Advanced Design of Ultrathin-Barrier AlN/GaN HEMTs

198
ADVANCED DESIGN OF ULTRA-THIN BARRIER AlN/GaN HEMTS; A STUDY OF DEVICE DESIGN, MODELING, AND ANALYSIS A Dissertation Submitted to the Graduate School of the University of Notre Dame in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy by David A. Deen, Huili Grace Xing, Director Graduate Program in Electrical Engineering Notre Dame, Indiana April 2011

Transcript of Advanced Design of Ultrathin-Barrier AlN/GaN HEMTs

ADVANCED DESIGN OF ULTRA-THIN BARRIER AlN/GaN HEMTS; A

STUDY OF DEVICE DESIGN, MODELING, AND ANALYSIS

A Dissertation

Submitted to the Graduate School

of the University of Notre Dame

in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

by

David A. Deen,

Huili Grace Xing, Director

Graduate Program in Electrical Engineering

Notre Dame, Indiana

April 2011

c© Copyright by

David A. Deen

All Rights Reserved

ADVANCED DESIGN OF ULTRA-THIN BARRIER AlN/GaN HEMTS; A

STUDY OF DEVICE DESIGN, MODELING, AND ANALYSIS

Abstract

by

David A. Deen

Of the III-Nitride family the AlN/GaN heterojunction has demonstrated the

largest combined polarization charge and energy band offsets available in the sys-

tem. Engineering the polarization fields through varying the AlN thickness leads

to two-dimensional electron gas densities (2DEGs) that may be tailored between

0.5 - 5 × 1013 cm−2. Furthermore, the ultra-thin (< 5 nm) barrier and excellent

transport properties of this all binary heterostructure make it well suited for high

electron mobility transistor applications where high frequency and high current

are required. This work encompasses various design aspects of GaN-based High

Electron Mobility Transistors (HEMTs) which ultimately result in the realization

of several generations that utilize the AlN/GaN heterostructure.

HEMTs fabricated from high-mobility, low sheet resistance heterostructures

have achieved drain current densities up to 2.3 A/mm and transconductance of

480 mS/mm, which set new benchmarks for GaN-based HEMTs. Ultra-thin pre-

metallization etching has been employed for the first time to reduce ohmic contact

resistance for AlN/GaN HEMTs and has enabled small signal frequency perfor-

mance in excess of 100 GHz. Moll’s method for delay time extraction has been uti-

lized to extract an effective electron velocity in the intrinsic region of the AlN/GaN

HEMT and was found to be ∼ 1.2 × 107 cm/s.

David A. Deen

By leveraging the allowable thickness window of the AlN barrier along with the

high density 2DEGs that result, several novel HEMT devices have been designed

and realized. High Al-content AlxGa1−xN back barriers have been employed for

improved 2DEG confinement in several new variations of the ultra-thin AlN/GaN

HEMT. A dual, parallel-channel AlN/GaN-based HEMT structure is designed and

realized for the first time as an epitaxial approach to mitigating DC-RF frequency

dispersion. These structures emphasize the facilitation of new device designs that

are made possible through the particular qualities the AlN/GaN heterostructure

possesses.

“... some things worth the tellin’

and you’d be right in guessin’

that each and every lesson,

they were hard won”

- Ray LaMontagne

ii

CONTENTS

FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi

TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

CHAPTER 1: INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . 11.1 III-Nitride HEMT applications . . . . . . . . . . . . . . . . . . . . 11.2 Polarization physics of the III-Nitride semiconductor system . . . 41.3 Background on the AlN/GaN heterostructure . . . . . . . . . . . 51.4 Scope of this work . . . . . . . . . . . . . . . . . . . . . . . . . . 101.5 Chapter 1 cited literature . . . . . . . . . . . . . . . . . . . . . . 11

CHAPTER 2: TRANSISTOR DESIGN CONSIDERATIONS . . . . . . . 132.1 The AlN/GaN HEMT structure . . . . . . . . . . . . . . . . . . . 142.2 Design considerations of the AlN/GaN HEMT . . . . . . . . . . . 17

2.2.1 Ohmic source-drain contacts . . . . . . . . . . . . . . . . . 182.2.2 Gate current suppression . . . . . . . . . . . . . . . . . . . 252.2.3 Gate capacitance model . . . . . . . . . . . . . . . . . . . 29

2.3 Chapter 2 cited literature . . . . . . . . . . . . . . . . . . . . . . 37

CHAPTER 3: THE SINGLE-HETEROJUNCTION AlN/GaN HEMT . . 413.1 The AlN/GaN HEMT . . . . . . . . . . . . . . . . . . . . . . . . 41

3.1.1 A comparative study on the thickness of the AlN barrier ondevice performance . . . . . . . . . . . . . . . . . . . . . . 42

3.1.2 AlN/GaN HEMTs grown by MBE on MOCVD GaN/SiCtemplates . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3.1.3 Ta2O5 insulated gate AlN/GaN HEMT . . . . . . . . . . . 513.1.4 First-generation insulated-gate AlN/GaN HEMTs . . . . . 56

3.2 Implications of peripheral resistance on HEMT intrinsic I-V char-acteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

iii

3.2.1 Ideal three-terminal current-voltage model . . . . . . . . . 603.2.2 Velocity saturation . . . . . . . . . . . . . . . . . . . . . . 633.2.3 External resistance . . . . . . . . . . . . . . . . . . . . . . 673.2.4 Experimental correspondence to model and its validity . . 69

3.3 Chapter 3 cited literature . . . . . . . . . . . . . . . . . . . . . . 76

CHAPTER 4: AlN/GaN HEMT SMALL-SIGNAL OPERATION . . . . . 804.1 Small signal model . . . . . . . . . . . . . . . . . . . . . . . . . . 814.2 De-embedding procedure . . . . . . . . . . . . . . . . . . . . . . . 884.3 Delay analysis of the insulated-gate AlN/GaN HEMT . . . . . . . 904.4 Points for improvement for RF operation of the AlN/GaN HEMT 1014.5 Chapter 4 cited literature . . . . . . . . . . . . . . . . . . . . . . 102

CHAPTER 5: HOT CARRIER EFFECTS IN SixNy-PASSIVATED Al-GaN/GaN HEMTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1055.1 Test Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1065.2 Experimental Procedure: Chynoweth model . . . . . . . . . . . . 1075.3 Discussion of charge dynamics . . . . . . . . . . . . . . . . . . . . 1105.4 Chapter 5 cited literature . . . . . . . . . . . . . . . . . . . . . . 113

CHAPTER 6: NOVEL HEMT DESIGNS UTILIZING THE AlN/GaN HET-EROSTRUCTURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1156.1 Polarization-engineered AlxGaN back barriers . . . . . . . . . . . 1166.2 Design of a dual-channel charge-screening AlN/GaN HEMT; an

epitaxial approach to mitigate frequency dispersion . . . . . . . . 1256.2.1 Multi-channel heterostructure design . . . . . . . . . . . . 1286.2.2 A dual-channel AlN/GaN HEMT . . . . . . . . . . . . . . 133

6.3 Chapter 6 cited literature . . . . . . . . . . . . . . . . . . . . . . 141

CHAPTER 7: SUMMARY AND FUTURE WORK . . . . . . . . . . . . . 1447.1 Summary of this work . . . . . . . . . . . . . . . . . . . . . . . . 1447.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1477.3 Chapter 7 cited literature . . . . . . . . . . . . . . . . . . . . . . 151

APPENDIX A: THE HETEROJUNCTION CAPACITOR;TERMAN’S METHOD . . . . . . . . . . . . . . . . . . . . . . . . . . 152A.1 Capacitance-Voltage Relationship . . . . . . . . . . . . . . . . . . 153

A.1.1 Ideal capacitive structure (Qit = 0) . . . . . . . . . . . . . 155A.1.2 Capacitive structure including interface traps (Qit 6= 0) . . 157A.1.3 Quantum capacitance . . . . . . . . . . . . . . . . . . . . . 160

A.2 Terman’s method for Dit determination . . . . . . . . . . . . . . . 161A.3 Discussion of Terman methodology . . . . . . . . . . . . . . . . . 163

iv

A.4 Appendix A cited literature . . . . . . . . . . . . . . . . . . . . . 165

APPENDIX B: DERIVATION OF CURRENT-VOLTAGE CHARACTER-ISTICS OF THE AlN/GaN HEMT BY GRADUAL CHANNEL AP-PROXIMATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166B.1 Linear I-V functionality . . . . . . . . . . . . . . . . . . . . . . . 166B.2 Transconductance . . . . . . . . . . . . . . . . . . . . . . . . . . . 172B.3 Current saturation . . . . . . . . . . . . . . . . . . . . . . . . . . 174B.4 External resistance . . . . . . . . . . . . . . . . . . . . . . . . . . 175B.5 Validity of the GCA . . . . . . . . . . . . . . . . . . . . . . . . . 176B.6 Appendix B cited literature . . . . . . . . . . . . . . . . . . . . . 177

APPENDIX C: GENERAL AlN/GaN HEMT PROCESS . . . . . . . . . . 178

v

FIGURES

1.1 Energy gap versus lattice spacing for the various semiconductorsystems. Wavelength corresponding to the energy gap can be seenon the right with the color scale for the visible spectrum plottedagainst the various material systems. . . . . . . . . . . . . . . . . 2

1.2 Example of a standard Al0.3GaN/GaN HEMT’s drain characteris-tics with specific qualities pointed out that help define the HEMT’sapplication. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 AlN/GaN heterojunction illustration showing the correspondingpolarization dipoles (a) and charge distributions (b) along with theenergy band diagram (c). . . . . . . . . . . . . . . . . . . . . . . . 6

1.4 2×2 µm atomic force micrographs of the surface of the AlN/GaNHEMTs for various AlN thicknesses. (images courtesy of Yu Cao,U. of Notre Dame) . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.5 2DEG properties of the AlN/GaN heterostructure. Images takenfrom reference [8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 Conduction band diagram for the insulator/AlN/GaN HEMT show-ing the charge distribution and the choice of electron wave function. 15

2.2 Current-voltage characteristics of the Zimmermann contacts takenfrom reference [15]. I-V characteristics of as-deposited contactsacross a 3 µm separation TLM structure for several different AlNthicknesses (left). Current level dependence on anneal temperaturefor the Zimmermann contact (right). . . . . . . . . . . . . . . . . 21

2.3 Current-voltage characteristics for pre-metallization etched contactsto AlN/GaN. Pre-etch to AlN/GaN (left) and the removal of theGaN cap on the GaN/AlN/GaN heterostructure (right). . . . . . 22

2.4 (left) Split wafer comparison of the effect of pre-etched contact ondrain characteristics. (right) Function of AlN thickness on contactresistance for various metal schemes attempted including the pre-metallization etched contacts. . . . . . . . . . . . . . . . . . . . . 23

vi

2.5 Effects of ALD and PECVD dielectrics on the AlN/GaN channelcharge characteristics µ and ns (left). Gated IV comparison of thedielectrics studied, taken on 100 µm diameter CV dots (right). . . 27

2.6 Modeled C-V dependence on various system parameters (left), andcalculated threshold voltage relationship dependance on the gatedielectric constant for varying AlN thicknesses (right). . . . . . . . 31

2.7 Modeled and measured C-V curves (bottom row) with the corre-sponding Dit spectrum (top row) for the insulated-gate AlN/GaNcapacitors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.8 HR-TEM image of a GaN/AlN/GaN structure with an ALD Al2O3

layer deposited on the surface. Image courtesy of D. Smith, ArizonaState University. . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3.1 2 × 2 µm AFM scans of varied AlN barrier layers. . . . . . . . . . 43

3.2 Room temperature sheet resistance determined from Hall effectmeasurement for varied AlN barrier thickness. . . . . . . . . . . . 44

3.3 Drain current characteristics for varied AlN barrier thicknesses. . 46

3.4 Transfer characteristics for varied AlN barrier thicknesses (left),and two-probe I-V measurements of insulated C-V dots for the var-ious AlN barrier thicknesses (right). . . . . . . . . . . . . . . . . . 46

3.5 Measured sheet resistance (a), 2D charge density (b), mobility (c),contact resistance (d), maximum drain current density (e), andtransconductance (f) as a function of AlN barrier thickness. . . . 48

3.6 DC transfer (a) and drain (b) characteristics (left) and small signalfrequency characteristics (right) of the AlN/GaN HEMT grown onGaN/SiC templates. . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.7 Drain characteristics of the Ta2O5/AlN/GaN HEMT (destructivebreakdown occurs at 96 V for VGS = -7 V). . . . . . . . . . . . . 53

3.8 (left) Drain and transfer characteristics and, (right) small signalfrequency performance of the Al2O3/AlN/GaN HEMT grown onGaN/SiC templates. . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.9 Drain (left) and transfer (right) characteristics of the AlN/GaNHEMT grown and fabricated at U. of Notre Dame. . . . . . . . . 58

3.10 AlN/GaN HEMT I-V characteristics showing the effect when twodifferent mobilities are considered. . . . . . . . . . . . . . . . . . . 62

3.11 Electron velocity versus electric field for various common semicon-ducting systems. The scattering limited velocity approximation isillustrated for the GaN curve. . . . . . . . . . . . . . . . . . . . . 64

vii

3.12 AlN/GaN HEMT I-V characteristics showing the effect when ve-locity saturation is included. . . . . . . . . . . . . . . . . . . . . . 66

3.13 GCA model results of imposing external source and drain resis-tance on drain I-V characteristics (internal source and drain volt-ages shown with respect to the applied external VDS on the right). 68

3.14 Progression of the AlN/GaN I-V characteristics for the intrinsicdevice (left), the addition of the VLSA (middle), and the final ad-dition of external resistance. . . . . . . . . . . . . . . . . . . . . . 70

3.15 GCA model fit to data (a) taken from Fig. 3.3 showing the low-fielddeviation due to the SLVA (b). GCA fit to transfer characteristics(c). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.16 Qualitative illustration of the conduction band diagram along thechannel showing the Dit-influenced channel choke leading to IDScompression at forward gate-source bias voltages. (a) Cross sec-tional conduction band diagram in “flatband” or pinched off condi-tion and, (b) band diagram of the forward gate bias condition withQit influence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.1 Transfer (left) and drain (right) characteristics of the Al2O3/AlN/GaNHEMT used for small signal model development. . . . . . . . . . . 82

4.2 Measured (circles) and modeled (lines) S-parameter data for boththe ColdFET (a) and maximum-gm (b) biasing regimes. . . . . . . 83

4.3 Schematic diagram of the small signal equivalent circuit model (a)showing the separation of the intrinsic, parasitic, and bond pad cir-cuits, “ColdFET” equivalent circuit model (b), and intrinsic circuitmodel (a) over-layed on a cross-sectional depiction of the insulatedgate HEMT. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

4.4 Flow diagram for the Y-subtraction method for parameter de-embeddingof a HEMT. Toggling the measured S-parameters between Y andZ parameters allows convenient reduction of pad and/or parasiticcomponent influence. . . . . . . . . . . . . . . . . . . . . . . . . . 89

4.5 Small signal frequency performance as measured (gray circles), mod-eled (line), and after de-embedding (blue circles) bonding pad in-ductance and capacitance. . . . . . . . . . . . . . . . . . . . . . . 90

4.6 Transfer (left) and drain (right) characteristics of the Ta2O5/AlN/GaNHEMT used for the delay analysis. . . . . . . . . . . . . . . . . . 92

4.7 Unity current gain, ft, as a function of VDS (left) and IDS (right). 95

4.8 Total delay as a function of VDS (left) and IDS (right) showing thedelay for the corresponding measurement. . . . . . . . . . . . . . 96

viii

4.9 Comparison of delay times between the AlGaN/GaN and AlN/GaNHEMTs extracted by Moll’s method. . . . . . . . . . . . . . . . . 98

5.1 Device cross section showing the Schottky and four MIS gates (a)with the corresponding band diagram along the channel direction(b) showing the impact ionization process and the relation to theKink effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

5.2 Transfer characteristics (left) of all gates on the HEMT prior to andafter bias-stress. Drain characteristics (right) of the AlGaN/GaNHEMT comparing the Schottky gate controlled (line) vs. the MISgate 2 controlled (open circles) family of curves. A clear kink inthe MIS gate controlled characteristics can be seen at ∼9 volts. . 109

5.3 (left) Gate current analysis for MIS gate 2 biased at VGS= -4 V.Linear character of the Ihole/IDS ratio indicates impact ionizationgenerated hole current. The inset shows the measured gate currentfor gate voltage steps of -2 V. (right) Time decay of the thresholdvoltage shift. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.1 Layer structure and band diagram in equilibrium of the AlGaN/AlN/GaNcontrol sample. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.2 Transfer (left) and drain (right) characteristics of the AlGaN/AlN/GaNHEMT control sample (without a back barrier). . . . . . . . . . . 119

6.3 (left) Layer structures, (center) band diagrams, and (right) 2×2µm2 AFM scans of the AlN surface for the AlGaN back barrierstructures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.4 (left) Calculated electron charge density from 1D Poisson Schrodingersolver, (middle) measured C-V data of the back barrier structures,and (right) calculated charge distribution from C-V data. . . . . . 122

6.5 drain (left) and transfer (right) characteristics comparison of backbarrier HEMTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.6 Virtual gate model showing the time plot of the surface charge andchannel depletion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.7 Cross section and corresponding band diagrams for the design ofthe charge screening HEMT. . . . . . . . . . . . . . . . . . . . . . 129

ix

6.8 Threshold voltage comparison. The top horizontal curve is for afixed charge density of the lower channel with a fixed 3 nm bar-rier and is for comparison to the access region Vth. The set of 3lower curves show the calculated threshold voltage for the accessregion with the 2nd 2DEG distribution for screening. Access regionthreshold voltage becomes increasingly more negative for increasedupper 2DEG density and this increase in amplified for thicker GaNspacer layers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

6.9 (right) 2×2 µm AFM scan of the AlN surface of the dual-channelstructure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

6.10 C-V analysis of the Al2O3/AlN/GaN/AlN/GaN heterostructure ca-pacitor (left) and the associated approximate charge profile (right). 135

6.11 (left) Transfer characteristics of both the dual-channel and charge-screening HEMT, (middle) drain characteristics of the charge-screeningHEMT, and (right) drain characteristics of the dual-channel HEMT.137

6.12 Small signal frequency performance of the charge screening HEMT(left) and the dual channel HEMT (right). . . . . . . . . . . . . . 138

6.13 Time profiles of the pulsed-gate lag measurements for (a) the dualchannel HEMT, (b) the charge-screening HEMT, and (c) a standardsingle heterojunction AlN/GaN HEMT for comparison. . . . . . . 140

A.1 Band diagram including interface trap states. . . . . . . . . . . . 155

B.1 Illustration of the active region in a HEMT where the GCA applies.External resistances are included to depict the quasi-circuit underconsideration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

B.2 Band diagram of the insulated gate AlN/GaN HEMT showing per-tinent parameters used for the GCA. . . . . . . . . . . . . . . . . 169

x

TABLES

2.1 PARAMETERS AFFECTING OPERATION OF THE AlN/GaNHEMT FOR THE DIELECTRICS CONSIDERED IN THIS WORK[40], [41]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.2 EXTRACTED VALUES OF INTERFACE TRAP DENSITY FORTHE DIELECTRICS EXPLORED IN THIS STUDY. . . . . . . . 36

4.1 EXTRACTED SMALL SIGNAL PARAMETERS FOR THE MOD-ELED INSULATED-GATE AlN/GaN HEMT . . . . . . . . . . . 86

4.2 EXTRACTED TIME DELAYS FOR THE Ta2O5 INSULATED-GATE AlN/GaN HEMT AND SCHOTTKY-GATED AlGaN/GaNHEMT. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

xi

ACKNOWLEDGMENTS

The path I have followed during this work has been all but ordinary and I have

enjoyed the fellowship of many instructors and colleagues along the way. It is my

pleasure to acknowledge those individuals who have been influential to me during

my graduate tenure.

I want to thank my advisor, Huili Grace Xing for her continued effort to push

me to be a better researcher. Her support in my decision to continue my graduate

work while at the Naval Research Laboratory allowed me to pursue my own life

path while also expanding my technical knowledge and exposure to new topics.

My gratitude is owed to her for such support.

I want to thank David Storm for his technical support with the MBE growths

he performed for me as well as our almost daily conversations over morning coffee.

It was a pleasure to have both a colleague and a friend at NRL. Patrick Fay with

his expertise in high frequency measurements and device physics was an integral

part to my technical development.

One of the most influential people I have had the privilege to learn from and

work with is DJ. His intrinsic joy for understanding the physical nature of things

overflows to where one cannot help but share his sentiment when around him.

I am indebted to my cohorts at the Naval Research Laboratory to which I owe

many thanks. Steve Binari taught by example the value of experience and careful

consideration. I am grateful to David Meyer and Mario Ancona for our many

xii

discussions and friendship. I want to thank Bob Bass for his persistence with the

demands that come with electron beam lithography along with Scott Katzer and

Chip Eddy for their help with growth. A great deal of my gratitude is owed to

James Champlain; an artistic yet analytical minded fellow, mentor, and friend.

James served as a great resource of technical knowledge and always had endless

patience for explanation.

I am grateful toward my Notre Dame colleagues for their multi-dimensional

influence that never lacked to motivate. I want to thank Yu Cao for his help with

sample growth. I want to thank Gouwang, Yong, and Zongyang for their help

with growth and device processing assistance. I am ever grateful to have such

close friends in Robin Joyce and Joey Herzog, who are also fellow witnesses to

my graduate experience. I feel my time during graduate school would have lacked

richness and color had these two individuals not been part of it.

I want to thank my colleagues at the other institutions I had the privilege of

working with including Theodosia Gougousi at UMBC and Greg Snider at Notre

Dame.

The validation, support, and love my immediate and extended family gave to

me during my graduate days will remain a dear memory to me of this experience.

To my parents, John and Pam for their relentless support and love, I will never

forget. To Jerry and Karen for their encouragement and support. And last but

foremost, my wife Angela for her very being that orchestrates the beauty in my

life.

xiii

CHAPTER 1

INTRODUCTION

1.1 III-Nitride HEMT applications

With direct band gap energies (0.9 - 6.2 eV) corresponding to wavelengths (λ

= 200 - 1380 nm) that range the entire visible spectrum (Fig. 1.1) and built-in

polarization fields that are tailorable through heterostructure engineering the III-

Nitride system lends itself to an almost endless cache of applications including the

High Electron Mobility Transistor (HEMT) that has yielded some of its highest

invested returns. Premier attributes inherent to GaN of high intrinsic breakdown

field (3 MV/cm) and the extraordinarily large electron densities achievable without

impurity doping have been core motivations of GaN-based HEMT development

toward high frequency and power applications. In less than 20 years the GaN

HEMT went from a nascent state to obtaining the highest current and power

density of any transistor and are beginning to keep pace with their narrow band

gap FET counterparts in terms of cut-off frequency.

The DC drain characteristics can offer a preliminary estimate for the qual-

ity of a device’s RF and power operation. As seen in Fig. 1.2 for a standard

depletion-mode AlGaN/GaN HEMT, the combination of a high (> 30 volts) off-

state breakdown voltage with a high maximum current density is desired for power

requirements. Additionally, electron velocity must be appreciable in order to ob-

tain high operational frequencies based on the simple relationship: ft = ve/2πLG.

1

Figure 1.1. Energy gap versus lattice spacing for the varioussemiconductor systems. Wavelength corresponding to the energy gapcan be seen on the right with the color scale for the visible spectrum

plotted against the various material systems.

2

Figure 1.2. Example of a standard Al0.3GaN/GaN HEMT’s draincharacteristics with specific qualities pointed out that help define the

HEMT’s application.

As an estimate of the best power density that could potentially be achieved in a

device the DC current-voltage relationship for power P = (∆I × ∆V )/8 or the

more precise form found in Fig. 1.2 can be used. However, this value is much

more difficult to attain in actual RF operation due to charging effects, destructive

breakdown field limitations of the materials used, and other electrical constraints

involved with the design of the device.

Efficient channel pinch-off with low sub-threshold current is necessary to im-

plement GaN-based HEMTs in the high frequency large signal applications they

are intended for. This requires efficient carrier confinement and a gate barrier

layer with high-valued breakdown properties that can be sustained amidst the

high electric fields present under large signal operation. Poor pinch-off properties

3

reduce the efficiency of the device for converting input power to amplified out-

put power, better known as Power Added Efficiency (PAE). To achieve higher

operational frequencies the gate length needs to be short. An often occurrence

with deep sub-micron gate lengths is pronounced short-channel effects that limit

scalability by impairing RF performance. All these issues have been addressed by

numerous different approaches with the III-Nitride HEMTs setting frequency and

power records along the way.

1.2 Polarization physics of the III-Nitride semiconductor system

The following will serve as a brief overview of the origin, manifestation, and

implication of the polar nature of the III-Nitride semiconductor system in order

to lay a foundation on which the subsequent research is built. For a rigorous

development of the polar physics involved with the III-Nitride system the reader

is referred to reference [1]. The III-Nitride system owes its spontaneous polar

nature to the ionicity of the constituent atoms (N, Ga, Al, In) grown in the

wurtzite crystal structure. Considering the case for a Ga-polar crystal (group III

species terminates the surface with crystal lattice orientation of (0001)), when

a heterojunction is built by the growth of a different group III species on top of

GaN, the difference in in-plane lattice constant of the two layers causes an in-plane

tensile strain and perpendicular compressive strain in the top layer that manifests

an additional piezoelectric polarization field to that of the inherent spontaneous

polarization. Fig. 1.3 (a) shows the case for the AlN/GaN heterostructure. The

combination of the polarization field in the AlN and conduction band discontinuity

at the AlN/GaN interface set up favorable conditions for the migration of mobile

charge to accumulate in the GaN layer at the heterointerface. The current theory

4

maintains that a donor-like surface state is the primary contributor of the mobile

electrons that accumulate at the AlN/GaN interface [2]. The resultant band

diagram is shown in Fig. 1.3 (c) showing the localized 2 dimensional charge

distribution that serves as the HEMT channel. The maximum polarization charge

is determined by the mole fraction content of group III atoms in the barrier while

the physical thickness of the barrier determines how much of that charge is ionized

from the surface donor. Therefore, the surface donor energy and the thickness of

the barrier in an AlN/GaN HEMT play the dominant roles in the formation of

the 2D channel.

1.3 Background on the AlN/GaN heterostructure

The inception of the AlN/GaN heterostructure began with the immediate ap-

plication to a MESFET device by Binari in 1994 [3]. Though at that time the

understanding of polarization-induced channel charge was still six years too young

[2]. Several reports followed in the same vain of device [4],[5],[6] but it wasn’t until

2000 that a study on high crystalline quality un-doped AlN/GaN was reported by

Smorchkova [7]. This turning point ushered in the realization through empirical

evidence that a fully strained AlN barrier to unintentionally-doped GaN could

achieve several factors larger charge density than the AlxGa1−xN/GaN structure

with comparable channel mobilities for corresponding x = 0.2 - 0.3. Moreover, this

could be done with an ∼5× reduction in barrier thickness implying a means to

vastly enhance scalability in a device structure through atomically smooth growth.

