Figure 6-2: Dark current density versus applied voltage for the seven devices at (a)
250 K and (b) 300 K.

6.2.3 Contribution of SRH process to dark current

When carrier transport is affected by the SRH process, the description based on

Equation 6-1 is prone to errors. In contrast, Equation 6-2 can account for combined effects

of various mechanisms. Inclusion of Auger and SRH mechanisms gives a thermal

generation rate that is expressed as [213-214]:

𝑛𝑖2 𝑛𝑖2 𝑛𝑖2


𝑔𝑡ℎ = + (𝑁 = (6-3 a)
𝑁𝑎 𝜏𝐴 𝑎 +𝑛𝑖 )𝜏𝑆𝑅𝐻 𝑁𝑎 𝜏

1 1 𝑁𝑎 1 1 1
= + (𝑁 ≈ + (6-3 b)
𝜏 𝜏𝐴 𝑎 +𝑛𝑖 ) 𝜏𝑆𝑅𝐻 𝜏𝐴 𝜏𝑆𝑅𝐻

where A represents the carrier lifetime due to the Auger mechanism and SRH is the SRH

carrier lifetime. The approximation made in Equation 6-3(b) is valid if the doping

concentration Na is much higher than the intrinsic carrier concentration ni. Therefore, one

can first extract gth from Equation 6-2 and then calculate the carrier lifetime  from

127
Equation 6-3, which covers various transport mechanisms and is more accurate than

Equation 6-1.

The contribution of SRH process in the seven devices can be indirectly assessed by

the activation energy Ea. Based on the temperature dependence of R0A [See Figure 6-3(a)],

the Ea in a temperature range of 200-340 K was 256, 252, 258, 254, 249, 256, 253 meV for

1S-1040, 1S-2340, M3S-312, M6S-312, N8S-312, M12S-156 and N16S-156, respectively.

These values of Ea were smaller than their bandgaps (~288 meV) at room temperature and

about 75% of the zero-temperature bandgap Eg (0) (~329 meV). This means that the SRH

processes were involved in the carrier transport besides the diffusion process [215]. The

bandgaps of the seven devices were estimated from the 100% cutoff wavelengths in their

responsivity spectra, which were very close at every temperature of interest and two were

presented in [151]. Particularly, the temperature dependence of the bandgap for M3S-312,

as well as the Varshni fitting, are presented in Figure 6-3(b).

T (K)
340320 300 280 260 240 220 200
103 330
(a)
2 320 (b)
10
310
R0A (.cm2)

101
Eg (meV)

1S-1040 300
100 1S-2340 Vashni parameters
M3S-312 290
-1 M6S-312 Eg (T=0)=329 meV
10
N8S-312
280 =0.364 meV/K
M12S-156 =241 K
10-2 N16S-156
270
3.0 3.5 4.0 4.5 5.0 100 150 200 250 300 350
1000/T (K-1) T (K)
Figure 6-3: (a) R0A of the seven devices in the temperature range of 200-340 K. (b)
Temperature dependence of bandgap for M3S-312. The fitting Varshni parameters
for the device are shown.

128
6.2.4 Linear fitting of dark current density

Since the carrier transport is affected by the SRH process, Equation 6-2 is

preferably used to extract gth for the seven devices at higher temperatures, which is simpler

with assuming constant parasitic resistances. The feasibility and validity of constant

parasitic resistances are supported by the observed linear relationship of current density

with voltage as shown in Figure 6-4. There are obvious linear regions of Jd for the five

multi-stage devices at large reverse bias starting from -1.5 V. This behavior was also

observed at other higher temperatures for the five devices. For the two single-stage devices,

their current density-voltage curves exhibited linear characteristics between about -1.5 and

-0.3 V as well. However, the current density increased sharply with reverse bias voltage

after -1.5 V, which was likely triggered by a substantial electric field in the absorber region

and the consequential tunneling of carriers through the bandgap. This is because the entire

voltage is applied exclusively on the single stage, while the multistage ICIPs have multiple

unipolar barriers to share and withstand the voltage. In this sense, this method of extracting

the thermal generation rate is particularly well suited for multistage ICIPs where the

linearity can be ensured in a wide range of reverse voltage.

Based on Equation 6-2, the thermal generation rate for the seven devices can be

extracted by linearly fitting the dark current density at larger reverse voltages with the

rearranged equation:

𝑉 𝑅𝑠
𝐽𝑑 = (−𝑒𝑔𝑡ℎ 𝑑1 + )⁄(1 + ) (6-4)
𝑅𝑠ℎ𝑢𝑛𝑡𝐴 𝑅𝑠ℎ𝑢𝑛𝑡

The lines that were linearly fit to the experimental data for the seven devices at 300 K are

shown in Figure 6-4. The linear fits were performed between -3 and -1.5 V for the five

multistage devices for good accuracy. By comparison, the linear fits for the two single-

129
stage devices were done in the voltage range of -1.5 to -0.3 V to circumvent the effect of

substantial tunneling of carriers as previously mentioned.

0.0
156
16S-
T=300 K N
-0.1

-0.2 M12S-156

Jd (A/cm2) -0.3
-312 0.0
M6S
-0.4 -0.5 40
-312 1S-10

Jd (A/cm2)
N8S -1.0
340
-0.5 12
-1.5 1S-2
S-3
M3 -2.0
-2.5
-0.6 -2.0 -1.5 -1.0 -0.5 0.0
Voltage (V)

-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0


Voltage (V)

Figure 6-4: Linear fitting (dashed) and experimental measurements (solid) of the
dark current density at reverse bias voltage for the five multistage devices at 300 K.
The inset shows the corresponding results of the two single-stage devices at 300 K.

6.2.5 Estimated thermal generation rate and carrier lifetime

Based on Equation 6-4, the thermal generation rate at 300 K found from the

intercept of the fitted line with the vertical axis was 3.1×1022 (1S-1040), 3.2×1022 (1S-

2340), 3.2×1022 (M3S-312), 3.6×1022 (M6S-312) and 4.0×1022 cm-3/s (N8S-312, M12S-

156 and N16S-156). Simultaneously, the shunt resistance obtained from slope of a fitted

curve was 1449, 1192, 3338, 5639, 5054, 13343 and 13361  for 1S-1040, 1S-2340, M3S-

312, M6S-312, N8S-312, M12S-156 and N16S-156, respectively. The Rs was extracted

from the differential resistance at large forward voltage, and was less than 10  at 300 K.

Since Rs was at least two orders of magnitude smaller than Rshunt, the term Rs/Rshunt in

Equation 6-4 can be ignored when extracting the thermal generation rate in the seven

devices.

130
With the extracted gth, the minority carrier lifetime can then be calculated from

Equation 6-3 in which the intrinsic carrier concentration is given by Equation 4-4. At 300

K, the calculated electron and hole effective masses of the T2SLs using a two-band k·p

model were 0.049m0 and 0.48m0, respectively. Note that the electron effective mass scales

linearly with the temperature-dependent bandgap according to Kane’s model [216]. At 300

K, the calculated intrinsic carrier concentration was 4.1×1015 cm-3, one order of magnitude

lower than the doping concentration in the absorbers. At a lower temperature (e.g. 200 K),

the calculated ni was 7.8×1013 cm-3 and the Fermi energy EF was 4.5 kbT higher than the

valence band edge Ev, implying that Equation 4-4 was still valid for the seven devices.

Hence, the extraction of carrier lifetime was carried out at 200-340 K.

Based on Equation 4-4 and 6-3, along with the extracted thermal generation rate,

the minority carrier lifetime at 300 K was estimated to be 22.9 (1S-1040), 22.1 (1S-2340),

22.3 (M3S-312) 19.7 (M6S-312), 17.8 (N8S-312), 17.8 (M12S-156) and 17.8 ns (N16S-

156). In the same manner with 300 K, the carrier lifetimes and thermal generation rates at

other higher temperatures were also obtaine as shown in Figure 6-5. For the seven devices

at 200-340 K, the extracted  was ranges between 167 and 8.5 ns depending on the material

quality, and monotonically decreased with increasing temperature. For example, compared

to M12S-312 and N16S-312 (which were grown later in a different growth campaign), the

longer  in M3S- is due to the better material and crystal structure quality (Table 6-1). For

the same reason, the lifetime of 1S-1040 was longer than that of 1S-2340. Also, the

extracted carrier lifetimes were similar between the two single-stage devices and the five

multistage devices. This supports the validity and feasibility of the developed method for

non-cascade photodetectors. At temperatures higher than 200 K, the extracted carrier

131
lifetime was somehow shorter than the values (of 135-108 ns between 200 and 300 K) that

were stated in [217] for T2SL ICIPs (with a cutoff wavelength near 5 m) based on the

fitting of the Jd-V curve to an equation similar to Equation 6-1. Also, the carrier lifetimes

exhibited a rapid decrease with increasing temperature, which was close to an exponential

relationship especially in the temperature range of 250-340 K. For instance, for M6S-312,

the carrier lifetime decreased from 132 ns to 10.6 ns while the temperature was increased

from 200 to 340 K. This dependence of lifetime on temperature was quite different from

the previous results obtained by optical and other electrical methods [192-193, 208], which

follow a T-1/2 law determined by the SRH mechanism [218]. Analogous to R0A, an effective

“activation energy” of ~150 meV was extracted for the seven devices at 250-340 K,

confirming an exponential relationship with inverse temperature (1/T). The sharp decrease

of the carrier lifetime with increasing temperature can be attributed to the growing

dominance of the Auger processes associated with the bandgap narrowing of the SL

absorber at high temperatures. The similar effect of Auger process has been analyzed by

others for InAs/InAsSb T2SLs [219-220]. Overall, the developed method to extract carrier

lifetime include contributions from various transport mechanisms such as Auger and SRH

processes as indicated in Equation 6-3, which should be effective in broader contexts and

closer to actual devices.

132
Temperature T (K)
340 320 300 280 260 240 220 200
1024

Thermal generation rate gth (s-1cm-3)


1023
100
85
70 1S-1040 1022

Lifetime  (ns)
55 1S-2340
M3S-312
40 M6S-312 1021
N8S-312
25 M12S-156
N16S-156 1020

1019
10
8.5
7 1018
3.0 3.5 4.0 4.5 5.0
1000/T (K-1)

Figure 6-5: The thermal generation rate and minority carrier lifetime for the five
multistage and two single-stage devices at high temperatures.

6.3 Interband cascade devices vs quantum cascade devices

6.3.1 Device structures

As estimated in the preceding section, the lifetime in IC devices (ICDs) has a tens

of nanosecond timescale at 300 K. This should be much longer than the intersubband

transitions occurring on a time scale of picosecond in QC devices (QCDs). It is generally

known that the relatively much longer carrier lifetime in ICDs has resulted in a significantly

lower threshold current density (Jth) and power consumption in ICLs at room temperature

(RT) compared to in QCLs. This has been demonstrated for a wide IR spectral region (2.7-

6 m) [73, 221]. Since the lasers normally operate under forward bias, the J-V

characteristics under reverse bias for extracting J0 are not readily available for QCLs.

Hence, the analysis of QCDs is mainly concentrated on RT QC detectors reported in the

literature [83-84, 96, 222-224]. Some ICDs included here are IC laser structures that were

reported previously [115, 120, 179, 225-226], while the others are IC light emitting devices

133
(LEDs). The active regions of all these ICDs consist of an asymmetric “W” quantum well

(QW) [227] with two InAs electron QW layers on both sides of the GaInSb hole QW layer.

The ICDs have numbers of cascade stages (Nc) ranging from 6 to 15. Besides, ICDs with

InAs/Ga(In)Sb T2SL absorbers that were designed as detectors and TPV cells are also

included here [137, 151, 159, 199, 228-231]. They will be denoted by “ICD_SL” to

differentiate from those having QWs in the active regions. Most of the ICDs were

processed into square mesa type devices as well as several broad area IC lasers.

6.3.2 Semi-empirical model for dark current density

In cascade devices, there is a potential barrier region formed between two ends of

adjacent cascade stages, since the electronic states near the two ends lie at a low energy

level on one end and a high energy level on the other. If a forward bias (positive on the

high energy end) is applied to a cascade stage, the number of available carriers being able

to overcome the potential barrier from the low energy end to the high energy end is

increased exponentially with the bias voltage. Consequently, the forward current density

will have an exponential increase with the bias voltage. Conversely, at reverse voltage, the

current density approximates to a constant (J0) value since the number of carriers that can

move from the high energy end does not increase with the reverse bias voltage. Hence,

semi-empirically, the current density-voltage (Jd-V) characteristic in a cascade device with

identical stages can be described by:

𝐽𝑑 (𝑉) = 𝐽0 (𝑒 𝑞𝑉⁄𝑁𝑐𝑘𝑏 𝑇 − 1) (6-5)

Qualitatively, this expression resembles the standard diode equation for a p-n junction.

Equation 6-5 can be derived from a fundamental level with lengthy mathematical

134
manipulations, as described in detail in [141] for ICDs and in [232-234] for QCDs. The

approach offered here grasps the main feature in cascade devices and offers a simple way

to derive Equation 6-5 for current-voltage characteristics in complicated cascade structures.

This approach has not been documented before should be beneficial in helping promote a

better understanding of complex cascade devices.

It has been shown that the value of J0 is proportional to the carrier concentration

and inversely proportional to carrier lifetime that can be affected by various scattering

mechanisms such as defects, doping, phonons and Auger recombination. This relationship

has been explored to extracted carrier lifetime in ICIPs as described in the preceding

section. From Equation 6-5, the R0A of a cascade device can be obtained as:

𝑁𝑐 𝑘𝑏 𝑇
𝑅0 𝐴 = (6-6)
𝑞𝐽0

In theory, the values of J0 for ICDs and QCDs can be extracted by fitting the measured Jd-

V curves to Equation 6-5. However, in an actual device, the parasitic series and shunt

resistance (Rs and Rshunt) are often presented. Considering these factors, the Jd-V curve of a

cascade device should be fitted to a modified equation:


𝑉−𝑅𝑠 𝐴𝐽𝑑
𝐽𝑑 (𝑉) = 𝐽0 (𝑒 𝑞(𝑉−𝑅𝑠 𝐴𝐽𝑑 )⁄𝑁𝑐𝑘𝑏𝑇 − 1) + (6-7)
𝑅𝑠ℎ𝑢𝑛𝑡 𝐴

6.3.3 Saturation current densities for cascade devices

From Equation 6-7, the three parameters, J0, Rshunt and Rs, can be extracted through

the least-square fitting method. In the fits, the values of Rshunt and Rs were kept in the range

of 103-104 and 1-10 , respectively. As an example, Figure 6-6 shows the measured Jd-V

curves and fitting results for a large area (400 m×400 m) eight-stage ICD (wafer R083)

[115] and a fifty-stage QCD (110 m×110 m) [96] at 300 K. The two devices have the

135
identical transition energy E of 0.23 eV in the active region at 300 K, which was the

bandgap for the ICD or the energy separation of the two involved conduction subbands for

the QCD. As shown in Figure 6-6, the magnitude of Jd is at least an order of magnitude

lower in the ICD than in the QCD. This difference is ascribed to the comparatively much

longer carrier lifetimes in the ICD. Also, the least-square fittings based on Equation 6-7

were in excellent agreement with measurements, supporting the validity of the semi-

empirical model. Specifically, the extracted J0 (Rshunt) obtained from the fitting procedure

is 0.017 A/cm2 (5945 ) and 1.8 A/cm2 (6772 ) for the eight-stage ICD and fifty-stage

QCD, respectively. The other fitting parameter Rs is 5  for the ICD, and 7  for the QCD

with a smaller device area.

101
Dark current density (A/cm2)

QCD
100

10-1
ICD
-2
10

10-3 E=230 meV line: simulated


circle: measured

-1.0 -0.5 0.0 0.5 1.0


Voltage (V)

Figure 6-6: The measured and fitted Jd-V curves for an 8-stage ICD and a 50-stage
QCD at 300 K. The ICD and QCD were mentioned in [115] (wafer R083) and [96],
respectively.

Aside from the two devices, the least-square fitting was also performed for other

ICDs [83-84, 96, 222-224] and QCDs [115, 120, 179, 225-226]. The extracted values of J0

at 300 K for these ICDs and QCDs are presented in Figure 6-7. As can be seen, the value

of J0 is more than one order of magnitude lower in ICDs than in QCDs with similar E.

136
This distinction of extracted J0 implies the significant effect of carrier lifetime on transport

current, consistent with threshold behavior in laser performance for a wide infrared spectral

region mentioned earlier. Also, Figure 6-6 shows that J0 tends to increase exponentially

with decreasing E for both ICDs and QCDs. It should be commented that ICDs are more

susceptible to surface leakage currents due to the existence of surface states in their

bandgap. Hence the extracted J0 in Figure 6-7 might be more overestimated for ICDs than

for QCDs. Since there is considerable variation in device area, the product of resistance

and area is a more appropriate quantity as used effectively in Equation 6-7. In general, the

value of RshuntA extracted from fitting is smaller for QCDs compared to ICDs. However,

the ratio of RshuntA to R0A is generally higher in QCDs than in ICDs, which suggests the

relatively lower percentage of surface leakage in QCDs than in ICDs. Moreover, the

material qualities and fabrication technologies may differ greatly between different groups

Overall, the extracted values of J0 are much lower in ICDs than in QCDs. This not only

manifests substantial difference of threshold current density in lasers between the two

families, but also yields considerable differences in detector and PV device performance

as will be discussed later. The vast gap of J0 between ICDs and QCDs is fundamentally

attributed to their distinctive carrier lifetimes since J0 is inversely proportional to the

lifetime. In ICDs, Auger and SRH (through defects) processes are the main scattering

mechanisms. In QCDs, longitudinal optical (LO) phonon scattering prevails and is fast (in

ps or shorter) between and within the conduction subbands. With interband transitions, the

carrier lifetime is in the nanosecond range, about three orders of magnitude slower than for

phonon scattering. The extracted J0 is much lower in ICDs than in QCDs, which

unambiguously proves the much longer lifetime in ICDs than QCDs.

137
Wavelength (m)
8 7 6 5 4
101

Saturation current density (A/cm2)


T=300 K

100 QCD

10-1

10-2 ICD

10-3
150 200 250 300 350
Transition energy (meV)

Figure 6-7: The extracted values of J0 for ICDs and QCDs at 300 K. Some ICDs have
been described previously in [83-84, 96, 222-224], while others are from our
unpublished studies. The QCDs are from [115, 120, 179, 225-226].

6.3.4 Effect of J0 on the performances of detectors

The saturation current density J0 is a measure of Johnson noise in a photodetector.

The R0A contained in Equation 6-6 is also reflected in the specific detectivity D*. As

described in Subsection 1.3.4, D* is essentially a measure of signal to noise ratio ─ the most

important figure of merit for photovoltaic photodetectors operating at zero bias. The

expression of D* is given by Equation 1-8 where Ri is the responsivity. Figure 6-8(a) shows

the measured peak Ri for ICDs and QCDs at 300 K. In addition to some of the ICDs

presented in Figure 6-7, another two ICDs (devices A and B) [136] and ICD_SLs from

[137, 151, 199, 229-231] are also included in Figure 6-8 (a) and (b). As shown in Figure

6-8(a), the peak Ri is generally higher in ICDs than in QCDs and is especially high in

ICD_SLs with enhanced absorption in SL absorbers. The lower Ri in QCDs is partly

because of the low escape probability that is proportional to the carrier lifetime [65, 83] for

QCDs, while this value is close to unity for ICDs with the much longer lifetime [99].

138
Another factor might be the polarization selection rule for intersubband transitions in

conduction band QWs [65, 79], which prohibits the absorption of normal incident light in

QCDs. This problem in QCDs is typically mitigated by making facets made by polishing

at an angle of 45o to the growth direction. In addition, an improved responsivity can be

achieved for intersubband photodetectors by using a photonic metamaterial to enhance the

light-matter interaction. This was demonstrated in a QWIP detector with photoconductive

gain near 9 m at RT [236]. in which the responsivity (~0.2 A/W) is comparable to those

in ICD_SLs as shown in Figure 6-8(a). However, due to substantial noise with a high dark

current density, its detectivity D* (~2.8×107 Jones) is about one order of magnitude lower

than that in ICD_SLs with similar E as shown in Figure 6-7(b).