The studies performed by Smorchkova have since been refined and expanded upon

by Cao who to date has demonstrated the highest 2D electron density achieved

in any single heterojunction leading to the lowest sheet resistance 2DEG in any

5

Figure 1.3. AlN/GaN heterojunction illustration showing thecorresponding polarization dipoles (a) and charge distributions (b) along

with the energy band diagram (c).

6

Figure 1.4. 2×2 µm atomic force micrographs of the surface of theAlN/GaN HEMTs for various AlN thicknesses. (images courtesy of Yu

Cao, U. of Notre Dame)

III-Nitride heterosystem [8]. This and a subsequent study have shown interesting

physics [9] that set aside the AlN/GaN HEMT structure as a means to harness

enhanced frequency/power operation in a conventional heterojunction FET struc-

ture through drastic down-scaling of device geometry. This physics will be the

subject of the following discussions.

AlN/GaN HEMT structures reported by Cao were grown on GaN templates

on Sapphire with a varied AlN barrier thickness between 2-7 nm [8]. The range

extremities were defined by the lack of 2DEG formation at the low end and strain

relaxation that led to cracking of the AlN barrier at the high end setting the

usable upper limit to ∼5 nm before mobility began to degrade. Atomic force mi-

croscopy (AFM) images for the various thicknesses are shown in Fig. 1.4 showing

micro-cracking beginning to occur for the 6 nm thick AlN originating at threading

dislocation centers. Clear cracks can be observed in the scan of the 8 nm thick

AlN.

Hall effect measurements on the series of samples with varied AlN thicknesses

reveal an increasing sheet charge density with increased AlN thickness (Fig. 1.5

7

Figure 1.5. 2DEG properties of the AlN/GaN heterostructure. Imagestaken from reference [8].

(a)). The charge density increase seems to persist up to the point of AlN lattice

relaxation with a maximum density very near the calculated charge polarization

limit which is dependent on the difference in polarization fields at the AlN/GaN

interface and assumes polarization from every atom at the interface [1]. Hall mo-

bility shows a maximum for smaller AlN thicknesses up to ∼5 nm where lattice

relaxation causes mobility of the 2DEG to begin to drop off (Fig. 1.5 (b)). The

combination of these trends show up in sheet resistance through the relationship

Rsh = 1/qµns for an n-type channel yielding a sheet resistance minima window

below 200 Ω/ for AlN thicknesses between 3 - 4 nm (inset to Fig. 1.5 (b)). Re-

cently, further growth refinements by Cao have led to a record low sheet resistance

of ∼100 Ω/ for a single junction AlN/GaN HEMT structure [12]

The AlN/GaN heterostructure being a binary junction eliminates the presence

of scattering from a disordered alloy such as in the case of an AlxGa1−xN barrier

and therefore greatly benefits the 2DEG low-field mobility. However, because

8

of the larger ∆Ec between AlN and GaN, 2DEG confinement is enhanced and

pushes the 2D electron charge distribution closer to the AlN/GaN interface. This

mechanism causes the atomic roughness at the interface to have an enhanced

effect on the 2DEG transport properties. The time for an electron to scatter from

its initial k -state to a different k -state due to its interaction with the atomically

rough interface is given by the following integral and has been calculated for the

AlN/GaN heterojunction by Cao as seen in Fig. 1.5 (d)

1

τIR=

∆2L2e4m∗

2ε2~3(ns

2

)2 ∫du

u4e−k2FL

2u2

(u+G(u) qTF

2kF)2√

1− u2

where the scattering potential for the rough interface is modeled as a coulombic

potential with a correlation length L and RMS roughness height ∆ [1]. The inte-

gration over this dimensionless integral gives the solution for the IR limiting com-

ponent to the momentum scattering time that can then be related to the IR scat-

tering limited mobility term through the Drude relationship by µIR = qτIR/m∗.

Other dominant scattering mechanisms that influence the total low-field mobility

are polar optical and acoustic phonon scattering. Cao demonstrated that IR scat-

tering is the dominant scattering mechanism at low temperature with combined

effects from PO and acoustic phonon contributions at room temperature, shown

in Fig. 1.5 (d). Therefore, it is expected that growth conditions are critical for

high mobility AlN/GaN structures and those that yield the smoothest AlN surface

(measurable by AFM) will subsequently demonstrate the highest mobility. The

final test for the AlN/GaN heterostructure is its implementation in a HEMT de-

vice. The integration of the high-quality, low-Rsh AlN/GaN heterostructure with

refined processing techniques will present HEMT technology with a candidate that

will surely rival if not surpass the best state-of-the-art HEMTs of today.

9

1.4 Scope of this work

This work explores the binary ultra-thin AlN/GaN heterostructure for the

purpose of HEMT devices that have operational frequencies up to 100’s of GHz.

All samples that will be discussed were grown by plasma-assisted molecular beam

epitaxy (MBE) at the Naval Research Laboratory unless otherwise noted.

Chapter 2 addresses the development of several key fundamental aspects in

the fabrication of AlN/GaN HEMTs. Namely, ohmic contacts and gate insulator

considerations as well as their implications on HEMT device operation.

Chapter 3 demonstrates the realization of the AlN/GaN HEMT through sev-

eral studies performed to analyze and improve DC and small signal RF perfor-

mance . The chapter concludes with an analytical discussion of the effects of

contact and access resistance on I-V characteristics as given by a physical model

based upon the gradual channel approximation found in Appendix B.

Chapter 4 develops a lumped element small signal model for the insulated-gate

AlN/GaN HEMT. Included in this chapter is transit and charging time analysis

following Moll’s method. Such analysis quantifies the limiting mechanisms in

small signal frequency performance and allows an extraction of saturated electron

velocity in the channel. Points of improvement for the small signal operation of

the AlN/GaN HEMT are also discussed.

Chapter 5 includes a separate study where charging effects of a thin SiN passi-

vation layer in the access regions of a typical AlGaN/GaN HEMT is investigated.

Through threshold voltage shifts after DC bias stressing the device, the concen-

tration of trapped charge is obtained.

Chapter 6 involves the exploitation of the thin AlN barrier in the development

of several novel HEMT device structures. Al-containing back barriers are explored.

10

A dual-channel AlN/GaN HEMT is designed and fabricated. The design aims to

leverage the high 2DEG density in the upper channel as a means to screen surface

potential fluctuations from the buried channel in an attempt to mitigate frequency

dispersion without the necessity of a passivation layer.

Chapter 7 concludes the work with closing remarks and suggestions for future

development utilizing the principles involved with the AlN/GaN HEMT.

1.5 Chapter 1 cited literature

1. D. Jena, “Polarization induced electron populations in III-V nitride semi-conductors; Transport, growth, and device applications,” Ph.D. dissertation,University of California, Santa Barbara, 2003.

2. J.P. Ibbetson, P.T. Fini, K.D. Ness, S.P. DenBaars, J.S. Speck, U.K. Mishra,“Polarization effects, surface states, and the source of electrons in AlGaN/GaNheterostructure field effect transistors,” App. Phys. Lett., no. 2, vol. 77,Jul. 2000.

3. S.C. Binari, L.B. Rowland, W. Kruppa, G. Kelner, K. Doverspike, D.K.Gaskill, “Mircrowave performance of GaN MESFETs,” Elec. Lett. vol. 30,no. 15, July 1994.

4. H. Kawai, M. Hara, F. Nakamura, S. Imanaga, “AlN/GAN insulated gateheterostructure FET with regrown n+GaN ohmic contact,” Elec. Lett no.6, vol. 34, Mar. 1998.

5. I. Daumiller, P. Schmid, E. Kohn, C. Kirchner, M. Kamp, K.J. Ebeling, L.L.Pond, C. Weitzel, “DC and RF characteristics of AlN/GaN doped channelheterostructure field effect transistor,” Elec. Lett. no. 18, vol. 35, Sept.1999.

6. E. Alekseev, A. Eisenbach, D. Pavlidis, “Low interface state density AlN/GaNMISFETs,” Elec. Lett. no. 24, vol. 35, Nov. 1999.

7. I.P. Smorchkova, S. Keller, S. Heikman, C.R. Elsass, B. Heying, P. Fini, J.Speck, U.K. Mishra, “Two-dimensional electron-gas AlN/GaN heterostruc-tures with extremely thin AlN barriers,” Appl. Phys. Lett. no. 24, vol. 77,2000.

11

8. Y. Cao, D. Jena, “High-mobility window for two-dimensional electron gasesat ultrathin AlN/GaN heterojunctions,” Appl. Phys. Lett. 90, 182112,2007.

9. Y. Cao, K. Wang, A. Orlov, H. Xing, D. Jena, “Very low sheet resistanceand Shubnikov-de-Haas oscillations in two-dimensional electron gasses atultrathin binary AlN/GaN heterojunctinos,” Appl. Phys. Lett. 92, 152112,2008.

10. I.P. Smorchkova, S. Keller, S. Heikman, C.R. Elsass, B. Heying, P. Fini, J.S.Speck, U.K. Mishra, “Two-dimensional electron-gas AlN/GaN heterostruc-tures with extrememly thin AlN barriers,” Appl. Phys. Lett. no. 24, vol.77, Dec. 2008.

11. I.P. Smorchkova, L. Chen, T. Mates, L. Shen, S. Heikman, B. Moran, S.Keller, S.P. DenBaars, J.S. Speck, U.K. Mishra, “AlN/GaN and (Al,Ga)N/AlN/GaNtwo-dimensional electron gas structures grown by plasma-assisted molecular-beam epitaxy,” J. Appl. Phys. no. 10, vol. 90, Nov. 2001.

12. C. Yu, “Study of AlN/GaN HEMTs: MBE growth, transport properties,and device issues,” Ph.D. dissertation, University of Notre Dame, 2010.

12

CHAPTER 2

TRANSISTOR DESIGN CONSIDERATIONS

The specific qualities afforded by the AlN/GaN heterostructure may be com-

bined in the HEMT structure to ultimately provide a technological vehicle to

access a high frequency amplification realm that is otherwise off-limits to the Al-

GaN/GaN HEMT. Owing to the thickness limitation set by strain relaxation of

the AlN barrier grown atop the GaN buffer, the AlN layer sets an upper limit in

barrier thickness of only 6 nm [13],[14]. This silver lining provides a means for

the reduction in gate-to-channel distance which ultimately sets a new benchmark

in feasibility of maximizing frequency metrics, ft and fmax through drastic barrier

scaling. Combined with a polarization charge limit of ∼6×1013 cm−2, a factor

of 3 higher than that of the AlxGaN/GaN heterostructure makes the AlN/GaN

HEMT a worthy candidate for a high frequency-power transistor.

While the III-Nitride system boasts many attractive qualities such as high

breakdown field (∼3 MV/cm), high peak electron velocity (∼2.5×107 cm/s), and

chemical stability all of which are results of its wide band gap nature, a number of

difficulties arise in their processing because of the same reason. The thin barrier

AlN/GaN HEMT seems to show enhanced difficulty at all the same design points

the AlGaN/GaN structure exhibits. The formation of low resistance (< 0.5 Ω-

mm) ohmic contacts remains elusive [15]. Schottky barrier gates to thin ( 2 nm)

AlN is exceedingly difficult due to the presence of high tunneling current through

the AlN barrier. Therefore, the insertion of a gate dielectric is required to suppress

13

gate current [16]. The proximity of the AlN surface to the 2DEG channel is only

a few nanometers. This provides a close charging surface that supports effective

gate length extension and current collapse during the RF operation of the HEMT.

Conventional passivation methods such as the deposition of a thick (∼100 nm)

PECVD SiN layer show improved RF performance but at the expense of a reduced

channel mobility due to plasma damage during the deposition. This ultimately

limits the maximum RF performance expected by such devices. These issues

amongst others serve as a reminder of the immature state of AlN/GaN HEMT

technology. This chapter addresses solutions to the problems listed toward the

advancement of the AlN/GaN HEMT.

2.1 The AlN/GaN HEMT structure

Shown in Fig. 2.1 is an illustration of the cross section of the AlN/GaN HEMT

device structure with the band diagram taken vertically through the insulated

gate region. By varying the AlN barrier thickness through atomically accurate

growth the polarization fields in the structure can be tailored to induce mobile

electron charge densities between the range of 5×1012 to 5×1013 cm−2 [14] with

the corresponding mobility trend as discussed in chapter 1.3.

The combination of the charge distribution and conduction band offsets present

in the AlN/GaN heterostructure set up an asymmetric potential profile comprising

the channel of the HEMT (as seen in Fig. 2.1) that can be approximately char-

acterized by a triangular potential well at the first sub-band energy level (dashed

line). The common practice is to use Airy functions to represent the electron

wave function in the triangular well in order to describe the spatially distributed

quasi-2DEG. However, according to charge neutrality the electric field deep in

14

Figure 2.1. Conduction band diagram for the insulator/AlN/GaNHEMT showing the charge distribution and the choice of electron wave

function.

15

the GaN buffer must go to zero and therefore is not accurately modeled by an

infinite triangular well. The Fang-Howard (F-H) variational wave function may

alternatively be used [17], [18] which accurately describes the wave function fall-

off deeper in the UID GaN layer where the confining potential becomes finite and

constant. The FH wave function follows

Ψ(y) =

√b3

2ye−

by2 , (2.1)

where b is the variational parameter and y is the depth from the AlN/GaN inter-

face. The wave function is used to determine the functional form of the spatially

distributed mobile charge through the relationship ns(y) = |n3D||Ψ(y)|2. The

mobile charge distribution function is therefore,

ns(y) =|n3D|b3

2y2e−by. (2.2)

The wave function describing the HEMT 2DEG channel may be approximated

as a weighted delta function for qualitative considerations or for calculations that

require only a knowledge of charge density and centroid location, < y >. However,

for cases that require a precise physical description of the charge-voltage relation-

ship the wave-function is necessary to provide the means to analytically describe

the evolution of mobile charge (density and spatial variation) when a gate voltage

is applied. By considering the conduction band profile in Fig. 2.1 a voltage loop

can be traced that gives the Vg-Vs-ns relationship for the insulated-gate AlN/GaN

HEMT

−VGS + ΦB + Vox + ΦoA − VAlN −∆ECq

+Eoq

+(EF − Eo)

q= 0, (2.3)

16

where VGS = VG − VS is the gate-to-source voltage difference, ΦB is the barrier

height of the gate metal to dielectric, Vox is the voltage drop within the oxide layer,

ΦoA is the effective conduction band offset between the oxide and AlN layers, VAlN

is the voltage drop within the AlN layer, ∆EC is the conduction band discontinuity

between AlN and GaN, Eo is the first sub-band energy, and EF is the Fermi level

in the GaN quantum well (QW). This is the charge-control equation (CCE) for

the insulator/AlN/GaN heterostructure. This governing equation describing the

electrostatics of the insulator/AlN/GaN structure immediately beneath the gate

will take various forms in subsequent sections. For the case of the AlN/GaN

heterostructure without an oxide cap the Vox term may be eliminated from the

equation. For a detailed development of the I-V relationship of the AlN/GaN

HEMT utilizing the FH wave function see Appendix B. A discussion of the effects

of access and contact resistance on the intrinsic I-V characteristics can be found

in section 3.2 of the following chapter.

2.2 Design considerations of the AlN/GaN HEMT

The AlN/GaN heterostructure has its share of caveats that must be considered

before it inherits the form of a transistor. The mobility-charge density trends

with respect to AlN barrier thickness, shown in Fig. 1.5 demonstrate the limited

usable range of AlN thickness. Confined between 2 nm and 5 nm the AlN barrier

introduces a myriad of issues the device designer must be cognisant of. Such

a thin barrier causes gate leakage current to be a concern for a Schottky-gated

depletion-mode transistor due to tunneling. This issue will be discussed in more

detail in section 2.2 of this chapter but an obvious solution is the insertion of a

dielectric layer to suppress gate leakage current [19]. The choice of insulator used

17

for the gate dielectric layer has major implications on device performance.

The employment of the AlN barrier has shown its utility as a thin (1 nm) in-

sertion layer in the AlGaN/GaN HEMT in order to create an all binary interface

with a large barrier height that would eliminate alloy disorder scattering by push-

ing the electron wave function tail out of the alloyed AlGaN barrier [20]. High

contact resistance is enhanced dramatically when the AlN layer is several nanome-

ters thick [15]. This extrinsic resistance further impairs DC and RF performance.

In this work we show several techniques have been used to reduce ohmic contact

resistances in the (> 2 nm) AlN barrier HEMTs (section 2.1).

The points to be discussed are in fact rather fundamental with respect to

device design and fabrication. However, in that they are fundamental they have

profound influence on the operation of the HEMTs. Therefore, a careful approach

must be taken when making design decisions for the sake of utilizing the high

frequency capability the AlN/GaN heterostructure offers.

2.2.1 Ohmic source-drain contacts

By and large the metallurgy of ohmic contact formation to the wide band gap

III-N system tends to be more experimentally based than scientifically understood.

To date the best ohmic contacts to AlGaN/GaN HEMTs are an order of magnitude

more resistive than those of their narrow band gap counterpart [21], [22]. Because

of this additional resistance, DC and RF performance is severely limited and the

intrinsic functionality of GaN-based devices is only modestly utilized.

Thermally annealed ohmic contacts to the AlGaN/GaN heterostructure have

received the most intense exploration to date since this heterostructure showed

immediate promise and therefore most attention upon its inception in the mid-

18

1990’s. The contact metal scheme to the AlGaN/GaN structure has seen a

fair number of different metal attempts but have mostly converged on the Ti-

tanium/Aluminum combination at the contact-semiconductor mating interface.

M.E. Lin et al. showed specific contact resistance of 8×10−6 cm−2 to GaN using a

Ti/Al stack and compared it to Al, Au, and a Ti/Au combination [23]. Contacts

have since been improved upon but still rely on the Ti/Al component [21].

The prevailing theory behind the operation of an annealed ohmic contact to

GaN maintains that a TiN alloy forms at the mating interface thereby leaving

behind an excess of N-vacancies that provide a low resistance interfacial layer

the contacts ultimately rely upon [24], [25]. A competing theory approaches the

issue from the standpoint of differences in work functions between the metal and

semiconductor [24]. Given that the work function of GaN is 4.1 eV, very near

that of Al the Al-containing contacts are ohmic due to the low barrier between

the two materials. In either case it is evident that the Al-content plays a large

role in the formation of low resistance ohmic contacts. However, the history since

the first theory was proposed has given reason to suggest it is the best since cross

sectional TEM has shown the formation of TiN alloyed layers and spiking at the

contact-semiconductor annealed interface [26], [27].

The choice for constituent metals involved in the ohmic metal stack follows

some general logic though it is largely accepted that the detailed picture of an-

nealed ohmic contact formation is still somewhat of a black box [28]. As discussed,

the Ti/Al component plays the main role in forming contacts. However, both met-

als have a high propensity for oxidation under elevated temperature such as an

anneal. Therefore, a low-resistive inert metal must be used to avoid the increase

in contact resistance from oxidation as well as reduce the overall lateral resistance

19

of the contact. Au is an obvious solution however, Au has a low melting point,

easily diffuses through GaN, is a deep level state in GaN, and causes instabilities

in the contacts. The solution used is to include a Au diffusion barrier layer to

retain the Au as the protective top layer. Au diffusion barriers such as Ni, Pt, Mo,

Ti, Ag and others have been studied. While no standard contact scheme has been

substantiated a popular layer structure has been accepted that uses the metal

sequence, Ti/Al/Ni/Au. With additional processing tricks this metal structure

has achieved as low as ∼ 0.2 Ωmm contact resistance in an AlGaN/GaN HEMT

[21], [29].

While the metallization scheme is critical another factor in ohmic contact for-

mation is the thickness of the barrier the metal stack must penetrate during the

anneal. In Ga-polar III-N heterostructures the barrier thickness is proportional to

the 2DEG density. For ohmic contacts to form the barrier must be thick enough

that a 2DEG is present to make contact to, but simultaneously thin enough for suf-

ficient metal diffusion to occur at 800 - 900C anneal temperatures. One popular

method adopted in research and industry for various material systems is a pre-

metallization etch to thin the barrier immediately underneath the contact metal

for improved diffusion. Buttari et al. showed for AlGaN/GaN HEMTs with a 25

nm AlGaN barrier and a 1 nm AlN insertion layer that a 7 nm deep pre-etch re-

duced ohmic contact resistance of the structure from 0.45 Ω-mm for an un-etched

structure down to 0.27 Ω-mm for the 7 nm pre-etched structure [21]. It should

be noted that the 1 nm AlN insertion layer generally causes an increase in con-

tact resistance so the pre-ohmic metallization etch offers an attractive candidate

for lowing contact resistance in the multi-nanometer AlN/GaN heterostructure.

However, because the AlN barrier is only 6 nm at maximum thickness it leaves

20

Figure 2.2. Current-voltage characteristics of the Zimmermann contactstaken from reference [15]. I-V characteristics of as-deposited contacts

across a 3 µm separation TLM structure for several different AlNthicknesses (left). Current level dependence on anneal temperature for

the Zimmermann contact (right).

little tolerance for error during the pre-etch.

The AlN/GaN heterostructure introduces the necessity for some changes to

the conventional thought and practice of the Ti/Al-based ohmic contact. Indeed,

the AlN barrier is much thinner than the AlGaN barrier already discussed but

it also sports a much wider bandgap (6.2 eV) with a strained lattice constant

that of GaN (3.2 A) making the anneal-diffusion of the contact metals a slightly

different scenario. A notable difference between the AlGaN/GaN and AlN/GaN

heterostructures in terms of contacts is that as-deposited contact metal prior to

any annealing showed near-ohmic to ohmic behavior depending on the thickness

of the AlN barrier (Fig. 2.2) [15]. Upon annealing, the current-voltage (I-V) char-

acteristics improved toward lower resistance but the ideal annealing temperature

for the best I-V characteristics was lower (400C) for a 3 nm AlN barrier than

21

Figure 2.3. Current-voltage characteristics for pre-metallization etchedcontacts to AlN/GaN. Pre-etch to AlN/GaN (left) and the removal of

the GaN cap on the GaN/AlN/GaN heterostructure (right).

that of the 20 nm AlGaN barrier. Higher temperatures led to a reduced current,

presumably due to the degradation of the AlN/GaN interface where the 2DEG

resides. If mobility is used as the metric for the crystalline quality of the AlN it

can be seen in Fig. 2.2 that the higher mobility structures cause more difficulty

in forming low resistance ohmic contacts. We speculate this is because of thick-

ness variations in the structure that leads to thin regions the ohmic metal can

diffuse through more readily making a better ohmic contact to a lower quality

heterostructures.

A number of metallization schemes have been investigated and the two most

successful follow a similar design. The first, a Ti/Al/Ti/Ni/Au metal series that is

electron beam deposited with thicknesses of 20/200/20/60/10 nm, respectively was

proposed by Zimmermann et al. [15]. The second, a Ti/Al/Ni/Au (30/200/40/10

nm) proved to be identical in I-V characteristics and contact resistance compared

22

Figure 2.4. (left) Split wafer comparison of the effect of pre-etchedcontact on drain characteristics. (right) Function of AlN thickness oncontact resistance for various metal schemes attempted including the

pre-metallization etched contacts.

to the Zimmermann contact suggesting, but not surprisingly, that the Ti content

is a major factor in ohmic formation. These two metal schemes were exclusively

used in HEMT fabrication.

For the AlN/GaN heterostructure a low Rsh window exists for AlN barrier

thicknesses between 3 and 5 nm as shown by Cao [14]. Therefore, in order to take

advantage of the low sheet resistance window while simultaneously achieving sub-1

Ω-mm contact resistance, a pre-metallization etch offers a viable solution. Indeed,

the most successful process for ohmic contact formation to an AlN/GaN 2DEG

was the employment of a modified ohmic pre-metallization etch by a BCl3/Cl2

plasma. The metal scheme used was either of the variations discussed previously.

An Oxford Systems inductively-coupled plasma (ICP) etcher was used for the pre-

etch with a plasma chemistry of Ar/BCl3/Cl2 at 100/25 watts upper and lower

electrode power, respectively. These conditions gave an AlN etch rate of ∼ 0.53

23

A/sec determined by atomic force microscopy.

The comparative evolution of IV characteristics of a pre-etched ohmic contact

and a standard ohmic contact on a 5 nm AlN barrier before and after annealing

can be seen in Fig. 2.3. The as-deposited pre-etched ohmic contact current is

drastically lower than that of the non-etched contact. This is attributed to the

thinning of the AlN barrier from the etch, thereby reducing the 2DEG density

along with probable plasma-induced damage to the 2DEG. However, after a 30

second anneal at 800C the I-V slope of the pre-etched contact is much steeper

than the non-etched contact indicating a lower contact resistance. This is in fact

verified with TLM measurements that yield on average contact resistances of ∼0.7

Ω-mm for the pre-etched contact in contrast with ∼1.6 Ω-mm for the non-etched

contact for a 5 nm AlN barrier. While these results are attractive they are not

remiss of problems associated with such a shallow etch. Because of details about

the etch that are difficult to determine (incubation time and etch details of the

plasma, true depth of the etch, etc.) the repeatability of obtaining an accurately

specified contact resistance is somewhat troublesome. This is further frustrated

when working with more “sophisticated” heterostructures such as GaN/AlN/GaN

and it’s derivatives.

Due to surface sensitivity of AlN/GaN structures, GaN capping layers were

also explored. However, it offers an additional degree of uncertainty in etch depth

during the ohmic pre-etch discussed above. Nonetheless, etching of the GaN cap

is essential to forming a contact which has a linear I-V characteristic for cap

(> 2 nm) and barrier (> 3 nm) thicknesses. It can be seen in Fig. 2.3 that a

GaN capped structure without the removal of the cap still shows a non-linear I-V

character even after annealing at 800C for 30 seconds despite the conventional

24

understanding that GaN allows easy penetration for ohmic metal diffusion. On

the other hand, when the GaN cap is etched away leaving only the AlN surface,

the I-V characteristics measure linear with high current in close agreement to the

non pre-etched contacts in Fig. 2.3 on the left.

Fig. 2.4 shows contact resistance for various AlN thicknesses for the AlN

HEMTs fabricated. A trend of increasing RC is observed for increasing AlN thick-

ness in agreement with previous work [15]. The linear trend line is used only as a

guide to the eye. Work on Ti/Al/Mo/Au contacts to AlN/GaN by Wang is plot-

ted (open square) for comparison. The average RC values obtained through the

pre-metallization etched ohmic contacts were notably lower than their non-etched

counterpartes. Despite the 5.5 nm thick AlN the contact resistances of the pre-

etched contacts were consistently below 0.9 Ω-mm with the lowest value for the 5

nm barrier thickness being 0.5 Ω-mm. Up to the time of this writing this value is

the lowest RC reported for the AlN/GaN HEMT and moreover this was achieved

in a relatively thick AlN barrier structure. Therefore, the pre-metallization etch

provides a means to reliably reduce contact resistance in AlN/GaN HEMTs.

2.2.2 Gate current suppression

High leakage currents in Schottky barrier gated AlN/GaN HEMTs exist due

to a limited barrier thickness (less than 6 nm) combined with the capability of

high (up to 6×10−13 cm−2) 2DEG charge densities1 [14]. Therefore, it is necessary

to use an insulating layer in between the gate metal and AlN surface to suppress

these otherwise high gate leakage currents.

The choice of insulating material as well as the deposition method has a critical

1The dependence of Fowler-Nordheim tunneling current on charge density follows: Jt ∝∫dEfST (E)(1−fM ) [30]. This shows that the higher charge population of the 2DEG causes an

increase in F-N assisted tunneling transport and therefore more unwanted gate leakage current.

25

influence on the operation of the HEMT. Higashiwaki et al. demonstrated a 3

nm SiN gate insulator deposited by Catalytic-CVD that served the simultaneous

purpose to passivate surface states in the access regions of the HEMT [31]. CVD

SiN has been employed as a “passivation” dielectric for AlGaN/GaN HEMTs since

2000 [32] and with the GaAs system before that [33] making this choice an obvious

one to explore. HRL researchers reported on the use of a GaN cap for gate leakage

current suppression [34]. Zimmermann et al. employed an e-beam deposited 3 nm

thick Al2O3 layer which showed sufficient gate current suppression [35]. Though

adequate for modulation of the reported device, the e-beam deposited oxide is a

rather crude method for oxidation of a surface. Pearton et al. utilized a UV-ozone

treatment prior to gate metallization that partially consumed the AlN barrier and

formed a mixed AlOxN1−x of unknown concentration [36]. This provided sufficient

gate current suppression as well as a positive threshold voltage. An attractive and

more advanced method for ex-situ oxidation of the AlN surface was reported by

Dabiran et al. with an Atomic Layer Deposited (ALD) Al2O3 gate insulator 30 nm

thick on a 4 nm AlN thick AlN/GaN HEMT [37]. The ALD deposition method

inherently allows atomic accuracy of the oxide providing a dense amorphous oxide

making this method a highly attractive oxidation process.

In order to qualify the gate insulator varieties available several metrics where

employed. A comparison of the changes in sheet resistance components ns and

µ, MOS diode IV characteristics, Capacitance-Voltage (CV) analysis, and ellip-

sometry were used to analyze the individual dielectrics for gate insulators. The

insulators explored in this study were PECVD deposited SiN, ALD Al2O3, HfO2,

and Ta2O5.

SiN was deposited by plasma-enhanced CVD using an alternating high/low

26

Figure 2.5. Effects of ALD and PECVD dielectrics on the AlN/GaNchannel charge characteristics µ and ns (left). Gated IV comparison of

the dielectrics studied, taken on 100 µm diameter CV dots (right).

plasma frequency excitation that reduces strain in the deposited layer. A nominal

thickness of 5 nm was deposited. Higashiwaki showed the progressive increase of

sheet resistance with increased SiN thickness where 5 nm gave the optimal condi-

tion for appropriate insulation, which was the basis for the choice of SiN thickness

in this work [38]. Ta2O5, Al2O3, and HfO2 were deposited by ALD, a very sim-

ilar method to the CVD process. Al2O3 films were deposited by means of the

precursors Trimethylaluminum (TMA, Al(CH3)3) and DI H2O at an optimal stoi-

chiometric deposition temperature of 250C that yielded a deposition rate of 1.05

A/cycle. HfO2 films were deposited2 using the precursors tetrakis (ethylmethyl)

amino hafnium (TEMAHf) and DI H2O at an optimal deposition temperature

of 250C that yielded a deposition rate of 1.2 A/cycle. All film thicknesses were

targeted at a nominal 5 nm and blanket deposited across the entire sample surface.

2HfO2 and Ta2O5 films were deposited by Dr. Theodosia Gougousi at the University ofMaryland Baltimore County Physics Department.

27

Prior to all dielectric depositions ohmic contacts were formed and mesa iso-

lation was made which include a Van der Pauw test structure for on-wafer Hall

effect measurement. Hall measurements were taken before and after the deposi-

tion of dielectrics to monitor changes in sheet resistance due to the presence of

the dielectric. Results of these measurements can be seen in Fig. 2.5. For the 5

nm PECVD deposited SiN a negligible increase in 2D sheet density was observed

at the expense of a drastic (28%) reduction in mobility. This reduction in mobil-

ity is attributed to plasma damage from ion bombardment of the heterostructure

during the deposition since the plasma recipe was not optimized for such a thin

barrier. On the other hand, for the ALD processes a very minor (1.6-2%) reduc-

tion in mobility was measured with a notable increase in charge density of 7% and

12%, respectively. The specific mechanism for the increase in charge density is not

completely understood but the argument has been made that for SiN [31] it is the

satisfaction of dangling surface bonds that otherwise cause partial depletion of the

2DEG. It is speculated that either SiN, Ta2O5, Al2O3, or HfO2 will provide such

surface state passivation which could account for the increase in charge density of

the 2DEG. This has been reported by other groups as well [37].

After the insulator and gate metal had been deposited two-probe current-

voltage measurements were taken on capacitance-voltage test structures to com-

pare insulating properties of the dielectrics. Metal-AlN Schottky diodes were

fabricated as a reference for the comparison to the different dielectrics. The com-

parative results are shown in Fig. 2.5. For the applied voltage range appropriate

for a standard gate voltage range the HfO2 in this study appeared to be the best

insulating dielectric of those listed earlier with a current density of 2×10−6 A/µm2

at a reverse bias of -6V [39].

28

2.2.3 Gate capacitance model

From the thin AlN barrier comes the necessity of a gate insulator where high-κ

materials have shown promise. A high “intrinsic” gate capacitance is desired from

the transconductance standpoint where its relationship to gate capacitance can

be found through the definitions as follows,

gm =∂ID∂VGS

=∂

∂VGS(qnve) = q

(ve

∂n

∂VGS+ n

∂ve∂VGS

), (2.4)

noting that gm is maximized when the device is biased in the saturated-velocity

regime where ve = vsat and is constant. Also, recalling that Q = qn = CV which

leads to q ∂n∂V

= C therefore,

gm|ve=vsat → qvsat∂n

∂VGS= qvsatCG, (2.5)

where CG is given by equation A.12 for the insulated gate HEMT as CG =

(1/Cox + 1/CA + 1/CQ)−1. A more exact form for transconductance is developed

in Appendix B but Eq. 2.5 is adequate to demonstrate that the transconductance

of the HEMT is directly proportional to the gate capacitance and therefore the di-

electric constant of the gate insulator. However, if the gate dielectric is deposited

over the entire device surface as in the case of the transistors shown in the fol-

lowing chapter, its high-κ value causes the enhancement of parasitic capacitances

CGSp and CGDp which are inversely influential on frequency performance. Because

of this reason, the components of the gate structure have significant implications

on the circuit model of the device and warrants a discussion of their benefits and

detriments on capacitance, transconductance, and ultimately frequency perfor-

mance (given in chapter 4).

29

TABLE 2.1

PARAMETERS AFFECTING OPERATION OF THE AlN/GaN

HEMT FOR THE DIELECTRICS CONSIDERED IN THIS WORK

[40], [41].

Eg ∆EC χe φNiB VBD εκ(lit.) deposition

[eV ] [eV ] [eV ] [eV ] [MV/cm]

GaN 3.4 – 4.1 1.0 5 8.9 MBE

AlN 6.2 2.1 0.6 2.0 1.8 8.5 MBE

Si3N4 5.3 1.3 2.1 1.98 4 7 (7) PECVD

Al2O3 8.8 2.16 1.0 3.5 10 8 (11) ALD

HfO2 6.0 1.09 2.4 3.2 8 11.8 (24) ALD

Ta2O5 4.4 0.1 3.3 0.72 4 11.7 (26) ALD

The simplest model for the capacitance structure below the gate (which ne-

glects any leakage or charging paths through the insulator) is the series combina-

tion of capacitances associated with each layer structure given by Eq. A.12. Both

Cox and CA are constant for fixed layer thicknesses and CQ is the semiconductor or

“quantum” capacitance that changes with a change in applied gate voltage3. The

analytical form of CQ for the AlN/GaN heterostructure is given by Eq. A.23. This

goes to show the total gate capacitance is largely dependent on the insulator/oxide

thickness and dielectric constant. The various properties of the gate dielectrics

3By definition, quantum capacitance is the change in charge of a semiconductor or metal dueto a change in local potential [42].

30

Figure 2.6. Modeled C-V dependence on various system parameters(left), and calculated threshold voltage relationship dependance on the

gate dielectric constant for varying AlN thicknesses (right).

explored during this work are given in Table 2.1. In the table all relevant dielec-

tric parameters are referenced to AlN. These values are used for modeling the

capacitance-voltage (C-V) and threshold voltage (Vth) characteristics, as well as

the I-V characteristics which are derived in Appendix B.

The modeled C-V functionality for an ideal insulator/AlN/GaN capacitor with

an insulator dielectric constant of 2 and 10 are shown in Fig. 2.6. This shows

the effect of dielectric constant on total capacitance and Vth. If a fixed (positive)

charge distribution is included in the barrier the dotted line shows the effect on

the C-V curve (negative voltage shift). The same result occurs for a change in

barrier height (ΦB or ΦoA) between the different layers. This is expected since

work function engineering at the metal-insulator interface is a common method

for threshold voltage control and was recently reported in GaN-based HEMTs by

Li et al. [43]. On the other hand, if there is a distribution of electronic states (say

at the insulator/AlN interface) that can (dis)charge based on Fermi level position

31

and proximity to a mobile charge supply (e.g. the 2DEG or ohmic contact metal),

a different effect occurs on the C-V curve due to these states (commonly called

interface states). For a uniformly distributed (in energy) trapped electron charge

distribution a “stretch-out” of the depletion portion of the C-V curve is expected

(not shown) [44]. If the trap distribution is non-uniform the depletion stretch-out

of the C-V curve is also non-uniform in correspondence with the energy level(s)

of the trap distribution(s).

The influence of interface charge trapping on HEMT I-V characteristics will

be discussed in the following chapter, but for now attention is on the electrostatics

of the multi-stack gate capacitor. Fig. 2.6 (right) shows the calculated thresh-

old voltage as a function of dielectric constant of the insulator for various AlN

thicknesses. The functionality of the threshold voltage, Vth follows:

Vth = ΘB −Qπnet

(1

Cox+

1

CAlN

)+Qπ

AlN

(1

Cox

), (2.6)

where ΘB = φB + φoA −∆EC/q is the sum of conduction band offsets as defined

previously. Equation 2.6 may be recast in terms of the 2DEG charge density may

be recast as4

Vth = −qns(

1

Cox+

1

CA+π~m∗

)−(

9π~q2ns8κGεo

√8m∗

)2/3

, (2.7)

where ns is the equilibrium 2D electron charge density. It was assumed that at Vth

the Fermi level at the AlN/GaN interface was equal to the conduction band energy,

Ec. The first form of Eq. 2.6 shows the dependence of Vth on the band offsets

4This form comes from the difference of the results of two different imposed conditions onEq. 2.3. The first conditions being VGS = 0, and the second being VGS = Vth where ns → 0(i.e. “flatband”).

32

and fixed polarization charge involved, whereas the second form, Eq. 2.7 shows

the functionality of Vth with the 2DEG charge density and avoids the necessity

of knowing barrier height and band offsets which may be difficult to determine

accurately. Unsurprisingly, both forms depend on the capacitance contributions

from the oxide and AlN layers and assume no additional fixed or trapped charge.

An obvious point to make with respect to the Vth-trend in Fig. 2.6 (right) is

the effect the high-κ dielectric has on Vth. Based on boundary conditions at the

insulator/AlN interface, Gauss’s law (εoxFox = εAlNFAlN) shows that the higher

the dielectric constant (εox) of the insulator the smaller the electric field in the

insulator layer. Thus, the higher-κ of the insulator, the more electric field from the

gate voltage is spread across the AlN and GaN layers which ultimately causes Vth

to be less negative. The effect can also be seen in either Eq. 2.6 or 2.7 through the

C−1ox term. Equations 2.6 and 2.7, which lead to Fig. 2.6 slightly underestimate

Vth as compared to the linear extrapolation of the transfer characteristic. This is

because in the case of an extrapolated threshold voltage, there remains residual

charge in the mostly depleted channel whereas the condition, ns = 0 cm-2 is forced

on the calculated Vth in Fig. 2.6. Additional, mild inaccuracy arises for practical

considerations of Vth since φB and φoA are both non-constant for all gate metal on

any dielectric. Nonetheless, the functionality of Vth on εr in Fig. 2.6 is appreciated.

In the probable event that the insulator/AlN interface accommodates trap

states that are chargeable, Terman’s high-frequency method [45] has been adapted

(Appendix A) for the insulator/AlN/GaN HEMT structure to analyze interfacial

trapped charge. Applying this analysis to the four dielectrics explored can be seen

in Fig. 2.7. The tabulated values of the extracted Dit spectrum for the dielectrics

explored in this work may be seen in Table 2.2. The bottom row contains the

33

Figure 2.7. Modeled and measured C-V curves (bottom row) with thecorresponding Dit spectrum (top row) for the insulated-gate AlN/GaN

capacitors.

34

C-V data showing the modeled and measured curves for each dielectric. From the

derivative of the difference of the two curves the Dit spectrum can be determined

following Dit = Cox/qkT ×∂VGS/∂ηit as given in Appendix A. The top row shows

the corresponding result of the Terman analysis with Dit plotted as a function of

energy away from the conduction band of the AlN barrier.

In all cases Dit is found to be ∼1013 cm−2eV −1. These values for interface state

density are quite high. However, the trap states would not have any significant

influence on the C-V curve if the trap density were not of a similar order of

magnitude as that of the mobile 2DEG density. It is instructive to take in mind

that the interface is comprised of a low-temperature deposited amorphous oxide

that is not necessarily similar in atomic ordering as the crystalline AlN to which

it is adhered. Additionally, the value of Dit shown in Fig. 2.7 also contains any

trapped charge in the oxide or buffer, imaged to the oxide/AlN interface, which

can further skew the true interfacial value of Dit towards a higher density.

Two final qualitative comments are worthy of mention regarding the extracted

Dit spectra. First, the calculated energetic location that corresponds to Dit is as

questionable as the electric field in the barrier and band offset inaccuracies. This is

true since the energy spectra is calculated from Eq. A.4 which directly depends on

those quantities. The less known the offsets are, the more inaccurate the location

in energy the corresponding Dit becomes. Secondly, the energy ranges shown in

Fig. 2.7 are rather narrow, only ∼30% at most, of the band gap of AlN. Taking in

mind, this value corresponds to a much larger applied gate voltage range. Given

such a narrow accessible energy range, other methods must be employed to probe

states that may occur deep within the band. Deep level trap spectroscopy (DLTS)

is ideally suited for this application.

35

TABLE 2.2

EXTRACTED VALUES OF INTERFACE TRAP DENSITY FOR THE

DIELECTRICS EXPLORED IN THIS STUDY.

Dielectric < Dit > Dit Range

[cm−2eV−1] [cm−2] [eV]

SiN 2.0×1013 3.0×1013 1.7

Al2O3 1.0×1013 0.9×1013 0.8

HfO2 3.0×1013 5.6×1013 1.0

Ta2O5 1.0×1013 1.7×1013 0.5

Figure 2.8. HR-TEM image of a GaN/AlN/GaN structure with an ALDAl2O3 layer deposited on the surface. Image courtesy of D. Smith,

Arizona State University.

36

Figure 2.8 shows a high-resolution transmission electron microscope (HR-

TEM) micrograph of a GaN/AlN/GaN (2/3/1000 nm) heterostructure with an

ALD Al2O3 oxide layer deposited after MBE growth. The interesting point to

note is the Dit extracted and given in Table 2.2 with respect to the micrograph.

Dit of Al2O3/GaN was found to be the lowest of the dielectrics applied. This may

be due to the possibility of a ∼1 nm thick ordered oxide nucleation layer that

forms at the Al2O3/GaN interface which is suggested in the micrograph and may

allow a lower density of interfacial trap states.

2.3 Chapter 2 cited literature

13. I.P. Smorchkova, S. Keller, S. Heikman, C.R. Elsass, B. Heying, P. Fini, J.S.Speck, U.K. Mishra, “Two-dimensional electron-gas AlN/GaN heterostruc-tures with extrememly thin AlN barriers,” Appl. Phys. Lett. no. 24, vol.77, Dec. 2008.

14. Y. Cao, D. Jena, “High-mobility window for two-dimensional electron gasesat ultrathin AlN/GaN heterojunctions,” Appl. Phys. Lett. 90, 182112,2007.

15. T. Zimmermann, D. Deen, Y. Cao, D. Jena, H. Xing, “Formation of ohmiccontacts to ultra-thin channel AlN/GaN HEMTs,” Phys. Stat. Sol. (c), no.6, vol. 5, Apr. 2008.

16. H. Xing, D. Deen, Y. Cao, T. Zimmermann, P. Fay, D. Jena, “MBE-GrownUltra-shallow AlN/GaN HFET Technology,” ECS Trans. vol. 11, 2007.

17. D. Jena, “Polarization induced electron populations in III-V nitride semi-conductors; Transport, growth, and device applications,” Ph.D. dissertation,University of California, Santa Barbara, 2003.

18. Herbert Kroemer, “Quantum Mechanics for Engineering, Materials Science,and Applied Physics”, Prentice Hall, New Jersey, 1994.

19. D. Deen, T. Zimmermann, Y. Cao, D. Jena, H. Xing, “2.3 nm barrierAlN/GaN HEMTs with insulated gates,” Phys. Stat. Sol. (c), no. 6,vol. 5, 2008.

37

20. L. Shen, S. Heikman, B. Moran, R. Coffie, N.Q. Zhang, D. Buttari, I.P.Smorchkova, S. Keller, S.P. DenBaars, U.K. Mishra, “AlGaN/AlN/GaNHigh-Power Microwave HEMT,” IEEE Elec. Dev. Lett. no. 10, vol. 22,Oct. 2001.

21. D. Buttari, A. Chini, G. Meneghesso, E. Zanoni, B. Moran, S. Heikman,N.Q. Zhang, L. Shen, R. Coffie, S.P. DenBaars, U.K. Mishra, “SystematicCharacterization of Cl2 Reactive Ion Etching for Improved Ohmics in Al-GaN/GaN HEMTs,” IEEE Elec. Dev. Lett. no. 2, vol. 23, Feb. 2002.

22. J.G. Champlain, R. Magno, J.B. Boos, “Low resistance, unannealed ohmiccontacts to n-type InAs0.66Sb0.34”, Elec. Lett. no. 23, vol. 43, Nov. 2007.

23. M.E. Lin, Z. Ma, F.Y. Huang, Z.F. Fan, L.H. Allen, H. Morkoc, “Lowresistance ohmic contacts on wide band-gap GaN,” Appl. Phys. Lett. no.8, vol. 64, Feb. 1994.

24. Z.M. Zhao, R.L. Jiang, P. Chen, W.P. Li, D.J. Xi, S.Y. Xie, B. Shen, R.Zhang, Y.D. Zheng, “Aluminum and GaN contacts on Si(111) and Sap-phire,” Appl. Phys. Lett. no. 20, vol. 77, Nov. 2000.

25. S. Ruvimov, Z. Liliental-Weber, J. Washburn, K.J. Duxstad, E.E. Haller,Z.F. Fan, S.N. Mohammad, W. Kim, A.E. Botchkarev, H. Morkoc, “Mi-crostructure of Ti/Al and Ti/Al/Ni/Au Ohmic contacts for n-GaN,” Appl.Phys. Lett. no. 11, vol. 69, Sept. 1996.

26. Y. Dora, A. Chakraborty, S. Heikman, L. McCarthy, S. Keller, S.P. Den-Baars, U.K. Mishra, “Effect of Ohmic Contacts on Buffer Leakage of GaNTransistors,” IEEE Elec. Dev. Lett. no. 7, vol. 27, July 2006.

27. L. Wang, I. Adesida, A.M. Dabiran, A.M. Wowchak, P.P. Chow, “Ti/Al/Mo/AuOhmic contacts to all-binary AlN/GaN high electron mobility transitors,”Appl. Phys. Lett. 93, 032109, 2008.

28. C.T. Lee, H.W. Kao, “Long-term thermal stability of Ti/Al/Pt/Au Ohmiccontacts to n-type GaN,” Appl. Phys. Lett. no. 17, vol. 76, Apr. 2000.

29. J.W. Chung, W.E. Hoke, E.M. Chumbes, T. Palacios, “AlGaN/GaN HEMTWith 300-GHz fmax”, IEEE Elec. Dev. Lett. Vol. 31, No. 3, Mar. 2010.

30. S.M. Sze, Physics of Semiconductor Devices, 2nd ed.,Wiley, New York, 1981.

31. M. Higashiwaki, T. Mimura, T. Matsui, “AlN/GaN INsulated-Gate HFETsUsing Cat-CVD SiN,” IEEE Elec. Dev. Lett. no. 9, vol. 27, Sept. 2006.

38

32. R. Vetury, N.Q. Zhang, S. Keller, UlK. Mishra, “The impact of surface stateson the DC and RF characteristics of AlGAN/GaN HFETs,” IEEE Trans.Electron Devices no. 3, vol. 48, Mar. 2001.

33. Y.C. Chou, R. Lai, G.P. Li, J. Hua, P. Nam, R. Grundbacher, H.K. Kim, Y.Ra, M. Biedenbender, E. Ahlers, M. Barsky, A. Oki, D. Streit, “innovativeNitrided Passivated Pseudomorphic GaAs HEMTs,” IEEE Elec. Dev. Lett.no. 1, vol. 24, 2003.

34. A.L. Corrion, K. Shinohara, D. Regan, I. Milosavljevic, P. Hashimoto, P.J.Willadsen, A. Schmitz, D.C. Wheeler, C.M. Butler, D. Brown, S.D. Burn-ham, and M. Micovic, “Enhancement-Mode AlN/GaN/AlGaN DHFET With700-mS/mm gm and 112-GHz ft”, IEEE Elec. Dev. Lett. 31, 10, 2010.

35. T. Zimmermann, D. Deen, Y. Cao, J. Simon, P. Fay, D. Jena, H. Xing,“AlN/GaN Insulated-Gate HEMTs with 2.3 A/mm Output Current and480 mS/mm Transconductance,” IEEE Elec. Dev. Lett. 7, 29, 2008.

36. C.Y. Change, S.J. Pearton, C.F. Lo, F. Ren, I.I. Kravchenko, A.M. Dabiran,A.M. Wowchak, B. Cui, P.P. Chow, “Development of enhancement modeAlN/GaN high electron mobility transistors”, Appl. Phys. Lett., 94, 263505,2009.

37. A.M. Dabiran, A.M. Wowchak, A. Osinsky, J. Xie, B. Hertog, B. Cui, D.C.Look, P.P. Chow, “Very high channel conductivity in low-defect AlN/GaNhigh electron mobility structures,” Appl. Phys. Lett. 93, 082111, 2008.

38. M. Higashiwaki, T. Mimura, T. Matsui, “Enhancement-Mode AlN/GaNHFETs Using Cat-CVD SiN,” IEEE Trans. Elec. Dev. no. 6, vol. 54,June 2007.

39. D.A. Deen, S.C. Binari, D.F. Storm, D.S. Katzer, J.A. Roussos, J.C. Hack-ley, T. Gougousi, “AlN/GaN insulated gate HEMTs with HfO2 gate dielec-tric,” IEE Elec. Lett. no. 8, vol. 45, Apr. 2009.

40. web address: http://www.ioffe.rssi.ru/SVA/NSM/Semicond.

41. J. Robertson and B. Falabretti, “Band offsets of high K gate oxides on III-Vsemiconductors”, J. Appl. Phys. 100, 014111, 2006.

42. S. Luryi, “Quantum capacitance device”, Appl. Phys. Lett., 52, 6, 1988.

43. G. Li, T. Zimmermann, Y. Cao, C. Lian, X. Xing, R. Wang. P. Fay, H.G.Xing, D. Jena, “Threshold Voltage Control in AlGaN/AlN/GaN HEMTs byWork-Function Engineering”, IEEE Elec. Dev. Lett., Vol. X, No. X, 2010.

39

44. E.H. Nicollian and J.R. Brews, “MOS (Metal Oxide Semiconductor) Physicsand Technology”, Wiley-interscience, New York, 1982.

45. L. M. Terman, “An investigation of surface states at a silicon/silicon oxideinterface smploying metal-oxide-silicon diodes”, Sol. Stat. Elec., Vol. 5,285, 1962.

40

CHAPTER 3

THE SINGLE-HETEROJUNCTION AlN/GaN HEMT

3.1 The AlN/GaN HEMT

AlN/GaN HEMTs were realized through a multi-faceted approach as a way

to study various specific aspects of their design1. In particular, a study focusing

on heterostructure and HEMT performance dependence on the AlN thickness

spectrum was performed (section 3.1.1). HEMTs grown on GaN templates on

SiC were also explored and showed some of the best frequency performance to

date for the insulated-gate AlN/GaN HEMT (section 3.1.2). The comparison of

various gate dielectrics discussed in the previous chapter led to the development

of the first Ta2O5/AlN/GaN HEMT with record off-state breakdown voltage for

the AlN/GaN-based HEMT (section 3.1.3). The “first generation” AlN/GaN

HEMTs fabricated at Notre Dame yielded record setting drain current density

and transconductance which subsequently set a new benchmark for the AlN/GaN

HEMT and is discussed in (section 3.1.4). Lastly, a discussion on the physics that

govern the current-voltage functionality is given as a means to better understand

the intrinsic operation of the insulated-gate AlN/GaN HEMT and relies heavily

on the analytical development of the I-V characteristics given in Appendix B.

1Heterostructures used during these studies were grown by MBE at the Naval ResearchLaboratory unless otherwise noted.

41

3.1.1 A comparative study on the thickness of the AlN barrier on device perfor-

mance

In an effort to systematically characterize the influence of the AlN cap thickness

on HEMT performance, a series of samples were grown with varied AlN barrier

thicknesses that encompassed the low Rsh window reported by Cao [46]. A similar

study was reported by Higashiwaki for varied barrier thicknesses in AlGaN/GaN

HEMTs [47]. The samples were grown on HVPE GaN substrates provided by

Kyma Technologies which have been shown to allow a reduced dislocation density

(∼ 106 cm−2) owing to the lattice match between the substrate and epitaxial

layers. The layer structure followed a 1.5 nm AlN nucleation layer, a 1 µm Be-

doped GaN, 300 nm UID GaN, and an AlN cap. The AlN cap thicknesses were

1.5, 2.1, 3.0, 3.6, 4.5, 6.0, and 7.5 nm and were chosen to capture operation from

both extremes of the functional range of AlN thickness. The thin AlN nucleation

layer was chosen based off work by Cao that showed a reduced buffer current

which detrimentally acts as a parallel conduction path to the 2DEG channel [48].

Homo-epitaxy growth of UID GaN on HVPE GaN has been shown to have a

high Oxygen density which acts as a shallow level donor species thus causing

buffer leakage current to be high. To mitigate high buffer currents Be-doping

has been shown to be successful as a compensating dopant species for n-type

unintentionally-doped (UID) GaN for the reduction of inter-device isolation [49].

Devices were fabricated using standard lithographic techniques and included

a shallow pre-metallization etch prior to contact metallization for Rc reduction.