Wavelength (m) Wavelength (m)


1210 8 6 4 1210 8 6 4
Peak responsivity (mA/W)

QWIP (a) (b)


Peak detectivity (Jones)

102 109

101 108
QWIP
ICD ICD
QCD QCD
100 ICD_SL 107 ICD_SL

100 150 200 250 300 350 400 450 100 150 200 250 300 350 400 450
Transition energy (meV) Transition energy (meV)

Figure 6-8: Measured peak (a) responsivities and (b) detectivities for ICDs, ICD_SLs
and QCDs at 300 K. In addition to some of the ICDs presented in Figure 6-6, two
ICDs (devices A and B) [136] and all ICD_SLs from [137, 151, 199, 229-231] are
included. One QWIP is from [236].

From Equation 1-8 and 6-6, the John-noise limited D* is inversely proportional to

the square root of J0. Figure 6-8(b) shows the measured peak D* values at 300 K for the

considered ICDs and QCDs. As can be seen, the values of D* are almost one order of

magnitude higher for ICDs than for QCDs. At 300 K, the achieved D* in most QCDs is less

139
than 3×107 Jones mainly because of a high J0, while most D* values for ICDs are higher

than 1×108 Jones and some even exceed 1×109 Jones. Also, the difference of D* between

ICDs and QCDs is more significant than the difference in Ri between the two families. This

arises from more than one order of magnitude lower Jo in ICDs than in QCDs, although

the number of cascade stages Nc (<15) in ICDs is less than in QCDs (≥30). If they had the

same Nc, the value of D* would be increasingly higher in ICDs than in QCDs.

According to Equation 1-8 and 6-6, D* is proportional to the square root of the

number of stages if Ri remains unchanged. This is roughly correct when individual

absorbers are only made of a pair QWs and kept thin, and the total absorber thickness does

not cause a substantial light attenuation [99]. Conversely, when the light absorption is

significant in individual absorbers (e.g. especially in ICD_SLs), the attenuation of light

intensity along the propagation direction needs to be considered in evaluating the Ri [141,

194, 232]. In this scenario, the D* for a non-current matched cascade device (e.g. with

identical absorbers) will reach a maximum value at a finite Nc as discussed in [141, 194].

This is particularly true for ICD_SLs where SLs are used as active absorbers to enhance

absorption and responsivity for attaining the highest value of D* among all devices, as

shown in Figure 6-8. Nevertheless, compared to ICDs, the additional increase in D* in

ICD_SLs is not as appreciable as the boost in the peak Ri. This is because the Jo is much

higher in ICD_SLs with thicker SL absorbers than in ICDs. Nevertheless, with two

adjustable parameters, the SL absorber thickness and the number of stages, ICD_SLs can

be optimized with more flexibilities to improve D* at high temperatures [141, 194].

In addition, if the detector has a voltage rather than a current output, one can define

its responsivity as the ratio of output voltage to power [9]. Analogous to p-n diodes,

140
neglecting shunt and series resistances, the net current density J in ICDs and QCDs (with

identical stages) under light illumination can be approximately written as:

𝐽 = 𝐽0 (𝑒 𝑞𝑉⁄𝑁𝑐𝑘𝑏𝑇 − 1) − 𝐽𝑝ℎ (6-8)

where the photocurrent density Jph is simply presumed to be bias independent. In an actual

device, the photocurrent may be bias dependent as described in Chapter 4 and 5 for some

ICTPV devices. Based on Equation 6-8, the open-circuit voltage Voc for a cascade device

can be expressed as:

𝑁𝑐 𝑘𝑏 𝑇 𝐽𝑝ℎ
𝑉𝑜𝑐 = 𝑙𝑛 ( + 1) (6-9)
𝑞 𝐽0

Hence from this equation one can see that the lower J0, the higher Voc would be.

Also, when the photocurrent is significantly lower than the dark current, which is

generally true in the detection of weak light at high temperatures. In this case, Equation 6-

9 can be approximated to first order as:

𝑁𝑐 𝑘𝑏 𝑇 𝐽𝑝ℎ
𝑉𝑜𝑐 = (6-10)
𝑞 𝐽0

which is linearly proportional to the number of stages and the ratio of photocurrent and

saturation current densities. Therefore, if the detector output is voltage, more stages and a

lower saturation current density will benefit the device performance. According to Figure

6-7 and 6-8, the voltage responsivity will be much higher in ICDs than in QCDs. The is

due to the higher photocurrent (proportional to responsivity) and the much lower J0 in ICDs

compared to QCDs. Overall, in terms of either current or voltage responsivity, ICDs will

maintain the advantages over QCDs.

141
6.3.5 Effect of J0 on the performances of photovoltaic cells

As for photovoltaic cells to convert light into electricity, the saturation current

density J0 remains an important parameter to evaluate device performance. As mentioned

in Chapter 1, the relatively low transition energy E in the active region makes ICDs and

QCDs more appropriate for TPV applications where the heat source temperatures are

generally at 1000-2000 K. In fact, TPV cells based on IC structures have been

experimentally demonstrated with high Voc that far exceeded the single bandgap value,

showing the cascade effect [159, 182, 190-191]. In contrast, TPV cells based on QC

structures have not been reported experimentally, possibly due to high values of J0 in

QCDs. Based on Equation 6-10 and the data in Figure 6-7 and 6-8, the open-circuit voltage

Voc can be calculated for cascade devices under light illumination at an incident power

density Pinc. Assuming Pinc=1 W/cm2, about ten times the average of solar radiation at the

surface of the earth, and the radiation peaks at the response wavelength (with spectral

control in a TPV system) for ICDs and QCDs so that Jph=Ri·Pinc, the Voc is estimated and

plotted in Figure 6-9 for the devices presented in Figure 6-8. As can be seen, the QCDs

have a very modest Voc (<3 mV) due to a high J0 even with many stages (≥30). This may

explain why QC TPV cells have not been demonstrated in experiment so far. In contrast,

the values of Voc for the ICDs are considerably higher (more than an order of magnitude in

most cases) than for QCDs. This mainly stems from the much lower J0 in ICDs than in

QCDs (Figure 6-7). Combined with the higher photocurrent density as indicated in Figure

6-8(a), the IC structure is more advantageous than the QC structure for TPV applications.

Note that, despite the much higher Ri, the Voc for ICD_SLs is similar with those for ICDs

because of the higher J0 in ICD_SLs. However, with higher Ri and Jph, ICD_SLs will have

142
a higher output power and conversion efficiency if they have the same number of cascade

stages.

Wavelength (m)
1210 8 6 4

102

Open-circuit voltage (mV)


101

100 ICD
QCD
ICD_SL

100 150 200 250 300 350 400 450


Transition energy (meV)

Figure 6-9: Estimated Voc at 300 K for the ICDs, ICD_SLs and QCDs shown in
Figure 6-8.

6.4 Summary and concluding remarks

In this chapter, firstly, an electrical method is developed to extract thermal generation

rate and minority carrier lifetime in in T2SL-based ICIPs. This method is more general and

considers the parasitic shunt and series resistances existed in practical devices. It can also

cover various transport mechanisms such as Auger and SRH processes. Based on this

method, the carrier lifetime at high temperatures (200-340 K) was evaluated to be between

8.5 and 167 ns, depending on the material quality. The extracted carrier lifetime displayed

a different temperature dependence from those previously obtained by other methods for

T2SL detectors, especially at high temperature range. Speculatively, such a temperature

dependence may be related to the growing dominance of the Auger process at high

temperatures. This method should also be applicable to detectors with other barrier

configurations, such as nBn, XBn and CBIRD [67, 81-82].

143
Secondly, the fundamental difference in carrier lifetime between ICDs and QCDs

is manifested by the saturation current density J0. By comparing and analyzing available

ICD and QCD data, it is shown that J0 can be used as a unified figure of merit to describe

both interband and intersubband cascade structures in terms of their device functionalities.

The significance of J0 on detector and PV cell performances was illustrated by comparing

the measured detectivity and the estimated open-circuit voltage, respectively. The extracted

values of J0 are more than one order of magnitude lower in ICDs than in QCDs with similar

transition energies. This result, in combination with the discussion of the consequences of

J0 on device performance, clearly revealed the advantages of IC configurations over

intersubband QC configurations based on the same framework. The overall picture for both

QCDs and ICDs sheds light from the perspective of a united figure of merit, which will

offer instructive guidance and stimulation to the future development of both ICDs and

QCDs. It is worth pointing out that both ICDs and QCDs have their respective merits. For

example, QCDs are based on more mature material systems. The epitaxial growth and the

device processing technologies are well-established. Consequently, at the present stage,

QCDs can have better uniformity as well as less surface leakage and higher output power

for lasers. Hence, both QCDs and ICDs will coexist for various applications with different

requirements.

144
7 Chapter 7: Long wavelength interband cascade infrared
photodetectors
7.1 Introduction

In Chapter 6, a new method was developed to extract the carrier lifetime in mid-

wavelength ICIPs and a common figure of merit (closely related to carrier lifetime) was

proposed to evaluate the performances of IC and QC devices. In this chapter, further

understandings in the operations and behaviors of ICIPs are presented. All the devices

included in this chapter operate in the LWIR band. However, the fundamental principles

revealed in this chapter are also suitable to ICIPs working in other spectral regions.

Compared to conventional single-absorber structures, the multistage configuration of ICIPs

provides more degrees of freedom for optimizing device performance. On the other hand,

this also complicates the design process and requires a more comprehensive understanding

of multiple factors in order to optimize device performance. For example, ICIPs can be

divided into two groups: current-matched ICIPs [142, 231, 237-239] where the

photocurrent is designed to equal in all stages, and noncurrent-matched ICIPs [99, 137,

198-199, 240-241] with identical stages, as shown in Figure 7-1. In a current-matched ICIP,

the absorbers in the optically deeper stages are made thicker to achieve an equal

photocurrent in all stages. This relies on the precise knowledge of material absorption

coefficients, which may vary with temperature and increase the difficulty in

implementation at different operating temperatures. By comparison, in a noncurrent-

matched ICIP, the individual absorber thicknesses are designed to be identical in each

stage. It’s simpler to implement but has a possible drawback of substantially reduced

responsivity due to light attenuation, especially with relatively thick absorbers [99, 141].

145
Figure 7-1: Schematic illustration of the multi-stage ICIP with (a) regular and (b)
reverse configurations. The two configurations can be realized by reversing the
growth order of layers in one structure without changing the light illumination
direction.

Although these two groups of ICIPs have been explored independently, they have

not been studied together in the same framework. To identify and understand their specific

features and differences in device performance, a comparative study of the electrical and

optical properties of several ICIPs with both absorber designs are presented in this chapter.

Electrical gains significantly exceeding unity are observed from noncurrent-matched

ICIPs. To further examine the preliminary findings on electrical gain and to better

understand how noncurrent-matched ICIPs can be designed for optimized device

performance, additional three ICIPs with varied absorber thicknesses and number of

cascade stages are studied and a theory is developed to quantitatively explain the electrical

gain. As will be discussed in detail in the Section 7.3, a reasonable agreement is obtained

between theoretical calculations and experimental results.

146
7.2 Current matched ICIPs vs noncurrent-matched ICIPs

7.2.1 Device structure, growth and fabrication

The two sets of four ICIP structures included in this section were designed to target

the LWIR region (8-12 m) with a reverse illumination configuration [142, 239]. The four

structures have different numbers of stages and variations of individual absorber

thicknesses, but they have identical electron and hole barriers and the same InAs/GaSb SL

composition. Each period (60 Å) of the SL absorber are made of layers: InSb (1.9 Å), InAs

(31 Å), InSb (1.9 Å) and GaSb (25.2 Å). The two thin InSb layers were inserted to balance

the strain from the InAs layer [168]. The absorbers in the four structures were p-doped to

2.6×1016 cm3 so that the electrons were the minority carriers. The electron barriers consist

of four GaSb/AlSb QWs with GaSb well thicknesses of 33/43/58/73 Å. The hole barriers

are seven digitally graded InAs/GaSb QWs and the InAs well thicknesses therein are

48/50/52/55/58/62/70 Å.

Set #1 includes two current-matched ICIP structures called Mat.-8S and Mat.-12S.

They have eight and twelve cascade stages, respectively. Mat.-8S was fabricated from

wafer S#4-8 that was described in detail in [239]. Mat.-12S is made up of 12 stages with

absorber thicknesses of 180, 192, 210, 228, 246, 264, 282, 306, 336, 366, 396, and 432 nm,

from the surface to the substrate (the direction of light illumination). Set #2 has two

noncurrent-matched ICIP structures, NMat.-16S and NMat.-20S, with sixteen and twenty

cascade stages, respectively. NMat.-16S has sixteen discrete identical stages with the

individual absorber thickness (222 nm) equal to that of the first-stage absorber in Mat.-8S.

NMat.-20S has twenty discrete identical stages with each absorber thickness (180 nm)

equal to that of the first absorber of Mat.-12S. The total absorber thickness in these four

147
ICIP structures is 2.29 m (Mat.-8S), 3.44 m (Mat.-12S), 3.55 m (NMat.-16S), and 3.60

m (NMat.-20S). Table 7-1 summarizes the design parameters, along with some key

material properties including cutoff wavelength 𝜆c, bandgap Eg and activation energy Ea

for the four ICIPs.

Table 7-1: Summary of material and design parameters for the four devices.

Total
Absorber # of 100% Eg (meV) Ea (meV) Ea (meV)
Device thickness
type stages 𝜆c (m) at 0K 78-125K 150-250K
(m)
NMat.-20S Identical 20 3.60 9.5 188 43 160
Current-
Mat.-12S 12 3.44 11.0 174 45 155
matched
NMat.-16S Identical 16 3.55 11.1 172 64 160
Current-
Mat.-8S 8 2.29 11.0 175 45 155
matched

The four ICIP structures were grown by molecular beam epitaxy (MBE) on p-type

GaSb substrates that were nominally undoped. After the MBE growth, the wafers were

processed into deep-etched square mesa devices with dimensions from 50 to 1000 m

using conventional contact UV photolithography and wet etching. A RF-sputter deposited

two-layer passivation (Si3N4 then SiO2) was used for improving overall stress management

and minimizing pin holes, and sputter deposited Ti/Au layers were used for top and bottom

contacts. Finally, the devices were mounted on heat sinks and wire bonded for

characterization.

7.2.2 Electrical properties

Electrical and optical properties of devices from these wafers were determined

through measurements of dark current density-voltage (Jd-V) characteristics and photo-

response spectra. From the measured Jd-V curves, the R0A were extracted for the four

148
representative devices as shown in Figure 7-2 at a wide temperature range. This allows to

obtain the activation energies by fitting R0A (1/T) to the following equation:

𝑅0 𝐴 = 𝐶𝑇 𝑏 𝑒 𝐸𝑎⁄𝑘𝑏 𝑇 (7-1)

where b and Ea are the two fitting parameters. In principle, the parameter b is expected to

be 1.5 if the dark current density scales with ni (SRH limited) and 3 if it scales with ni2

(diffusion limited). The extracted Ea values are shown in Table 7-1, where q=0 was used

at 78-125 K and q=2 was used at 150-250 K. From the extracted Ea, the carrier transport in

these devices at high temperatures (>150 K) is diffusion limited. This is because the

extracted Ea is nearly equal to their zero-temperature bandgaps Eg (T=0), which can be

determined by fitting Eg (T) to the Varshni formula:

𝛼𝑇 2
𝐸𝑔 (𝑇) = 𝐸𝑔 (𝑇 = 0) − (7-2)
𝛽+𝑇

where  and  are the Varshni parameters. The evaluated Eg (T=0) based on Equation 7-2

was 188, 174, 172 and 165 meV for NMat.-20S, Mat.-12S, NMat.-16S and Mat.-8S,

respectively.

The diffusion limited carrier transport can be further examined by comparing the

experimentally extracted R0A with the theoretical projections of a diffusion transport

model, which is given by Equation 2-8. According to this equation, R0A is larger for

detectors with more cascade stages, but lower for detectors with thicker absorbers. This

feature is corroborated by Figure 7-2, where the values of R0A for NMat.-20S and NMat.-

16S are higher than Mat.-12S and Mat.-8S thanks to the larger number of stages and the

thinner individual absorbers for all stages. Note that the thermal generation rate in Equation

2-8 is given by Equation 4-3, which implies that it scales with e− E g / kbT
.

149
T (K)
250 200 150 100

105 R0 A = CT − q e Ea kb T

104

103

R0A (.cm2)
102
NMat.-20S
101
Mat.-12S
100 NMat.-16S
Mat.-8S
10-1
4 6 8 10 12
-1
1000/T (K )
Figure 7-2: Extracted R0A of the four representative devices at various temperatures.

Based on Equation 2-8, the ratio of R0A between NMat.-20S (NMat.-16S) and Mat.-

12S (Mat.-8S) can be obtained from:

𝑅0 𝐴𝑁𝑀𝑎𝑡.−20𝑆 ∑𝑁𝑀𝑎𝑡.−20𝑆(tanh𝑑𝑚 ⁄𝐿 )−1


= 𝑚
∑𝑀𝑎𝑡.−12𝑆(tanh𝑑𝑚 ⁄𝐿 )−1
𝑒 ∆𝐸𝑔 ⁄𝑘𝑏𝑇 (7-3 a)
𝑅0 𝐴𝑀𝑎𝑡.−12𝑆 𝑚

𝑅0 𝐴𝑁𝑀𝑎𝑡.−16𝑆 ∑𝑁𝑀𝑎𝑡.−16𝑆(tanh𝑑𝑚 ⁄𝐿 )−1


= 𝑚
∑𝑀𝑎𝑡.8𝑆(tanh𝑑𝑚 ⁄𝐿 )−1
𝑒 ∆𝐸𝑔 ⁄𝑘𝑏𝑇 (7-3 b)
𝑅0 𝐴𝑀𝑎𝑡.−8𝑆 𝑚

where 𝛥Eg is the bandgap variation between two devices. Here, the diffusion length and

carrier lifetime were assumed to be same for the four wafers, which is reasonable because

they were designed with nominally identical SL absorber periods and grown in a close time

interval. Since the cutoff wavelength of NMat.-20S was shorter than the other three devices

that had a nearly equal bandgap, one needs to account the bandgap difference between

NMat.-20S and Mat.-12S in Equation 7-3(a), while 𝛥Eg can be neglected for NMat.-16S

and Mat.-8S in Equation 7-3(b). Based on Equation 7-3, the calculated ratios of R0A as a

function of diffusion length at 300 K are shown in Figure 7-3. As can be seen, if the

diffusion length far exceeds absorber thickness, the two R0A ratios approach a saturation

value of 6.44 and 2.50 for R0ANMat.-20S/R0AMat.-12S and R0ANMat.-16S/R0AMat.-8S, respectively.

150
7

T=300 K
5

R0A ratio
NMat.-20S/Mat.-12S
4
NMat.-16S/Mat.-8S

2
0 500 1000 1500 2000
Diffusion length (nm)

Figure 7-3: The theoretical R0A curves at T=300K. The device dark current was
dominated by the diffusion process at this temperature.

Table 7-2 shows the experimentally obtained R0A ratios and the theoretically

calculated R0A ratios by assuming the diffusion length is appreciably longer than the

individual absorber thickness (i.e. L ≫ dm). The variations in the calculated values of

R0ANMat.-20S/R0AMat.-12S with temperature resulted from the exponential term exp[Eg /(kbT)]

in Equation 7-3, where Eg was determined from the experimental data with certain

uncertainty. The experimentally obtained values used in Table 7-2 are for bulk R0A,

obtained by excluding the surface leakage contribution based on Equation 4-6. The non-

monotonic temperature dependence of the theoretical and experimentally extracted R0A

ratios may be caused by the uncertainty of Eg as mentioned above. Nevertheless, as shown

in Table 7-2, the experimentally extracted R0A ratios are in good agreement with theoretical

calculations at these high temperatures, confirming the diffusion limited carrier transport.