After mesa isolation a 7.4 nm thick blanket layer of ALD HfO2 was deposited and

1 µm long optical gates were applied. Room temperature on-wafer Hall measure-

ments were taken before and after HfO2 deposition and the results are shown in

42

Figure 3.1. 2 × 2 µm AFM scans of varied AlN barrier layers.

Fig. 3.2 alongside similar results reported by Cao [46]. Prior to HfO2 deposition

sheet resistance (Rsh), Hall mobility (µ), and sheet charge density (ns) were in

excellent agreement with previous reports [46], [50]. Namely, the 1.5 and 2.1 nm

AlN barrier samples show low charge density and mobilities that suffer from opti-

cal phonon, remote ionized impurity, and remote surface roughness scattering as

shown by Cao [51] which yield sheet resistances which were too high to accurately

measure. Additionally, lattice relaxation was observed through cracks observed

by AFM (Fig. 3.1) starting with the 4.5 nm AlN barrier and became worse for

thicker barriers causing a drastic reduction in mobility. Interestingly, after the

HfO2 was deposited a significant increase in 2DEG charge density was observed

for the 1.5 and 2.1 nm AlN barrier samples which allowed an appreciable reduction

in Rsh as shown by the arrows in Fig. 3.2. It appears this effect is in alignment

43

Figure 3.2. Room temperature sheet resistance determined from Halleffect measurement for varied AlN barrier thickness.

44

with the partial satisfaction of dangling surface bonds which cause depletion of

the 2DEG (e.g. surface state passivation). In both cases, before and after HfO2,

the low Rsh window was observed in very close agreement with Cao’s report. A

record low RT sheet resistance of 130 Ω/ (µ = 1800 cm2/Vs, ns = 2.3×1013

cm−2) was measured after HfO2 deposition within the window, specifically at the

3.0 nm AlN thickness. Moreover, both the mobility and sheet resistance trends

with AlN barrier thickness are in close agreement with Cao’s report with the only

notable exceptions being, 1) the upper limit for the Rsh window is ∼0.5 nm less

making the total window for the HVPE GaN sample set slightly narrower than

that shown by Cao and, 2) the maximum achieved mobilities for the samples

grown on HVPE GaN substrates are ∼100 cm2/Vs higher than Cao’s structures

grown on GaN/Sapphire templates. Both of these observations are speculated to

be due to the homo-epitaxial growth on HVPE GaN which results in a significantly

lower dislocation density in the buffer causing reduced dislocation scattering of

the 2DEG electrons but simultaneously less strain relief of the AlN/GaN hetero-

junction, causing precipitous cracking of the AlN barrier layer. Further study is

necessary to verify these mechanisms.

Drain current density characteristics for the five barrier thicknesses which most

closely encompassed the low Rsh window is shown in Fig. 3.3. Corresponding

transfer characteristics are shown in Fig. 3.4 (left). From the figures it is appar-

ent that the optimal AlN thickness for maximizing drain current and transconduc-

tance occurs at ∼ 3 nm where 2D charge density and mobility are both maximized.

However, if considering only extrinsic operation, contact resistance is also a prac-

tical consideration that serves to reduce extrinsic gm and forces a larger knee

voltage.

45

Figure 3.3. Drain current characteristics for varied AlN barrierthicknesses.

Figure 3.4. Transfer characteristics for varied AlN barrier thicknesses(left), and two-probe I-V measurements of insulated C-V dots for the

various AlN barrier thicknesses (right).

46

Figure 3.5 shows comparatively, the trends for various device parameters with

the varied AlN barrier thickness. Contact resistance scaled non-linearly with in-

creasing AlN barrier thickness in fair agreement for obtained RC values with the

ohmic contact study discussed in the previous chapter. Contact resistance has a

strong dependency on AlN thickness but is found to be slightly higher for thin-

ner as-grown barriers. This occurrence is speculated to be due to a lower 2D

charge density below the contacts for thinner barriers resulting in a higher RC

despite more favorable metal diffusion conditions. The combination of the RC

and Rsh-trends ultimately resulted in drain (Imax, Vknee) and transfer (V th, gm)

characteristics trends that followed similarly to RC-Rsh functionality. Therefore,

the 3 nm AlN barrier thickness offers a layer structure that is electronically opti-

mized for the highest current density and transconductance within the minimum

Rsh window.

It is interesting to note that gate leakage current was higher for the thicker

(4.5+ nm) barriers, causing a reduced gm and impairing drain I-V curves at forward

gate bias. This is thought to be due to cracking (i.e. lattice relaxation) in the

barrier layer for the thicker barriers which provide current paths through the

barrier along dislocation boundaries for increased gate current leakage despite the

HfO2 insulation. Besides mobility degradation, this mechanism is the cause for

an upper limit in the range of AlN barrier thicknesses when fabricating HEMTs

from these layer structures. Furthermore, two-probe I-V measurement results

of insulated C-V dots are shown in Fig. 3.4 (right). The 3.6 nm and 4.5 nm

barrier thicknesses showed an abrupt increase in reverse-bias gate current at -6V

suggesting that even though the total gate-to-channel thickness is increasing, the

gate current is a function of the cracking that occurs in the thicker AlN barriers.

47

Figure 3.5. Measured sheet resistance (a), 2D charge density (b),mobility (c), contact resistance (d), maximum drain current density (e),

and transconductance (f) as a function of AlN barrier thickness.

48

In terms of DC characteristics the final result is that the 3.0 nm AlN barrier

sample showed a current density of 1.8 A/mm and transconductance of nearly

300 mS/mm for 1 µm long gates (LSD = 5 µm). These values are the highest

of their kind for this gate length. It is worthwhile to note that Zimmermann et

al. [52] demonstrated comparable DC performance with a 4 nm thick AlN barrier

and 3 nm thick SiN gate insulation. This shows that by leveraging the high

charge density (∼ 3 × 1013 cm−2) and improvements in mobility (1800 cm2/Vs)

by homo-epitaxial growth on HVPE GaN substrates, the low sheet resistance

window demonstrated by Cao and Jena [46] can be exploited to yield exceptional

DC HEMT characteristics from the AlN/GaN HEMT.

3.1.2 AlN/GaN HEMTs grown by MBE on MOCVD GaN/SiC templates

The most typical substrates used for epitaxial growth are sapphire or SiC. SiC

has it’s advantage over sapphire from the standpoint that its thermal conductivity

is much higher than sapphire’s. This allows heat generated by the HEMT channel

during current conduction and high electric field to be transported away more

efficiently and subsequently the avoidance of the detrimental effect of self-heating

which causes unwanted negative output conductance (through a reduction in high-

T mobility/electron velocity) and a reduction of output power.

GaN was grown by MOCVD on a 2-inch 4H-SiC substrate and followed by

the growth of the AlN/GaN (3 / 234 nm) heterostructure by MBE2 as a way to

utilize the beneficial thermal properties of SiC. An 11 nm thick Al2O3 was blanket

deposited over the entire sample surface by ALD and both optical and e-beam

defined gates followed. Various gate lengths of 1.5 µm, 250, 200, and 145 nm were

2MBE growth was performed by Y. Cao at the University of Notre Dame and is gratefullyacknowledged.

49

Figure 3.6. DC transfer (a) and drain (b) characteristics (left) and smallsignal frequency characteristics (right) of the AlN/GaN HEMT grown on

GaN/SiC templates.

written by electron-beam lithography and electron-beam metal deposition. After

Al2O3 deposition the room-temperature Hall properties of the 2DEG were found

to be ns = 2.8 × 1013 cm−2, µ = 1770 cm2/Vs which yielded a sheet resistance of

Rsh = 165 Ω/. C-V analysis revealed a lower than expected relative dielectric

constant for the Al2O3 of 3.6. It is speculated that this occurred due to an aging

batch of TMA precursor used for the ALD deposition. Circular transmission line

method (CTLM) measurements showed Rc = 0.9 - 1.1 Ω-mm and Rsh values that

were commensurate with those of Hall measurements.

The outstanding Hall characteristics resulted in HEMT devices that showed

maximum drain current density of 1.5 A/mm and 325 mS/mm extrinsic transcon-

ductance (Fig. 3.6). Since the UID GaN buffer was grown homo-epitaxially on

the GaN-terminated template surface without a pseudomorphic nucleation layer,

the regrowth interface electrical properties (dislocation density, additional charge

distribution, etc.) were unknown. Although the gate current is sufficiently low, as

50

can be seen in Fig. 3.6, it does not match the drain current in sub-threshold which

indicates a rather large parallel conduction current through the buffer. Further

C-V analysis did not show a buried charge distribution at the regrowth interface.

Despite a low dielectric constant of the Al2O3 gate insulator, small signal

frequency performance was notable with ft/fmax of 92/115 GHz, respectively.

A ground-signal-ground open calibration with the geometry and layer structure

equivalent to the HEMTs measured, was simulated using ADS/Momentum3. From

the simulated S-parameters, data was fit using AWR Microwave Office to a 3-

capacitor circuit which signifies the pad parasitic capacitances CGSp, CGDp, and

CDSp [53]. These three components were de-embedded from the measured S-

parameter data which yielded values for ft and fmax of 104 and 115 GHz, re-

spectively. The pre and post results of de-embedding can be seen in Fig. 3.6

(right) which highlights the improvement in frequency performance from remov-

ing the bonding pad impedance. These values are demonstrative of some the best

frequency performance for an insulated-gate AlN/GaN HEMT to date [54], [55].

The resulting device performance is a testament to the quality of material and

geometric scalability which leads to the remarkable frequency performance capa-

bility of the AlN/GaN HEMT. Further improvement in DC and RF performance

is anticipates with the reduction of Rc and higher dielectric constant of the gate

insulator.

3.1.3 Ta2O5 insulated gate AlN/GaN HEMT

In order to take advantage of the outstanding current density capability of

these devices for high-power applications the accessible bias voltage range must

be increased. HEMTs designed with a thin (< 6 nm) AlN barrier require a gate

3S-parameter simulation was performed by Y. Tang at the University of Notre Dame.

51

insulator for gate leakage current suppression due to an otherwise prevalent tunnel-

ing current. Recent reports have demonstrated the use of atomic layer deposited

(ALD) Al2O3 or HfO2 for gate insulators to the AlN/GaN HEMT [56, 57]. Yet

the off-state breakdown voltage has remained low due primarily to gate insulator

failure leading to premature off-state breakdown. One of the highest breakdown

voltage reports to date for an AlN/GaN HEMT stated a maximum VDS of 45 V

and a resulting output power density of 850 mW/mm at 2 GHz with VDS = 15 V

[58]. However, the reported devices showed soft pinched-off characteristics and a

low drain current density of only 380 mA/mm.

One common technique used for increasing breakdown voltage in HEMTs is

the use of field plates.[59] However, metallurgical gate extensions drastically in-

crease parasitic capacitance in the device and thus hamper frequency performance.

Significant progress has been made with ALD oxides, including Ta2O5, for Si

and GaN-based metal-oxide-semiconductor test structures[60] since they provide

a high dielectric constant film and are electronically robust. However, Ta2O5 has

yet to be investigated for the purpose of gate insulation in GaN HEMT struc-

tures. In an effort to improve the range of applied bias voltage of an insulated-

gate AlN/GaN HEMT while simultaneously taking advantage of the structure’s

inherent scalability, a 6 nm thick ALD Ta2O5 film was deposited for gate current

suppression on an AlN/GaN structure with a 3.5 nm thick AlN cap (Fig. 1.2).

With an effective band-gap of ∼4.4 eV,[61] a relative dielectric constant value [60]

as high as 20, and a critical breakdown field of approximately 4 MV/cm, Ta2O5

appears favorable for gate insulation where vertical down-scaling imposes high

electric fields conditions.

III-Nitride epitaxial layers were grown by plasma assisted molecular beam

52

Figure 3.7. Drain characteristics of the Ta2O5/AlN/GaN HEMT(destructive breakdown occurs at 96 V for VGS = -7 V).

epitaxy (MBE) on a 2-inch semi-insulating 6H-SiC substrate using procedures

similar to those described previously [49]. Three cycles of Ga deposition and

desorption like that described by Brandt et al.,[62] were performed in the growth

chamber prior to growth. A 60 nm AlN nucleation layer was grown first, followed

by a 1 µm GaN buffer and a 3.5 nm AlN barrier layer. The AlN barrier thickness

was chosen on the basis of work by Cao et al. who showed a minimum in sheet

resistance in single heterostructure AlN/GaN for AlN thicknesses between 3 - 4.5

nm [46]. All layers were grown without interrupts or doping and at a substrate

temperature of 730 C.

Device processing was initiated by an ohmic-first sequence. Ohmic contacts

were formed by a shallow pre-metallization etch prior to the deposition of a

53

Ti/Al/Ni/Au metal stack and an 800 C anneal [63]. This process yielded a con-

tact resistance (Rc) of 0.7 Ω-mm as measured by the transmission line method. A

6 nm ALD Ta2O5 layer was blanket deposited following a 100 nm deep, Cl-based,

mesa isolation etch. Ta2O5 films were prepared using an ALD process utilizing

pentakis(dimethylamino)tantalum (PDMAT) and water as reagents at 250 C.

The films were deposited in a hot wall flow tube type reactor [64]. A constant

flow of ultrahigh purity N2 gas (15 sccm) was used to maintain a flow tube process-

pressure of 200 mTorr during deposition and a 30 second purge time was used to

separate the reagent pulses. Under these conditions linear growth was observed

and a growth rate of 0.6 A/cycle was measured using spectroscopic ellipsometry.

As-deposited films were found to be amorphous and slightly over-oxidized (O:Ta

= 3.0 ± 0.3) with a bonded carbon content of ∼5 at.%. Atomic force microscopy

of the Ta2O5 film deposited on the AlN surface had a RMS roughness of 0.58

nm for the 1 × 1 µm scan shown in Fig. 1.2 (RMS roughness of the as-grown

AlN surface was 0.45 nm). Ta2O5 films received no post-deposition thermal treat-

ments. Electron-beam lithographic definition of sub-micron gates with a T-profile

concluded the device processing. Pertinent geometric dimensions of the HEMTs

were source-drain separation (LSD) of 3 µm, gate width (Wg) of 2×25 µm, and

gate footprint length (Lg) of ∼150 nm.

On-wafer room-temperature Hall measurements were taken after the Ta2O5

deposition and sheet resistance was found to be 356 Ω/ with a 2D electron

density and mobility of 2.2 × 1013 cm−2 and 800 cm2/Vs, respectively.

Drain characteristics are shown in Fig. 3.7 with a maximum current density

of 1.37 A/mm at VGS = 4 V. Corresponding transfer characteristics were taken

at VDS = 8 V and shown in Fig. 3.8 (a). A maximum extrinsic transconductance

54

Figure 3.8. (left) Drain and transfer characteristics and, (right) smallsignal frequency performance of the Al2O3/AlN/GaN HEMT grown on

GaN/SiC templates.

(gm) was measured to be 315 mS/mm. Taking into account source contact (0.7

Ω-mm) and access (0.21 Ω-mm) resistance, this value corresponds to an intrinsic

gintm ∼ 450 mS/mm. Gate current was found to be the limitation in off-state drain

current (Fig. 3.8 (b)). Low (10−8 A/mm) parallel conduction in the buffer layer

was measured.

Off-state breakdown voltage is often defined by the criterion that 1 mA/mm

drain current density is reached in the sub-threshold state. According to this

metric, VBD ∼30 V. However, the Ta2O5-insulated devices ultimately sustained

drain-source voltages of up to 96 V under off-state gate-source bias (VGS = -7

V) before destructively breaking down (Fig. 3.7) demonstrating an ∼5X increase

of this particular characteristic over other reports [67], [55], [56], [54], [57]. It is

noted that after ∼60 V drain voltage sweep in off-state conditions, the device’s

low-voltage IDS and gm degraded by 30 - 50%. The pre-DBD degradation is

speculated to be the result of a localized breakdown of the AlN barrier at the

55

drain edge of the gate under reverse bias forcing the Ta2O5 to sustain the entire

voltage drop. This would still maintain transistor action after AlN breakdown

but with degraded transport across the gated region. It is therefore plausible

that DBD occurs after AlN breakdown when the local electric field is of sufficient

intensity to cause electrical failure of the Ta2O5 film.

Terman’s method implicitly probes interfacial charge states which are below

the ac perturbation frequency thus, fast states are not accounted for. These

“slow” states apparently do not impose deleterious effects on small signal fre-

quency performance as can be seen in Fig. 3.8 (b). Small signal S-parameter

characteristics were taken at the bias point that yielded maximum gm. Extrinsic

values of unity current gain frequency (ft) and maximum frequency of operation

(fmax) were measured to be 55 GHz and 115 GHz, respectively. By employing the

ColdFET and standard Y-subtraction technique,[65] pad parasitics were modeled

(table inset, Fig. 3.8 (b)) and the de-embedded value of ft was determined to be

75 GHz. According to these results, Ta2O5 may offer a nascent solution to enable

the insulated-gate AlN/GaN HEMT to perform at frequency levels dictated by

it’s scalability with power handling capability that can contend with that of the

AlGaN/GaN hetero-system.

3.1.4 First-generation insulated-gate AlN/GaN HEMTs

While the generational improvements at NRL are noteworthy, the caliber of

AlN/GaN device performance achieved at Notre Dame is of the highest reported

to date in terms of maximum current density and transconductance. As discussed

previously, the AlN/GaN HEMT has the highest polarization charge of the III-

Nitride HEMTs which lends itself to the highest 2D mobile electron densities.

56

Given appreciable mobility (∼1600 cm2/Vs) the 2D conductivity yields the lowest

of any III-Nitride heterojunction measured or modeled. With such low conductiv-

ity a very high current density is expected (see next section for discussion). The

highest current density to date by any FET is shared by two structures in the

III-Nitride system, the InAlN/GaN HEMT [66] and the AlN/GaN HEMT [67].

Results published in reference [67] are now discussed.

MBE growth of the structure was initiated by the growth of a 200 nm thick

UID GaN buffer on a GaN template on sapphire. Following buffer growth was

a 3.5 nm pseudomorphic AlN cap. The structure resulted in an as-grown sheet

resistance of 166 Ω/ (ns = 2.75×1013 cm−2, µ = 1367 cm2/Vs). Standard

lithographic and III-Nitride processing techniques were used and Ti/Al/Ni/Au

contacts were formed by rapid thermal annealing. Transmission line measurements

yielded values for Rc and Rsh of 1.1 Ω-mm and 171 Ω/, respectively. Sub-

micron gate (Lg = 250 nm) definition was defined through e-beam lithography

and a sequential oxide-metallization of Al2O3/Ni/Au was deposited by e-beam

evaporation. HEMT access regions were left un-passivated.

Shown in Fig. 3.9 are the drain (left) and transfer (right) characteristics of an

AlN/GaN HEMT device grown and fabricated at the University of Notre Dame

[67]. A maximum extrinsic transconductance of 480 mS/mm was measured with

a maximum current density of 2.3 A/mm for a LG = 250 nm HEMT. A conser-

vative estimate for intrinsic gm taking only source contact resistance into account

yielded ∼0.8 S/mm. Pulsed-drain I-V measurements (Fig. 3.9 (left)) showed only

minor self heating for the structures grown on GaN/sapphire templates although

the pulse width was rather long at 500 µs. A 3 nm thick electron beam deposited

Al2O3 gate dielectric was used for gate current insulation. A major caveat to

57

Figure 3.9. Drain (left) and transfer (right) characteristics of theAlN/GaN HEMT grown and fabricated at U. of Notre Dame.

this structure was the presence of a large buffer current that led to a high ∼200

mA/mm off-state current at VGS = -10 V and VDS = 8 V due to parallel con-

duction at the GaN re-growth interface which was located ∼200 nm below the

channel. This resulted in a larger negative threshold voltage in order to pinch off

the HEMT. This growth issue has since been resolved by a novel approach that

utilizes a dopant-free polarization engineered AlN nucleation layer [48]. Small

signal frequency performance was measured to yield an extrinsic ft = 52 GHz and

fmax = 60 GHz. These values are lower than expected when considering such high

transconductance and current density. It was found to be due to the gate bonding

pad positioned such that it was connected to the leaky buffer layer. Thus, the RF

gate signal was forced to modulate the HEMT channel as well as suffer from the

lossy buffer below it.

The AlN/GaN heterojunction has the capability to achieve a charge density

in excess of 2 times higher than its only contender, the InAlN/AlN/GaN het-

erostructure. Accompanied by a high mobility (1000 - 1800 cm2/Vs for devices

58

demonstrated in this work), the AlN/GaN HEMT has the capacity to supersede

all other III-N heterostructures in current density and transconductance. Cur-

rently, there is a medley of other issues (high contact resistance, gate insulation,

low breakdown voltage, etc.) that are hampering the AlN/GaN HEMT to come

to full technological maturity. Nonetheless, it is anticipated that upon resolve

of these issues, a full appreciation of this heterostructure for HEMT applications

may be realized, as has been partially demonstrated by the outstanding HEMT

device performance shown.

3.2 Implications of peripheral resistance on HEMT intrinsic I-V characteristics

Shockley’s 1952 theoretical development of the current-voltage operation for a

uniformly doped channel FET [68] was the first physical model-based analytical

assessment for a solid-state FET. In his work he utilized a gradually changing

potential profile along the transistor’s channel which allowed the simplification

of the mathematics involved and so termed it the gradual channel approximation.

Supporting work by Grebene and Ghandhi followed in 1969 with the generalization

of Shockley’s original theory by developing the pinched channel characteristics for

an arbitrary doping profile [69]. The combination of these two seminal works have

become the benchmark for understanding the physics of FET operation. It is

worthwhile to note that in both of these works the field effect device had a doped

channel from whence the mobile carriers were derived, which is in contrast to the

polarization-doped channel employed by the III-Nitride HEMT. This difference

does little to change the ultimate I-V functionality between the doped-channel

FET and polarization-doped HEMT but is significant with respect to the I-V

derivation for the HEMT which is presented in Appendix B.

59

The drift velocity of electrons in the channel is of fundamental importance as

it directly determines the current density and speed at which a transistor may

operate. A back of the envelope calculation for current density assuming typical

values for an AlN/GaN HEMT with charge density of 2.5× 1013 cm−2 and electron

velocity of 1 × 107 cm/s gives a current density (J = qnsve) of 4 A/mm. This

value is a factor of 2 higher than the highest value reported for a III-Nitride HEMT

based on today’s literature [67],[66]. Therefore, the question arises; from where

does this discrepancy originate? The following pedagogical discussion seeks to

shed light on this issue.

3.2.1 Ideal three-terminal current-voltage model

The basis for the analytical development of the polarization-doped HEMT be-

gins with Ohm’s law ( J = σF||) along the channel, in combination with Poisson’s

equation (O2V = σ/ε) approximated to 1D in the vertical dimension. The latter

approximation is based off the assumption that the change in potential along the

channel is much less than the change taken vertically through the gated region

(∂V/∂x << ∂V/∂y), and is the well known gradual channel approximation (GCA)

[70],[68],[69].

Poisson’s equation may be utilized to obtain a relationship for the charge

control relationship between applied voltage and charge below the gate. The

result has been given in Chapter 2 but is restated here in terms of the charges

present in the heterosystem

− VGS + V (x) +

(ΦB + ΦoA −

∆Ecq

)+QπA

Cox−Qπ

net

(1

Cox+

1

CAlN

)+ qns

(1

Cox− 1

CAlN+

π~qm∗

)+ (qns)

2/3

(9π~q

8εG√

8m∗

)2/3

= 0, (3.1)

60

where ΦB and ΦoA are the metal-oxide barrier and oxide-AlN offset, respectively,

∆Ec is the AlN/GaN conduction band discontinuity, εG = εrεo is the dielectric

constant of GaN, Cox and CAlN are the fixed capacitances per unit gate area asso-

ciated with the gate oxide and AlN barrier, respectively, m∗ is the effective mass,

QπA is the fixed polarization (spontaneous and piezoelectric) charge of the strained

AlN, Qπnet is the net polarization charge at the AlN/GaN interface, and qns is the

2D mobile charge comprising the channel. V (x) accounts for the distributed volt-

age drop across the channel due to the voltage difference between the source and

drain contacts. This relation is used to determine the I-V functionality in general.

Due to the cubic nature of ns the roots are difficult to handle and the final form

for the drain current, ID = f(VG, VD, VS) is most conveniently written in terms

of 2D charge density, ID = f(ns(VG, VD, VS)). For the “ideal” III-Nitride HEMT

(i.e. no external resistance, electron velocity limitation, charge unaccounted for)

the drain current density is written,

JDS =qµ

Lg

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)n2s +

2

5

(9π~q2

8εG√

8m∗

)2/3

n5/3s

]ns,D

ns,S

, (3.2)

where JDS is the drain current per unit gate width, Lg is the metallurgical gate

length, µ is the two-dimensional mobility and is assumed to be constant, and the

drain-source-gate voltage functionality is contained in the determination of ns as

given in Eq. A.4.

Equation 3.2 shows linear dependency on mobility and gate length, as ex-

pected. The typical square law dependence of drain current on VGS found in the

doped long-channel HFET derivation is not clearly apparent for the polarization-

doped HEMT. Square law dependence is indeed contained in Eq. 3.2 through the

n2 term (when Eq. A.4 is also considered) but with the addition of the n5/3 term

61

Figure 3.10. AlN/GaN HEMT I-V characteristics showing the effectwhen two different mobilities are considered.

that arises from the integration of the charge of the first populated sub-band.

This second term frustrates the clarity of simple square law behavior but does not

diminish its functionality.

A comparison of simulated I-V characteristics is given in Fig. 3.10 showing the

result of two different mobilities, 1600 2/V s (left) and 800 cm2/V s (right). No

external resistance or velocity saturation was assumed. Other parameters used

were those typical of the AlN/GaN HEMT; Lg = 1 µm, oxide thickness of 10 nm,

AlN thickness of 3 nm, and a metal work function chosen such that a zero gate

bias 2D charge density of 2.13 × 1013 cm−2 was present in the channel. These

values will be used for the following representative simulations.

It is obvious that mobility simply scales current density. However, the values

shown for current density in Fig. 3.10 are far higher than the highest current

62

density for the AlN/GaN HEMT reported in literature [67],[52]. This is because

the derivation above is for the intrinsic portion of the HEMT which does not take

into account external resistance and is ideal from the standpoint that no limitation

on current transport (i.e. electron velocity saturation) is imposed other than the

fixed low field mobility. The following sections discuss the integration of these

necessary practicalities.

3.2.2 Velocity saturation

A key assumption made by Shockley [68] and also Grebene [69] is that drain

current saturation in a HEMT is the direct result of the electron drift velocity be-

coming limited (saturation) at a specified lateral electric field within the transistor

channel. Velocity saturation occurs when the measured electron drift velocity be-

gins to deviate from the linear low electric field relationship (ve = µF||). This

property is a phenomenological occurrence at high electric field that arises from

the various scattering elements within the crystal. When the electric field achieves

such a level that scattering prohibits further increases in velocity with higher fields,

velocity saturation has occurred. Figure 3.11 shows velocity-field curves for several

different semiconductor systems. The detriment that velocity saturation imposes

is that it prevents the “peak” electron velocity from being achieved (which has

been calculated to be as low as ∼1.5 × 107 cm/s for GaN [71]) and therefore

sets a restraint on current density, transconductance, and ultimately frequency

performance of a GaN-based transistor.