This also implies that the diffusion length is indeed longer than the individual absorber

thicknesses, though there may be minor inaccuracies in experimental data related to

variations of their bandgaps and parasitic series resistances. From Table 7-2 and Figure 7-

3, it can be inferred that the diffusion length in the four devices is finite, but probably longer

151
than 500 nm at 300 K.

Table 7-2: Theoretical calculated and experimental extracted values of R0A ratios at
high temperatures.
T (K) 280 300 320

(𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟐𝟎𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟏𝟐𝑺 )𝒎𝒆𝒂𝒔𝒖𝒓𝒆𝒅 6.2 6.1 6.3


(𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟐𝟎𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟏𝟐𝑺 )𝒕𝒉𝒆𝒐𝒓𝒚 6.3 6.4 6.4
(𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟏𝟔𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟖𝑺 )𝒎𝒆𝒂𝒔𝒖𝒓𝒆𝒅 2.4 2.5 2.3
(𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟏𝟔𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟖𝑺 )𝒕𝒉𝒆𝒐𝒓𝒚 2.5 2.5 2.5

7.2.3 Responsivity

The optical response of the ICIPs was collected using a FTIR spectrometer and then

calibrated with a 600 K blackbody source (aperture diameter of 0.762cm) with a 2π field

of view (FOV). Due to efficient carrier collection in these ICIPs with thin individual

absorbers, the photocurrent is insensitive to bias voltage. The zero-bias responsivity spectra

of the four representative devices at 200-300 K are shown in Figure 7-4. As can be seen,

the current-matched ICIPs have higher responsivities than the noncurrent-matched ICIPs

at all temperatures of interest. The responsivity of the noncurrent-matched ICIPs was only

about 60% of that obtained from the corresponding current-matched ICIPs with the same

absorber thickness (180 or 222 nm) in the first stage. This relation is exemplified in Table

7-3, where the value of Ri was taken at 7 m for NMat.16S, Mat.12S and Mat.8S ICIPs,

and at 5 m for NMat.-20S since its cutoff wavelength was about 2 m shorter than other

three detectors. These data clearly evidence the necessity of current match for optimal

responsivity, and substantial light attenuation in the optically deeper stages. This

conclusion can be further examined and illustrated by considering the temperature

dependence of responsivity, as shown in Figure 7-5.

152
3 4 5 6 7 8 9 10 11 12
0.20 NMat.-20S
Mat.-12S
0.15 NMat.-16S
0.10 Mat.-8S

0.05 T=200K

Zero-bias responsivity (A/W)


0.00
0.20

0.15

0.10

0.05 T=250K
0.00
0.20

0.15

0.10 T=300K

0.05

0.00
3 4 5 6 7 8 9 10 11 12
Wavelength (m)

Figure 7-4: Zero-bias responsivity spectra for the four devices at different
temperatures.

Table 7-3: Experimentally obtained ratio of responsivity for ICIPs at different


temperatures.

T (K) 200 250 280 300 320

𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟐𝟎𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟏𝟐𝑺 0.66 0.62 0.63 0.61 0.63

𝑹𝟎 𝑨𝑵𝑴𝒂𝒕.−𝟏𝟔𝑺 ⁄𝑹𝟎 𝑨𝑴𝒂𝒕.−𝟖𝑺 0.63 0.61 0.62 0.59 0.57

As shown in Figure 7-5, the responsivities of the four devices exhibited similar

trends with temperature as they peaked at certain temperatures and then fell off with further

increasing temperature. The observed trends were linked with variations of absorption

coefficient, diffusion length, and current match with temperature. As discussed earlier, the

diffusion length (>500 nm at 300 K) was likely longer than or comparable to individual

absorber thicknesses throughout the entire temperature range of interest. Accordingly, the

153
collection of photogenerated carriers would not be affected in these ICIPs at various

temperatures. Hence, the temperature dependence of responsivity resulted mainly from the

increase of absorption coefficient due to bandgap narrowing at higher temperatures and the

consequential change in current match. In other words, the responsivity initially increased

with enhanced absorption as the temperature was raised, and then decreased when the more

substantial light attenuation in the optically deeper stages began to disrupt the current

match. This was more significant for devices with relatively thick absorbers. For instance,

since the first-stage absorber of Mat.-8S and NMat.-16S is thicker (222 nm) than that (180

nm) of Mat.-12S and NMat.-20S, their responsivities peaked at lower temperatures (280

and 250 K) compared to the peak locations (300 and 320 K) for Mat.-12S and NMat.-20S.

This fact once again demonstrates the existence of substantial light attenuation and the need

of current match in achieving optimal responsivity. Note that the cutoff wavelength of

NMat.-20S was much shorter than the other three devices and approached 7 m at low

temperatures. Thus, the light absorption (and attenuation) was small at this wavelength.

This yielded a relatively rapid increase of the corresponding responsivity with temperature

up to 280 K and the peak at 320 K, as shown in Figure 7-5.

0.20
=7m
Zero-bias responsivity (A/W)

0.15

0.10

NMat.-20S
0.05
Mat.-12S
NMat.-16S
Mat.-8S
0.00
150 200 250 300 350
Temperature (T)
Figure 7-5: Temperature-dependent responsivity of the four devices at 7 m.

154
7.2.4 Electrical gain

To perform a further quantitative analysis of current-matched and noncurrent-

matched ICIPs, the absorption coefficients of the SL absorbers were measured at room

temperature as shown in Figure 7-6. Based on the measured absorption coefficient, the

evaluated responsivity was much lower than the values in Figure 7-4 for noncurrent-

matched ICIPs, indicating possible electrical gain (G) exceeding unity. Theoretically, the

responsivities of current-matched and noncurrent-matched ICIPs are expressed as [242]:


1.24
𝑅𝑖 (𝜆) = (1 − 𝑅)(1 − 𝑒 −𝛼𝑑1 )𝐺 (7-4 a)
𝜆

1.24
𝑅𝑖 (𝜆) = (1 − 𝑅)𝑒 −(𝑁𝑐−1)𝛼𝑑1 (1 − 𝑒 −𝛼𝑑1 )𝐺 (7-4 b)
𝜆

where R is the from surface reflectance taken to be 0.31 for an InAs cap layer, and d1 is the

absorber thickness in the first stage. Only the first stage was considered in Equation 7-4(a)

for the current matched ICIPs owing to an equal photocurrent in every stage. All stages

were considered with Equation 7-4(b) for noncurrent-matched ICIPs because the

photocurrent is the smallest in the last stage. According to Equation 7-4, the electrical gain

can be estimated from the measured responsivities and absorption coefficients for the four

devices.

Figure 7-6 shows the estimated electrical gain at room temperature for the four

devices. As can be seen, the electrical gain for the ICIPs exceeds the unity when the

absorption coefficient is higher than a certain value (e.g. >1500 cm-1). As the absorption

coefficient further increases at the higher photon energies, G increases for noncurrent-

matched ICIPs, but remains nearly unchanged in current-matched ICIPs. This is because

the enhanced absorption at a larger photon energy attenuates the light intensity in the last

stage, which then necessities a large electrical gain to maintain current continuity. In

155
contrast, in first stage of the current-matched ICIPs, the increase of electrical gain is not

required since the photocurrent is highest among all the stages. Also, to maintain current

continuity, the electrical gain is required to be higher in ICIPs with thinner absorbers to

make up for a shorter absorption length. This is revealed in Figure 7-6, where the G is

higher in Mat.-12S than Mat.-8S, and is higher in NMat.-20S compared to NMat.-16S

when the photon energy is higher than 0.2 eV. Note that the value of G could vary greatly

in different cascade stages with substantial light attenuation. Gain exceeding unity was also

observed in single-absorber T2SL detectors (>5) [242] and in other MWIR ICIPs [151,

165], although the mechanism was not fully understood. The underlying mechanism and

the relevant theory of electrical gain in ICIPs will be described in detail in Section 7.3.

Wavelength (m)
1110 9 8 7 6 5 4
3.5
3.5 NMat.-20S
Absorption coefficient (103 cm-1)

Mat.-12S 8S 3.0
3.0 NMat.-16S at.-
,M
Mat.-8S 12S
t.-
2.5 , Ma 2.5
Electrical gain
6 S 0S
at.-1 at.-2
2.0 NM NM
2.0
1.5
1.5
1.0

0.5 1.0

0.0 0.5
0.15 0.20 0.25 0.30 0.35
Photon energy (eV)
Figure 7-6: Absorption coefficient and electrical gain at room temperature. The dips
near 4.2 m in the gain curves were due to CO2 absorption in the response spectra.

7.2.5 Johnson-noise limited detectivity

Overall, the generated electrical gain in ICIPs can partly compensate for the light

attenuation in an optically deeper stage. As such, the responsivity in noncurrent-matched

156
ICIPs can be appreciable although not as impressive as in the current-matched ICIPs. Given

much higher R0A (Figure 7-2) and suppressed noise as shown in cleaner response spectra

(Figure 7-4), noncurrent-matched ICIPs may achieve detectivities comparable to current-

matched ICIPs. Also, due to substantial electrical gain, perfect current match is not a must

in ICIPs, which offers great flexibility in design and practical implementation.

Based on the measured responsivity and R0A, the estimated Johnson-noise limited

detectivities for the four devices are presented in Figure 7-7. The general advantage

provided by ICIPs with more stages (theoretically discussed in Chapter 2) can be seen from

the maximum values of D* for NMat.-20S. For example, at 250K, the Johnson-noise-

limited D* at =7 m (with a FOV of 2) were 6.05×108, 5.12×108, 4.51×108 and 4.56×108

Jones for NMat.-20S, Mat.-12S, NMat.-16S and Mat.-8S, respectively. At a higher

temperature (e.g. 300K), the corresponding Johnson-noise limited D* are 2.40×108 (NMat.-

20S), 1.77×108 (Mat.-12S), 1.48×108 (NMat.-16S) and 1.40×108 (Mat.-8S) Jones. These

values of D* significantly exceeds the claimed value (e.g. ≥ 4.0×107 Jones with a FOV

between /2 and 2) for commercial uncooled MCT detectors [91]. The significantly

higher D* for NMat.-20S was partially due to the relatively shorter cutoff wavelength

compared to the other three devices. Nevertheless, with a similar cutoff wavelength, the D*

of NMat.-16S is slightly higher than Mat.-8S with same first-stage absorber thickness, even

though the responsivity is lower in NMat.-16S. Hence, in terms of detectivity, noncurrent-

matched ICIPs with appropriate designs can have comparable or even better performance

over current-matched ICIPs. In fact, there is still room for improvement of the performance

for noncurrent-matched ICIPs. When the stages of an ICIP are made identical, there is a

tradeoff between reduced signal and suppression of noise with increasing stages. Adding

157
more stages to a noncurrent-matched ICIP reduces the thermal noise, but also compromises

the signal current, due to light attenuation in the optically deeper stages. Hence, an

optimized number of cascade stages may exist for maximizing D* based on the absorption

coefficient and absorber thickness [141]. If, however, the electrical gain is considered, the

optimal number of stages will change as discussed in next section.

T=200K NMat.-20S
Mat.-12S
Detectivity (cm.Hz1/2/W)

NMat.-16S
Mat.-8S
109 T=250K

T=300K
108

3 4 5 6 7 8 9 10
Wavelength (m)
Figure 7-7: Johnson-noise limited D* spectra of the four devices at various
temperatures.

7.3 A comprehensive study of electrical gain in ICIPs

7.3.1 Device structure, growth and fabrication

To fully unlock the mechanism and theory of the electrical gain observed in ICIPs,

apart from the two ICIP structures (NMat.-16S and NMat.-20S) in the preceding section,

another three noncurrent-matched structures are studied and compared in this section,.

Hence, there are in total five noncurrent-matched ICIPs quoted in this section. The three

structures were grown using GENxplor MBE system on nominally-undoped p-type GaSb

(001). The electron barriers, the hole barriers, the InAs/GaSb SL composition and the

doping concentration in them are the same with those in NMat.-16S and NMat.-20S.

158
However, they have different numbers of stages and variations of individual absorber

thicknesses. The three structures have 15, 23 and 28 cascade stages, and the corresponding

individual absorber thicknesses are 180, 180 and 150 nm, respectively. For convenience,

the three structures are denoted as I15S-180, I23S-180 and I28S-150. Also, for consistency,

NMat.-16S and NMat.-20S are designated afresh here as I16S-222 and I20S-180,

respectively. In the notations, the “I” indicates the identical-stage design. The total absorber

thicknesses are 2.70 (I15S-180), 3.55 (I16S-222), 3.60 (I20S-180), 4.14 (I23S-180) and

4.20 m (I28S-150). The absorption is insignificant in the electron and hole barriers, since

they are composed of semiconductor QWs with bandgaps that are much wider than the

absorber bandgap.

Table 7-4 summarizes key design and material parameters, including defect density

and perpendicular (⊥) lattice mismatch of the five wafers, which have comparable material

and crystal structural quality. After the MBE growth, the wafers were processed into square

mesa devices with dimensions from 50 to 1000 m using standard contact UV

photolithography followed by wet-chemical etching. A RF-sputter deposited two-layer

passivation (Si3N4 then SiO2) was used to improve overall stress management and

minimize pin holes. Sputter deposited Ti/Au layers provided top and bottom contacts.

Finally, the devices were mounted on heat sinks and wire bonded for characterization.

159
Table 7-4: Summary of the design and material parameters of the five wafers.

# of Individual SL Total Defect density ⊥ lattice


Device
stages thickness (nm) periods thickness (m) (cm-2) mismatch
I15S-180 15 180 30 2.70 5.5×104 -0.394%
I16S-222 16 222 37 3.55 5.0×104 0.043%

I20S-180 20 180 30 3.60 3.3×104 0.061%


I23S-180 23 180 30 4.14 6.4×104 -0.378%
I28S-150 28 150 25 4.20 4.7×104 -0.369%

7.3.2 Responsivity

The optical response of the ICIPs was characterized following the same procedure

described in the beginning of Subsection 7.2.23. The calibrated responsivities of

representative devices (200×200 m2) from the five wafers at 200-300 K are shown in

Figure 7-8(a). As shown, at 300 K, I15S-180, I23S-180 and I28S-150 have a nearly

identical cutoff wavelength (10.6 m), which is longer than for I20S-180 (9.5 m) but

slightly shorter than for I16S-222 (11.1 m). As descried in the Section 7.2, the

responsivities of these ICIPs are relatively small due to the thin individual absorbers,

especially for noncurrent-matched ICIPs because of light attenuation. On the other hand,

the shot and Johnson noises are suppressed for thinner individual absorbers and a larger

number of cascade stages. As shown in Figure 7-8(a), the responsivity spectra for the five

ICIPs at high temperatures are low but clear. Although the spectra were red shifted with

temperature due to bandgap narrowing, the peak responsivity was either nearly unchanged

or raised slightly (<10%) with increasing temperature. This is because the light absorption

and attenuation in multiple stages limit the maximal value of QE and increasing the

absorption coefficient beyond a certain value does not enhance QE, as shown in Figure 7-

160
8(b) for the five devices.

In general, the QE of a noncurrent-matched ICIP is determined by the last stage

with minimum number of photogenerated carriers, as expressed by Equation 7-4(b). The

calculated QEs for the five devices as a function of absorption coefficient are shown in

Figure 7-8(b). The quite small values (<1.8%) agree with the relatively low responsivities

shown in Figure 7-8(a). Also, the order of the calculated QEs of the five devices is nearly

the same as for the measured responsivities. The peak values of QEs are1.75%, 1.64%,

1.30%, 1.13% and 0.92% that occur at an absorption coefficient of 3527, 2737, 2567, 2287

and 2119 cm-1 for I15S-180, I16S-222, I20S-180, I23S-180 and I28S-150, respectively.

This is because of a tradeoff between the light absorption and attenuation in the last

individual stage. From Equation 7-4(b), for a given absorption coefficient, it is anticipated

that the device with thinner individual absorbers and thicker total absorber will have a

smaller QE, and thus a lower responsivity. For instance, with similar cutoff wavelengths,

the responsivity of I28S-150 is lower than I23S-180 and I28S-180 at each temperature of

interest. Specifically, at T=300 K and =7 m, the responsivity of I15S-180, I23S-180 and

I28S-150 is 0.098, 0.078 and 0.065 A/W, respectively. Note that the lower responsivity in

I28S-150 does not necessarily result in a lower detectivity since it also relies on the noise

as will be discussed later.

161
3 4 5 6 7 8 9 10 11 12
1.8
0.12 (a) T=200K
0.10 (b) I15S-180
I15S-180 1.6
0.08
I16S-222 I16S-222
0.06 I20S-180
1.4
Zero-bias Responsivity (A/W)
0.04 I23S-180
0.02 I28S-150
1.2 I20S-180
0.12
0.10 T=250K
I23S-180
0.08 1.0

QE(%)
0.06
0.04 0.8 I28S-150
0.02
0.12 0.6
0.10 T=300K
0.08 0.4
0.06
0.04 0.2
0.02
0.00
3 4 5 6 7 8 9 10 11 12 0 1000 2000 3000 4000 5000
Wavelength (m) Absorption coefficient (cm-1)

Figure 7-8: (a) Zero-bias responsivity spectra for the five devices at different
temperatures. (b) Theoretically calculated external quantum efficiency of the five
devices vs. absorption coefficient.

As mentioned above, there is a tradeoff between light attenuation and absorption

related to the last stage in a noncurrent-matched ICIP. This is revealed by the trends of the

calculated QE with absorption coefficient for the five devices. As can be seen in Figure 7-

8(b), the calculated QE curves of the five devices exhibit similar and nearly parallel

patterns with increasing absorption coefficient. They all peak at a certain absorption

coefficient and then fall off with further increases. Although the individual absorbers of

I16S-222 are thicker than I15S-180, the calculated QE of I16S-222 was smaller when the

absorption coefficient exceeds 2600 cm-1, due to more substantial light attenuation in the

last stage of I16S-222. This explains the measured lower responsivity of I16S-222

compared to I15S-180 at shorter wavelengths (e.g.  4 m). In the opposite case where the

light absorption had a greater effect than the attenuation in the last stage, the responsivity

162
of I16S-222 was higher than I15S-180, as manifested in the longer-wavelength region.

Note that the analyses have not accounted for the effect of electrical gain on responsivity.

In fact, in this context, the responsivity of a device follows the same sequence as the

photocurrent in the last stage, which will be discussed in the Subsection 7.3.5.

7.3.3 Electrical gain

Like the devices described in Section 7.2, the estimated responsivities with the

measured absorption coefficient for the five devices are smaller than the values shown in

Figure 7-8. This means that the electrical gain (G) exceeds unity in the five ICIPs. Based

on Equation 7-4(b), the G can be extracted from the experimentally measured absorption

coefficient and responsivities. Figure 7-9 shows the estimated G, along with the measured

absorption coefficients at room temperature. The electrical gain of the five noncurrent-

matched ICIPs exhibits a monotonic increase with absorption coefficient and when the

absorption coefficient is higher than a certain value (e.g. 1500 cm-1), the electrical gain

exceeds unity. For I23S-180 and I28S-150, G can be as high as ~4 at an absorption

coefficient of 4800 cm-1, which is expected to compensate for more significant light

attenuation when the absorption is increased. Thanks to the high G, the Johnson-noise

limited detectivity of the two devices can exceed that of I15S-180 at 300 K, as will be

discussed in Subsection 7.3.7.

Equation 7-4(b) states that the electrical gain in noncurrent-matched ICIPs is to

compensate for the attenuation of incident light in the last stage due to absorption in the

preceding stages. Hence, the G needs to be higher in ICIPs with thinner individual absorber

and thicker total absorber to make up for the shorter absorption length and larger

attenuation in the last stage. This inference from a physical viewpoint agrees with the

163
estimated G for the five devices. As shown Figure 7-9, the five devices in ascending order

of G are I15S-180, I16S-222, I20S-180, I23S-180 and I28S-150. This sequence is exactly

in ascending order of the total absorber thickness. The higher G in I28S-150 compared to

the other four devices was also partially because of a shorter absorption length with a

thinner individual absorber. In addition, the G was slightly higher in I16S-222 compared

to I15S-180 because there was more substantial light attenuation in the last stage of I16S-

222, even though the thicker individual absorbers enabled more light absorption in the last

stage. Accordingly, although both light attenuation and absorption in the last stage were

relevant, the attenuation outweighed the absorption in the five devices when determining

G. In fact, the G differs between stages in a noncurrent matched ICIP due to different light

attenuations. The optically deeper stages have higher G to compensate for the more

significant light attenuation. Consequently, the G depends on the number of cascade stages

and is not the same for all stages.

Wavelength (m)
10 8 6 4
5.5
4500 I15S-180
5.0
I16S-222
Absorption coefficient (cm-1)

4000 I20S-180 50 4.5


1
I23S-180 S- 2
3500 ,I
28 -22
I28S-150 0 I 16S 4.0
-18
Electrical gain

3000 23S 3.5


0 ,I
2500 -18 80 3.0
5S S-1
I1 I20
2000 2.5
1500 2.0
1000 1.5
500 1.0
0 0.5
0.15 0.20 0.25 0.30 0.35 0.40
Photon energy (eV)

Figure 7-9: Absorption coefficient and electrical gain at room temperature. The dips
near 4.2 μm in the gain curves were due to CO2 absorption in the response spectra.

164
7.3.4 Underlying mechanism of electrical gain

As initially proposed in Refs. [141], the electrical gain in ICIPs stems from the

adjustment of the electric potential over every cascade stage to maintain current continuity.

In a noncurrent-matched ICIP, since the light is partially absorbed in the preceding stages

and attenuates along the propagation direction, the number of photogenerated carriers (or

the photocurrent) will not be the same in each stage. To fulfil the same current flow in each

stage, the large photocurrent in the front stages (near the top surface), must be

counterbalanced by an injection current induced by a forward electric potential. Contrarily,

the small photocurrent in the back stages (near the bottom) must be supplemented by a

thermal generation current resulted from a reverse electric potential. The total electric

potential over all of stages equates zero or the external voltage if a bias is applied on the

device. At high temperatures, the thermal generation current is high and therefore

significant gain can be obtained in the back stages, as illustrated in the current five devices.

In next subsection, a theory is developed to quantitatively describe the measured

photocurrent and the electric potentials over each stage in these ICIPs.

As per Planck’s law and standard theories for barrier detectors [141, 244], the

photocurrent in the mth stage (Iphm) of a noncurrent-matched ICIP receiving the radiation

from a standard blackbody is given by:

2𝜋𝑞𝐴𝑜𝑝𝑡 𝑟𝑎 2 ∞ 𝐸2 𝐸2
𝐼𝑝ℎ = ( ) ∫𝐸 𝑄𝐸𝑚 ( 𝐸⁄𝑘𝑏𝑇𝑏𝑏 − 𝐸⁄𝑘𝑏 𝑇𝑎𝑚𝑏 ) 𝑑𝐸 (7-5 a)
ℎ3𝑐 2 𝑑𝑠𝑑 𝑔 𝑒 𝑒

𝑄𝐸𝑚 = (1 − 𝑅)𝑇𝑤𝑖𝑛 𝑒 −(𝑚−1)𝛼𝑑 𝑄𝐸𝑑 (7-5 b)

𝛼𝐿 𝛼𝐿exp(−𝛼𝑑)
𝑄𝐸𝑑 = 1−(𝛼𝐿)2 × [tanh(𝑑 ⁄𝐿) + − 𝛼𝐿] (7-5 c)
cosh(𝑑 ⁄𝐿 )

where Aopt is the optical area of the device, ra is the radius of the aperture of the blackbody

165
source, dsd is the distance between the blackbody source and the device, Eg is the bandgap

of the absorber, QEm is the effective quantum efficiency in the mth stage, Tbb is the

blackbody temperature (set to 600 K), Tamb is the ambient temperature (~297 K), E is the

photon energy, Twin (~0.7) is the transmittance of the cryostat window (ZnSe), and QEd is

the individual quantum efficiency, which is equal in each stage. The collection probability

of photogenerated carriers is imbedded in Equation 7-5(b) and Equation 7-4(b) corresponds

to the limiting case of Equation 7-5(b) where the diffusion length is much longer than the

individual absorber thickness, leading to complete collection of photo-generated carriers.

As mentioned before, the responsivities of the five devices have weak bias dependence.

This conveys that the photo-generated carriers are efficiently collected in the five devices

due to thin individual absorbers. Hence, there is no essential difference between the two

equations, and the choice of diffusion length (typically <2 m at room temperature) is

inconsequential to the calculation of QEm; here Ln was taken to be 0.7 m. The optical loss

due to the reflection of cryostat window was considered during the calibration of

responsivity, hence Equation 7-5(b) only accounts for reflectance at the top surface of the

device. Based on Equation 7-5, the calculated photocurrent in each stage of the five devices

at room temperature is shown in Figure 7-10(a). As shown, the calculated individual

photocurrent decreases with stage number, in agreement with the attenuation of light

intensity. The first stage is unaffected by light attenuation, therefore the photocurrent in

this stage only depends upon the absorption coefficient and individual absorber thickness.

Among the five devices, I16S-222 has the highest photocurrent in the first stage since it

has the thickest individual absorber. The I20S-180 device has the lowest first-stage

photocurrent because it has the largest bandgap. Additionally, the order of the five devices,

166
in ascending photocurrent in the last stage, is nearly consistent with the order according to

the responsivity spectra [Figure 7-8(a)]. As will be illustrated later, the signal current in the

context of electrical gain, follows the same sequence as well. The calculated photocurrents

of I15S-180 and I23S-180 overlap as expected, because they have the same individual

absorber thickness, cutoff wavelength, and absorption coefficient.

9
60 I15S-180
8 (a) I15S-180 (b)
I16S-222 I16S-222
I20S-180 40 I20S-180

Electric potential (nV)


7
I23S-180
Photocurrent (nA)

I23S-180
6 I28S-150 20 I28S-150

5 0

4 -20
3
-40
2
0 5 10 15 20 25 0 5 10 15 20 25
Stage number Stage number

Figure 7-10: Theoretically calculated photocurrent based on Equation 7-5 and (b)
electric potential calculated based on Equation 7-7 for each stage of the five devices
at room temperature.

7.3.5 Net effect of electrical gain

Based on the mechanism discussed above, with electrical gain, the signal current Is

can be expressed as:

𝐼𝑠 = 𝐼𝑝ℎ1 − 𝐼0 (𝑒 𝑞𝑉1⁄𝑘𝑏𝑇 − 1) = 𝐼𝑝ℎ𝑚 − 𝐼0 (𝑒 𝑞𝑉𝑚 ⁄𝑘𝑏 𝑇 − 1) (7-6)

where I0 is the saturation dark current, which is identical in each stage for a noncurrent-

matched ICIP, and Vm is the electric potential across the mth stage. At zero external bias,

the sum of the electric potential across each stage is zero: V1+V2+···+ VNc-1+ VNc=0. At high

temperatures, I0 is much higher than the photocurrent, thus the magnitude of the electric

167
potential will be quite small and a first-order approximation in Vm can be used. Equation

(7-6) plus the condition of zero total electrical potential, to the first-order approximation,

leads to the expression of Vm:

1 𝑘𝑏 𝑇 𝑐 𝑁
𝑉𝑚 = (𝑁𝑐 𝐼𝑝ℎ𝑚 − ∑𝑖=1 𝐼𝑝ℎ𝑖 ) (7-7)
𝑁𝑐 𝑞𝐼0

where i denotes the stage number. Based on this equation, the calculated electric potential

across each stage at room temperature for the five devices is shown in Figure 7-10(b). As

can be seen, the individual electric potential is very small as it ranges from several to tens

of nV. Hence, the first-order approximation is appropriate when estimating the signal

current in the five ICIPs. In a certain stage, the electric potential shifts from positive to

negative. This means that the electrical gain is above unity in the subsequent stages.

By replacing Vm with Equation 7-7, the signal current in Equation 7-6 can be

modified to:

𝑓𝑖𝑟𝑠𝑡 𝑜𝑟𝑑𝑒𝑟 𝑞𝑉𝑚


𝐼𝑠 = 𝐼𝑝ℎ𝑚 − 𝐼0 (𝑒 𝑞𝑉𝑚 ⁄𝑘𝑏 𝑇 − 1) → 𝐼𝑝ℎ𝑚 − 𝐼0
𝑘𝑏 𝑇

𝑞𝐼0 1 𝑘𝑏 𝑇 𝑐 𝑁 𝑐 𝑁 𝐼𝑝ℎ𝑖
= 𝐼𝑝ℎ𝑚 − (𝑁𝑐 𝐼𝑝ℎ𝑚 − ∑𝑖=1 𝐼𝑝ℎ𝑖 ) = ∑𝑖=1 (7-8)
𝑘𝑏 𝑇 𝑁𝑐 𝑞𝐼0 𝑁𝑐

This equation states that the signal current in a noncurrent-matched ICIP will be the

average of the photocurrents in each stage, provided that the dark current is much higher

than the photocurrent. The net effect of electrical gain is to raise the signal current from

the minimum photocurrent in the last stage to the average photocurrent over all the stages.

Figure 7-11 shows the calculated and the measured signal currents for the five devices in a

temperature range of 200-300 K. The calculations agree well with the experimental values

for the five devices, considering some inaccuracies and uncertainties in the absorption

coefficients and possible underestimates for I16S-222 at high temperatures with a small

168
resistance. Also, the device sequences according to the calculated last-stage photocurrents

[Figure 7-10(a)], and the calculated and measured signal currents have almost the same

order. The theory predicts that the photocurrent should increase with temperature, since the

number of photogenerated carriers increases due to bandgap narrowing. However, for

I16S-222 and I15S-180, the measured photocurrent slightly decreased while the device

temperature was raised from 280 to 300 K. This was probably caused by an error from the

small resistances or other factors that have not been understood yet, which deserve future

investigation.

6.0
5.5
5.0
Signal current (nA)

4.5
4.0
3.5
I15S-180
3.0 I16S-222
2.5 Solid symbol: measured I20S-180
2.0 Open symbol: calculated I23S-180
I28S-150
1.5
200 220 240 260 280 300
T (K)
Figure 7-11: Theoretically calculated and experimentally measured signal current for
the five devices.

Thanks to the electrical gain, the signal current is enlarged. Likewise, the spectral

responsivity is enhanced and can be expressed by the average value of QEm in each stage:

𝜆
𝑅𝑖 (𝜆) = 1.24 (1 − 𝑅) [𝑄𝐸𝑑 + 𝑒 −𝛼𝑑 𝑄𝐸𝑑 + ⋯ + 𝑒 −(𝑁𝑐−1)𝛼𝑑 𝑄𝐸𝑑 ]⁄𝑁𝑐

𝜆 (1−𝑅)𝑄𝐸𝑑 (1−𝑒 −𝑁𝑐 𝛼𝑑 )


= (7-9)
1.24 𝑁𝑐 (1−𝑒 −𝛼𝑑)

This expression of Ri () can be further simplified for ICIPs with thin absorbers. The QEd

in the numerator can be canceled with the term (1-e-ad) in the denominator when the

169
photogenerated carriers are fully collected. Therefore, this equation indicates that, for

noncurrent-matched ICIPs with thin absorbers, Ri () should monotonically increase with

absorption coefficient at high temperatures. This is consistent with the calculated

temperature dependence of the signal current as shown in Figure 7-11. However, when 

is large at a photon energy well above the bandgap, the exponential term exp(-Ncd) in the

numerator in Equation 7-9 is small and negligible. Consequently, the Ri () reaches its

saturation value, as observed in Figure 7-8(a) where the peak responsivities are almost

insensitive to temperature.

Based on Equation 7-9, the simulated responsivity spectra for I20S-180 and I23S-

180 at 250 K are shown in Figure 7-12. Also displayed are the calculations without

considering the gain, experimental results with the regular mode of the IR source (inside

the Nicolet 8700 FTIR spectrometer) and experimental results with a standard blackbody

radiation source (model IR-563 from Infrared Systems Development Corporation) at 800

and 1200 K. In comparison with the regular theory without the gain, the calculation based

on Equation 7-9 agrees much better with the experimental results. However, there are some

deviations from the experimental results at high photon energies. Also, the real responsivity

spectrum depends on the light source, while the calculated responsivity cannot express this

feature. The effect of the light source is significant when it radiates more photons at high

energies, which is evidenced by the higher responsivity at short wavelengths measured

with the IR source (which has more high energy photons than the 1200 K blackbody

source) and with the blackbody source at different temperatures. This means that the gain

spectrum has some dependence on the incident photon distribution and the real response

spectrum might not exactly follow with Equation 7-9, especially when the incident light

170
has a broad energy distribution with a large percentage of high energy photons. One

interpretation of this phenomenon is that larger electrical gains are required to compensate

for the increasing light attenuation at high photon energies and it turns out to be more

dominant with the increased proportion of high energy photons.

0.10 0.10
I20S-180 I23S-180
0.08 T=250K 0.08 T=250K
Responsivity (A/W)

Responsivity (A/W)
0.06 0.06
(a) (b)
0.04 800 K 0.04 800 K
1200 K 1200 K
IR Source IR Source
0.02 0.02
Theory-with gain Theory-with gain
Theory-without gain Theory-without gain
0.00 0.00
3 4 5 6 7 8 9 10 3 4 5 6 7 8 9 10
Wavelength (m) Wavelength (m)

Figure 7-12: Theoretical and experimental responsivity spectra for two devices at 250
K with the IR source and a standard blackbody radiation source at 800 and 1200 K.

7.3.6 Electrical characteristics

The electrical properties of the ICIPs were characterized at 78-340 K. The measured

dark current densities at -50 mV and the R0A of the five devices are shown in Figure 7-13.

At 300 K, the Jd at -50 mV was 0.95, 1.46, 0.32, 0.56and 0.43 A/cm2 for I15S-180, I16S-

222, I20S-180, I23S-180 and I28S-150, respectively. These values of Jd are nearly two

orders of magnitude lower than that (50-70 A/cm2) stated by the “Rule 07” for HgCdTe

detectors [244]. Table 7-5 presents the activation energies extracted from the temperature

dependence of R0A, along with the zero-temperature bandgaps for the five devices. For

I16S-222 and I20S-18, the carrier transport is diffusion limited since the activation energies

approach the zero-temperature bandgaps. In contrast, for the other three devices, the

extracted Ea is 50%-100% of zero-temperature bandgap, suggesting the involvement of

171
both the diffusion and the SRH processes in carrier transport. As can be seen in Table 7-5,

the relatively larger perpendicular lattice mismatch may lead to a somewhat poorer material

quality for these three devices compared to I16S-222 and I20S-180. Theoretically, given

diffusion-limited carrier transport, the R0A of a noncurrent-matched ICIP can be expressed

by Equation 2-9. This equation indicates that, with a similar cutoff wavelength, the

noncurrent-matched ICIP with more stages and thinner individual absorber will have a

larger R0A. This correlation is directly proved by the ascending order of R0A of I15S-180,

I23S-180 and I28S-150, although the carrier transport was partially affected by the SRH

process. With similar cutoff wavelengths at 300 K, I28S-150 had the largest R0A (1.12×10-

1
.cm2), followed by I23S-180 (8.43×10-2 .cm2) and then I15S-180 (4.78×10-2 .cm2).

The largest R0A (1.48×10-1 .cm2 at 300 K) of I20S-180 among the five devices was

ascribed to the shortest cutoff wavelength. On the same account, the R0A (3.15×10-2 .cm2

at 300 K) of I16S-222 was smallest, as a result of the longest cutoff wavelength as well as

the thickest individual absorber among the five devices.

T (K) T (K)
340 320 300 280 260 240 220 200 340 320 300 280 260 240 220 200

I15S-180 10 I15S-180
(a) I16S-222 I16S-222 (b)
1 I20S-180 I20S-180
Jd @ -50 mV (A/cm2)

I23S-180 I23S-180
R0A (.cm2)

I28S-150 1 I28S-150

0.1

0.1

0.01

0.01
3.0 3.5 4.0 4.5 5.0 3.0 3.5 4.0 4.5 5.0
1000/T (K-1) 1000/T (K-1)

Figure 7-13: Arrhenius plot of dark current density (measured at -50 mV) and R0A
of the five devices in the temperature range of 200-340 K.

172
Table 7-5: Comparison of electrical parameters of the five ICIPs.

I15S-180 I16S-222 I20S-180 I23S-180 I28S-150


R0A (cm2) @ 200 K 3.68 2.35 14.30 5.76 6.43
R0A (10-2 cm2) @ 300 K 4.78 3.15 14.8 8.43 11.2
Eg (meV) @ 0 K 176 172 188 174 174
Ea (meV) 150-250 K 132 160 160 132 102

7.3.7 Johnson-noise limited detectivity

The estimated Johnson-noise limited detectivities for the five devices are shown in

Figure 7-14. Because of significant electrical gain, in terms of detectivity, these noncurrent-

matched ICIPs can outperform the commercially viable uncooled HgCdTe detectors with

a similar cutoff wavelength. For instance, at T=250 K, the Johnson-noise limited D* values

(for 𝜆=7 m and a FOV of 2) were 5.34×108 (I15S-180), 4.41×108 (I16S-222), 5.91×108

(I20S-180), 5.28×108 (I23S-180) and 5.45×108 (I28S-150) Jones. At a higher temperature

(e.g. 300 K), the corresponding Johnson-noise limited D* were 1.66×108, 1.46×108,

2.37×108, 1.84×108 and 1.87×108 Jones, for I15S-180, I16S-222, I20S-180, I23S-180 and

I28S-150, respectively. By comparison, the stated D* (FOV between /2 and 2 ) for

commercial uncooled MCT detectors is about 4.0×107 Jones [91]. The significantly higher

D* of I20S-180 was partially due to the relatively shorter cutoff wavelength than the other

four devices. By the same token, the lowest D* of I16S-222 was partly because of the

longest cutoff wavelength among the five devices. With similar cutoff wavelengths, despite

the lower responsivities, the D* of I23S-180 and I28S-150 are slightly higher than that of

I15S-180 at 300 K, due to the larger R0As of these two devices than that of I15S-180 (Table

7-5). The Johnson-noise limited D* (𝜆=7 m) and the 100% cutoff wavelength (at 300 K)

for the five devices are summarized in Table 7-6.

173
T=200K I15S-180
I16S-222
I20S-180
I23S-180

Detectivity (Jones)
109 T=250K I28S-150

108 T=300K

3 4 5 6 7 8 9 10
Wavelength (m)
Figure 7-14: Johnson-noise limited D* spectra of the five devices at various
temperature.

Table 7-6: Comparison of D* at 𝜆=7 m, along with the 100% cutoff wavelengths at
300 K, for the five devices.
I15S-180 I16S-222 I20S-180 I23S-180 I28S-150
100% cutoff (m) @ 300 K 10.6 11.1 9.5 10.6 10.6
* 8
D (10 Jones) @ 250 K 5.34 4.41 5.91 5.28 5.45
D* (108 Jones) @ 300 K 1.66 1.46 2.37 1.84 1.87

In fact, there is still room for improvement of D* in noncurrent-matched ICIPs. As

mentioned in Section 7.2, the tradeoff between reduced signal and suppressed noise as the

number of stages increases implies that there is an optimal number of stages that maximizes

D* based on the absorption coefficient. The optimal number depends on the electrical gain

is considered or not since it alleviates the signal current compromise. If G is accounted,

according to Equation 7-9, the Ri (𝜆) will be equal to the average value of all the stages. If,

however, the gain is excluded, the Ri (𝜆) will be determined by the value of the last stage.