For GaN channels, the room temperature saturation velocity has been shown

to be reduced from theoretically calculated levels by the interplay between optical

phonon scattering and the hot phonon bottleneck restricting phonon decay [72, 73].

63

Figure 3.11. Electron velocity versus electric field for various commonsemiconducting systems. The scattering limited velocity approximation

is illustrated for the GaN curve.

This sets a “choke” on the attainable electron velocity. Assuming only optical

phonon scattering limited velocity, the saturation velocity is taken to be vsat ≈

µOPFcrit, where Fcrit is the critical electric field that dictates the onset of phonon-

limited velocity saturation. Taking the optical phonon scattering time to be τOP

= 0.3 ps sets an optical phonon limited mobility ceiling4 of ∼ 2600 cm2/Vs and a

scattering limited velocity saturation of ∼ 1.7 × 107 cm/s [73].

A common approximation made on the velocity-field characteristic is called

the scattering-limited velocity approximation (SLVA) [70],[74],[75] and is given by

4The “ceiling” is meant as if only optical phonons limited electron momentum scattering time.In reality, and according to the drude model, multiple scattering mechanisms add to further limitelectron drift velocity and subsequently mobility. The OP-limited ceiling is assumed here as azero-order approximation for velocity saturation in GaN.

64

ve =µF

1 + µFvsat

, (3.3)

where vsat is the saturated drift velocity and µ is the low field mobility. The ap-

proximation can be seen in Fig. 3.11 where the intersection of the low-field veloc-

ity tangent and the saturated velocity lines define the approximated velocity-field

curve. This approximation has been used to explain some short-channel and low

barrier aspect ratio effects [70],[74] where frequency performance deviates from

the typical long-channel rule, ft ∼ 1/Lg. The approximation seeks to give a phe-

nomenological functionality to electron velocity saturation and can be understood

by considering the case where the saturation velocity is very large compared to

the drift velocity (vsat >> µE). This is the low-field condition. Then the drift

velocity approximates to the low-field velocity. Conversely, when the electric field

is such that the drift velocity becomes comparable to vsat, the electron velocity

approaches the saturation velocity ceiling, as illustrated in Fig. 3.11.

By the inclusion of the SLVA in the GCA development of the drain current, a

modified low-field I-V functionality is obtained,

JDS =qµ

Lg + µVDS

vsat

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)n2s +

2

5

(9π~q2

8εG√

8m∗

)2/3

n5/3s

].

(3.4)

From Eq. 3.4 the original form of Eq. 3.2 is maintained through the bracketed

term. However, it is now scaled by a term that amounts to an effective gate length

(Leff = Lg+µVDS/vsat) which is saturation velocity-dependent and arises directly

from the inclusion of the SLVA. Therefore, it is reasonable to expect a reduction

in drain current when a saturated velocity is assumed. This can be seen in Fig.

3.12 where ideal I-V curves are shown as before and compared to the vsat-limited

65

Figure 3.12. AlN/GaN HEMT I-V characteristics showing the effectwhen velocity saturation is included.

66

HEMT I-V characteristics where vsat was assumed to be 1.7 × 107 cm/s. A drastic

reduction in current is calculated and shown. Interestingly, by including the SLVA

the flat saturated current characteristic no longer maintains its constant value with

respect to an increasing VDS but rather demonstrates a “negative conductance”

nature. This is due to Leff being VDS-dependent and having a greater scaling effect

at higher VDS-bias. Based on this observation and taking in mind self-heating

being a common argument for negative drain conductance, it seems reasonable

that some portion of negative conductance observed in measured data may be

due to the physical manifestation of a saturated electron velocity rather than

strictly self heating.

3.2.3 External resistance

The occurrence of high contact resistance (Rc) has already been discussed from

a practical standpoint in Chapter 2 and is a major limitation for the GaN-based

HEMT from reaching its intrinsic potential. The primary effect of external source

and drain resistance is to form a voltage divider which causes a reduced voltage

drop across the intrinsic region of the HEMT and thus, a reduced drain-source

current for a given external VDS bias. However, since the intrinsic portion of the

device becomes non-linear around the knee voltage, a non-linear voltage division

occurs across the intrinsic device. This effect is exacerbated by the typical asym-

metry in source and drain access resistances that arise from an offset gate. Access

resistance is taken as the additive combination of contact (Rc) and sheet resis-

tance in the access region (Rsh) of the device. Therefore, source access resistance

follows RS,a = Rc + (LSG/WG)Rsh and drain access resistance follows RD,a =

Rc + (LDG/WG)Rsh, where LSG and LDG are the source-to-gate and drain-to-gate

67

Figure 3.13. GCA model results of imposing external source and drainresistance on drain I-V characteristics (internal source and drain

voltages shown with respect to the applied external VDS on the right).

distances, respectively.

The determination of the resistance-afflicted I-V characteristics is found nu-

merically since the I-V functionality becomes recursive when the intrinsic cur-

rent is determined by the voltage drop across the access resistance. This can

be understood by considering the voltage drops in the resistor-HEMT circuit:

V intDS = V ext

DS − IDS(RS,a +RD,a). Since IDS is found by applying V intDS , the I-V rela-

tionship cannot be found in closed form. Therefore, the values for V intD and V int

S

are found through an iterative calculation (Jacobi method) using the equations:

V intD = V ext

D −RD,aIDS (3.5a)

68

V intS = V ext

S −RS,aIDS (3.5b)

The result is the reduction of the voltage drop across the intrinsic portion of the

HEMT. Figure 3.13 shows the result of imposing symmetric source and drain re-

sistances of 0.5 Ω-mm to the ideal I-V characteristics previously discussed. The

intrinsic drain and source voltages calculated for this external resistance can been

seen in the rightmost plot. It is noted that VS does not maintain a linear change

with the applied external voltage as does VD. This is due to VS not being implicitly

dependent on VD,ext but rather increasing only as a function of VS,ext. The reduced

voltage drop across the intrinsic portion causes less current to be carried through

the device and an apparent drop in maximum current. Additionally, since the in-

trinsic source voltage has changed, the intrinsic VGS has changed correspondingly.

The effect of a changed channel potential causes the obvious knee-voltage walkout

as well as a reduction in ∆IDS/∆VGS (i.e. transconductance).

3.2.4 Experimental correspondence to model and its validity

Two strong limiting factors, electron velocity saturation and external resis-

tance, have been discussed individually with respect to their implications on

HEMT I-V characteristics. The integration of these effects into one complete

analytical model will give insight into how to improve the device based on specific

criteria. This is provided that good agreement is had between the model and

measured results. The primary goal, however, is to reconcile the discrepancy of

the AlN/GaN HEMT between the anticipated current density based off a simple

consideration of Ohm’s law and those obtained in practice.

Figure 3.14 demonstrates quantitatively, the progression of the I-V charac-

69

Figure 3.14. Progression of the AlN/GaN I-V characteristics for theintrinsic device (left), the addition of the VLSA (middle), and the final

addition of external resistance.

70

teristics of the AlN/GaN HEMT when each limiting component is added to the

model. The leftmost plot shows the familiar ideal I-V calculation, the middle plot

includes the VLSA, and the rightmost plot includes both the VLSA and symmet-

ric external drain and source resistances of 0.5 Ω-mm. It is apparent that the

inclusion of a saturated electron velocity drastically reduces the maximum satu-

rated drain current and further so does the addition of external resistances. The

latter also shifts the knee voltage of the drain characteristics to higher values. The

choice of reasonably realistic parameter values for µ, Rc, and vsat, albeit some-

what arbitrary, causes an ∼ 80% drop in extrinsic drain current density down

to a realistic range (∼3 A/mm). Therefore, the question now is if by including

only the two limitations discussed, does the model show correspondence with the

characteristics of a measured HEMT?

The GCA model with the SLVA is plotted with the measured data of the

HfO2/AlN/GaN HEMT with the 3 nm AlN cap (shown in Fig. 3.15). The pa-

rameter values used in the fit were vsat = 1.3 × 107 cm/s, µ = 1840 cm2/Vs, RD

= 0.7 Ω-mm, RS = 0.2 Ω-mm, and ΦB was adjusted such that at VGS = 0V, the

value of ns corresponded to that measured by Hall. It can be seen that the near-

threshold characteristics show close correspondence with the model’s current level

in the saturation regime. In this range, the modeled current levels show agreement

with the measured data and the point of threshold for the model and measured

data correspond well. However, two noticeable deviations are recognized. The

first and most apparent is the deviation of the family of curves at higher (less

negative) gate-source voltages. The second is the deviation in the linear regime of

the model from the measured data where the knee voltage is approximated to oc-

cur at higher drain-source voltages. The dotted lines in Fig. 3.15 (a) demonstrate

71

Figure 3.15. GCA model fit to data (a) taken from Fig. 3.3 showing thelow-field deviation due to the SLVA (b). GCA fit to transfer

characteristics (c).

the case when the SLVA maintains it’s bias voltage dependence beyond pinch-off.

The solid lines indicate a bias-independent SLVA beyond the knee voltage. The

latter allows a more accurate fit closer to threshold.

The model appears to over-estimate the drain current for less negative VGS

biases. The over-estimation may be a partial result of a mechanism referred to as

“source starvation” [76]. If the model transfer characteristic is considered (Fig.

3.15 (c)) it is observed that the model drain current continually rises without

ceasing when VGS increases positively. This is untrue for a practical device. One

possibility for this occurrence is that when the gate-source bias in a HEMT in-

creases positively, the energy barrier along the channel between the gate and

source decreases to a point that it allows free flow of electrons from the source

contact to the drain contact. Beyond this gate-source bias, any additional VGS

will only minimally increase the electric field between the source and gate that

72

is responsible for electron flow in that region. Thus, a “compression” in drain

current occurs for increases in VGS.

While source starvation has been shown to limit further increases of IDS in

HFETs the primary cause of the IDS compression shown in this work is suggested

to be the interplay between the high interface state density observed (Ch. 2) and

the physical proximity of the 2DEG to the charging surface in these extremely

shallow buried channel HEMTs. It is speculated that due to the large charging

capacity of the interface states at the oxide/AlN interface, the Fermi level under

and at the drain edge of the gate is quasi-pinned and the pinning is dictated by

both the 2DEG density as well as the interface trap state density. When the traps

fill, the charge densities are overwhelming and VGS no longer causes a sufficient

change in ns and subsequently, IDS no longer increases. This “choked channel”

mechanism is illustrated by the band diagram in Fig. 3.16. To add to the issue

source resistance also exacerbates IDS compression, further limiting maximum

drain current and causes severe transconductance fall-off in HEMTs.

The nature of the SLVA accounts for the phenomenological occurrence of ve-

locity saturation, but in doing so imposes a deviation of the ve-E slope at low field

(i.e. mobility) causing an apparent lower mobility than the true low-field mobility

in a HEMT. This causes a positive shift in knee voltage (or underestimation of

the drain current at Vknee) and the “fanning out” of the drain characteristics at

low field that is usually not observed in an actual HEMT. This is the cause of the

non-correspondence of the model to the measured data at low field approaching

pinch-off5. This deviation is a direct result of an inaccuracy of the SLVA. Numeri-

5The usage of the word “pinch-off” is taken as originally used by Shockley and indicates thetransition point of the channel from linear operation to saturated operation where the mobilecharge in the channel is minimized. This is in contrast to the loosely used terminology oftentaken to mean sub-threshold or off-state characteristic.

73

Figure 3.16. Qualitative illustration of the conduction band diagramalong the channel showing the Dit-influenced channel choke leading toIDS compression at forward gate-source bias voltages. (a) Cross sectionalconduction band diagram in “flatband” or pinched off condition and, (b)

band diagram of the forward gate bias condition with Qit influence.

74

cal methods for improving this inaccuracy can be performed but have no physical

meaning and therefore are not used here.

As discussed in Chapter 2, one reality that follows from using an ex-situ oxide

deposited on the terminal AlN surface is the presence of a high interface state

trap density at the oxide/AlN interface. These traps may have implications on

RF performance but they predominantly serve to cause a more gradual slope of

the DC transfer characteristic of a HEMT (i.e. DC transconductance). Trap

states were not accounted for in the model given.

The occurrence of self-heating or heat-reduced mobility/electron velocity from

internal power consumption an often widely used explanation for negative drain

conductance in GaN-based HEMTs. As stated in Appendix B, the model assumes

a constant mobility for simplicity and does not take into account the effect of self

heating.

The model does a decent job to reconcile near-threshold I-V characteristics of

the insulated-gate AlN/GaN HEMT using parametric values extracted from mea-

surements (e.g. mobility, sheet resistance, contact resistance, etc.). In doing, so

the model, by including only two constraints, provides reconciliation for the lesser-

than-expected current density compared with theoretical estimations. Therefore,

it is concluded that the GCA and SLVA are applicable as shown above, but under

the limitation that only near-threshold values for VDS and VGS are used.

Other mechanisms have been discussed that account for non correlations be-

tween the model and measured I-V characteristics. Through this discussion several

points for improvement have been touched upon for the AlN/GaN HEMT. Con-

tact resistance is a major limitation causing knee voltage walkout, reduction in

internal VDS and therefore IDS, and enhancement of IDS compression in the sat-

75

uration regime. Reduction of contact resistance will greatly benefit both DC and

RF extrinsic device performance. Interface trap states act to reduce the transfer

characteristic slope and thus peak transconductance as well as promote transcon-

ductance roll-off and thus, ft linearity. The minimization of trap density should

improve, gm. Maximum drain current is determined by channel charge density,

mobility, contact resistance, and electron velocity in the intrinsic channel (i.e. be-

low the gate). For the AlN/GaN HEMT it has been shown that charge density

may be increased up to the theoretical polarization limit [46] by increasing the

AlN barrier thickness. Therefore, this parameter is not easily improved since it is

already at a ceiling value. However, mobility (as a measure of electron velocity)

can be improved by growth methods. Cao et al. has shown the sensitivity of mo-

bility to growth parameters and barrier thickness of these ultra-shallow channel

structures [51]. Section 3.1 in this chapter demonstrated a preferred AlN window

with respect to processing considerations and a 3 nm AlN barrier was determined

to be ideal. Further improvements in growth (substrate crystal type, parameters)

should serve to further increase mobility and thus saturated electron velocity.

3.3 Chapter 3 cited literature

46. Y. Cao and D. Jena, “High-mobility window for two-dimensional electrongases at ultrathin AlN/GaN heterojunctions”, Appl. Phys. Lett., 90, 182112,2007.

47. M. Higashiwaki and T. Matsui, “Barrier Thickness Dependence of ElectricalProperties and DC Device Characteristics of AlGaN/GaN HeterostructureField-Effect Transistors Grown by Plasma-Assisted Molecular-Beam Epi-taxy”, Jap. J. Appl. Phys., Vol. 43, No. 9, 2004.

48. Y. Cao, T. Zimmermann, H. Xing, and D. Jena, “Polarization-engineeredremoval of buffer leakage for GaN transistors”, Appl. Phys. Lett., Vol. 96,No. 4, 042102, 2010.

76

49. D.F. Storm, D.S. Katzer, S.C. Binari, B.V. Shanabrook, L. Zhou, D.J.Smith, “Effect of Al/N ratio during nucleation layer growth on Hall mo-bility and buffer leakage of molecular-beam epitaxy grown AlGAN/GANheterostructures”, Appl. Phys. Lett. 85, 3786, 2004.

50. I.P. Smorchkova, S. Keller, S. Heikman, C.R. Elsass, B. Heying, P. Fini, J.S.Speck, U.K. Mishra, “Two-dimensional electron-gas AlN/GaN heterostruc-tures with extrememly thin AlN barriers”, Appl. Phys. Lett., 77, 24, 2008.

51. Y. Cao, “Study of AlN/GaN HEMTs: MBE Growth, Transport Properties,and Device Issues”, Ph.D. Dissertation, University of Notre Dame, 2010.

52. T. Zimmermann, Y. Cao, D. Jena, H.G. Xing, “4-nm AlN barrier all binaryHFET with SiNx gate dielectric”, Int. J. High Spd. Elec. Sys., 19, 153,2009.

53. L.F. Tiemeijer and R.J. Havens, “A Calibrated Lumped-Element De-EmbeddingTechnique for On-Wafer RF Characterization of High-Quality Inductors andHigh-Speed Transistors”, IEEE Trans. Elec. Dev., Vol. 50, No. 3, 2003.

54. N. Onojima, N. Hirose, T. Mimura, T. Matsui, “Ultrathin AlN/GaN het-erostructure field-effect transistors with deposition of Si atoms on AlN bar-rier surface”, Appl. Phys. Lett. 93, 223501, 2008.

55. M. Higashiwaki, T. Mimura, T. Matsui, “AlN/GaN Insulated-Gate HFETsUsing Cat-CVD SiN”, IEEE Elec. Dev. Lett., Vol. 27, No. 9, 2006.

56. A.M. Dabiran, A.M. Wowchak, A. Osinsky, J. Xie, B. Hertog, B. Cui, D.C.Look, P.P. Chow, “Very high channel conductivity in low-defect AlN/GaNhigh electron mobility transistor structures”, Appl. Phys. Lett., 93, 082111,2008.

57. D. Deen, S. Binari, D. Storm, D. Katzer, J. Roussos, J. Hackley, T. Gougousi,“AlN/GaN insulated gate HEMTs with HfO2 gate dielectric”, Elec. Lett.,45, 8, 2009.

58. S. Seo, G.Y. Zhao, D. Pavlidis, Elec. Lett. 44, 3 (2008).

59. H. Xing, Y. Dora, A. Chini, S. Heikman, S. Keller, U. Mishra, IEEE ElectronDevice Lett. 25, 161 (2004).

60. L. W. Tu, W. C. Kuo, K. H. Lee, P. H. Tsao, C. M. Lai, Appl. Phys. Lett.77, 3788 (2000).

61. J. Robertson and B. Falabretti, J. Appl. Phys. 100, 014111 (2006).

77

62. O. Brandt, R. Muralidharan, P. Waltereit, A. Thamm, A. Trampert, H. vonKiedrowski, and K.H. Ploog, Appl. Phys. Lett. 75, 4019 (1999).

63. D.A. Deen, D.F. Storm, D.S. Katzer, D.J. Meyer. S.C. Binari, “Depen-dence of ohmic contact resistance on barrier thickness of AlN/GaN HEMTstructures”, Sol. Stat. Elec. 54, 613, 2010.

64. J.C. Hackley, J.D. Demaree, T. Gougousi, “Nucleation of HfO2 atomic layerdeposition films on chemical oxide and H-terminated Si”, J. Appl. Phys.102, 034101, 2007.

65. G. Dambrine, A. Cappy, F. Heliodore, E. Playez, “A New Method for Deter-mining the FET Small-Signal Equivalent Circuit”, IEEE Trans. Elec. Dev.,Vol. 36, No. 7, 1988.

66. F. Medjdoub, J.F. Carlin, M. Gonschorek, E. Feltin, M.A. Py, D. Ducatteau,D. Gaquiere, N. Grandjean, and E. Kohn, “Can InAlN/GaN be an alter-native to high power/high temperature AlGaN/GaN devices?” in IEDMTech., pp. 1-4, 2006.

67. T. Zimmermann, D. Deen, Y. Cao, J. Simon, P. Fay, D. Jena, and H.G.Xing, “AlN/GaN Insulated-Gate HEMTs With 2.3 A/mm Output Currentand 480 mS/mm Transconductance”, IEEE Elec. Dev. Lett. Vol. 29, No.7, 2008.

68. W. Shockley, “A Unipolar ”Field-Effect” Transistor”, Proc. of I.R.E., pp.1365-1376, 1952.

69. A.B. Grebene, S.K. Ghandhi, “General theory for pinched operation of thejunction-gate FET”, Sol. Sta. Elec., vol. 12, pp. 573-589, 1969.

70. S.M. Sze, in Physics of Semiconductor Devices, 2nd ed., Wiley Publishers,1981.

71. A. Matulionis, J. Liberis, I. Matulioniere, M. Ramonas, L.F. Eastman, J.R.Shealy, V. Tilak, A. Vertiatchikh, “Hot-phonon temperature and lifetimein a biased AlxGa1−xN/GaN channel estimated from noise analysis”, Phys.Rev. B 68, 035338, 2003.

72. D. Jena, “Polarization induced electron populations in III-V nitride semi-conductors; Transport, growth, and device applications”, Ph.D. dissertation,University of California, Santa Barbara, 2003.

73. J. Khurgin, Y.J. Ding, D. Jena, “Hot phonon effect on electron velocitysaturation in GaN: A second look”, Appl. Phys. Lett. 91, 252104 (2007).

78

74. P. Gray, P. Hurst, S. Lewis, and R. Meyer, “Analysis and Design of AnalogIntegrated Circuits”, Fourth Ed., New York, John Wiley and Sons, 2001.

75. F. Ali and A. Gupta, “HEMT Device Physics and Models”, in HEMTs andHBTs: Devices, Fabrication, and Circuits, Artech House, Inc., Norwood,MA, 1991.

76. M.V. Fischetti, S. Jin, T.W. Tang, P. Asbeck, Y. Taur, S.E. Laux, M.Rodwell, N. Sano, “Scaling MOSFETs to 10 nm: Coulomb effects, sourcestarvation, and virtual source model”, J. Comp. Elec. 8, 60, 2009.

79

CHAPTER 4

AlN/GaN HEMT SMALL-SIGNAL OPERATION

The central tool used to analyze the ac-functionality of any transistor is

its small signal equivalent circuit model. The small signal equivalent circuit is

developed using the ColdFET technique and standard biasing regimes for the

insulated-gate AlN/GaN HEMT. From the measured ColdFET S-parameters the

bias-independent equivalent circuit parameters can be modeled. These parame-

ters provide a means to accurately de-embed the effects of bonding pad impedance

elements from the small signal frequency performance data using the standard S-

parameter reduction technique [77]. The small signal frequency metrics that are

most commonly used for RF and microwave transistor amplifiers are the unity

current gain frequency, ft, defined as the frequency at which the current gain

becomes unity and the unity power gain frequency, fmax, defined likewise for the

power gain of the transistor. A simple lumped element expression for ft in terms

of the equivalent circuit elements typically follows, ft = gm× (2π(CGS +CGD))−1,

but has been more accurately expressed [78] in terms of both intrinsic and extrinsic

elements by,

ft =gm/2π

[(CGS + CGD)(1 + (RS +RD)/RDS)] + gmCGD(RS +RD). (4.1)

The de-embedding of bonding pad impedances from measured data become

even more necessary when the electron-beam defined sub-micron gate length re-

80

duces intrinsic gate capacitance and causes the combination of bonding pad and

parasitic access capacitance to dominate RF operation. The presence of a confor-

mal high-κ Al2O3 in the access and inter-pad regions of the device further increases

these parasitic capacitances thus causing a pronounced effect. It is shown that for

120 nm long T-gates and a 10 nm thick blanket deposition of the Al2O3 dielectric,

bonding pad capacitance is enhanced and reduces extrinsic ft by 10-15 GHz.

The small signal unity current gain frequency, ft can be separated into a sum

of transit and charging times involved with the constituent components of the

HEMT. Further ac-analysis includes Moll’s method [79] for measuring these char-

acteristic times. By source and drain resistance extraction the transit times may

further be compartmentalized [80]. Such analysis provides a way to quantify

the specific ac-limiting components of the HEMT design and is performed on

insulated-gate AlN/GaN HEMTs with a 6 nm thick high-κ ALD Ta2O5 gate in-

sulator.

4.1 Small signal model

DC drain and transfer characteristics were first taken to establish a bias point

for small signal characterization that maximized DC transconductance. Figure

4.1 shows DC characteristics for a typical AlN/GaN HEMT (Lg = 120 nm, Wg

= 150 µm) with maximum current density of 1.3 A/mm at a forward-bias gate

voltage (VGS) of +4 V and a maximum transconductance of 225 mS/mm at VDS

= 10 V. Off-state breakdown (VGS = -8 V) for the devices occurred at a drain

bias (VDS) between 18 - 20 V and was a result of gate insulator breakdown. The

typical threshold voltage occurred at VGS = -6.5 V (VDS = 10 V) and maximum

DC gm occurred at VGS = -5.5 V which was the small signal bias point chosen to

81

Figure 4.1. Transfer (left) and drain (right) characteristics of theAl2O3/AlN/GaN HEMT used for small signal model development.

maximize ft.

Dambrine et al. introduced an alternative technique (“ColdFET”) to deter-

mine the bias-independent equivalent circuit parameters in the absence of on-wafer

calibration structures [77], which has become widely used [81]. The principle gov-

erning the ColdFET measurement follows that when the drain-source bias is held

at 0V and the gate is reverse-biased into deep pinch-off, the intrinsic device can be

modeled as a simple RC network, allowing an accurate calculation of the induc-

tance and capacitance associated with the device’s bonding pads. The ColdFET

technique is employed as the initial step in the modeling sequence for the AlN/GaN

HEMT. For the analytical modeling portion the commercial software Microwave

Office is used. Both ColdFET and standard S-parameter measurements were

taken with a frequency range over 0.2 - 50 GHz.

Shown in figure 4.2(a) is the measured (circles) and modeled (line) S-parameters

82

Figure 4.2. Measured (circles) and modeled (lines) S-parameter data forboth the ColdFET (a) and maximum-gm (b) biasing regimes.

for our HEMT under ColdFET bias (VDS = 0V, VGS = -8V) showing an ex-

cellent fit-agreement. The corresponding equivalent circuit parameter values ex-

tracted from this measurement fit, following after Dambrine [77], include the bias-

independent parameters: gate inductance (LG), drain inductance (LD), source

inductance (LS), extrinsic (bonding pad) gate-to-source capacitance (CGSp), ex-

trinsic (bonding pad) drain-to-source capacitance (CDSp), extrinsic (bonding pad)

gate-to-drain capacitance (CGDp), gate charging resistance RG, and the bias-

dependent intrinsic RC network parameters CA, CB, CC associated with the

pinched-off depletion region, source resistance (RS), and drain resistance (RD)

as is illustrated in Fig. 4.3 (b). The capacitive network set up by CA, CB, and CC

is the converted delta-wye representation of the pinched-off channel capacitances

CGS, CGD, and CDS. Dambrine originally proposed the equivalent ColdFET cir-

cuit where CA was not accounted for. White suggested a more accurate RC net-

83

work by introducing the third capacitor that accounts for source-drain depletion

and using symmetry arguments made the claim that all depletion capacitances

must be equal [82]. While the addition of a source-drain capacitance in the Cold-

FET model provides a more accurate determination of drain pad capacitance,

the requirement of equal depletion capacitances is not necessary. For vertically

scaled devices where the aspect ratio of the gate depletion region can be large and

therefore cause the source-drain depletion capacitance to be relatively large, an

inequality in depletion capacitances (CA = CB 6= CC) for the ColdFET bias is

more likely.

S-parameters taken at maximum gm (VGS = -4 V, VDS = 10 V, IDS = 51

mA) and used to model the small signal equivalent circuit are shown in Fig. 4.2

((a), circles). The S-parameters were modeled using the equivalent circuit shown

in Fig. 4.3(a) with fit criteria of 1% error in optimization goals for both phase

and magnitude of the modeled S-parameters. The corresponding fitted parameter

values are listed in Table 4.2. A modification to the intrinsic region equivalent

circuit is made for higher fit accuracy in S21. As can be seen in Fig. 4.2, the

modeled S-parameters show close agreement with the measured data indicating

very good model correspondence. Note that the modeled parameter values are

not unique and will vary with the geometry and layer structure of the HEMT

involved.

An accurate small signal model hinges on the intrinsic region of the HEMT.

Owing to GaN’s ability to sustain large lateral electric fields, the VDS-VGS bias in

the saturated velocity region causes the gate-drain depletion region of the HEMT

to take the form shown in Fig. 4.3 (c) [83]. The intrinsic region defined by the

gate boundaries is most accurately described by an RC network on the gate-to-

84

Figure 4.3. Schematic diagram of the small signal equivalent circuitmodel (a) showing the separation of the intrinsic, parasitic, and bondpad circuits, “ColdFET” equivalent circuit model (b), and intrinsic

circuit model (a) over-layed on a cross-sectional depiction of theinsulated gate HEMT.

85

TABLE 4.1

EXTRACTED SMALL SIGNAL PARAMETERS FOR THE

MODELED INSULATED-GATE AlN/GaN HEMT

parameter units value

LG [pH] 65

LS [pH] 0

LD [pH] 13

CGDp [pF] 0.018

CDSp [pF] 0.028

CGSp [pF] 0.019

RS [Ω] 7.1

RD [Ω] 300

RG [Ω] 5.5

Ri [Ω] 36

RDS [Ω] 320

CGS [pF] 63

CGD [pF] 6.5

CDS [pF] 8.7

gintm [S/mm] 0.49

86

source (G-S) end with the standard RC network in parallel with a gain element on

the drain end. Considering the compressed drain depletion width, the connective

nodes of this network can be seen in Fig. 4.3 (c) and have a significant effect

on the fitting of the model. The G-S RC network can be reduced by parallel

arguments with only slight loss of accuracy to a single RC network [84] where the

resistance and capacitance involved is the input resistance Ri and gate-to-source

capacitance CGS, respectively. For the HEMT discussed (Rsh > 400 Ω/), Ri =

39 Ω for the given bias conditions. Furthermore, it is speculated that because of

the curvature of the gate-drain depletion edge the reference voltage, VGS, must

be taken across both CGS and Ri of the intrinsic G-S network. For material

systems that offer lower sheet resistance, Ri reduces, the gate-drain depletion

region has milder curvature, and the intrinsic network then takes the form of the

conventional HFET network [81]. Such an equivalent circuit configuration has

implications on the behavior of |h21|. The presence of additional impedance in

the intrinsic network through a multi-pole circuit can be manifested in |h21| by a

deviation of the 20 dB/dec roll-off at higher operational frequencies and has been

discussed by Steer and Trew [85].

A consistency check on the modeled parameter values can be made based on

results of DC characterization, namely CTLM, Hall, CV, and drain characteristics.

Access resistance which is composed of both contact and channel sheet resistance

was calculated for the 150 µm wide HEMTs to be 4.4 Ω and 9.7 Ω, respectively.

Using these results as lower limits in a biased device the modeled RS (6.8 Ω) closely

agrees. Based on geometry and dielectric constant considerations the calculated

Cg ∼ 90 - 110 pF. Considering CGS and CGD along with the parasitic capacitances

CGSp and CGDp make up total gate capacitance, the combined measured values

87

agree relatively well with this calculated value. Finally, intrinsic transconductance

follows the relationship gintm = gextm /(1-RSgextm ), where RS is the source resistance

(RS ∼ RC+RshLSG/4WG), where the factor of 4 accounts for the gate offset closer

to the source contact edge. Intrinsic gm calculated from DC measurements is 460

mS/mm which agrees well with the modeled gm and verifies agreement with the

modeled |h21|.

4.2 De-embedding procedure

The widely conventional method for reducing the measured S-parameters by

the proportion of the pad impedance (i.e. de-embedding) is shown by the flow

diagram given in Fig. 4.4. By removing the detrimental influence of the probe

pad impedance a more realistic measure of the active device may be had. In the

ideal situation on-wafer calibration standards (open, short, and load) would be

available to measure side-by-side with the actual HEMT. This would provide the

most accurate measurement of the unknown pad parasitic impedance as well as the

specific influence of the underlying layer structure (buffer, nucleation, substrate,

etc.). However, in the case where on-wafer calibration standards are not available

the ColdFET provides an alternative means to measure the impedance and is the

method used here. The Y-subtraction method for de-embedding is the same for

both cases.

The result of de-embedding pad inductance and capacitance on small signal

frequency performance can be seen in Fig. 4.5 along with the modeled |h21| result.

The measured |h21| yielded an extrapolated ft of 45 GHz. Upon removing LG, LD,

LS, CGDp, CGSp, and CDSp through the de-embedding procedure mentioned pre-

viously, an extrapolated ft of 56 GHz was obtained. On average across the wafer,

88

Figure 4.4. Flow diagram for the Y-subtraction method for parameterde-embedding of a HEMT. Toggling the measured S-parameters betweenY and Z parameters allows convenient reduction of pad and/or parasitic

component influence.

89

Figure 4.5. Small signal frequency performance as measured (graycircles), modeled (line), and after de-embedding (blue circles) bonding

pad inductance and capacitance.

values for ft where found to be 10 - 15 GHz higher after removing bonding pad in-

fluences (LG, LD, LS, CGDp, CGSp, and CDSp). This discrepancy is expected to be

more exaggerated for devices capable of higher operational frequencies. A blanket

deposition of high-κ oxide causes the electronic coupling between bonding pads to

act as reactive “inertia” on channel modulation frequency for the insulated-gate

AlN/GaN HEMT.

4.3 Delay analysis of the insulated-gate AlN/GaN HEMT

III-Nitride based HEMTs quickly gained interest for high frequency and power

amplifiers due to the high critical breakdown field of GaN with the combined

advantage of a theoretical/simulated peak electron velocity (∼ 2.5 × 107 cm/s)

exceeding that of GaAs (2 × 107 cm/s) [86],[87]. However, the electron velocity

90

inferred from small signal cut-off frequency measurements (ftLg) of AlGaN/GaN

HEMTs show a discrepancy between it and the expected peak electron velocity

calculated from Monte Carlo simulation [86]. A clue to this discrepancy might be

found in the low-field 2D mobility. For the AlN/GaN heterostructure the dominant

scattering mechanisms at room temperature are optical (OP), acoustic phonons

(AP), and interface roughness (IR) as well as remote surface roughness (RSR)

[88] under certain conditions as discussed in Chapter 1. Assuming a low-field 2D

mobility of 1800 cm2/Vs the associated intrinsic scattering time for the AlN/GaN

HEMT would be 0.2 pico-seconds (τ = m∗µ/q) which would correspond to an

intrinsic ft of ∼800 GHz, almost an order of magnitude higher than any report to

date1 [89],[90],[91]. Therefore, the question remains: what mechanism is limiting

the saturated electron velocity in the AlN/GaN HEMT? Or reconsidering the

question: where in our interpretation of the inferred velocity from the ftLg product

may things be going awry? The following analysis based on cut-off frequency

measurements of a “short”-gate HEMT seeks to shed light on these questions.

The frequency performance metric, ft, can be used as an analysis tool for ex-

tracting the delay times associated with a HEMT’s design. The original methodol-

ogy was developed by Moll in 1988 [79] and has gained wide acceptance as a means

to further characterize fundamental transport properties such as the effective elec-

tron velocity, ve,eff, within a transistor. In the original work Moll demonstrated

that ft can be cast by the inverse sum of three distinct times: the drain delay (τdd)

which is the time an electron takes to traverse the drain depletion region of the

channel which may include the depletion extension due to the virtual gate effect,

the channel charging time (τr) which is an RC charging time associated with the

1This comparison is not entirely fair since any reported ft values invariably contain moredominant delay and charging times. This simplistic comparison is only to point out that low-field mobility is a measure of the electron saturation velocity.

91

Figure 4.6. Transfer (left) and drain (right) characteristics of theTa2O5/AlN/GaN HEMT used for the delay analysis.

resistance of the access regions and therefore the channel charge density, and the

intrinsic delay (τi) which is a charging time associated with the RC coupling of

the metal gate capacitance (parasitic and intrinsic). Bolognesi et al. [92, 93] and

in a separate work Inoue et al. [80] extended the delay analysis further for Al-

GaN/GaN HEMTs by acknowledging the series resistances of the access regions

play a significant role in GaN HEMTs. In Bolognesi’s first work [93] he referred to

Moll’s channel charging delay as τCH and intrinsic delay as the “transistor” delay,

τTR. His contribution however, was that he differentiated τTR into two constituent

delay components such that τTR = τ1 + τRSD where τRSD was again associated

with the access region channel charge and therefore is a function of contact and

access resistances RS and RD, and τ1 is the transit time across the intrinsic chan-

nel region under the gate. By this separation of terms the ft dependence on varied

92

access region lengths could be measured and τRSD determinable by extrapolation.

Finally, τ1 could be determined by self consistency of the total measurement which

yields an estimation for ve,eff by Lg/τ1.

In this work a slightly different notation is used for the delay time naming

convention. The drain delay is given by τdd, the channel charging time by τcc,

the intrinsic (Moll) or transistor (Bolognesi) delay by τRC , and a fourth charging

time, τpd which accounts for the frequency loss due to the bonding pad impedance

which is determined by ColdFET measurement and S-matrix reduction technique

discussed previously. Therefore, the extrinsic ft is given by ft = 1/2πτtot where

τtot = τdd + τcc + τRC + τpd. Since a variation in the access region separations, LGS

and LGD were not available at the time of measurement, τRC (τTR in Bologesi’s

work) could not be separated into the constituent times as was done by Bolognesi.

So as an approximate means to extract a meaningful value for ve,eff , the Cold-

FET model was used to determine both bonding pad impedance as well as the

access region resistances (RS and RD) which was then subsequently removed from

|h21| by S-matrix reduction. This method enabled the approximation of ve,eff

for the AlN/GaN HEMT as well as a generic AlGaN/GaN HEMT (via τRC), for

comparison under the assumption that RS and RD are indeed bias-independent.

One issue that perpetually arose was that of anomalous drain-current degrada-

tion for the NRL-deposited Al2O3 insulated-gate AlN/GaN HEMTs. Most often

times, the initial maximum drain current and threshold voltage would reduce af-

ter VDS was swept to 10V or higher. This degradation is very undesirable in such

devices and work is ongoing to improve the Al2O3 which is at this point the sus-

pect cause of the problem. This point is mentioned because it was observed that

when these devices went under test for delay analysis, their frequency performance

93

would degrade in-situ making it impossible to measure self-consistency between

the VDS and IDS bias sweeps. Therefore, HEMTs with a Ta2O5 gate oxide were

chosen for delay measurement and analysis since they were much less prone to

failure under the bias range needed (DC I-V characteristics shown in Fig. 4.6).

The following properties were measured for the Ta2O5/AlN/GaN HEMT: µ = 400

cm2/Vs, ns = 2.2 × 1013 cm−2, Rsh = 567 Ω/, Rc = 0.7 Ω-mm, LSD = 3 µm,

and Lg ∼150 nm. The data shown pertains to these HEMTs.

As a control sample, a standard heterostructure Al0.3GaN/GaN HEMT with a

barrier thickness of 25 nm was measured using the same measurement flow as the

AlN/GaN HEMT. This allowed a side-by-side comparison of a known structure

to that of the AlN/GaN HEMT. The pertinent properties of the AlGaN/GaN

HEMT were µRT = 1040 cm2/Vs, ns = 1.06 × 1013 cm−2, Rsh = 567 Ω/, Rc

= 0.35 Ω-mm, LSD = 3 µm, and Lg ∼160 nm. All measurements described that

were applied to the AlN/GaN HEMT were likewise applied to the AlGaN/GaN

HEMT for comparison.

Figure 4.7 shows the result of the ft trend as a function of VDS (left) and IDS

(right) for both the AlN/GaN and AlGaN/GaN HEMT after the pad impedances

were de-embedded. Clear maximums can be seen for both spectra. The occurrence

of the linearly decreasing ft with VDS and VGS biases is consistent with other

reports [79],[93],[94]. These trends yield the information sought (the reduction

in ft is undesirable in terms of amplifier frequency linearity). The unity current

gain frequency, ft = 1/2πτtot, where τtot is the total delay after de-embedding

pad impedance and is given by the sum of all the constituent delays (e.g. τtot =

τdd + τcc + τRC)2. By inverting the ft values shown in Fig. 4.7 and plotting

2The total delay as a sum of the constituent delays is similar to the work by Bolognesi [93].In his work the total delay was given by τtot = τD +τCH +τTR = 1/2πft where τTR = τ1+τRSD.

94

0 2 4 6 8 1 0 1 2 1 4 1 6 1 8 2 02 22 42 62 83 03 23 43 63 84 04 24 44 64 85 0

m a x i m u m e x t r i n s i c f t A l G a N / G a N A l N / G a N

A l G a N / G a N( V G S = - 2 . 5 V )

A l N / G a N( V G S = - 2 V )f t (G

Hz)

V D S ( V )0 . 0 0 0 . 0 1 0 . 0 2 0 . 0 3 0 . 0 4 0 . 0 5 0 . 0 60

1 0

2 0

3 0

4 0

5 0A l G a N / G a N( V D S = 1 0 V )

A l N / G a N( V D S = 7 V )

f t (GHz)

I D S ( A )

Figure 4.7. Unity current gain, ft, as a function of VDS (left) and IDS(right).

with respect to the channel voltage or inverse drain current density gives the

trends shown in Fig. 4.8. By removing the voltage drop through the contacts

the channel voltage may be obtained (Fig. 4.8 (left)). It is noted that there is

a minuscule difference in τtot,min for the two measurements and is due to slight

device degradation during the measurement however, the difference amounts to

less than 3 GHz in frequency performance and is therefore accepted.

The linear functionality of the τtot - Vch curve is associated with the increasing

drain depletion region. The linear extrapolation to Vch = 0 gives the total delay

excluding τdd. Therefore, the difference in the minimum total delay and the linear

extrapolation to Vch = 0 yields the drain delay component, τdd. Similarly, the

linear functionality of the τtot - WG/IDS curve is associated with the channel

charging delay, τcc. Therefore, the difference in the minimum total delay and the

linear extrapolation to WG/IDS → 0 yields the channel charging delay, τcc. Finally,

95

Figure 4.8. Total delay as a function of VDS (left) and IDS (right)showing the delay for the corresponding measurement.

since the total delay equals the sum of all constituent delays, the RC-charging

delay, which is the delay associated with the RC charging of the gate-to-channel,

may be found through the difference in total delay and the channel charging and

drain delay components. Self consistency may also be checked between the two

measurements since both involve the intrinsic delay component.

A separate de-embedding-oriented analysis was performed to give further in-

sight into the delay components involved with these HEMTs and the results are

included in Table 4.2. As discussed, the ColdFET measurement/fitting procedure

was originally developed for the purpose of extracting the bias-independent pa-

rameters which also include access resistances RS and RD. By further reducing

the measured S-parameters beyond simply the pad impedance through the same

Y-parameter subtraction technique discussed, by the amounts of the modeled ac-

cess resistances, an “intrinsic” delay, which is the delay associated with electron

drift across only the gated region may be inferred. This delay is denoted as τint

96

TABLE 4.2

EXTRACTED TIME DELAYS FOR THE Ta2O5 INSULATED-GATE

AlN/GaN HEMT AND SCHOTTKY-GATED AlGaN/GaN HEMT.

AlGaN/GaN AlN/GaN

parameter delay [ps] delay [ps] definition

τpd 0.948 0.768 bonding pad delay

τdd 0.322 0.681 drain depletion transit time (VDS)

τcc 0.428 0.466 channel charging delay (1/IDS)

τRC 2.752 2.565 transistor delay

τtot (no pads) 3.464 3.712 total delay excluding τpd

τtot (extrinsic) 4.413 4.521 total delay (= 1/2πft)

τRa 1.207 2.442 access resistance delay

τint 2.257 1.269 intrinsic region time

and the delay associated with the reduction of the access resistances is denoted

as τRa.

The collective delay totals are listed in Table 4.2 and plotted graphically in Fig.

4.9. It is apparent that the dominant delay limitation for both HEMTs resides

in the RC charging delay. However, it is also noted that the pad impedance

imposes a significant delay on both transistors as well. This being a necessary

detriment in order to probe the HEMTs to take such measurements. In both

cases the channel charging delays are comparable owing to similar sheet resistances

between the HEMTs. Interestingly, the AlN/GaN HEMT has a drain delay that is

97

Figure 4.9. Comparison of delay times between the AlGaN/GaN andAlN/GaN HEMTs extracted by Moll’s method.

a factor of 2 larger than that of its AlGaN/GaN counterpart. This is speculated

to be due to the successful mitigation of virtual gating by SiN passivation for

the AlGaN/GaN HEMT reducing the gate length extension and thus the time it

takes electrons to traverse this length. Pulsed gate-lag results of the AlN/GaN

HEMTs with the thin high-κ gate dielectric proved that such a thin layer does not

effectively passivate the AlN surface from the standpoint of large signal operation

(not shown). Although, the thin film deposition of the various high-κ oxides on

the AlN does cause a change in 2D channel charge, as previously discussed. SiN

passivation for the access regions of the AlN/GaN HEMT was observed to be

less effective by pulsed-gate lag measurements indicating a remaining virtual gate

extension issue. Therefore, a means to successfully passivate the AlN/GaN HEMT

remains a subject of further study.

98

The intrinsic delay, τint, is used as an approximate means at arriving at an

effective electron velocity. This being the velocity immediately below the gate if

one assumes the drift velocity to be constant throughout. This assumption is not

accurate given the changing electric field profile along the channel below the gate

but allows an approximate velocity to be obtained. This velocity is also taken to be

the effective saturation velocity since the data was taken while the HEMTs were in

their respective saturation regimes of bias. Taking the equality, ve,eff = Lg/τint, it

is found that ve,eff ∼ 0.7 × 107 cm/s and ∼ 1.18 × 107 cm/s for the AlGaN/GaN

and AlN/GaN HEMTs, respectively. These values are reasonable compared to

recent reports yet are lower than the peak electron velocity calculated by Monte

Carlo simulation. It is speculated that the lower electron velocities arise due to

more severe scattering as indicated by the lower valued mobilities in both HEMTs.

Moreover, it is noted that despite a higher room temperature mobility for the

AlGaN/GaN HEMT, the effective electron velocity in the AlN/GaN HEMT had

a higher extracted value. The cause of this disparity is still not well understood but

it is acknowledged that the procedure used to extract the effective velocities needs

refinement and may introduce some inexactness. The procedure relies on knowing

exact values for the pad impedance, RS, and RD while under bias which was in fact

not exactly known. Nonetheless, the values for parasitic and access impedances

were modeled with the accuracy that the ColdFET method can provide and gives

a fair estimation of these parameter values. In this light, it is interesting that

the AlN/GaN HEMT has the higher estimated ve,eff and is promising for this

heterostructure’s involvement in high frequency amplification. More investigation

is needed on this topic to iron out exact values for the electron velocity within the

intrinsic region in the AlN/GaN HEMT while under bias.

99

It is necessary to take in mind the caveats involved with using the ColdFET

modeling approach to approximate the bias-independent model parameters for

data extraction. The ColdFET fit as it was originally conceived by Dambrine

[77] used only two intrinsic capacitances, CGS and CGD. White argued that using

Dambrine’s circuit topology caused an over-estimation of the drain pad capac-

itance [82]. The answer to the problem was the inclusion of a third intrinsic

capacitance which aided in the proper capacitive partitioning during S-parameter

fitting of ColdFET data. White’s topology is used throughout this work.

A deficiency of the ColdFET measurement for extrinsic parameter determi-

nation lies in the assumption that RS and RD are bias-independent. It has been

argued by Palacios et al. that a differential RS occurs in AlGaN/GaN HEMTs that

is bias-dependent due to exterior gate velocity saturation. Palacios argues that

this causes a non-linear reduction in extrinsic transconductance and ft [94]. Fur-

thermore, DiSanto and Bolognesi demonstrated the occurrence of a RF-modulated

RS which caused subsequent error in the determination of the intrinsic small signal

parameters for AlGaN/GaN HEMTs. This further aided in the under-estimation

of RS for RF modulated HEMTs [92]. This under-estimation led to the under-

estimation of intrinsic gm, intrinsic ft, and subsequently the intrinsic delay com-

ponent which allows the estimation of saturated electron velocity in the HEMT

channel. Therefore, the implicit assumption of the ColdFET measurement, that

all extrinsic small signal elements are “static” when VDS = 0V and the gate is

sufficiently reverse-biased so that the intrinsic impedance from gate capacitance

becomes negligible, is shown to be untrue for the GaN-based HEMT. Subsequently,

the use of the ColdFET method for series access resistance determination leads to

an under-estimation of the detrimental influence RS has on the HEMT’s dc and

100

RF operation.

These deficiencies are acknowledged when considering the data shown above

and therefore some error is accepted in the results discussed. It is further noted

that the ideal scenario would be to have on-wafer calibration test structures (open,

short, through, and load), but were not available from the mask set used during

HEMT fabrication.

4.4 Points for improvement for RF operation of the AlN/GaN HEMT

The large delay time associated with the access region resistance indicates the

necessity to reduce this extraneous resistance as much as possible. After all, it

acts to essentially extend the resistance from the contact electrodes. Minimization

of this resistance will lead to significant improvement in ft. It is noted that for

the devices measured in this study, Rsh ∼ 450 Ω/ due to a low mobility of only

∼ 500 cm2/Vs for the AlN/GaN HEMT. Lower sheet resistance obtainable in

the AlN/GaN heterostructure are anticipated to lead to a reduced access region

charging time, τRa.

The delay time associated with RC and is a fractional portion of τRa is also

long and may be reduced by improving contact resistance. This issue has been

a bane for GaN-based HEMTs from their inception. Pre-metallization etching or

n+ ion-implanted contacts are anticipated to further lower RC and thus improve

ft. During the writing of this work the successful demonstrate of n+ regrown

contacts to an ultra-thin AlN/GaN HEMT was reported which had a measured

contact resistance of 50 Ω− µm [95].

Since low-field mobility gives a measure of electron velocity, the increase in

mobility will play a major role in improving ft through the delay components τint,

101

τcc, and τRC For GaN-based HEMTs if a choice must be made between increasing

2D charge density or mobility, mobility should be the choice made given that it

will yield the greatest improvement on RF performance since it affects more delay

components.

GaN-based HEMTs have long showed severe current collapse and frequency

dispersion owing to a highly charge-susceptible surface within the access regions

(particularly the drain edge of the gate). These issues necessitate the use of a

surface state passivation layer within the access regions of the HEMT. The blanket

deposition of a high-κ dielectric for gate insulation was often made in this study

for the AlN/GaN HEMT as it was more convenient during processing. It is also

favorable as a way to assist in the elimination of dispersion effects. However, the

byproduct of a blanket deposited high-κ dielectric is an enhanced RC charging

time, τRC , since parasitic CGS and CGD become large. It would therefore be

appropriate to utilize a process which allows the preferential deposition of a high-

κ dielectric only under the gate metal while in the access regions a low-κ dielectric

suitable to be used for a passivation layer. The mechanics of frequency dispersion

are discussed in Chapter 6 with a novel approach for the mitigation of DC-RF

dispersion.

4.5 Chapter 4 cited literature

77. G. Dambrine, A. Cappy, F. Heliodore, E. Playez, “A New Method for Deter-mining the FET Small-Signal Equivalent Circuit”, IEEE Trans. Elec. Dev.,Vol. 36, No. 7, 1988.

78. P.J. Tasker and B. Hughes, “Importance of Source and Drain Resistance tothe Maximum ft of Millimeter-Wave MODFET’s”, IEEE Elec. Dev. Lett.,Vol. 10, No. 7, 1989.

79. N. Moll, M.R. Hueschen, A. Fishcer-Colbrie, “Pulse-Doped AlGaAs/InGaAs

102

Pseudomorphic MODFETs”, IEEE Trans. Elec. Dev., Vol. 35, No. 7, 1988.

80. T. Inoue, Y. Ando, K. Kasahara, Y. Okamoto, T. Nakayama, H. Miyamoto,M. Kuzuhara, “Advanced RF Characterization and Delay-Time Analysis ofShort Channel AlGaN/GaN Heterojunction FETs”, IEICE Trans. Elec.,Vol. E86-C, No. 10, 2003.

81. S. Nuttinck, E. Gebara, J. Laskar, J. Shealy, M. Harris, “Improved RF Mod-eling Techniques for Enhanced AlGaN/GaN HFETs”, IEEE Micr. Wire.Comp. Lett., 13: 140, 2003.

82. P.M. White, R.M. Healy, “Improved Equivalent Circuit for Determinationof MESFET and HEMT Parasitic Capacitances from “Coldfet” Measure-ments”, IEEE Micr. Guid. Wave Lett. 3:12:453-454, 1993.

83. Karmalkar S, Mishra U, IEEE Trans. Elec. Dev. 48:8, 2001.

84. Ali F, Gupta A, HEMTs & HBTs: Devices, Fabrication, and Circuits,Artech House; 1991.

85. Steer M, Trew R, IEEE Elec Dev Lett 1986, 7:11.

86. TH. Yu, K.F. Brennan, “Monte Carlo calculation of two-dimensional elec-tron dynamics in GaN-AlGaN heterosturctures”, J. Appl. Phys., 91, 3730,2002.

87. Ioffe Institute: http://www.ioffe.ru/SVA/NSM//Semicond/

88. Yu Cao, “Study of AlN/GaN HEMTs: MBE growth, transport properties,and device issues”, Ph.D. dissertation, 2010.

89. T. Zimmermann, Y. Cao, D. Jena, H. Xing, “4-nm AlN barrier all binaryHFET with SiNx gate dielectric”, Int. J. High Speed Elec. and Syst., 19,153, 2008.

90. M. Higashiwaki, T. Mimura, T. Matsui, “AlN/GaN Insulated-Gate HFETsUsing Cat-CVD SiN”, IEEE Elec. Dev. Lett., Vol. 27, No. 9, 2006.

91. N. Onojima, N. Hirose, T. Mimura, T. Matsui, “Ultrathin AlN/GaN het-erostructure field-effect transistors with deposition of Si atoms on AlN bar-rier surface”, Appl. Phys. Lett. 93, 223501, 2008.

92. D.W. DiSanto, C.R. Bolognesi, “At-Bias Extraction of Access Parasitic Re-sistances in AlGaN/GaN HEMTs: Impact on Device Linearity and ChannelElectron Velocity”, IEEE Trans. Elec. Dev., Vol. 53, No. 12, 2006.

103

93. C.R. Bolognesi, A.C. Kwan, and D.W. DiSanto, “Transistor delay analysisand effective channel velocity extraction in AlGaN/GaN HFETs”, IEDMTech. Dig., pp. 879-886, 2002.