In [141] and [245], the optimizations of D* ignored the effect of G and consequently the

optimized D* (and corresponding Nc) was underestimated.

174
If the electrical gain is considered, based on Equation 2-9 and 7-9, the Johnson-

noise limited detectivity of a noncurrent-matched ICIP can be estimated by the following

equation:

𝜆 𝑄𝐸𝑑 [1−exp(−𝑁𝑐 𝛼𝑑)]


𝐷∗ = (7-10)
ℎ𝑐 √𝑁𝑐 [1−exp(−𝛼𝑑)]√4𝑔𝑡ℎ 𝐿 tanh(𝑑 ⁄𝐿 )

where QEd is the individual quantum efficiency and is given by Equation 7-5(c). The

calculated D* as a function of the number of stages for different individual absorber

thicknesses are shown in Figure 7-15. Both cases are considered, where the gain is included

or excluded. In the calculation, the absorption coefficient was taken to be 2000 cm-1,

closely corresponding to =7 m (Figure 7-9), and the diffusion length was assumed to 0.7

m. As can be seen in Figure 7-15, the calculated D* peaks at a certain number of stages

and then decreases with more stages, as anticipated from the tradeoff between signal and

noise mentioned above. However, with certain individual absorber thickness, the D* peaks

at a higher value and at a larger number of stages when the gain is considered. For instance,

for d=0.5L, the calculated optimal number of stages is 18 when the gain is considered,

while it is 7 when the gain is ignored. This is consistent with the previous statement that

the gain alleviates the effect of light attenuation, thus bringing an upward shift of the

optimal number of stages. It was also reflected by a modest drop of D* after the peak value,

as distinguished from the sharp decrease in the case without the gain. Adding many stages

in a noncurrent-matched ICIP could make D* approach zero if the gain is absent. However,

this could occur only at a significantly larger number of stages if the gain is included. The

peak value of D* is raised by about 40% with the gain for each given absorber thickness.

But, in both cases, the peak D* has a weak dependence on the absorber thickness, especially

when the absorber in each stage is made thin.

175
1.0 d/L=0.25

0.8 d/L=0.50

0.25 d/L=0.75
0.6

D* (a.u.)
d/L=1.00
0.50
0.4
0.75
0.2 1.00
solid: with gain
0.0 dashed: without gain
0 10 20 30 40 50
Number of stages
Figure 7-15: Detectivity derived from Equation 7-10 versus the number of stages with
various ratios of the individual absorber thickness to the diffusion length (d/L), which
are labeled near the curves in the two cases.

7.4 Summary and concluding remarks

In this chapter, a comparative study of four LWIR ICIPs with current-matched and

noncurrent-matched configurations is presented. It is demonstrated that current match is

necessary to maximize the utilization of absorbed photons for optimal responsivity. The

reduced responsivity in noncurrent-matched ICIPs is correlated with light attenuation in

the optically deeper stages. Based on the extracted R0As for these LWIR ICIPs, the

diffusion length is evaluated to be longer or comparable to 0.5 m at various temperatures

of interest. In addition, electrical gain above unity is observed, which is more substantial

in noncurrent-matched ICIPs for maintaining current continuity. The significant electrical

gain enabled an appreciable responsivity in noncurrent-matched ICIPs, although still not

comparable with current-matched ICIPs. This, combined with the large R0A, resulted in

Johnson-noise limited detectivities (>1.4×108 Jones at 300 K) comparable to or even better

than in current-matched ICIPs.

176
To fully explain the observed electrical gain, additional three noncurrent-matched

structures are included and studied, which shows that the electrical gain commonly exists

in noncurrent-matched ICIPs. Furthermore, a theory is developed to quantitatively explain

the electrical gain in ICIPs. The calculations based on this theory exhibit good agreement

with experimental results. Also, on this basis, insights and guidance to optimize the

Johnson-noise limited detectivities in noncurrent-matched ICIPs are provided. This theory

on electrical gain should also be applicable to other types of multistage photodetectors such

as QWIPs [65, 78] and QCDs [83, 84]. This is because, even with distinctive transition

mechanisms from ICIPs, these types of multistage detectors are also limited by light

attenuation in the optically deeper stages, especially when the total absorbers are made

thick. Likewise, the electric potential across each stage in QWIPs and QCDs will be self-

adjusted to maintain current continuity and electrical gain will supplement the photocurrent

in the optically deeper stages.

177
8 Chapter 8: Concluding notes and future work

8.1 Dissertation summary

The aim of this dissertation research was to identify and understand specific factors

that affect narrow bandgap TPV cell performance and investigate how interband cascade

(IC) structures can improve thermophotovoltaic (TPV) cells and infrared detectors, as well

as to gain further understanding of relevant device physics and operations. IC devices are

unique because of their multistage and multifactor nature in design, which was made

feasible largely thanks to the type-II broken gap alignment between InAs and GaSb. For

example, electron inter-stage transport profits much from this alignment as it enables the

smooth transition of electrons from the valence band in GaSb layer to the conduction band

in InAs layer without any considerable resistance. Through this process, electrons recycle

themselves between stages with a transport path that consists of a series of interband

excitation and collection events.

A consequence of the multistage strategy is the reduction of quantum efficiency (or

photocurrent) due to the fact that multiple photons are required for an electron to traverse

between the contacts. Nevertheless, the quantum efficiency is no longer an appropriate

measurement for multistage structures where the particle conversion efficiency is more

appropriate and is higher in IC devices. The multistage design uses thin absorbers in all

stages to ensure efficient collection of photogenerated carriers before they recombine;

while utilizing multiple stages to absorb incident photons to the maximal extent. This

results in advantages such as enhanced open-circuit voltage and suppressed noise in ICTPV

cells and IC infrared photodetectors (ICIPs), respectively. Ultimately, these advantages

178
enable the higher conversion efficiency and detectivity in multistage ICTPV cells and

ICIPs compared to conventional single-absorber TPV cells and detectors.

In chapter 3, compelling theoretical arguments are provided to underpin the

advantages of multistage ICTPV devices over single-absorber TPV devices. This chapter

begins with the identifications of the limiting factors that have driven low efficiencies in

single-absorber TPV devices. These factors are closely integrated with the high dark

saturation current density, short carrier lifetime, small absorption coefficient and limited

diffusion length. Their impact on conversion efficiency was illustrated in T2SL based TPV

devices in view of several scenarios with different values of L. It is shown that the

multistage IC structure can eliminate the diffusion length limitation that affects single-

absorber devices. As such, the particle conversion efficiency can approach 100%, and the

conversion efficiency can be increased by about 10% in a wide range of L values and

bandgaps.

In chapter 4, a fair amount of experimental evidence is presented to illustrate and

confirm the theoretically projected advantage of multistage ICTPV devices. This is done

by a comparative study of three narrow bandgap (~0.2 eV at 300 K) TPV devices with a

single stage, and three and five cascade stages. Based on the measured quantum efficiency

(QE), the diffusion length is extracted to be ~1.5 m at 300 K, which severely limited the

collection efficiency of photogenerated carriers in the single-absorber device (<20%).

Instead, the extracted collection efficiency in multistage devices approach 100%, thus its

conversion efficiency is greatly improved compared single-absorber TPV devices (3.6%

vs 0.9%).

179
Chapter 5 deals with the detailed characterization and performance analysis in

narrow bandgap (0.22-0.25 eV at 300 K) multistage ICTPV devices with increased number

of stages (i.e. 6, 7, 16, and 23 stages). It is found that current mismatch between stages

could be significant with more stages due to the variation of absorption coefficient. In

contrast, the collection efficiency of photogenerated carriers can be much improved with

thinner individual absorbers and more stages. Also, the carrier lifetime is extracted from

the dark current density to evaluate the material quality. Moreover, the effects of material

quality, current mismatch and collection efficiency on device performance are quantified.

The quantitative analysis shows that the material quality has the most significant impact

on the device performance among the three factors.

Starting from Chapter 6, experimental studies of IC structures for infrared

photodetection are provided. In this chapter, a novel and simple method is developed to

extract the thermal generation rate and minority carrier lifetime in in T2SL-based ICIPs.

This method is more general and can cover various transport mechanisms such as Auger

and SRH processes. Based on this method, the carrier lifetime at high temperatures (200-

340 K) is extracted to be 8.5-167 ns, which turns out to be affected by the material and

structural quality. The exponential temperature dependence of carrier lifetime was

speculated due to the growing dominance of the Auger process at high temperatures. In

addition, in this chapter, fundamental difference in carrier lifetime between IC devices

(ICDs) and quantum cascade devices (QCDs) is apparent from the saturation current

density J0. The extracted values of J0 are more than one order of magnitude lower in ICDs

than in QCDs with similar transition energies. Also, it is shown that J0 can be used as a

common figure of merit to describe cascade structures in terms of the device

180
functionalities. The significance of J0 on detector and PV cell performances was revealed

by the measured detectivity and the estimated open-circuit voltage, respectively.

Chapter 7 attempts a comparative study of four LWIR ICIPs with current-matched

and noncurrent-matched configurations. The cutoff wavelength of these ICIPs is around 11

m at 300 K. It is formally shown that current match is necessary to maximize the

utilization of absorbed photons for optimal responsivity. Also, the reduced responsivity in

noncurrent-matched ICIPs is strongly linked with the light attenuation in the optically

deeper stages. These ICIPs feature a substantial electrical gain, especially for noncurrent-

matched configurations. The significant electrical gain boosts the responsivity in

noncurrent-matched ICIPs, although it is still less than that in the current-matched ICIPs.

This, combined with the large R0A, results in Johnson-noise limited detectivities (>1.4×108

Jones at 300 K) comparable to that in current-matched ICIPs. The values of detectivity in

these LWIR ICIPs are better than that (~4.0×107 Jones) for uncooled state-of-the-art MCT

detectors with similar cutoff wavelengths. Hence, ICIPs can be positioned to be a

prospective candidate for replacing the commercially available MCT detectors in the

LWIR regime.

In Chapter 7, to gain an exhaustive understanding of the observed electrical gain,

three additional LWIR noncurrent-matched ICIPs are studied to allow a possible-in-depth

comparison. The study shows that the electrical gain universally exists in noncurrent-

matched ICIPs. Furthermore, a theory is developed to quantitatively elucidate the electrical

gain in ICIPs. The calculations based on this theory exhibit good agreement with

experimental results. On such a basis, insights and guidance to optimize the Johnson-noise

limited detectivities in noncurrent-matched ICIPs are provided.

181
8.2 Future works

As repeatedly stated in Chapter 3, 4, and 5, in the current phase, the relatively low

conversion efficiency in ICTPV devices is primarily due to the high saturation dark current

density coupled with a short carrier lifetime and narrow bandgap. Hence, grand structural

modifications or/and improvements in material quality are required. Otherwise for the

normal ICTPV structure with current InAs/GaSb SL materials, an attractive energy

efficiency would continue to be an unrealistic goal. The reduced dark current with

increased carrier lifetime will be equally beneficial for detector performance as the

dominating thermal noise is reduced. From this perspective, several means for objectively

reducing the dark current can be employed alone or in combination. For example, to

increase carrier lifetime, one feasible direction to pursue is to replace the InAs/GaSb SL

absorbers with gallium free InAs/InAsSb SLs with a relatively longer carrier lifetime. This

would be somewhat challenging with zero experience in incorporating this type of SL and

IC scheme together. The difficulty also lies in the possible substantial strain released from

the InAsSb layers in the SL.

Alternatively, one can improve the performance from the perspective of raising

photocurrent rather than reducing the dark current. This relies on a special technique to

enhance the light absorption, such as using plasmonic structures for achieving strong light

focusing at a certain wavelength [246-247]. Plasmons can create very strong local fields

around particle and can be guided along the interface in the form of traveling wave, known

as a surface plasmon-polariton. The enhanced absorption can only occur at the plasmonic

resonance wavelength, resulting in extremely narrow response spectra of the integrated IC

devices. For ICTPV cells, this would require an optimal spectral match between the

182
radiation spectrum of the selective emitter (or filter) and the plasmonic resonance.

However, for ICIPs, this feature would restrict them to applications in only very limited

areas.

In addition, other issues in IC devices are not fully resolved at this moment. From

the extracted activation of energy, the SRH process is identified to affect the dark current

in the form of G-R current, whose occurrence can only be in the depletion region. In the

quasi-neutral absorber region in IC structures, the current arising from the SRH process

essentially is still diffusion current. This goes counter to the ideal situation where depletion

regions are fully eliminated in IC devices since no p-n junction exists therein. It would be

meaningful to locate the depletion regions and remove them from IC devices, and

eventually to reduce the dark current. Another not fully appreciated problem is the

significant surface leakage, especially in IC devices with relatively small sizes as discussed

in Chapter 4. The ongoing fabrication research of IC devices is mainly dedicated to

dielectric passivation (SiO2 and SiNx) to improve the surface quality, which however seems

to be less than ideal. Other passivation techniques such as MBE regrowth of a wide-

bandgap semiconductor layer and deposition of a sulfide-based layer can be explored as

well to reduce the dark current. In addition, as raised in Chapter 4, the surface leakage tends

to cause less additional dark current in IC devices with more stages. This conflicts with the

larger resistance with more stages and consequentially more shunting current through the

parallel surface path, which needs to be understood in the future as well.

Another interesting subject of further investigation is the voltage dependent

collection efficiency of photogenerated carriers in ICTPV devices under laser illumination.

All the ICTPV devices in this dissertation have this feature in common. The remaining gap

183
about this subject is to theoretically simulate the collection efficiency while acknowledging

the effect of the applied external voltage. Specifically, one needs to build a reliable

mathematical model that can accurately describe the transport of electrons through the

diffusion process under an electric filed. In addition, future effort needs to be directed

toward explaining the observed exceptionally high collection efficiency for IC devices with

many stages (e.g. the 16- and 23-stage devices in Chapter 5). This result is intuitively not

surprising since more stages consume the applied voltage. However, it also might be that

the model used to extract collection efficiency (Equation 4-2) has limited power in ICTPV

devices with many stages as it is based on two idealized assumptions. Hence, additional

factors need to be considered in the future to improve the model’s performance.

Finally, research into improving the source and spectral shaping technology is

ongoing, but not in the MWIR regime. A good selective emitter that is able to convert the

radiation emitted from a broadband source to a narrow spectral band make the spectral

splitting approach unnecessary. However, there is a lack of effort into the development of

selective emitters whose radiation spectrum would match with the response of ICTPV

devices. Therefore, a reasonable next step in ICTPV research may be to utilize absorbers

with different bandgaps in order to achieve spectral splitting. This can be useful only if the

radiation received by the cell has a broadband spectral distribution. Because there are

already many inherent losses in a TPV system, this may be the most promising path

towards an efficient system.

184
9 References
[1] J. Byrnes, “Unexplored Ordnance Detection and Mitigation,” Springer, 2009.

[2] OxMachina, available: https://medium.com/0xmachina/applying-infrared-


thermography-for-coronavirus-screening-ee5b2dc8a6cb

[3] C. L. DeBellis, M. V. Scotto, L. Fraas, J. Samaras, R. C. Watson, S. W. Scoles,


“Component Development for 500watt Diesel Fueled Portable Thermophotovoltaic
(TPV) Power Supply,” AIP Conference Proceedings, 460, p. 362, 1999.

[4] T. J. Coutts, “A review of progress in thermophotovoltaic generation of electricity”,


Renewable and Sustainable Energy Review, 3, pp. 77-184, 1999.

[5] T. Bauer, “Thermophotovoltaics: Basic Principles and Critical Aspects of System


Design,” Springer, 2011.

[6] J. Zhao, A. Wang, M. A. Green, “19.8% efficient “honeycomb” textured


multicrystalline and 24.4% monocrystalline silicon solar cells,” Applied Physics
Letters, 73, p. 1991, 1998.

[7] J. Nelson, “The Physics of Solar Cells,” Imperial College Press, 2003.

[8] O. Nesher, S. Elkind, A. Adin, I. Nevo, A. B. Yaakov, S. Raichshtain, et al.,


“Digital cooled InSb detector for IR detection,” SPIE Proceedings, 5074, Infrared
Technology and Applications, 2003.

[9] A. Rogalski, “Infrared detectors,” CRC Press Inc, 2010.

[10] N. K. Dhar, R. Dat, A. K. Sood, “Advances in Infrared Detector Array


Technology,” IntechOpen, 2012.

[11] M. A. Khalighi, M. Uysal, “Survey on Free Optical Communication: A


Communication Theory Perspective,” IEEE Communications Surveys & Tutorials,
16, pp. 2231-2258, 2014.

[12] R. E. Nelson, “A brief history of thermophotovoltaic development,” Semiconductor


Science and Technology, 18, p. S141, 2003.

[13] M. Yamaguchi, T. Takamoto, K. Araki, N. Ekins-Daukes, “Multi-junction III-V


solar cells: current status and future potential,” Solar Energy, 79, pp. 78-85, 2005.

[14] R. R. King, D. C. Law, K. M. Edmondson, C. M. Fetzer, G. S. Kinsey, H. Yoon, et


al., “40% efficienct metamorphic GaInP/GaInAs/Ge multijunction solar cells,”
Applied Physics Letters, 90, p. 183516, 2007.

[15] J. E. Raynolds, “Enhanced electro-magnetic energy transfer between a hot and cold
body at close spacing due to evanescent fields,” Proceedings of the 4th NREL

185
conference on thermophotovoltaic generation of electricity, American Institute of
Physics, pp. 49-57, 1999.

[16] M. D. Whale, E. G. Cravalho, “Modeling and Performance of Microscale


Thermophotovoltaic Energy Conversion Devices,” IEEE Transactions on Energy
Conversion, 17, p.130, 2002.

[17] M. Laroche, R. Carminati, J.-J. Greffet, “Near-Field Thermophotovoltaic Energy


Conversion”, Journal of Applied Physics, 100, p. 063704, 2006.

[18] L. M. Fraas, J. E. Avery, H. X. Huang, R. U. Martinelli, “Thermophotovoltaic


system configurations and spectral control,” Semiconductor Science and
Technology, 18, pp. 165–173, 2003.

[19] O. M. Nielsen, L. R. Arana, C. D. Baertsch, K. F. Jensen, M. A. Schmidt,


“Thermophotovoltaic micro-generator for portable power applications,” 12th
International Conferences on Solid-State Sensors, Actuators and Microsystems, 1,
pp. 714–717, 2003.

[20] W. R. Chan, P. Bermel, R. C. N. Podgurski, C. H. Marton, K. F. Jensen, J. J.


Senkevich, et al., “Toward high-energy-density, high-efficiency, and moderate-
temperature chip-scale thermophotovoltaics,” Proceedings of the National
Academy of Sciences of the United States of America, 110, pp. 5309–5314, 2013.

[21] B. Bitnar, W. Durisch, J. C. Mayor, H. Sigg, H .R. Tschudi, G. Palfinger, et al.,


“Record electricity-to-gas power efficiency of a silicon solar cell based TPV
system,” Proceedings of the 5th conference on thermophotovoltaic generation of
electricity, American Institute of Physics, pp 18–28, 2003.

[22] L. Fraas, J. Samaras, J. Avery, L. Minkin, “Antireflection coated refractory metal


matched emitters for use with GaSb thermophotovoltaic generators,” Proceedings
of the 28th IEEE photovoltaic specialists conference, pp 1020–1023, 2000.

[23] B. Wernsman, R. R. Siergiej, S. D. Link, R. G. Mahoeter, M. N. Palmisano, R. J.


Wehrer, et al., “Greater than 20% radiant heat conversion efficiency of a
thermophotovoltaic radiator/module system using reflective spectral control,”
IEEE Transactions of Electronic Devices, 51, pp. 512–515, 2004.

[24] D. Wilt, D. Chubb, D. Wolford, P. Magari, C. Crowley, “Thermophotovoltaics for


space power applications,” Proceedings of the 7th world conference on
thermophotovoltaic generation of electricity, American Institute of Physics, pp
335–345, 2007.