94. T. Palacios, S. Rajan, A. Chakraborty, S. Heikman, S. Keller, S. P. Den-Baars, U. K. Mishra, “Influence of the Dynamic Access Resistance in thegm and ft Linearity of AlGaN/GaN HEMTs”, IEEE Trans. On ElectronDevices, vol. 52, pp. 2117-2123, 2005.

95. A.L. Corrion, K.Shinohara, D. Regan, I. Milosavljevic, P. Hashimoto, P.J.Willadsen, A. Schmitz, D.C. Wheeler, S.M. Butler, D. Brown, S.D. Burn-ham, M. Micovic, “Enhancement-Mode AlN/GaN AlGaN DHFET with 700mS/mm gm and 112 GHz ft”, International Workshop on Nitride Semicon-ductors (IWN), 2010.

104

CHAPTER 5

HOT CARRIER EFFECTS IN SixNy-PASSIVATED AlGaN/GaN HEMTS

Hot carrier effects are a common degradation mechanism in FET technologies

that can compromise both performance and reliability. The literature on Si and

SOI technologies provide a wealth of information about such phenomena as impact

ionization, hot carrier injection, and electron and hole trapping [96],[97]. Fair et al.

demonstrated that bias-dependent hot-hole emission into a SiO2 gate oxide modi-

fied the threshold characteristics of the MOSFET [97]. Compound semiconductors

are also known to exhibit hot electron effects in the form of impact ionization and

hot carrier injection that cause back-gating and current-voltage nonlinearities that

impair performance [98],[99],[100]. Huang et al. showed negative charge trapping

at the surface and within the barrier region of a GaAs HEMT causing degrada-

tion in current-voltage characteristics [98]. In the GaN-based system, Mittereder

et al. demonstrated that short-term dc bias stress of AlGaN/GaN HEMTs can

induce trapping effects as manifested by current collapse [101]. Meneghesso et al.

recently related an anomalous kink in drain current in an AlGaN/GaN HEMT

to electron de-trapping [102] while Neuberger et al. showed electron trapping

within the passivation layer of an AlGaN/GaN HEMT [103]. This study investi-

gates hot carrier effects in III-Nitride HEMTs using a specialized multi-gate test

structure. In particular, evidence for hole trapping in SiN passivated HEMTs is

presented. Furthermore, it is asserted that the source of the holes arises from

impact ionization within the channel of the HEMT, implied by a kink in the drain

105

Figure 5.1. Device cross section showing the Schottky and four MISgates (a) with the corresponding band diagram along the channel

direction (b) showing the impact ionization process and the relation tothe Kink effect.

characteristics. Finally, a qualitative model has been developed to analyze aspects

of the trapped hole charge.

5.1 Test Structure

Sample growth was performed by RF-plasma assisted MBE on free-standing,

semi-insulating, HVPE-grown Fe:GaN [105]. The MBE-grown layer structure con-

sisted of a 1 µm thick 1019 cm−3 doped Be:GaN isolation layer, a 0.5 µm undoped

106

GaN buffer, and a 250 A AlxGa1−xN barrier, where x ∼0.27. The MBE growth

was performed in a Vacuum Generators V80H MBE system. The free-standing

GaN substrates were grown by HVPE at Kyma Technologies, Inc. Nominal dislo-

cation densities of the substrates were estimated by cathodoluminescence imaging

to be lower than 107 cm−2.

Processing was initiated by the photolithographic definition and deposition of

Ti/Al/Ni/Au contacts that were annealed at 860C for 30 seconds. This was fol-

lowed by a mesa isolation etch in a BCl3/Cl2 plasma. E-beam lithography was

used to define a 300 nm long gate located 1.85 µm from the source contact (LSG).

The Ni/Au gate metal stack was deposited by e-beam evaporation. A 50 nm SiN

layer was deposited by PECVD that employed an alternating high/low frequency

plasma assist. A set of four auxiliary Ni/Au-insulator-semiconductor (MIS) test

gates with gate lengths of 300 nm and period of 600 nm (LP ) were defined by e-

beam lithography 700 nm from the gate edge (Lsep). The source-drain separation

was nominally 6 µm (LSD in Fig. 5.1 (a)). Windows were opened in the SiN by a

SF6 plasma and an overlay metal stack of Ti/Pt/Au was deposited on all contact

pads for probing. Hall effect measurements were taken after the SiN deposition,

and a 2DEG charge density of 1 × 1013 cm−2 and a mobility of 1400 cm2/Vs at

room temperature were measured.

5.2 Experimental Procedure: Chynoweth model

Transfer characteristics were first measured for all gates (Schottky and MIS)

on the HEMT. Next the devices were subjected to a 2 hour bias stress in open

channel conditions via the Schottky gate (VDS = 20 V, VGS = +3 V, ID = 820

mA/mm). Immediately following bias stressing, transfer characteristics on all

107

gates were again taken. This history is shown in Fig. 5.2 (left). The threshold

voltage, Vth, is defined in the usual way by an extrapolation of the linear region of

the transfer characteristics. After bias stress the threshold voltage of the Schottky

gate remained the same as before stress, but the MIS gates all shifted to more neg-

ative voltages (∆Vth = 1.3 V) indicating the presence of positive trapped charge.

It is noted that this contrasts with Neuburger who observed a positive Vth shift,

indicative of negative trapped charge [103]. This discrepancy is not well under-

stood at this time. The transfer characteristics of both the Schottky and MIS gates

demonstrated similar slopes (transconductance) despite the additional 20 nm sep-

aration of the MIS gates by the SiN passivation layer. It is not well understood

at this time the origin of this but based off gate-to-gate current measurements it

is believed to be due to a degree of electronic coupling (conduction) between the

different electrodes. Attention is also drawn to the increased maximum current in

the MIS-gated transfer characteristics after bias stress. This increase in current

reflects the decreased access resistance as a result of the induced electron density

in the channel due to the trapped positive charge. That no shifts are observed

under the Schottky gate suggests the charge trapping affects only the access re-

sistance. Through successive measurements after bias stress of the MIS transfer

characteristics at room temperature the positive shifts were observed to return

slowly to unstressed levels with full recovery after ∼24 hours (Fig. 5.3 (right)).

Although the apparent buildup in positive charge could originate from electron

de-trapping [102], a more compelling hypothesis is that the increase in positive

charge is due to hole trapping. As to the source of the holes in these n-channel

devices, impact ionization is the likely origin. Strong evidence for this is found in

the kink observed in the drain current of the MIS-gated HEMT (Fig. 5.2 (right)).

108

Figure 5.2. Transfer characteristics (left) of all gates on the HEMT priorto and after bias-stress. Drain characteristics (right) of the AlGaN/GaNHEMT comparing the Schottky gate controlled (line) vs. the MIS gate 2controlled (open circles) family of curves. A clear kink in the MIS gate

controlled characteristics can be seen at ∼9 volts.

Such kinks are a known byproduct of impact ionization within the channel of the

HEMT, which arises from source-gate barrier lowering due to impact ionization-

generated hole accumulation in the source-gate region thereby abruptly increasing

channel conductivity (Fig. 5.1 (b)) [99],[100],[107],[108].

A second crucial piece of evidence that holes are not only being generated but

also are being injected into the gate barrier comes from the measurement of gate

current. As seen in the inset of Figure 5.3 (left) for the second MIS gate, there is

an additional current which adds to the already existing but minimal gate leakage

current (Itot = Ileak + Ihole). The VDS dependence of the extra current strongly

suggests that it is a hole current. For a more quantitative analysis we focus on

the corresponding gate current characteristics, specifically the VGS = -4 V curve

in the inset of Fig. 5.3 (left). Both leakage and hole currents can be extracted

from the measured gate current following the technique originally developed by

109

Chynoweth [104] and can be seen in Fig. 5.3. This technique has been used to

characterize the kink effect in GaAs MESFETs by Hui et al. [100] and also in

AlGaN/GaN HEMTs by Dyankonova et al. [106]. The gate hole current can be

approximated as a function of drain current (IDS) and the electric field in the high

field region ((VDS − Vsat)/Leff ) by

Ihole = αIDSLeffe−EiLeff

(VDS−Vsat) , (5.1)

where Vsat is the voltage at the onset of current saturation, Leff is the effective

gate length, and the electric field at the onset of impact ionization (Ei), where

the impact ionization coefficient, α is a constant for low fields. The gate leakage

current component is small compared to the hole current after onset, but most

notable is that the gate leakage current component is only a weak function of

VDS. Figure 5.3 shows the logarithmic gate current ratios as a function of recip-

rocal field. The region of interest is the linear regime on the logarithmic scale of

the hole current (Ihole/IDS) which validates impact ionization as the source of the

gate extracted hole current. From the slope of a linear fit, Ei can be extracted

for our system as ∼1 MV/cm. This value of the field needed for impact ioniza-

tion is slightly less than that reported in literature [107] and is exemplary of the

approximate nature of the analysis.

5.3 Discussion of charge dynamics

The band diagram seen in the inset of Fig. 5.2, is of the SiN/AlGaN/GaN

heterostructure cross section taken vertically through one of the MIS gates. The

relatively low valence band barrier and favorable fields inside the AlGaN support

the picture of hole injection. Furthermore, the fields in the SiN and AlGaN layers

110

Figure 5.3. (left) Gate current analysis for MIS gate 2 biased at VGS=-4 V. Linear character of the Ihole/IDS ratio indicates impact ionizationgenerated hole current. The inset shows the measured gate current forgate voltage steps of -2 V. (right) Time decay of the threshold voltage

shift.

combined with the valence band discontinuity suggests that holes will accumulate

in the vicinity of the SiN/AlGaN interface. From this we speculate the charge

becomes trapped at the SiN/AlGaN interface but is also possible that they remain

mobile. The relationship between the threshold voltage shift and trapped charge

is [109],

∆Vth =q

εoεSiN

∫ t

0

ypi(y)dy =qpiεo

(tSiNεSiN

), (5.2)

where the integral is over the combined SiN and AlGaN thicknesses (t), pi is

the 2D trapped hole charge density, tSiN is the SiN thickness, and εSiN is the

dielectric constant of SiN. The second equality follows if one assumes a delta

function distribution of trapped charge at the SiN/AlGaN interface. It is noted

that Neuburger’s formalism assumed a gaussian-type trapped charge distribution

that was displaced from the SiN/AlGaN interface within the passivation layer.

111

However, no reason was given for such an assumption therefore, it not used here.

Moreover, as discussed, the electric field profile shown in the inset in Fig. 5.2

suggests favorable conditions for the interface made by the SiN and AlGaN layers

to be the location where positive trapped charge would reside. Based on this

analysis, our measured results for MIS gate 2 of ∆Vth = 1.3 V from Fig. 5.2

(right), we find a trapped charge density of 9.6 × 1011 cm−2. The rightmost

plot in Fig. 5.3 shows the time-elapsed result of the threshold voltage shift as

the trapped charge was allowed to relax (i.e. all biases removed from the test

structure). It was observed that a long decay time constant occurred for the

trapped charge and followed an exponential relaxation as can be seen by the fit

line. The long decay time constant suggests that for time scales on the order of

operational frequencies, the trapping time constant imposes no direct deleterious

effects in terms of ac operation. This point is in agreement with the virtual gate

model which maintains that a charged length of the access region beyond the gate

cannot modulate at the same high frequency as the gate signal but acts to increase

access resistance within the gate-to-drain region. This topic will be discussed in

more detail in the following chapter. The inset to Fig. 5.3 shows the calculated

threshold voltage shift as a function of trapped positive charge density for various

thickness of SiN passivation layers since typical passivation layers are generally

thicker than those used in this study.

Lastly, we consider the magnitude of the threshold voltage shift of each of

the MIS gates. One could expect the individual shifts to differ due to a differing

trapped charge density in accord with the density and location of impact ionization

generated EHPs along the channel. This is indeed observed in the slight expansion

of the threshold voltage spread of the MIS gates after bias stress and is additional

112

support for our entire interpretation. However, because the MIS gates in the device

used in this study were grouped closely together, the voltage spreading does not

fully highlight the distribution of holes trapped along the channel direction in

comparison to if the MIS gates were positioned evenly between the Schottky gate

and drain, which is the eventual goal of this test structure.

5.4 Chapter 5 cited literature

96. N.S. Saks, P.L. Heremans, L. Van Den Hove, H.E. Maes, R.F. DeKeers-maecker, G.J. Declerck, “Observation of Hot-Hole Injection in NMOS Tran-sistors Using a Modified Floating-Gate Technique,” IEEE Trans. Elec. Dev.,vol. ED-33, no. 10, 1986.

97. R.B. Fair and R.C. Sun, “Threshold-Voltage Instability in MOSFET’s Dueto Channel Hot-Hole Emission,” IEEE Trans. Elec. Dev., vol. ED-28, no.1, 1981.

98. J.C. Huang, G.S. Jackson, S. Shanfield, A. Platzker, P.K. Saledas, C. We-ichert, “An AlGaAs/InGaAs Pseudomorphic High Electron Mobility Tran-sistor With Improved Breakdown Voltage for X- and Ku-Band Power Ap-plications,” IEEE Trans. on Microwave Theory and Tech., Vol. 41, No. 5,1993.

99. B. Brar and H. Kroemer, “Influence of Impact Ionization on the Drain Con-ductance in InAs-AlSb Quantum Well Heterostructure Field-Effect Transis-tors,” IEEE Elec. Dev. Lett., vol. 16, no. 12, 1995.

100. K. Hui, C. Hu, P. George, P.K. Ko, “Impact Ionization in GaAs MESFET’s,”IEEE Elec. Dev. Lett., vol. 11, no. 3, 1990.

101. J.A. Mittereder, S.C. Binari, P.B. Klein, J.A. Roussos, D.S. Katzer, D.F.Storm, D.D. Koleske, A.E. Wickenden, R.L. Henry, “Current collapse in-duced in AlGaN/GaN high-electron-mobility transistors by bias stress,”Appl. Phys. Lett., vol. 83, no. 8, 2003.

102. G. Meneghesso, F. Zanon, M.J. Uren, E. Zanoni, “Anomalous Kink Effectin GaN High Electron Mobility Transistors,” IEEE Elec. Dev. Lett., vol.30, no. 2, 2009.

113

103. M. Neuburger, J. Allgaier, T. Zimmermann, I. Daumiller, M. Kunze, R.Birkhahn, D.W. Gotthold, E. Kohn, “Analysis of Surface Charging Effectsin Passivated AlGaN-GaN FETs Using a MOS Test Electrode,” IEEE Elec.Dev. Lett., vol. 25, no. 5, 2004.

104. A.G. Chynoweth, “Ionization Rates for Electrons and Holes in Silicon”,Phys. Rev., Vol. 109, No. 5, 1958.

105. D.F. Storm, D.S. Katzer, J.A. Roussos, J.A. Mittereder, R. Bass, S.C. Bi-nari, Lin Zhou, David J. Smith, D. Hanser, E.A. Preble, and K.R. Evans,“Microwave performance and structural characterization of MBE-grwon Al-GaN/GaN HEMTs on low dislocation density GaN substrates,” J. Cryst.Growth, vol. 305, no. 340, 2007.

106. N. Dyankonova, A. Dickens, M.S. Shur, R. Gaska, J.W. Yang, “Temperaturedependence of impact ionization in AlGaN-GaN heterostructure field effecttransistors,” Appl. Phys. Lett., vol. 72, no. 20, 1998.

107. C.H. Lin, W.K. Wang, P.C. Lin, C.K. Lin, Y.J. Chang, Y.J. Chan, “Tran-sient pulsed Analysis on GaN HEMTs at Cryogenic Temperatures,” IEEEElec. Dev. Lett., vol. 26, no. 10, 2005.

108. S. Nuttinick, S. Pinel, E. Gebrara, J. Laskar, M. Harris, J.R. Shealy, “Floating-Body Effects in AlGaN/GaN Power HFETs,” in Gallium Arsenide applica-tions symp., Milano, 2002.

109. S.M. Sze, “JFET and MESFET” in Physics of Semiconductor Devices, 2nd

ed., New York, Wiley, 1981, ch. 6, pg. 319-322.

114

CHAPTER 6

NOVEL HEMT DESIGNS UTILIZING THE AlN/GaN HETEROSTRUCTURE

The ultra-thin AlN barrier heterostructure makes feasible an endless assort-

ment of novel HEMT designs depending on the desired application. In the past

10 years a collection of devices have been demonstrated that utilize some form of

the AlN/GaN heterostructure. Bayram et al. demonstrated double-barrier reso-

nant tunneling diodes [110]. Digital-alloy barriers have been utilized for HEMT

barriers [111]. Not to mention the plethora of HEMT designs already mentioned

[112], [113], [114].

This chapter highlights the realization of two independent but novel design

concepts that utilize the AlN/GaN heterostructure. First, a set of three varying

HEMT heterostructures that employ an AlxGaN back barrier for short channel

effect mitigation is presented. Such structures utilize 74 - 100% Al-containing top

barriers and an Al0.13GaN back barrier, the highest Al-containing back barrier in

a Ga-polar HEMT to date. Secondly, a dual-channel AlN/GaN HEMT design is

proposed and realized for the first time. Its purpose, to exploit the top channel as a

field-screening layer for surface potential fluctuations and is an epitaxial approach

for the mitigation of DC-RF dispersion. These new-generation AlN/GaN HEMTs

attest to the structure’s merit for advanced GaN HEMT designs.

115

6.1 Polarization-engineered AlxGaN back barriers

A distinctive characteristic of the GaN-based HEMT is that due to the very

wide band gaps involved and subsequently the large conduction band discontinu-

ities utilized in the constituent heterostructure along with the polarization sheet

charges that form, an energy back barrier is not necessary for sufficient 2DEG

confinement. However, there have been several reports, that have incorporated

several differing varieties of back barriers in a GaN-based HEMTs. Palacios pro-

posed and realized the use of a thin (∼1 nm) InGaN layer close to the AlGaN/GaN

interface on the GaN side. By utilizing the conduction band offset between GaN

and InGaN, an effective back barrier of ∼0.2 eV occurs [115]. Mikovic has utilized

an Al0.04GaN back barrier reporting improvements in device performance owing to

the incorporation of the low Al percentage back barrier [116]. This structure was

also used for a polarization-engineered enhancement-mode AlGaN/GaN HEMT

[117]. Rajan and co-workers have successfully demonstrated GaN/AlGaN HEMTs

fabricated from structures grown in the Nitrogen-face crystal polarity which in-

corporates a built-in Al0.3GaN back barrier [118], [119].

The value in such structures that incorporate some form of energetic back

barrier lies in the fact that the structure sets up a rigid interface that enhances

termination of the electron distribution and allows more efficient charge control

of the 2DEG by the gate bias voltage. It is also plausible that the incorporation

of an AlGaN buffer may reduce residual buffer charge from unintentional dopants

due to its wider band gap making the thermodynamics favorable for reduced

ionization, but this is speculative at this point. Jessen et al. suggested a gate

length to barrier thickness ratio of 15 for GaN HEMT technology in order to avoid

short channel effects, increasing the commonly accepted value of 3 - 6 from GaAs

116

Figure 6.1. Layer structure and band diagram in equilibrium of theAlGaN/AlN/GaN control sample.

technology by a factor of ∼3 [120]. While this stricter criterion is unfavorable, it

was statistically measured across a wide range of vendors’ devices and found to

be very systematic. One caveat to the study was that all GaN HEMTs measured

utilized the conventionally thick GaN buffer without the employment of a back

barrier. Based on the above argument, it is anticipated that the incorporation of

an energy back barrier will relax Jessen’s criterion, thus allowing a reduced aspect

ratio before short channel effects become an issue. It is this point that is at the

heart of the following study while simultaneously demonstrating some of the first

high Al-content AlGaN back barrier HEMTs.

During the time of this writing, HRL researchers reported on a similar device

consisting of a GaN/AlN/GaN multi-heterostructure that included a Al0.8GaN

back barrier grown on SiC. By also including n+ ohmic contact regrowth, enhancement-

mode HEMTs fabricated from this structure demonstrated outstanding small sig-

nal frequency performance of 120 GHz and 200 GHz for ft and fmax, respectively,

117

which set a record for the ultra-thin AlN/GaN HEMT [114]. Furthermore, a

forthcoming report will be presented on depletion-mode AlN/GaN HEMTs which

include an Al0.8GaN back barrier that have achieved record small signal perfor-

mance for both the ultra-thin AlN/GaN HEMT and the GaN-based HEMT in

general. Small signal performance of 220 GHz and 400 GHz for ft and fmax is

shown for 40 nm gate length HEMTs, respectively [121]. Such outstanding per-

formance is a testament to the capability of back barrier technology to mitigate

short-channel effects in deep sub-micron gate length transistors and the ultra-thin

AlN/GaN HEMT in general.

In this study three different structures were grown by plasma-assisted MBE at

the University of Notre Dame1. The first structure contained an Al0.72GaN/AlN/GaN

active layer without a back-barrier, similar to what has been reported elsewhere

[122] and was intended as a calibration and control sample for comparison to the

back barrier samples. The layer structure, band diagram, and AFM scan of the

AlGaN surface can be seen in Fig. 6.1. The additional structures include an

Al0.72GaN/AlN/GaN active layer and AlN/GaN active layer with both including

the Al0.13GaN back barrier. Schematic cross-sections of these structures including

their respective equilibrium band diagrams and surface AFM are shown in Fig.

6.3.

Processing followed the typical AlN/GaN process flow given in Appendix C

with the exception that the ohmic pre-etch times were all increased by a factor of

3 to account for material differences between U. Notre Dame and NRL. An ohmic

first process was employed with the mesa etch following which resulted in contact

resistances of 0.7 - 2.2 Ω-mm as measured by circular TLM. A blanket deposition

1Guowang Li is gratefully acknowledged for performing the epitaxial growth of these struc-tures.

118

Figure 6.2. Transfer (left) and drain (right) characteristics of theAlGaN/AlN/GaN HEMT control sample (without a back barrier).

of ALD Al2O3 of 5 nm thick was then deposited at the U. of Notre Dame. Finally,

gates were defined by optical lithography and deposited and a patterned buffered

HF etch was performed to open access regions to the ohmic contacts. Pertinent

device geometries include LSD = 5 µm, WG = 150 µm, and LG = 1 µm. Samples

are currently undergoing additional processing for the application of sub-micron

T-gates but were not available at the time of this writing.

Hall effect measurements were taken on the as-grown sample as well as after

Al2O3 deposition and results were found to change from 405 Ω/ (ns = 1.33 ×

1013 cm−2, µ = 1160 cm2/Vs) to 975 Ω/ (ns = 0.68 × 1013 cm−2, µ = 950

cm2/Vs), respectively. The reduction in 2D charge density is speculated to be due

to partial relaxation of the strained barrier layer from multiple anneals at 860 oC

in order to achieve low-resistance ohmic contacts. This also may explain the slight

119

Figure 6.3. (left) Layer structures, (center) band diagrams, and (right)2×2 µm2 AFM scans of the AlN surface for the AlGaN back barrier

structures.

120

reduction in mobility. Transfer and drain characteristics of the AlGaN/AlN/GaN

HEMTs without the back barrier are shown in Fig. 6.2. Maximum drain current

of 0.93 A/mm at VDS = 8 V and VGS = +7 V was achieved with an accompanying

extrinsic transconductance of ∼175 mS/mm. The higher contact resistance was

responsible for the lower than expected extrinsic transconductance. Nonetheless,

these characteristics demonstrate a factor of two improvement over similar devices

reported [122] in terms of drain current density and contact resistance as well as

comparable transconductance.

The two layer structures which include the AlxGaN back barrier are shown

schematically in Fig. 6.3 which also includes the respective equilibrium band

diagram and 2×2 µm2 AFM scan of the AlGaN and AlN as-grown surfaces. Figure

6.3(a) shows the Al0.72GaN/AlN/GaN structure including the 5 nm thick Al2O3

film while Fig. 6.3(b) shows the AlN/GaN structure. For the AlGaN/AlN/GaN

structure prior to oxide deposition the on-wafer RT Hall results yielded Rsh = 261

Ω/ (ns = 3.3 × 1013 cm−2, µ = 725 cm2/Vs) whereas after oxide deposition they

were found to be Rsh = 305 Ω/ (ns = 3.3 × 1013 cm−2, µ = 614 cm2/Vs). The

drop in mobility is not well understood at this time. For the AlN/GaN structure

prior to oxide deposition the RT Hall results yielded Rsh = 152 Ω/ (ns = 4.29

× 1013 cm−2, µ = 958 cm2/Vs) whereas after oxide deposition they were found to

be Rsh = 170 Ω/ (ns = 4.3 × 1013 cm−2, µ = 852 cm2/Vs).

Figure 6.4 shows (left) the electron charge density as a function of distance

from the AlN/GaN interface for the three heterostructures as calculated by a 1D

Poisson Schrodinger solver and (middle) the results of C-V measurements taken

at room temperature. The gray area in Fig. 6.4 (left and right) represents the

physical location of the GaN/AlGaN heterojunction that results in the energy

121

Figure 6.4. (left) Calculated electron charge density from 1D PoissonSchrodinger solver, (middle) measured C-V data of the back barrierstructures, and (right) calculated charge distribution from C-V data.

back barrier while the white region represents the GaN channel region. It can be

seen that without the back barrier (grey dashed line) the charge density profile is

calculated to persists deep within the GaN buffer but when the AlGaN back barrier

is present the charge density profiles quickly reduce below 1013 cm−2 before the

junction location. As expected, the AlN/GaN top heterojunction causes a higher

2D charge concentration in the GaN quantum well (channel) but still maintains

a more abrupt decrease by the inclusion of the back barrier.

C-V profiling is a convenient and often used technique for the determination

of the mobile charge “profile” and is an obvious tool to extract the effect of the

back barrier on the charge profile in these structures. Figure 6.4 (middle) shows

the comparative results between all three HEMT structures of room temperature

C-V measurements on 100 µm diameter dots with an oscillation frequency of 10

122

Figure 6.5. drain (left) and transfer (right) characteristics comparison ofback barrier HEMTs

MHz. A practical constraint that arose was the onset of phase reduction early

in the voltage sweep that prevented a clear depletion portion of the back barrier

structures. However, the control structure without a back barrier demonstrated a

very clear C-V curve with minuscule hysteresis (∼0.08 V) indicating a high quality

oxide without oxide charge. Shown in the right-most plot of Fig. 6.4 are the

resulting charge profiles as a function of distance from the top AlN/GaN interface

calculated from the C-V curves by the typical ns ∝ (∂C−2/∂V )−1 relationship.

It can be seen that the charge fall-off within the buffer region of the back-barrier

structures is much faster than the structure without the back barrier, as expected.

It is noted that the C-V profiling technique tracks the centroid of the mobile

charge density therefore the right-most and left-most plots in Fig. 6.4 are not to

be taken as equivalent but rather two different ways of viewing the charge profile

in the structure.