[25] A. Datas, C. Algora, “Development and experimental evaluation of a complete


solar thermophotovoltaic system,” Progress of Photovoltaics: Research and
Applications, 21, pp.1025–1039, 2013.

[26] C. Ungaro, S. Gray, M. Gupta, “Solar thermophotovoltaic system using

186
nanostructures,” Opt. Express, 23, pp. A1149–A1156, 2015.

[27] W. Yang, S. Chou, C. Shu, H. Xue, Z. Li, “Research on micro-thermophotovoltaic


power generators with different emitting materials,” Journal of Micromechanics
and Microengineering. 15, p. S239, 2005.

[28] K. Qiu, A. Hayden, E. Entchev, “TPV power generation system using a high
temperature metal radiant burner,” AIP Conference Proceedings, 890, pp. 27–36,
2007.

[29] M. G. Krishna, M. Rajendran, D. R. Pyke, A. K. Bhattacharya, “Spectral emissivity


of ytterbium oxide-based materials for application as selective emitters in
thermophotovoltaic devices,” Solar Energy Materials and Solar Cells, 59, pp. 337–
348, 1999.

[30] W. Tobler, W. Durisch, “Plasma-spray coated rare-earth oxides on molybdenum


disilicide–high temperature stable emitters for thermophotovoltaics,” Applied
Energy, 85, pp. 371–383, 2008.

[31] S. Molesky, C. J. Dewalt, Z. Jacob, “High temperature epsilon-near-zero and


epsilon-near-pole metamaterial emitters for thermophotovoltaics”, Optical
Express, 21, pp. A96-A110, 2013.

[32] C. Argyropoulos, K. Q. Le, N. Mattiucci, G. D’Aguanno, A. Alù, “Broadband


absorbers and selective emitters based on plasmonic Brewster metasurfaces”,
Physics Review B, 87, p. 205112, 2013.

[33] P. Nagpal, S. E. Han, A. Stein, D. J. Norris, “Efficient Low-Temperature


Thermophotovoltaic Emitters from Metallic Photonic Crystals”, Nano Lett. 8 (10),
pp. 3238–3243, 2008.

[34] R. M. Swanson, “Silicon photovoltaic cells in thermophotovoltaic energy


conversion,” 1978 International Electron Devices Meeting, pp. 70-73, 1978.

[35] Fernández J, Dimroth F, Oliva E, Hermle M, Bett AW, “Back-surface Optimization


of Germanium TPV Cells,” Proceeding of the 7th world conference on
thermophotovoltaic generation of electricity, pp. 190–197, 2007.

[36] O. V. Sulima, A. W. Bett, “Fabrication and simulation of GaSb thermophotovoltaic


cells,” Solar Energy Materials and Solar Cells. 66, pp. 533-540, 2001.

[37] B. Wernsman, R. G. Mahorter, R. Siergiej, S. D. Link, R. J. Wehrer, S. J. Belanger,


P. Fourspring, S. Murray, F. Newman, D. Taylor, T. Rahmlow, “Advanced
thermophotovoltaic devices for space nuclear power systems,” AIP Conference
Proceedings, 746, p. 1441, 2005.

[38] R. S. Tuley, J. M. S. Orr, R. J. Nicholas, D. C. Rogers, P. J. Cannard, S. Dosanjh,


“Lattice-matched InGaAs on InP thermophotovoltaic cells,” Semiconductor

187
Science and Technology, 28, p. 015013, 2012.

[39] M. W. Dashiell, J. F. Beausang, H. Ehsani, G. Nichols, D. M. Depoy, L. R.


Danielson, et al, “Quaternary InGaAsSb Thermophotovoltaic Diodes,” IEEE
Transactions on Electron Devices, 53, p. 2879, 2006.

[40] C. A. Wang, H. K. Choi, S. L. Ransom, G. W. Charache, L. R. Danielson, D. M.


DePoy, “High-quantum-efficiency 0.5 eV GaInAsSb/GaSb thermophotovoltaic
devices,” Applied Physics Letters, 75, p. 1305, 1999.

[41] V. Sundaram, S. Saban, M. Morgan, W. Horne, B. Evans, J. Ketterl, et al., "GaSb


based ternary and quaternary diffused junction devices for TPV applications," AIP
Conference Proceedings, 401, pp. 105-115, 1997.

[42] P. Dutta, J. Borrego, H. Ehsani, G. Rajagopalan, I. Bhat, R. Gutmann, et al., "GaSb


and Ga1-xInxSb thermophotovoltaic cells using diffused junction technology in bulk
substrates," Fifth Conference on Thermophotovoltaic Generation of Electricity, pp.
392-401, 2003.

[43] K. J. Cheetham, P. J. Carrington, N. B. Cook, A. Krier, “Low bandgap GaInAsSbP


Pentanary thermophotovoltaic diodes”, Solar Energy Materials and Solar Cells,
95, pp. 534-537, 2011.

[44] M. Mauk, O. Sulima, J. Cox, R. Mueller, “Low-bandgap (0.3 to 0.5 eV) InAsSbP
thermophotovoltaics: assessment for open-circuit voltage improvements,” 3rd
World Conference on Photovoltaic Energy Conversion, 2003.

[45] A. Krier, M. Yin, A.R.J. Marshall, and S.E. Krier, “Low Bandgap InAs-Based
Thermophotovoltaic Cells for Heat-Electricity Conversion”, Journal of Electronic
Materials, 45, p. 2826, 2016.

[46] Q. Lu, X. Zhou, A. Krysa, A. Marshall, P. Carrington, C-H. Tan, A. Krier, “InAs
thermophotovoltaic cells with high quantum efficiency for waste heat recovery
applications below 1000 °C”, Solar Energy Materials and Solar Cells, 179, p. 334,
2018.

[47] D. Cakiroglu, J. Perez, A. Evirgen, C. Lucchesi, P. Chapuis, T. Taliercio, E.


Tournié, R. Vaillon, “Indium antimonide photovoltaic cells for near-field
thermophotovoltaics,” Solar Energy Materials and Solar Cells, 203, p. 110190,
2019.

[48] X-L. Zhang, A-B. Huang, C-Y. Tian, Y. Wang, Yi-Yi Lou, “Thermophotovoltaic
generation of electricity with InAs0.91Sb0.09 device”, IEEE Transactions on Electron
Devices, 65, p. 4429, 2018.

[49] T. J. Coutts, J. S. Ward, “Thermophotovoltaic and Photovoltaic Conversion at


High-Flux Densities,” IEEE Transactions on Electron Devices, 46, pp. 2145-2153,
1999.

188
[50] P. F. Baldasaro, J. E. Raynolds, G. W. Charache, D. M. DePoy, C. T. Ballinger, T.
Donovan and J. M. Borrego, “Thermodynamic analysis of thermophotovoltaic
efficiency and power density tradeoffs”, Journal of Applied Physics, 89, pp. 3319-
3327, 2001.

[51] G. D. Cody, “Theoretical maximum efficiencies for thermophotovoltaic devices,”


Proceedings of the 4th NREL Conference on Thermophotovoltaic Generation of
Electricity, American Institute of Physics, pp. 58-67, 1999.

[52] L. D. Woolf, “Optimum efficiency of single and multiple bandgap cells in


thermophotovoltaic energy conversion,” Solar Cells, 19, pp. 19–38, 1986.

[53] M. W. Wanlass, K. A. Emery, T. A. Gessert, G. S. Horner, C. R. Osterwald, T. J.


Coutts, “Practical considerations in tandem cell modelling,” Sol Cells, 27, pp. 191–
204, 1989.

[54] A. Caruso, G. Piro, “Theoretical efficiency of realistic solar cells intended for
thermophotovoltaic applications,” Sol Cells, 19, pp. 123–130, 1986

[55] N. P. Harder, P. Würfel, “Theoretical limits of thermophotovoltaic solar energy


conversion.” Semiconductor Science and Technology, 18, pp.151–157, 2003.

[56] R. L. Bell, “Concentration ratio and efficiency in thermophotovoltaics,” Solar


Energy, 23, pp. 203–210, 1979.

[57] V. Badescu, “Thermodynamic theory of thermophotovoltaic solar energy


conversion,” Applied Physics, 90, pp. 6476–6486, 2001.

[58] W. Shockley, H. J. Queisser, “Detailed balance limit of efficiency of p-n junction


solar cells,” Journal of Applied Physics, 32, p. 510, 1961.

[59] T. H. Johnson, “Lead Salt Detectors and Arrays PbS and PbSe,” SPIE Proceedings,
0443, Infrared Detectors, 1983.

[60] J. Werner, M. Oehme, M. Schmid, M. Kaschel, A. Schirmer, E. Kasper, et al.


“Germanium-tin p-i-n photodetectors integrated on silicon grown by molecular
beam epitaxy,” Applied Physics Letters, 98, p. 061108, 2011.

[61] D. N. Hall, R. S. Aikens, R. Joyce, T. W. McCurnin, “Johnson Noise Limited


Operation of Photovoltaic InSb Detectors,” Applied Optics, 14, pp. 450-453, 1975.

[62] F. Shepherd, A. Yang, “Silicon Schottky Retinas for Infrared Imaging,” IEDM
Technical Digest, pp. 310–13, 1973.

[63] A. Rogalski, “HgCdTe infrared detector material: history, status and outlook,”
Reports on Progress in Physics, 68, p. 2267, 2005.

[64] A. Lacaita, F. Zappa, S. Cova, P. Lovati, “Single-photon detection beyond 1 m:

189
performance of commercially available InGaAs/InP detectors,” Applied Optics, 35,
pp. 2986-2996, 1996.

[65] H. Schneider, H. C. Liu, “Quantum Well Infrared Photodetectors,” Springer, 2007.

[66] H. C. Liu, M. Gao, J. McCaffrey, Z. R. Wasilewski, S. Fafard, “Quantum dot


infrared photodetectors,” Applied Physics Letters, 78, p. 79, 2000.

[67] S. Maimon and G. W. Wicks, “nBn detector, an infrared detector with reduced dark
current and higher operating temperature,” Applied Physics Letters, 89, p. 151109,
2006.

[68] D. L. Smith and C. Mailhiot, “Proposal for strained type II superlattice infrared
detectors,” Journal of Applied Physics, 62, p. 2545, 1987.

[69] HITRAN Database, available: https://www.cfa.harvard.edu/hitran/

[70] J. Faist, F. Capasso, D. L. Sivco, C. Sirtori, A. L. Hutchinson, A. Y. Cho,


“Quantum Cascade Laser,” Science, 264, pp. 553-556, 1994.

[71] J. Faist, “Quantum Cascade Lasers,” Oxford University Press, 2013.

[72] R. Q. Yang, “Interband Cascade (IC) Lasers,” in Semiconductor lasers:


Fundamentals and applications. A. Baranov. E. Tournie (EDs). Woodhead
Publishing, Chap. 12, 2013.

[73] I. Vurgaftman, R. Weih, M. Kamp, J. R. Meyer, C. L. Canedy, C. S. Kim, M. Kim,


W. W. Bewley, C. D. Merritt, J. Abell, S. Höfling, “Interband cascade lasers,”
Journal of Physics D: Applied Physics, 48, p. 123001, 2015.

[74] E. L. Dereniak, G. D. Boreman, “Infrared Detectors and Systems,” Wiley, 1996.

[75] P. W. Kruse, “Uncooled Thermal Imaging: Arrays, Systems, and Applications,”


SPIE Press, 2001.

[76] A. Rogalski, “History of infrared detectors,” Opto-Electronics Review, 20, pp. 270-
308, 2012.

[77] Hope-Wish, available: http://hpthermalcamera.com/cooled-vs-uncooled-thermal-


cameras-long-range-surveillance/

[78] B. F. Levine, “Quantum-well infrared photodetectors,” Journal of Applied Physics,


74, p. R1, 1993.

[79] R. Q. Yang, J. M. Xu, M. Sweeny, “Selection rules of intersubband transitions in


conduction-band quantum wells,” Physics Review B, 50, p. 7474, 1994.

[80] A. Yoshizawa, H. Tsuchida, “A 1550 nm Single-Photon Detector Using a


Thermoelectrically Cooled InGaAs Avalanche Photodiode,” Japanese Journal of

190
Applied Physics, 40, p. 200, 2001.

[81] P. Klipstein, “XBn barrier photodetectors for sensitivity and high operating
temperature infrared sensors,” SPIE proceedings, 6940, p. 69402U, 2008.

[82] D. Z.-Y. Ting, C. J. Hill, A. Soibel, S. A. Keo, J. M. Mumolo, J. Nguyen, et al., “A


high-performance long wavelength superlattice complementary barrier infrared
detector”, Appllied Physics Letter, 95, p. 023508, 2009.

[83] F. R. Giorgetta, E. Baumann, M. Graf, Q. Yang, C. Manz, K. Köhler, H. E. Beere,


D. A. Ritchie, E. Linfield, A. G. Davies, et al., “Quantum Cascade Detectors,” IEEE
Journal of Quantum Electronics, 45, p. 1039, 2009.

[84] P. Reininger, B. Schwarz, H. Detz, D. MacFarland, T. Zederbauer, A. M. Andrews,


W. Schrenk, O. Baumgartner, H. Kosina and G. Strasser, “Diagonal-transition
quantum cascade detector,” Applied Physics Letters, 105, p. 091108, 2014.

[85] J. F. Klem, J. K. Kim, M. J. Cich, G. A. Keeler, S. D. Hawkins, T. R. Fortune,


“Mesa-isolated InGaAs photodetectors with low dark current,” Applied Physics
Letters, 95, p. 031112, 2009.

[86] J. Masek, A. Ishida, H. Zogg, C. Maissen, S. Blunier, “Monolithic Photovoltaic


PbS-on-Si Infrared-Sensor Array,” IEEE Electron Device Letters, 11, pp. 12-14,
1990.

[87] V. Kasiyan, Z. Dashevsky, C. M. Schwarz, M. Shatkhin, E. Flitsiyan, L. Chernyak,


et al., “Infrared detectors based on semiconductor p-n junction of PbSe,” Applied
Physics Letters, 112, p. 086101, 2012.

[88] R-M. Lin, S-F. Tang, S-C. Lee, C-H. Kuan, G-S. Chen, T-P. Sun, et al., “Room
Temperature Unpassivated InAsp-i-n Photodetectors Grown by Molecular Beam
Epitaxy,” IEEE Transactions on Electron Devices, 44, pp. 209-213, 1997.

[89] A. G. Foyt, W. T. Lindley, J. P. Donnelly, “n-p junction photodetectors in InSb


fabricated by proton bombardment,” Applied Physics Letter, 16, p. 335, 1970.

[90] H. Shao, W. Li, A. Torfi, D. Moscicka, W. I. Wang, “Room-Temperature InAsSb


Photovoltaic Detectors for Mid-Infrared Applications,” IEEE Photonics
Technology Letters, 18, pp. 1756-1758, 2006.

[91] Catalogue, Vigo System S.A.

[92] A. M. Hoang, G. Chen, A. Haddadi, S. Abdollahi Pour, M. Razeghi,


“Demonstration of shortwavelength infrared photodiodes based on type-II
InAs/GaSb/AlSb superlattices,” Applied Physics Letters, 100, p. 211101, 2012.

[93] S. A. Pour, E. K. Huang, G. Chen, A. Haddadi, B.-M. Nguyen, M. Razeghi, “High


operating temperature midwave infrared photodiodes and focal plane arrays based

191
on type-II InAs/GaSb superlattices,” Applied Physics Letters, 98, p. 143501, 2011.

[94] A. Haddadi, G. Chen, R. Chevallier, A. Hoang, and M. Razeghi, “InAs/InAs1-xSbx


type-II superlattices for high performance long wavelength infrared detection,”
Applied Physics Letters, 105, p. 121104, 2014.

[95] A. Haddadi, G. Chen, R. Chevallier, A. M. Hoang, and M. Razeghi,


“InAs/InAs1−xSbx type-II superlattices for high performance long wavelength
infrared detection,” Applied Physics Letter, 105, p. 121104, 2014.

[96] T. Dougakiuchi, K. Fujita, T. Hirohata, A. Ito, M. Hitaka and T. Edamura, “High


photoresponse in room temperature quantum cascade detector based on coupled
quantum well design,” Applied Physics Letter, 109, p. 261107, 2016; Erratum, 110,
p. 109902, 2017.

[97] P. Reininger, B. Schwarz, H. Detz, D. MacFarland, T. Zederbauer, A. M. Andrews,


W. Schrenk, O. Baumgartner, H. Kosina and G. Strasser, “Diagonal-transition
quantum cascade detector,” Applied Physics Letter, 105, p. 091108, 2014.

[98] R. Q. Yang, “Infrared laser based on intersubband transitions in quantum wells,”


Superlattices and Microstructures, 17, pp. 77-83, 1995.

[99] R. Q. Yang, Z. Tian, Z. Cai, J. F. Klem, M. B. Johnson, H. C. Liu, “Interband-


cascade infrared photodetectors with superlattice absorbers,” Journal of Applied
Physics, 107, p. 054514, 2010.

[100] R. Q. Yang, Z. Tian, J. Klem, T. D. Mishima, M. B. Santos, M. B. Johnson,


“Interband cascade photovoltaic devices”, Applied Physics Letters, 96, p. 063504,
2010.

[101] R. Q. Yang, J. M. Xu, “Bound and quasibound states in leaky quantum wells,”
Physics Review B, 46, p. 6969, 1992.

[102] R. Q. Yang, J. M. Xu, “Leaky quantum wells: a basic theory and application,”
Canadian Journal of Physics, 70, p. 1153, 1992.

[103] L. Esaki, L. L. Chang, E. E. Mendez, “Polytype superlattices and multi-


heterojunctions”, Japanese Journal of Applied Physics, 20, p. L529, 1981.

[104] B. M. Nguyen, D. Hoffman, P. Y. Delaunay, M. Razeghi, “Dark current


suppression in type II InAs/GaSb superlattice long wavelength infrared
photodiodes with M-structure barrier”, Applied Physics Letters, 91, p. 163511,
2007.

[105] I. Vurgaftman, E. H. Aifer, C. L. Canedy, J. G. Tischler, J. R. Meyer, J. H. Warner,


et al., “Graded band gap for dark-current suppression in long-wavelength infrared
W-structure type-II superlattice photodiodes,” Applied Physics Letters, 89, p.
121114, 2006.

192
[106] O. Salihoglu, A. Muti, K. Kutluer, T. Tansel, R. Turan, Y. Ergun, A. Aydinli, “N”
structure for type-II superlattice photodetectors, Applied Physics Letters, 101, p.
073505, 2012.

[107] L. F. Luo, R. Beresford, W. I. Wang, “Resonant tunneling in AlSb/InAs/AlSb


double-barrier heterostructures,” Applied Physics Letters, 53, p. 2320, 1988.

[108] E. R. Brown, J. R. Söderström, C. D. Parker, L. J. Mahoney, K. M. Molvar, T. C.


McGill, “Oscillations up to 712 GHz in InAs/AlSb resonant-tunneling diodes,”
Applied Physics Letters, 58, p. 2291, 1991.

[109] J. Devenson, R. Teissier, O. Cathabard, A. N. Baranov, “InAs-based quantum


cascade lasers,” SPIE Proceedings, 6909, p. 69090U, 2008.

[110] C. Cathabard, R. Teissier, J. Devenson, J. C. Moreno, A. N. Baranov, “Quantum


cascade lasers emitting near 2.6 m,” Applied Physics Letters, 96, p. 141110, 2010.

[111] D. Z. Garbuzov, H. Lee, V. Khalfin, R. Martinelli, J. C. Connolly, G. L. Belenky,


“2.3-2.7-m room temperature CW operation of InGaAsSb-AlGaAsSb broad
waveguide SCH-QW diode lasers,” IEEE Photonics Technology Letters, 11, pp.
794-796, 1999.

[112] J. G. Kim, L. Shterengas, R. U. Martinelli, G. L. Belenky, D. Z. Garbuzov, W. K.


Chan, “Room-temperature 2.5 m InGaAsSb/AlGaAsSb diode lasers emitting 1 W
continuous waves,” Applied Physics Letters, 81, p. 3146, 2002.

[113] D. L. Partin, “Lead salt quantum well diode lasers,” Superlattices and
Microstructures, 1, pp. 131-135, 1985.

[114] Z. Shi, M. Tacke, A. Lambrecht, H. Böttner, “Midinfrared lead salt multi-quantum-


well diode lasers with 282 K operation,” Applied Physics Letters, 66, p. 2537, 1995.

[115] R. Q. Yang, L. Li, L. Zhao, Y. Jiang, Z. Tina, H. Ye, et al., “Recent progress in
development of InAs-based interband cascade lasers,” SPIE Proceedings, 8640, p.
86400Q, 2013.

[116] J. Scheuermann, R. Weih, M. von Edlinger, L. Nähle, M. Fischer, J. Koeth, et al.,


“Single-mode interband cascade lasers emitting below 2.8 m,” Applied Physics
Letters, 106, p. 161103, 2015.

[117] M. Kim, C. L. Caney, W. W. Bewley, C. S. Kim, J. R. Lindle, J. Abell, et al.,


“Interband cascade laser emitting at =3.75 m in continuous wave above room
temperature,” Applied Physics Letters, 92, p. 191110, 2008.

[118] W. W. Bewley, C. L. Canedy, C. S. Kim, M. Kim, C. D. Merritt, J. Abell, et al.,


“Continuous-wave interband cascade lasers operating above room temperature at λ
= 4.7-5.6 μm,” Optics Express, 20, pp. 3235-3240, 2012.

193
[119] L. Li, Y. Jiang, R. Q. Yang, T. D. Mishima, M. B. Santos, M. B, Johnson, “Low-
threshold InAs-based interband cascade lasers operating at high temperatures,”
Applied Physics Letters, 106, p. 251102, 2015.

[120] S. M. Rassel, L. Li, Y. Li, R. Q. Yang, J. A. Gupta, X. Wu, et al., “High temperature
and low-threshold interband cascade lasers at wavelengths longer than 6 m,”
Optical Engineering, 57, p. 011021, 2017.

[121] S. Höfling, R. Weih, M. Dallner, J. Scheuermann, M. V. Edlinger, L. Nähle, et al.,


“Mid-Infrared (∼2.8 m to ∼7.1 m) interband cascade lasers,” SPIE Proceedings,
9550, p. 95599F, 2015.

[122] Z. Tian, R. Q. Yang, T. D. Mishima, M. B. Santos, M. B. Johnson, “Plasmon-


waveguide interband cascade lasers near 7.5 m,” IEEE Photonics Technology
Letter, 21, pp. 1588–190, 2009.

[123] Z. Tian, L. Li, Y. Hao, R. Q. Yang, T. D. Mishima, M. B. Santos, and M. B.


Johnson, “InAs-based interband cascade lasers with emission wavelength at 10.4
µm,” Electronics Letters, 48, p. 113, 2012.

[124] R. Q. Yang, L. Li, W. Huang, S. M. Rassel, J. A. Gupta, X. Wu, et al., “InAs-Based


Interband Cascade Lasers,” IEEE Journal of Selected Topics in Quantum
Electronics, 25, p. 1200108, 2019.

[125] R. Q. Yang, C. J. Hill, K. Mansour, Y. Qiu, A. Soibel, R. E. Muller, et al.,


“Distributed feedback mid-infrared interband cascade lasers at thermoelectric
cooler temperatures,” IEEE Journal of Selected Topics in Quantum Electronics, 13,
pp. 1074–1078, 2007.

[126] I. Vurgaftman, W. W. Bewley, C. L. Canedy, C. S. Kim, M. Kim, C. D. Merritt, et


al., “Rebalancing of internally generated carriers for mid-infrared interband
cascade lasers with very low power consumption,” Nature Communications, 2, p.
585, 2011.

[127] G. A. Sai-Halasz, R. Tsu, L. Esaki, “A new semiconductor superlattice,” Applied


Physics Letters, 30, p. 651, 1977.

[128] E. R. Youngdale, J. R. Meyer, C. A. Hoffman, F. J. Bartoli, C. H. Grein, P. M.


Young, et al., “Auger lifetime enhancement in InAs-Ga1-xInxSb superlattices,”
Applied Physics Letter, 64, p. 3160, 1994.

[129] H. Mohseni, V. I. Litvinov, M. Razeghi, “Interface-induced suppression of the


Auger recombination in type-II INAs/GaSb superlattices,” Physics Review B, 58,
p. 15378, 1998.

[130] D. R. Rhiger, “Performance comparison of long-wavelength infrared type II


superlattice devices with HgCdTe,” Journal of Electronic Materials, 40, pp. 1815–
1822, 2011.

194
[131] C. M. Ciesla, B. N. Murdin, C. R. Pidgeon, R. A. Stradling, C. C. Phillips, M.
Livingstone, et al., “Suppression of Auger recombination in arsenic-rich InAs-
1xSbx,” Journal of Applied Physics, 80, p. 2994, 1996.

[132] B. Tang, Y. Q. Xu, Z. Q. Zhou, R. T. Hao, G. W. Wang, Z. W. Ren, and Z. C. Niu,


“GaAs Based InAs/GaSb Superlattice Short Wavelength Infrared Detectors Grown
by Molecular Beam Epitaxy,”Chinese Physics Letters, 26, p. 028102, 2009.

[133] A. M. Hoang, G. Chen, A. Haddadi, S. A. Pour, and M. Razeghi, “Demonstration


of shortwavelength photodiodes based on type-II InAs/GaSb/AlSb superlattices,”
Applied Physics Letter, 100, p. 211101, 2012.

[134] B.-M. Nguyen, M. Razeghi, V. Nathan, G. J. Brown, “Type-II M structure


photodiodes: an alternative material design for mi-wave to long wavelength
infrared regions,” SPIE Proceedings, 6479, p. 64790S, 2007.

[135] B. M. Nguyen, D. Hoffman, P. Y. Delaunay, E. K. Huang, M. Razeghi, J.


Pellegrino, “Band edge tunability of M-structure for heterojunction design in Sb
based type II superlattice photodiodes,” Applied Physics Letter, 93, p. 163502,
2008.

[136] J. V. Li, R. Q. Yang, C. J. Hill, S. L. Chuang, “Interband cascade detectors with


room temperature photovoltaic operation,” Applied Physics Letter, 86, p. 101102,
2005.

[137] Z. Tian, R. T. Hinkey, R. Q. Yang, D. Lubyshev, Y. Qiu, J. M. Fastenau, et al.,


“Interband cascade infrared photodetectors with enhanced electron barriers and p-
type superlattice absorbers,” Journal of Applied Physics, 111, p. 024510, 2012.

[138] V. Andreev, V. Khvostikov, V. Kalinovsky, V. Lantratov, V. Grilikhes, V.


Rumyantsev, et al., “High current density GaAs and GaSb photovoltaic cells for
laser power beaming,” 3rd World Conference on Photovoltaic Energy Conversion,
pp. 761-764, 2003.

[139] E. Oliva, F. Dimroth, A. W. Bett, “GaAs converter for high power densities of laser
illumination,” Progress in Photovoltaic: Research and Applications, 16, pp. 289-
295, 2008.

[140] C. Donolato, “A reciprocity theorem for charge collection,” Applied Physics Letter,
46, pp. 270-272, 1985.

[141] R. T. Hinkey, R. Q. Yang, “Theory of Multiple-Stage Interband Photovoltaic


Devices and Ultimate Performance Limit Comparison of Multiple-Stage and
Single-Stage Interband Infrared Detectors,” Journal of Applied Physics, 114, p.
104506, 2013.

[142] H. Lotfi, L. Lei, L. Li, R. Q. Yang, J. C. Keay, M. B. Johnson, et al., “High-


temperature operation of interband cascade infrared photodetectors with cutoff

195
wavelengths near 8 μm,” Optical Engineering, 54, p. 063103, 2015.

[143] B. Satpati, J. B. Rodriguez, A. Trampert, E. Tournié, A. Joullié, P. Christol,


“Interface analysis of InAs/GaSb superlattice grown by MBE,” Journal of Crystal
Growth, 301-302, pp. 889-892, 2007.

[144] Z.-B Tian, E. A. Plis, R. T. Hinkey, S. Krishna, “Influence of composition in


InAs/GaSb type-II superlattices on their optical properties,” IEEE Electronics
Letters, 50, pp. 1733-1734, 2014.

[145] Y. Livneh, P. C. Klipstein, O. Klin, N. Snapi, S. Grossman, A. Glozman, et al., “k.p


model for the energy dispersions and absorption spectra of InAs/GaSb type-II
superlattices,” Physics Review B, 86, p. 235311, 2012.

[146] I. Vurgaftman, G. Belenky, Y. Lin, D. Donetsky, L. Shterengas, G. Kipshidze, et


al., “Interband absorption strength in long-wave infrared type-II superlattices with
small and large superlattice periods compared to bulk materials,” Applied Physics
Letters, 108, p. 222101, 2016.

[147] H. Mohseni, E. Michel, J. Sandoen, M. Razeghi, “Growth and characterization of


InAs/GaSb photoconductors for long wavelength infrared range,” Applied Physics
Letters, 71, p. 1403, 1997.

[148] R. Rehm, M. Walther, J. Schmitz, F. Rutz, J. Fleibner, R. Scheibner, et al.,


“InAs/GaSb superlattices for advanced infrared focal plane arrays,” Infrared
Physics and Technology, 52, pp. 344-347, 2009.

[149] Y. Wei, A. Hood, H. Yau, A. Gin, M. Razeghi, M. Z. Tidrow, et al., “Uncooled


operation of type-II InAs/GaSb superlattice photodiodes in the midwavelength
infrared range”, Appl. Phys. Lett. 86, p. 233106, 2005.

[150] J. V. Li, C. J. Hill, J. Mumolo, S. Gunapala, S. Mou, and S-L. Chuang, “Midinfrared
type-II InAs/GaSb superlattice photodiodes toward room temperature operation”
Applied Physics Letters, 93, p. 163505, 2008.

[151] L. Lei, L. Li, H. Lotfi, H. Ye, R. Q. Yang, T. D. Mishima, et al., “Mid-wavelength


interband cascade infrared photodetectors with superlattice absorbers and gain”,
Optical Engineering, 57, p. 011006, 2018.

[152] R. Rehm, M. Walther, J. Schmitz, F. Rutz, J. Fleibner, R. Scheibner, et al.,


“InAs/GaSb superlattices for advanced infrared focal plane arrays,” Infrared
Physics and Technology, 52, pp. 344-347, 2009.

[153] H. Lotfi, L. Li, L. Lei, H. Ye, S. M. S. Rassel, Y. Jiang et al., “High-frequency


operation of a mid-infrared interband cascade system at room
temperature,” Applied Physics Letters, 108, p. 201101, 2016.

[154] Y. Chen, X. Chai, Z. Xie, Z. Deng, N. Zhang, Y. Zhou, “High Speed Mid-Infrared

196
Interband Cascade Photodetector Based on InAs/GaSb Type-II Superlattice”, IEEE
Journal of Lightwave Technology, 38, pp. 939-945, 2019.

[155] A. Y. Cho, J. R. Arthur, “Molecular beam epitaxy,” Progress in Solid State


Chemistry, 10, pp. 157-191, 1975.

[156] M. A. Herman, H. Sitter, “Molecular Beam Epitaxy: Fundamentals and Current


Status,” Springer, 2012.

[157] H. Ye, H. Lotfi, L. Li, R. T. Hinkey, R. Q. Yang, L. Lei, et al., “Multi-stage


interband cascade photovoltaic devices with a bandgap of 0.23 eV operating above
room temperature,” Chinese Science Bulletin, 59, pp. 950-955, 2014.

[158] W. Huang, L. Li, L. Lei, Jeremy A. Massengale, H. Ye, Rui Q. Yang, et al.,
“Minority carrier lifetime in mid-wavelength interband cascade infrared
photodetectors,” Applied Physics Letters, 112, p. 251107, 2018.

[159] W. Huang, L Lei, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, et al.,


“Enhanced collection efficiencies and performance of interband cascade structures
for narrow bandgap semiconductor thermophotovoltaic devices,” Journal of
Applied Physics, 124, p. 023101, 2018.

[160] P. Asbeck, “Self-absorption effects on the radiative lifetime in GaAs–GaAlAs


double heterostructures,” Journal of Applied Physics, 48, p. 820, 1977.

[161] A. Marti, J. L. Balenzategui, R. F. Reyna, “Photon recycling and Schockley’s diode


equation,” Journal of Applied Physics, 82, p. 4067, 1997.

[162] W. van Roosbroeke, W. Shockley, “Photon-radiative recombination of electrons


and holes in germanium,” Physical Review, 94, p. 1558, 1954.

[163] E. Blandre, P-O. Chapuis, R. Vaillon, “High-injection effects in near-filed


thermophotovoltaic devices,” Scientific Report, 7, p. 15860, 2017.

[164] M. A. Green, “Limiting Photovoltaic Monochromatic Light Conversion


Efficiency,” Progress in Photovoltaic: Research and Applications, 9, p. 257, 2001.

[165] L. Lei, L. Li, H. Lotfi, Y. Jiang, R. Q. Yang, M. B. Johnson, “Mid-wave interband


cascade infrared photodetectors banded on GaInAsSb absorbers,” Semiconductor
Science and Technology, 31, p. 105014, 2016.

[166] E. H. Steenbergen, B. C. Connelly, G. D. Metcalfe, H. Shen, M. Wraback, D.


Lubyshev, et al., “Significantly improved minority carrier lifetime observed in a
long-wavelength infrared III-V type-II superlattice cmprised of InAs/InAsSb,”
Applied Physics Letters, 99, p. 251110, 2011.

[167] B. V. Olson, E. a. Shaner, J. K. Kim, J. F. Klem, S. D. Hawkins, L. M. Murray, et


al., “Time-resovled optical measurements of minority carrier recombination in a

197
mid-wave infrared InAsSb alloy and InAs/InAsSb superlattice,” Applied Physics
Letters, 101, p. 092109, 2012.

[168] H. Ye, L. Li, H. Lotfi, L. Lei, R. Q. Yang, J. C. Keay, et al., “Molecular beam
epitaxy of interband cascade structures with InAs/GaSb supperlattice absorbers for
long-wavelength infrared detection,” Semiconductor Science and Technology, 30,
p. 105029, 2015.

[169] R. T. Hinkey, R. Q. Yang, “Theoretical comparison of performance of limits of


single- and multiple-stage photovoltaic devices,” Semiconductor Science and
Technology, 30, p. 015013, 2015.

[170] W. Huang, R. Q. Yang, “Limiting factor and efficiencies of narrow bandgap single-
absorber and multi-stage interband cascade thermophotovoltaic cells under
monochromatic light illumination,” Journal of Applied Physics, 126, p. 045714,
2019.

[171] S. Hegedus, D. Desai, C. Thompson, “Voltage dependent photocurrent collection


in CdTe/CdS solar cells,” Progress of Photovoltaics: Research and Applications,
15, pp. 587-602, 2017.

[172] R. Crandall, “Modeling of thin film solar cells: uniform field approximation,”
Journal of Applied Physics, 54, pp. 7176-7186, 1983.

[173] X. X. Liu, J. R. Sites, “Solar-cell collection efficiency and its variation with
voltage,” Journal of Applied Physics, 75, pp. 577-581, 1994.

[174] S. S. Hegedus, “Current-voltage analysis of a-Si and a-SiGe solar cells including
voltage-dependent photocurrent collection,” Progress of Photovoltaics: Research
and Applications, 5, pp. 151-168, 1997.

[175] D. A. Fardig, J. E. Phillips, “Characterization of CdTe/CdS Solar Cells,”


Proceedings of 22nd IEEE Photovoltaic Specialists Conference, pp. 1146-1150,
1991.

[176] S. M. Arnab, M. Z. Kabir, “Modeling of the effects of charge transport on voltage-


dependent photocurrent in ultrathin CdTe solar cells,” Journal of Vacuum Science
and Technology A, 31, p. 061201, 2013.

[177] M. Eron, A. Rothwarf, “Effects of a voltage-dependent light-generated current on


solar cell measurements: CuInSe2/Cd(Zn)S,” Applied Physics Letters, 44, pp. 131-
133, 1984.

[178] J. Phillips, J. Titus, D. Hoffman, “Determining the voltage dependence of the light
generated current in CuInSe2-based solar cells using I-V measurements made at
different light intensities,” Proceedings of 26th IEEE Photovoltaic Specialists
Conference, pp. 463-466, 1997.

198
[179] R. Q. Yang, C. J. Hill, L. E. Christensen, C. R. Webster, “Mid-IR type-II interband
cascade lasers and their applications,” SPIE Proceedings, 5624, pp. 413-422, 2005.

[180] R. Häcker, A. Hangleiter, “Intrinsic upper limits of the carrier lifetime in silicon,”
Journal of Applied Physics, 75, p. 7570, 1994.

[181] A. Cuevas, D. Macdonald, “Measuring and interpreting the lifetime of silicon


wafers,” Solar Energy, 76, pp. 255-262, 2004.

[182] H. Lotfi, L. Li, L. Lei, R. Q. Yang, J. F. Klem, M. B. Johnson, et al., “Narrow-


bandgap interband cascade thermophotovoltaic cells,” IEEE Journal of
Photovoltaics, 7, pp. 1462-1468, 2017.

[183] M. Kinch, “State-of-the-Art Infraed Detector Technology,” SPIE Press, 2014.

[184] E. A. Plis, M. N. Kutty, S. Krishna, “Passivation tehcniques for InAs/GaSb strained


layer superlatccies detectors,” Laser & Photonics Reviews, 7, p. 45, 2013.

[185] H. K. Chung, M. A. Rosenberg, P. H. Zimmermann, “Origin of 1/f noise observed


in Hg0.7Cd0.3Te variable area photodiode arrays,” Journal of Vacuum Science and
Technology A, 3, p. 189, 1985.

[186] R. Q. Yang, H. Lotfi, L. Li, R. T. Hinkey, H. Ye, J. F. Klem, et al., “Quantum-


engineered interband cascade photovoltaic devices,” SPIE Proceedings, 8993, p.
899310, 2014.

[187] W. Huang, L Lei, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, at al.,


“Enhanced collection efficiencies and performance of interband cascade structures
for narrow bandgap semiconductor thermophotovoltaic devices,” Journal of
Applied Physics, 124, p. 023101, 2018.

[188] W. Huang, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, M. B. Santos,


“Investigation of Narrow Bandgap Interband Cascade Thermophotovoltaic Cells,”
SPIE Proceedings, 10913, p. 1091317, 2019.

[189] R. T. Hinkey, Z. Tian, S. M. Rassel, R. Q. Yang, J. F. Klem, M. B. Johnson,


“Interband cascade photovoltaic devices for conversion of mid-infrared radiation,”
IEEE Journal of Photovoltaics, 3, p. 745, 2013.

[190] H. Lotfi, R. T. Hinkey, L. Li, R. Q. Yang, L. Lei, M. B. Johnson, et al., “Narrow-


bandgap photovoltaic devices operating at room temperature and above with high
open-circuit voltage,” Applied Physics Letters, 102, p. 211103, 2013.

[191] W. Huang, J. A. Massengale, Y. Lin, L. Li, R. Q. Yang, T. D. Mishima, M. B.


Santos, “Performance analysis of narrow-bandgap interband cascade
thermophotovoltaic cells,” Journal of Physics D: Applied Physics, 53, p. 175104,
2020.

199
[192] B. C. Connelly, G. D. Metcalfe, H. Shen, M. Wraback, “Direct minority carrier
lifetime measurements and recombination mechanism in long-wave infrared type-
II superlattices using time-resolved photoluminescence,” Applied Physics Letters,
97, p. 251117, 2010.

[193] D. Wang, D. Donetsky, S. Jung, G. Belenky, “Carrier lifetime measurements in


long-wave infrared InAs/GaSb superlattices under low excitation conditions”,
Journal of Electronic Materials, 41, pp. 3027–3030, 2012.

[194] W. Huang, L. Li, L. Lei, J. A. Massengale, R. Q. Yang, T. D. Mishima, et al.,


“Electrical gain in interband cascade infrared photodetectors”, Journal of Applied
Physics, 123, p. 113104, 2018.

[195] M. M. Alkaisi, N. A. Aldawody, “Factors affecting the hot spot efficacy in


photovoltaic arrays,” Solar Cells, 28, pp. 11-17, 1990.

[196] D. Roche, H. Outhred, R. J. Kaye, “Analysis and Control of Mismatch Power Loss
in Photovoltaic Arrays,” Progress in Photovoltaic: Research and Applications, 3,
pp. 115-127, 1995.

[197] W. Huang, S. Rassel, L. Li, J. Massengale, R. Q. Yang, T. D. Mishima, et al., “A


Unified Figure of Merit for Interband and Intersubband Cascade Devices,” Infrared
Physics and Technology, 96, p. 298, 2019.

[198] Z. Tian, S. E. Godoy, H. S. Kim, T. Schuler-Sandy, J. A. Montoya, S. Krishna,


“High operating temperature interband cascade focal plane arrays,” Applied
Physics Letters, 105, p. 051109, 2014.

[199] N. Gautam, S. Myers, A.V. Barve, B. Klein, E.P. Smith, D.R. Rhiger, L.R. Dawson,
and S. Krishna, “High operating temperature interband cascade midwave infrared
detector based on type-II InAs/GaSb strained layer superlattice,” Applied Physics
Letters, 101, p. 021106, 2012.

[200] D. Donetsky, S. P. Svensson, L. E. Vorobjev, G. Belenky, “Carrier lifetime


measurements in short-period InAs/GaSb strained-layer superlattice structures,”
Applied Physics Letters, 95, p. 212104, 2009.

[201] D. Donetsky, G. Belenky, S. Svensson, S. Suchalkin, “Minority carrier lifetime in


type-2 InAs-GaSb strained-layer superlattices and bulk HgCdTe materials,”
Applied Physics Letters, 97, p. 052108, 2010.

[202] L. M. Murray, K. S. Lokovic, B. V. Olson, A. Yildirim, T. F. Boggess, J. P. Prineas,


“Effects of growth rate variations on carrier lifetime and interface structure in
InAs/GaSb superlattices,” Journal of Crystal Growth, 386, p. 194, 2014.

[203] B. Klein, N. Gautam, E. Plis, T. Schuler-Sandy, T. J. Rotter, Sanjay Krishna,


“Carrier lifetime studies in midwave infrared type-II InAs/GaSb strained layer
superlattice,” Journal of Vacuum Science & Technology B, 32, p. 02C101, 2014.

200
[204] S. Bandara, P. Maloney, N. Baril, J. Pellegrino, M. Tidrow, “Doping dependence
of minority carrier lifetimes in long-wave Sb-based type-II superlattice infrared
detector materials,” Optical Engineering, 50, p. 061015, 2011.

[205] D. R. Rhiger, “Performance of Comparison of Long-Wavelength Infrared Type II


superlattice Devices with HgCeTe,” Journal of Electronic Materials, 40, p. 1815,
2011.

[206] Q. K. Yang, C. Pfahler, J. Schmitz, W. Pletschen, F. Fuchs, “Trap cneters and


minority carrier lifetimes in InAs/(GaIn)Sb superlattice long wavelength
photodetectors,” SPIE Proceedings, 4999, p. 448, 2003.

[207] J. Pellegrino, R. DeWames, “Minority carrier lifetime characteristics in type II


InAs/GaSb LWIR superlattice n+πp+ photodiodes,” SPIE Proceedings, 7298, p.
72981U, 2009.

[208] M. Delmas, J. B. Rodriguez, P. Christol, “Electrical modeling of InAs/GaSb


superlattice mid-wavelength infrared pin photodiode to analyze experimental dark
current characteristics,” Journal of Applied Physics, 116, p. 113101, 2014.

[209] C. Cervera, K. Jaworowicz, H. Aït-Kaci, R. Chaghi, J.B. Rodriguez, I. Ribet-


Mohamed, et al., “Temperature dependence performances of InAs/GaSb
superlattice photodiode,” Infrared Physics and Technology, 54, p. 258, 2011.

[210] D. K.Schroder, “The Concept of Generation and Recombination Lifetimes In


Semiconductors,” IEEE Transactions on Electron Devices, 29, p. 1336, 1982.

[211] J. Yin, R. Paiella, “Multiple-junction quantum cascade photodetectors for


thermophotovoltaic energy conversion,” Opt. Express, 18, p. 1618, 2010.

[212] J. Yin, R. Paiella, “Limiting performance analysis of cascaded


interband/intersubband thermophotovoltaic devices,” Applied Physics Letters, 98,
p. 041103, 2011.

[213] M.A. Kinch, F. Aqariden, D. Chandra, P-L Liao, H.F. Schaake, H.D. Shih,
“Minority Carrier Lifetime in p-HgCdTe,” Journal of Electronic Materials, 34, p.
880, 2005.

[214] A. Rogalski, P. Martyniuk, M. Kopytko, “InAs/GaSb type-II superlattice infrared


detectors: Future prospect,” Applied Physics Reviews, 4, p. 031304, 2017.

[215] G. Marre, B. Vinter, V Berger, “Strategy for the design of a non-cryogenic quantum
infrared detector,” Semiconductor Science and Technology, 18, p. 284, 2003.

[216] E. O. Kane, “Semiconductors & Semimetals,” chapter 3, vol. 1, New York:


Academic, 1966.

[217] K. Hackiewicz, P. Martyniuk, J. Rutkowski, A. Kowalewski, “Calculation of dark

201
current in interband cascade type-II infrared InAs/GaSb superlattice detector,” Acta
Physica Polonica A, 132, p. 1415, 2017.

[218] R. K. Ahrenkiel, “Semiconductors and Semimetals,” R. K. Ahrenkiel and M. S.


Lundstrom (ETs), 39, pp. 39–150, New York: Academic, 1993.

[219] B. V. Olson, E. A. Shaner, J. K. Kim, J. F. Klem, S. D. Hawkins, M. E. Flatté, et


al., “Identification of dominant recombination mechanism in narrow-bandgap
InAs/InAsSb type-II superlattices and InAsSb alloys,” Applied Physics Letters,
103, p. 052106, 2013.

[220] L. Hőglund, D. Z. Ting, A. Khoshakhlagh, A. Soibel, C. J. Hill, A. Fisher, et al.,


“Influenence of radiative and non-radiative recombination on the minority carrier
lifetime in the midewave infrared InAs/InAsSb supperlatices,” Applied Physics
Letters, 103, p. 221908, 2013.

[221] I. Vurgaftman, W. W. Bewley, C. D. Merritt, C. L. Canedy, M. V. Warren, C. S.


Kim, et al., “Sensitive Chemical Detection with Distributed Feedback Interband
Cascade Lasers”, Encyclopedia of Analytical Chemistry, John Wiley & Sons, Ltd.,
2016.

[222] S.-Q. Zhai, J.-Q. Liu, K. Ning, F.-Q. L, L. Li, L.-J. Wang, et al., “Strin-composed
InP-based InGaAs/InAsAs Quantum Cascade Infrared Detectors for 3~5 m
Atmospheric Window,” SPIE Proceedings, 8193, p. 81931X, 2011.

[223] X. Wang, J. Liu, S. Zhai, F. Liu, Z. Wang, “Room temperature quantum cascade
detector operating at 4.3 m,” Journal of Semiconductors, 35, p. 104009, 2014.

[224] P. Reininger, T. Zederbauer, B. Schwarz, H. Detz, D. MacFarland, A. M. Andrews,


et al., “InAs/AlAsSb based quantum cascade detector,” Applied Physics Letters,
107, p. 081107, 2015.

[225] T. Dougakiuchi, K. Fujita, T. Hirohata, A. Ito, M. Hitaka, T. Edamura, “High


photoresponse in room temperature quantum cascade detector based on coupled
quantum well design,” Applied Physics Letters, 109, p. 261107, 2016; Erratum,
110, p. 109902, 2017.

[226] Y. Jiang, L. Li, R. Q. Yang, J. A. Gupta, G. C. Aers, E. Dupont, et al., “Type-I


interband cascade lasers near 3.2 m,” Applied Physics Letters, 106, p. 041117,
2015.

[227] Y. Li, L. Li, W. Huang, R. Q. Yang, J. A. Gupta, X. Wu, “Low-threshold InAs-


based interband cascade lasers near 6.3 m,” 14th International Conference on
Mid-IR Optoelectronics: Materials and Devices MIOMD-XIV, 2018.

[228] J. R. Meyer, C. A. Hoffman, F. J. Bartoli, L. R. Ram-Mohan, “Type-II quantum-


well lasers for the mid-wavelength infrared,” Applied Physics Letters, 67, p. 757,
1995.

202
[229] Z.-B. Tian, T. Schuler-Sandy, S. Krishna, “Electron barrier study of mid-wave
infrared interband cascade photodetectors,” Applied Physics Letters, 103, p.
083501, 2013.

[230] H. Lotfi, L. Lei, L. Li, R. Q. Yang, J. C. Keay, M. B. Johnson et al., “High-


temperature operation of interband cascade infrared photodetectors with cutoff
wavelength near 8 m”, Optical Engineering, 54, p. 063103, 2015.

[231] H. Lotfi, L. Li, L. Lei, Y. Jiang, R.Q. Yang, J.F. Klem, et al., "Short-wavelength
interband cascade infrared photodetectors operating above room temperature",
Journal of Applied Physics, 119, p. 023105, 2016.

[232] W. Huang, L. Lei, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, et al.,


“Current-matching versus non-current-matching in long wavelength interband
cascade infrared photodetectors,” Journal of Applied Physics, 122, p. 083102,
2017.

[233] C. Koeniguer, G. Dubois, A. Gomez, V. Berger, “Electronic trnasport in quantum


cascade strucutures at equilibrium,” Physics Review B, 74, p. 235325, 2006.

[234] A. Buffaz, A. Gomez, M. Carras, L. Doyennette, V. Berger, “Role of subband


occupancy on electronic transport in quantum cascade detectors,” Physics Review
B, 81, p. 075304, 2010.

[235] A. Delga, L. Doyennette, M. Carras, V. Trinite, P. Bois, “Johnson and shot noises
in intersubband detectors,” Applied Physics Letters, 102, p. 163507, 2013.

[236] D. Palaferri, Y. Todorov, A. Bigioli, A. Mottaghizadeh, D. Gacemi, A. Calabrese,


et al., “Room-temperature nine-m-wavelength photodetctors and GHz-frequency
heterodyne receivers,” Nature, 556, pp. 85-88, 2018.

[237] H. Lotfi, L. Li, H. Ye, R. T. Hinkey, L. Lei, R. Q. Yang, et al., “Interband cascade
infrared photodetectors with long and very-long cutoff wavelengths,” Infrared
Physics and Technology, 70, p. 162, 2015.

[238] Y. Zhou, J. Chen, Z. Xu, L. He, “High quantum efficiency mid-wavelength


interband cascade infrared photodetectors with one and two stages,” Semiconductor
Science and Technology, 31, p. 085005, 2016.

[239] L. Lei, L. Li, H. Ye, H. Lotfi, R. Q. Yang, M. B. Johnson, et al., “Long wavelength
interband cascade infrared photodetectors operating at high temperatures,” Journal
of Applied Physics, 120, p. 193102, 2016.

[240] W. Pusz, A. Kowalewski, P. Martyniuk, W. Gawron, E. Plis, S. Krishna, et al.,


“Mid-wavelength infrared type-II InAs/GaSb superlattice interband cascade
photodetectors,” Optical Engineering, 53, pp. 043107-043107, 2014.

[241] Z. Tian, S. Krishna, “Mid-Infrared Interband Cascade Photodetectors With

203
Different Absorber Designs,” IEEE Journal of Quantum Electronics, 51, pp. 1-5,
2015.

[242] A. Soibel, D. Z. Ting, C. J. Hill, M. Lee, J. Nguyen, S. A. Keo, et al., “Gain and
noise of high-performance long wavelength superlattice infrared detectors,”
Applied Physics Letters, 96, p. 111102, 2010.

[243] M. B. Reine, A. K. Sood, T. J. Tredwell, “Semiconductors and Semimetals,” R. K.


Willardson and A. C. Beer (ETs), Vol. 18, p. 201, Academic: New York, 1981.

[244] W. E. Tennant, “”Rule 07” Revisited: Still a good Heuristic Predictor of p/n
HgCdTe Photodiode Performance?,” Journal of Electronic Materials, 39, p. 1030,
2010.

[245] Klaudia Hackiewicz, Jaroslaw Rutkowski, Piotr Martyrium, Tetina Manyk,


“Calculation of optimal absorber thickness in interband cascade type-II infrared
InAs/GaSb super lattice photodetectors,” SPIE Proceedings, 10433, p. 104330Z,
2017.

[246] S. A. Maier, “Plasmonics: fundamentals and applications,” Springer, 2007.

[247] R. Stanley, “Plasmonics in the mid-infrared,” Nature photonics, 6, pp. 409-411,


2012.

204
10 Appendix A: Publications list

Refereed journal articles


[1] W. Huang, J. A. Massengale, Y. Lin, L. Li, R. Q. Yang, T. D. Mishima, M. B.
Santos, “Performance analysis of narrow-bandgap interband cascade
thermophotovoltaic cells”, Journal of Physics D: Applied Physics 53,175104
(2020).

[2] W. Huang, R. Q. Yang, “Limiting Factors and Efficiencies of Narrow Bandgap


Thermophotovoltaic Cells under Monochromatic Light Illumination”, Journal of
Applied Physics 126, 045714 (2019)

[3] W. Huang, S. M. Rassel, L. Li, J. A. Massengale, H. Ye, R. Q. Yang, T. D.


Mishima, M. B. Santos, “A unified figure of merit for interband and intersubband
cascade devices”, Infrared Physics and Technol. 96, 298 (2019)

[4] W. Huang, L. Lei, L. Li, J. A. Massengale, H. Ye, R. Q. Yang, T. D. Mishima, M.


B. Santos, “Enhanced collection efficiencies and performances of interband
cascade structures for narrow bandgap semiconductor thermophotovoltaic
devices”, Journal of Applied Physics 124, 023101 (2018)

[5] W. Huang, L. Li, L. Lei, J. A. Massengale, R. Q. Yang, T. D. Mishima, M. B.


Santos, “Electrical gain in interband cascade infrared photodetectors”, Journal of
Applied Physics 123, 113104 (2018)

[6] W. Huang, L. Li, L. Lei, J. A. Massengale, H. Ye, R. Q. Yang, T. D. Mishima, M.


B. Santos, “Minority carrier lifetime in mid-wavelength interband cascade infrared
photodetectors”, Applied Physics Letters 112, 251107 (2018)

[7] W. Huang, L. Lei, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, M. B.


Santos, “Current-matching versus non-current-matching in long wavelength
interband cascade infrared photodetectors”, Journal of Applied Physics 122,
083102 (2017)

[8] Y. Lin, J. A. Massengale, W. Huang, R. Q. Yang, T. D. Mishima, M. B. Santos,


“Examination of the durability of interband cascade lasers against structural
variations”, Journal of Infrared and Millimeter Waves, 39, 137-141, (2020)

[9] R. Q. Yang, L. Li, W. Huang, S. M. Rassel, J. A. Gupta, A. Bezinger, X. Wu, S.


G. Razavipour, G. C. Aers, “InAs-based Interband Cascade Lasers”, IEEE Journal
of Selected Topics in Quantum Electronics, 25, 1200108 (2019)

[10] L. Lei, W. Huang, J. A. Massengale, H. Ye, H. Lotfi, R. Q. Yang, T. D. Mishima,


M. B. Santos, M. B. Johnson, “Resonant tunneling and multiple differential
conductance features in long wavelength interband cascade infrared
photodetectors”, Applied Physics Letters, 111, 113504 (2017)

205
Conference presentations and proceedings

[1] W. Huang, R. Q. Yang, “Efficiencies and limiting factors of narrow bandgap


thermophotovoltaic cells”, talk 11275-36 at Physics, Simulation, and Photonic
Engineering of Photovoltaic Devices IX at SPIE. Photonics West, San Francisco,
California, Feb. 1-6, 2020

[2] W. Huang, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, M. B. Santos,


“Multistage Interband Cascade Thermophotovoltaic Devices with ~0.2 eV
Bandgap”, at Photovoltaic Specialists Conference, Chicago, Illinois, June16-21,
2019

[3] W. Huang, L. Li, J. A. Massengale, R. Q. Yang, T. D. Mishima, M. B. Santos,


“Investigation of narrow bandgap interband cascade thermophotovoltaic cells”,
talk 10913-42 at Physics, Simulation, and Photonic Engineering of Photovoltaic
Devices VIII at Photonics West, San Francisco, CA, Feb. 2-7, 2019 (in Proc. SPIE.
10913, 1091317)

[4] J. A. Gupta, X. Wu, G. C. Aers, Y. Li, L. Li, W. Huang, R. Q. Yang, “Low-


threshold InAs-based interband cascade lasers with room-temperature emission at
6.3 μm” (Invited), talk 10939-33 at Novel In-Plane Semiconductor Lasers XVIII
at Photonics West, San Francisco, CA, Feb. 2-7, 2019

[5] W. Huang, S. Rassel, L. Li, J. Massengale, Y. Li, R. Q. Yang, T. D. Mishima, M.


B. Santos, “A Unified Figure of Merit for Interband and Intersubband Cascade
Devices”, at 14th International Conference on Mid-IR Optoelectronics: Materials
and Devices MIOMD-XIV (MIOMD 2018), Flagstaff, AZ, Oct. 7-10, 2018

[6] W. Huang, L. Lin, L. Li, J. Massengale, R. Q. Yang, T. D. Mishima, M. B. Santos,


“Collection Efficiency and Device Performance in Narrow Bandgap
Thermophotovoltaic Cells Based on Interband Cascade Structures”, at 14th
International Conference on Mid-IR Optoelectronics: Materials and Devices
MIOMD-XIV (MIOMD 2018), Flagstaff, AZ, Oct. 7-10, 2018

[7] Y. Li, L. Li, W. Huang, R. Q. Yang, J. A. Gupta, X. Wu, G. Aers, “Low-threshold


InAs-based Interband Cascade Lasers near 6.3 μm”, at 14th International
Conference on Mid-IR Optoelectronics: Materials and Devices MIOMD-XIV
(MIOMD 2018), Flagstaff, AZ, Oct. 7-10, 2018

[8] J. A. Gupta, A. Bezinger, S.G. Razavipour, X. Wu, G. C. Aers, Y. Li, L. Li, W.


Huang, R. Q. Yang, “Long-Wavelength InAs-based Interband Cascade Lasers
Grown by MBE”, paper TuM6 at 34th North American Conference on Molecular
Beam Epitaxy (NAMBE 2018), Banff, Canada, Sept. 30-Oct. 5, 2018

[9] W. Huang, L. Li, L. Lei, J. A. Massengale, H. Ye, R. Q. Yang, T. D. Mishima, M.


B. Santos, “Carrier Lifetime in Mid-Infrared Type-II Superlattice Photodetectors”,

206
at 31th Annual Conference of the IEEE Photonics Society, Reston, VA, Sept. 30-
Oct. 4, 2018

[10] R. Q. Yang, W. Huang, L. Li, L. Lei, J. A. Massengale, T. D. Mishima, and M. B.


Santos, “Gain and resonant tunneling in interband cascade IR photodetectors”
(Invited), talk 10543-13 at Quantum Sensing and Nanophotonic Devices XV at
Photonics West, San Francisco, CA, Jan. 27- Feb.1, 2018 (in Proc. SPIE. 10540,
105400E)

207

You might also like