Comparative transfer and drain characteristics of the back barrier HEMTs are

shown in Fig. 6.5. The lower sheet resistance held by the AlN/GaN HEMT is rep-

123

resented in the higher maximum drain current of 1.2 A/mm over 1.15 A/mm of

the AlGaN/AlN/GaN HEMT. The higher charge density of the AlN/GaN HEMT

is manifested by a more negative threshold voltage seen in the transfer character-

istic. The AlN/GaN HEMT showed a higher extrinsic transconductance of ∼150

mS/mm over ∼120 mS/mm for the AlGaN/AlN/GaN HEMT despite the higher

contact resistance. Conservative estimates of intrinsic gm for the respective de-

vices was gintm ∼ 214 mS/mm (RS ≈ Rc = 2 Ω-mm) and gintm ∼ 155 mS/mm (RS

≈ Rc = 1.9 Ω-mm). Since the dielectric/barrier stacks were of comparable capac-

itance for both structures it is speculated that the higher gm for the AlN/GaN

HEMT is due to a higher saturation velocity as inferred by the higher low-field

mobility of the AlN/GaN HEMT than the AlGaN/AlN/GaN HEMT. This is left

to be confirmed by velocity extraction through a varied LG experiment that is to

be carried out after sub-micron gate application.

The successful demonstration of ultra-thin high Al-content AlGaN/AlN/GaN

and AlN/GaN HEMTs with Al0.13GaN back barriers has been presented for the

first time. Maximum drain current densities exceeded 1A/mm for all device struc-

tures. High contact resistance prevented the sub-10 nm barrier thickness to be

advantaged for improvement of extrinsic gm but still maintained respectable levels

greater than 100 mS/mm. It is further anticipated that these structures will allow

for a reduced gate length/barrier thickness ratio requirement below 15 which was

originally suggested for GaN HEMT technology that lacked the back barrier [120].

Through this particular advancement, GaN HEMT technology is expected to rival

if not surpass its GaAs contenders in frequency performance.

124

6.2 Design of a dual-channel charge-screening AlN/GaN HEMT; an epitaxial

approach to mitigate frequency dispersion

A number of multi-channel HEMT designs have been proposed and attempted,

usually as a means to increase drain current density [117], [115]. One of the more

unconventional and interesting multi-channel HEMT devices was developed by

Capasso [123]. In this work, two channels were present only below the gate and

showed a resonant tunneling peak in the transfer characteristic. The operation of

these designs are typically foiled by the difficulty of making ohmic contact to the

buried channel. Therefore, they have remained a research concept and have yet

to graduate to become practically useful devices.

Surface state passivation refers to the modification of the donor surface state

such that the surface is less prone to charging. The most popular method for

passivation of surface states is achieved by the deposition of a dielectric such as

SiN2 within the access regions of a FET but has also been achieved by the intro-

duction of Flourine ions [115] and other chemical treatments. However, another

issue becomes apparent when the surface is passivated by a dielectric. The surface

of the dielectric can act as a new charging surface which sets a minimum passi-

vation material thickness limit. Palacios suggests a SiN thickness of 100 nm [115]

while a recent study by Pei demonstrated improved power performance in an Al-

GaN/GaN HEMT using 120 nm SiN [124]. In either case it seems a SiN thickness

exceeding 80 nm is required. A common byproduct of the use of a passivation

material is the observation of an increase in 2DEG charge density. This was dis-

cussed previously in Chapter 2 for various dielectric materials on the AlN/GaN

heterostructure. However, this is not to say that an increase in 2DEG density as

2SiN has had a longer history in its role as a passivation dielectric for GaAs-based devices[125]. Largely because of this reason was it implemented with GaN-based devices.

125

a result of a deposited dielectric is indicative of a sufficiently passivated surface.

The primary reason for the implementation of surface passivation to HEMTs

is that it minimizes the deviation of the dynamic I-V characteristics of a transistor

from the DC I-V characteristics. This deviation between the two different opera-

tion regimes is commonly known as DC-to-RF dispersion, or simply dispersion and

is a major issue for the III-Nitride HEMT family. The detailed manner by which

surface charging causes the deviation between DC and RF characteristics can be

understood by considering the method by which a charge becomes trapped. If an

electron is introduced to a surface state that willingly accepts negative charge it

can then become “trapped” in that state. For it to be “de-trapped” requires a pre-

scribed amount of energy that can be gained thermally, electrically, or optically.

Despite the method there is a certain amount of time the trapped electron takes

to become un-trapped. If this amount of time exceeds the duration that energy

is supplied, the trapped charge stays trapped and vice versa. Therefore, during

DC operation trapped charges in a HEMT have a significant amount of time (and

energy) to become trapped and de-trapped but this is not the case for RF oper-

ation when the HEMT undergoes periodic voltage oscillations. These oscillations

can have too short of a period to de-trap charge thus causing the compression

of I-V characteristics which ultimately reduces the device’s achievable RF output

performance.

One model that has gained significant popularity for III-Nitride HEMTs is the

virtual gate model [126]. The concept of the virtual gate extension is illustrated

in Fig. 6.6 for the AlN/GaN HEMT. Three charged states can be defined during

a single gate pulse cycle as demonstrated in the figure. In state 1 the device is

biased under sub-threshold conditions with a drain-source bias applied. This state

126

Figure 6.6. Virtual gate model showing the time plot of the surfacecharge and channel depletion.

is given by the intentional depletion of the channel through the gate bias. Because

of the large fields encountered in this state the surface of an unpassivated device

becomes charged in a region between the gate and drain where the most intense

electric field occurs. This region causes the gate to look electronically longer and

is labeled “Leff” for “effective gate length” in Fig. 6.6. Once the gate voltage is

made to open the channel such as in stage 2, the drain current cannot respond

immediately due to the fact that it takes a discrete amount of time to discharge the

access region trap states. The charged region between the gate and drain remains

partially charged depending on the length of the gate voltage pulse. Herein lies

the reason for the “virtual gate extension” naming convention since this region

serves to keep the channel partially depleted along a length slightly larger than

the metallurgical gate length. Finally, the drain current reaches its full value for

the gate pulse width (stage 3) after a sufficient amount of time for the traps to

127

discharge.

The steps just outlined describing the concept behind the virtual gate model

also describes a measurement known as gate lag. Gate lag measures the ratio of

the deviation in the measured open channel DC drain current to that of a gate

pulsed drain current. A ratio less than 1 indicates the severity of surface charging

in the HEMT. Because of its simplicity, this measurement is widely used as a quick

informational feedback measurement indicating the degree to which large-signal

performance might be hindered. The time-dependent reduction in the GL ratio is

indicative of additional charging which could be due to oxide-AlN interface states.

This sinking in drain current with time under bias is indicative of trap charging

causing partial depletion and can be a severe detriment to the HEMT’s power

performance. This mechanism is a tell-tale sign of what is termed “power drift”

and a good qualitative explanation can be found in reference [127]. These issues

are at the heart of HEMT design for high frequency power performance.

6.2.1 Multi-channel heterostructure design

The first question sought to answer in the following study is whether the

formation of two parallel channels is achievable with an AlN/GaN/AlN/GaN het-

erostructure. If so, the second is whether the top channel can be successfully

exploited as a charge supply layer that will effectively screen the electric field

extension arising from surface charging near the drain edge of the gate. The

thicknesses of the active layers are of critical importance in the design and thus,

the band diagram is a necessary design tool.

Akin to the method used in Chapter 2 and the Appendices, a voltage loop may

be summed about the conduction band profile for an oxide/AlN/GaN/AlN/GaN

128

Figure 6.7. Cross section and corresponding band diagrams for thedesign of the charge screening HEMT.

multi-heterostructure (MH) which yields the following,

− VG + ΦB − VT −∆Ecq

+ V1 +∆Ecq− VB −

∆Ecq

+ E02 + (Ef − E02) + VS = 0, (6.1)

where V1 accounts for the voltage drop in the GaN spacer that harbors the upper

2DEG. Assuming single sub-band population of both quantum wells and using

Gauss’s Law, the voltage loop may be written in terms of the applied voltage and

129

charges in the structure,

VGS = (ΦB −∆Ecq

) + (9π~q

8εG√

8m∗)2/3[(qns1)

2/3 + (qns2)2/3]

−Qnet(tT + tBεA

) + qns1(tTεA

+π~2

m∗)

+ qns2(tT + tBεA

+tGεG

+π~2

m∗), (6.2)

where ΦB and ∆Ec are the metal-oxide barrier height and AlN/GaN conduction

band offsets, respectively; ns1 and ns2 are the top and bottom channel mobile

2DEG densities, respectively; Qnet is the net fixed polarization charge at the

AlN/GaN interface taking into account both spontaneous polarization of both

materials as well as piezoelectric polarization of the strained AlN lattice, tT is the

thickness of the top AlN cap, tB is the thickness of the bottom AlN layer, and the

relative dielectric constants of the respective layers take the form, εx = εr,xεo.

As per the discussion in the introduction of this section, the cause of DC-

RF frequency dispersion is commonly taken to be due to the virtual gate effect

or access region surface charing between the gate and drain of a HEMT. Based

on this principle, the design logos of the surface potential fluctuation screening

HEMT is to sacrifice the top 2DEG as a screening layer for the bottom channel

be protected from surface fluctuations in the gate-to-drain access region. This

sets the governing design criterion, the HEMT needs a more negative Vth in the

access region (Fig. 6.7 (b)) than in the gated intrinsic region (Fig. 6.7 (a)).

This will allow surface charging to have no effect on the underlying channel due

to the effective screening from the upper channel. Because the 2DEG density

is dependent on AlN thickness, device dimensions can be tailored to achieve the

desired effect.

130

The gated region will therefore require a recess etch through the top AlN and

GaN layers such that when a gate voltage is applied, it modulates only the bottom

channel. The gate voltage-2DEG charge dynamics are given in Chapter 2 by Eq.

2.6 or 2.7. Maintaining the original MH in the access regions will be necessary

in achieving the desired screening effect. In the access region on the drain side

of the gate Eq. 6.2 can be used to determine pinch-off of both channels. It is

assumed for the sake of simplicity that both wells are decoupled from each other

(in equilibrium both wells are simultaneously coupled to the same Fermi level).

Imposing flatband conditions on Eq. 6.2 and grouping charge terms the threshold

voltage in the access region can be simplified to,

Vth,access(ns) = −(

9π~q8εG√

8m∗

)2/3

[(qns1)2/3 + (qns2)

2/3]− qns1(tTεA

+π~2

qm∗

)− qns2

(tT + tBεA

+tGεG

+π~2

m∗

). (6.3)

It needs to be ensured that the access region threshold voltage is much more neg-

ative than underneath the gate. Shen discussed this design concept for thick GaN

capped AlGaN/GaN HEMTs in detail [128]. This condition allows a numerical

design constraint for the structure and can be written: Vth[Channel 2, (b)] <<

Vth[Channel 2, (a)]. Figure 6.8 shows a comparison between the threshold voltage

within the gated region versus the threshold voltage within the access region for

various GaN spacer thicknesses ranging between 10 to 20 nm. The horizontal

dashed line, indicating threshold voltage for the gated region, does not vary as

the lower AlN barrier is taken to be fixed at 3 nm (this threshold voltage is for

comparison). 2DEG charge density of the top channel is plotted along the x-axis

as an indicator of the top AlN cap thickness. Threshold voltage refers to complete

131

Figure 6.8. Threshold voltage comparison. The top horizontal curve isfor a fixed charge density of the lower channel with a fixed 3 nm barrierand is for comparison to the access region Vth. The set of 3 lower curvesshow the calculated threshold voltage for the access region with the 2nd

2DEG distribution for screening. Access region threshold voltagebecomes increasingly more negative for increased upper 2DEG density

and this increase in amplified for thicker GaN spacer layers.

132

2DEG charge elimination of both channels for the bottom three curves. Interface

trap state charging is not taken into account. It can be seen qualitatively, that

the total threshold voltage for the access region falls off very quickly in negative

voltage for increasing charge density (top AlN cap thickness) and more so for a

thicker GaN spacer. Based off these results a device has been designed, grown,

fabricated, and analyzed. The following section involves this study.

6.2.2 A dual-channel AlN/GaN HEMT

The dual-channel structures were grown by plasma-assisted MBE by David

Storm at the Naval Research Laboratory following similar growth details given

in Chapter 3. Semi-insulating HVPE GaN substrates were used as the template

for growth. After a 60 second nitridation clean of the substrate a 1.5 nm AlN

nucleation layer was grown followed by a 1.3 µm thick layer of Be-doped GaN as

a compensation dopant to oxygen impurities. These layers were followed by 200

nm unintentionally doped GaN, 3 nm AlN, 15 nm GaN, and the final 3 nm AlN

cap. All layers were grown without interrupts and the Ga shutter was left open

during all layer growths.

The as-grown structure after Ga-droplet removal resulted in a sheet resistance

of 210 Ω/ measured by a contactless sheet resistance measurement. AFM indi-

cated a smooth surface (RMS roughness = 0.64 nm for a 2 × 2 µm2 scan) without

indication of lattice relaxation of the AlN layers (Fig. 6.9). Sample processing

followed an ohmic first process however, during this stage the entire top AlN and

GaN layers were removed by a BCl3/Cl2 ICP/RIE etch. The design does not

require low resistance ohmic contact to the upper 2DEG since the main concern

is the threshold voltages referenced to the Fermi level deep into the GaN buffer.

133

Figure 6.9. (right) 2×2 µm AFM scan of the AlN surface of thedual-channel structure.

No sidewall protection was used during the e-beam deposition of the ohmic metal

stack. An 860 C rapid thermal anneal was performed for 30 seconds to diffuse

the contact metal and make contact with the lower channel. This resulted in

contact resistances of ∼2 Ω-mm as determined by circular TLM. A 50 nm deep

mesa isolation etch was performed in a Cl-based plasma. E-beam lithography was

used to define the 200 nm long gate recess etch lines in some of the devices where

others were left unwritten so that both the screening HEMT and a dual channel

HEMT could simultaneously be realized on the same wafer. After a ∼17 nm gate

recess etch the samples were sent to U. Notre Dame for ALD Al2O3 gate insulator

deposition and received a 7 nm thin film3. Finally, 300 - 400 nm long e-beam gates

were written and deposited using a Ni/Au metal stack and holes in the oxide were

etched in a buffered HF acid in order access the ohmic pads.

Post oxidation Hall effect measurements were performed and the sheet resis-

3Zongyang Hu is gratefully acknowledged for performing the ALD oxide depositions at theU. Notre Dame.

134

Figure 6.10. C-V analysis of the Al2O3/AlN/GaN/AlN/GaNheterostructure capacitor (left) and the associated approximate charge

profile (right).

135

tance and constituent parameters were determined to be Rsh = 180 Ω/, ns =

2.3 × 1013 cm−2, µ ∼ 1500 cm2/Vs. The measured 2DEG density is lower than

expected if one assumes a 2X increase in the single 3 nm thick AlN barrier het-

erojunction. However, given the thick GaN spacer, second AlN layer, and the

ensuing polarization charge densities, it is assumed that the lower 2DEG density

is small. C-V measurements on 100 µm diameter dots at 10 MHz showed two dis-

tinct plateaus indicating two separate charge distributions in the heterostructure

(Fig. 6.10). The plot shows three curves with corresponding sweep directions.

Curve (1) was the initial sweep while curve (2) followed in the same direction and

curve (3) swept from low to high gate bias voltage. The curve evolution from (2)

to (3) shows some hysteresis indicating oxide charge or possibly confined mobile

charge in the GaN layer. It is hard to say the origin of the hysteresis in this

structure given its complexity.

A clear deviation is seen between curves (1) and subsequent curves and is

speculated to be due to a “burn in” effect that may amount to simply the charging

of interface and other traps in the hetero-system. Nonetheless, it is assumed that

the first curve gives the best evidence of the pristine heterostructure. Thus, the

integration of the smaller capacitance plateau yields a charge density of 1.5 × 1013

cm−2 associated with the lower 2DEG. The integration of the second capacitance

plateau yields a combined charge density of 2.6 × 1013 cm−2 which is in decent

agreement with the charge density measured by Hall effect. The difference tells

the upper 2DEG density and was found to be 1.1 × 1013 cm−2. It is noted that

this is the first demonstration of two parallel 2DEG channels that are formed in

an all binary AlN/GaN heterostructure.

Both dual-channel HEMTs and charge screening HEMTs were fabricated on

136

Figure 6.11. (left) Transfer characteristics of both the dual-channel andcharge-screening HEMT, (middle) drain characteristics of the

charge-screening HEMT, and (right) drain characteristics of thedual-channel HEMT.

the same wafer and their DC transfer (left) and drain (right) characteristics are

shown in Fig. 6.11. Interestingly, the dual channel HEMT appeared to have

a maximum drain current of ∼2X that of the charge screening HEMT which

indicates there was in fact efficient contact made to the etched sidewalls of the

top channel. Therefore, it as taken that both channels are contacted by the source

and drain contacts and may pass current to some extent. The quality of contact

to the upper channel is unknown. The dual channel HEMT was anticipated to

shown a dual region transfer characteristic based on the C-V results but little if

any was observed. However, the charge screening HEMT showed peculiar transfer

characteristics in its lack of a well defined maximum current density and non-

linear characteristic. A multi-peaked gm was observed, although unexpected. It is

speculated that there is potential communication between channels in the access

137

Figure 6.12. Small signal frequency performance of the charge screeningHEMT (left) and the dual channel HEMT (right).

regions that frustrate the typical linear transfer characteristic, but this is yet to

be determined. Nonetheless, the measured transfer and drain characteristics show

modulating channels that may be biased into a sub-threshold regime. Therefore,

a test for DC-RF dispersion is warranted and can be carried out through pulsed

gate lag measurements as discussed above.

Small signal performance was measured up to 50 GHz of both the dual channel

HEMT and the charge screening HEMT and is shown in Fig. 6.12. With gate

lengths being ∼ 300 nm in length for both device structures, ft and fmax were

measured to be 27 GHz and 46 GHz, respectively for the charge screening HEMT

and 48 GHz and 60 GHz, respectively for the dual channel HEMT. The cause for

the dual channel HEMT having better frequency performance is not well under-

stood but is likely to be due, at least in part, to a lower access resistance from

138

the two parallel channels. Nonetheless, the frequency performance measured is

respectable for these “first of their kind” devices and demonstrates the potential

of these designs for high frequency amplification.

The results of pulsed gate lag measurements are shown in Fig. 6.13 for the (a)

dual channel AlN/GaN HEMT, the (b) charge-screening HEMT, and (c) a stan-

dard single channel AlN/GaN HEMT for comparison. The dual channel HEMT in

(a) demonstrated a gate lag ratio (GLR) of 0.74 which reduced over time while the

HEMT was biased in sub-threshold to 0.67 for subsequent pulses. It is assumed

that this reduction is mainly due to surface state charging and corresponding de-

pletion of the upper channel since it was shown that the upper channel makes up

a large portion of the total drain current for these dual-channel HEMTs. After

restarting the measurement the open channel drain current was found to have

diminished by 82 mA/mm which is indicative of 2DEG channel depletion and

possibly some buffer trapping. The charge-screening HEMT in (b) demonstrated

an unchanging GLR of 0.86 for all pulses. While a GLR ratio of 1.0 is indica-

tive of no channel depletion and is preferred, there was some charging that led

to channel depletion as evidenced by the 1.0 GLR. The origin of this trapping

is unknown at this time but could be due to interface trap state charging of the

GaN spacer/AlN interface of the lower channel since the upper channel 2DEG

could act as a charge supply for this interface traps. This condition is very un-

desirable since it would trump the very purpose of the design. However, it is

only speculated at this time whether this interface is the source of trapping. An

important positive result of the charge-screening HEMT is that after the GLR

sequence was stopped and restarted, it was found that the drain current density

had not diminished in the slightest. The constant GLR of 0.86 is relatively good

139

Figure 6.13. Time profiles of the pulsed-gate lag measurements for (a)the dual channel HEMT, (b) the charge-screening HEMT, and (c) a

standard single heterojunction AlN/GaN HEMT for comparison.

140

given no passivation other than a thin 7 nm Al2O3 layer was deposited. As well as

no slump in the open channel drain current [127]. It is therefore concluded that

the method of employing a sacrificial charge screening layer for the mitigation of

DC-RF frequency dispersion is a novel design that has indeed shown promise in its

functionality through proof-of-concept devices shown here. Further refinements

in the contacts, gating process, and layer structure are anticipated to advance the

design to perfectly remove the detrimental frequency dispersion observed in these

ultra-shallow channel AlN/GaN HEMTs.

6.3 Chapter 6 cited literature

110. C. Bayram, Z. Vashaei, M. Razeghi, “AlN/GaN double-barrier resonanttunneling diodes grown by metal-organic chemical vapor deposition”, Appl.Phys. Lett. 96, 042103, 2010.

111. S. Yagi, X.-Q. Shen, Y. Kawakami, T. Ide, M. Shimizu, “Demonstrationof Quasi-AlGaN/GaN HFET Using Ultrathin GaN/AlN Superlattices as aBarrier Layer”, IEEE Elec. Dev. Lett. 31, 9, 2010.

112. T. Zimmermann, D. Deen, Y. Cao, J. Simon, P. Fay, D. Jena, and H.G.Xing, “AlN/GaN Insulated-Gate HEMTs With 2.3 A/mm Output Currentand 480 mS/mm Transconductance”, IEEE Elec. Dev. Lett. Vol. 29, No.7, 2008.

113. C.Y. Chang, S.J. Pearton, C.F. Lo, F. Ren, I.I. Kravchenko, A.M. Dabiran,A.M. Wowchak, B. Cui, P.P. Chow, “Development of enhancement modeAlN/GaN high electron mobility transistors”, Appl. Phys. Lett. 94, 263505,2009.

114. A.L. Corrion, K. Shinohara, D. Regan, I. Milosavljevic, P. Hashimoto, P.J.Willadsen, A. Schmitz, D.C. Wheeler, C.M. Butler, D. Brown, S.D. Burn-ham, M. Micovic, “Enhancement-Mode AlN/GaN/AlGaN DHFET with 700mS/mm gm and 112 GHz ft”, International Workshop on Nitride Semicon-ductors, 2010.

115. T. Palacios, ”Optimization of the High Frequency Performance of Nitride-Based Transistors,” Ph.D. Dissertation, University of California, Santa Bar-bara, 2006.

141

116. M. Micovic, P. Hashimoto, M. Hu, I. Milosavljevic, J. Duvall, P.J. Willadsen,W.-S. Wong, A.M. Conway, A. Kurdoghlian, P.W. Deelman, J.-S. Moon, A.Schmitz, M.J. Delaney, “GaN Double Heterojunction Field Effect TransistorFor Microwave and Millimeterwave Power Applications”, IEEE IEDM 4,807, 2004.

117. R. Chu, Z. Chen, S.P. DenBaars, U.K. Mishra, “V-Gate GaN HEMTs WithEngineered Buffer for Normally Off Operation”, IEEE Elec. Dev. Letts.,Vol. 29, No. 11, 2008.

118. S. Rajan, M. Wong, Y. Fu, F. Wu, J.S. Speck, U.K. Mishra, “Growth andElectrical Characterization of N-face AlGaN/GaN Heterostructures”, Jap.J. of Appl. Phys., Vol. 4, No. 49, 2005.

119. S. Rajan, “Advanced Polarization Engineering for GaN-based Transistors”,Ph.D. Dissertation, University of California Santa Barbara, 2006.

120. G.H. Jessen, R.C. Fitch, J.K. Gillespie, G. Via, A. Crespo, D. Langley, D.J.Denninghoff, M. Trejo, E.R. Heller, “Short-Channel Effect Limitations onHigh-Frequency Operation of AlGaN/GaN HEMTs for T-Gate Devices”,IEEE Trans. Elec. Dev., Vol., No. 2007.

121. K. Shinohara, A. Corrion, D. Regan, I. Milosavljevic, D. Brown, S. Burn-ham, P.J. Willadsen, C. Butler, A. Schmitz, D. Wheeler, A. Fung, M. Mi-covic, “220 GHz ft and 400 GHz fmax in 40-nm DH-HEMTs with Re-grownOhmic”, International Electron Devices Meeting, San Francisco, CA 2010.

122. G. Li, T. Zimmermann, Y. Cao, C. Lian, X. Xing, R. Wang, P. Fay, H.G.Xing, D. Jena, “Threshold Voltage Control in Al0.72Ga0.28N/AlN/GaN HEMTsby Work-Function Engineering”, IEEE Elec. Dev. Letts., Vol. 31, No. 9,2010.

123. F. Capasso, S. Sen, F. Beltram, A.Y. Cho, “Resonant Tunneling Gate FieldEffect Transistor”, IEE Elec. Lett, No. 5, Vol. 23, 1987.

124. Y. Pei, S. Rajan, M. Higashiwaki, Z. Chen, S.P. DenBaars, U.K. Mishra, “Ef-fect of Dielectric Thickness on Power Performance of AlGaN/GaN HEMTs”,IEEE Elec. Dev. Lett. No. 4, Vol. 30, 2009.

125. Y.C. Chou, R. Lai, G.P. Li, J. Hua, P. Nam, R. Grundbacher, H.K. Kim, Y.Ra, M. Biedenbender, E. Ahlers, M. Barsky, A. Oki, D. Streit, “InnovativeNitride Passivated Pseudomorphic GaAs HEMTs”, IEEE Elec. Dev. Lett.No. 1, Vol. 24, 2003.

142

126. R. Coffie, “Characterizing and Suppressing DC-to-RF Dispersion in Al-GaN/GaN High Electron Mobility Transistors”, Ph.D. dissertation, Uni-versity of California Santa Barbara, 2003.

127. J.C.M. Hwang, “Relationship between gate lag, power drift, and powerslump of pseudomorphic high electron mobility transistors”, Sol. Stat. Elec.,Vol. 43, pp. 1325-1331, 1999.

128. L. Shen, “Advanced Polarization-Based Design of AlGaN/GaN HEMTs”,Ph.D. Dissertation, University of California, Santa Barbara, 2004.

143

CHAPTER 7

SUMMARY AND FUTURE WORK

7.1 Summary of this work

This work has sought to develop and advance GaN-based HEMT technology

through the ultra-thin barrier AlN/GaN HEMT along with some novel deriva-

tives. Furthermore, several methodical adaptions have been developed as theo-

retical contributions which assist to better understand the various aspects of the

AlN/GaN HEMT.

The work in Chapter 2 laid the foundation for the design of several detailed

points of the AlN/GaN HEMT. An ultra-shallow pre-metallization etch was de-

veloped for the reduction of Rc. This method allowed the reduction of contact

resistance down to ∼0.3 Ω-mm, the lowest values for post-growth processed con-

tacts. A variety of gate insulators were investigated and their performance was

gauged by several standard metrics. In particular, Terman’s high frequency C-V

method for Dit extraction at the oxide/semiconductor interface was adapted for

the heterojunction capacitor and applied to the oxide/AlN/GaN heterostructure

capacitors under comparison. It was found that a high Dit was determined for

as-deposited films.

A gamut of insulated-gate AlN/GaN HEMT experiments were carried out and

in the process, several records were set for different performance metrics. In partic-

ular, the highest drain current density at the time of its inception of 2.3 A/mm by

144

a GaN-based HEMT was obtained by the AlN/GaN HEMT. The off-state break-

down voltage for the ultra-thin barrier AlN/GaN HEMT was extended to 96V by

the employment of an ALD Ta2O5 gate dielectric. A question that has persisted

regarding the maximum theoretical drain current density of the AlN/GaN HEMT

was addressed. Given such high 2DEG density with an accompanying high mobil-

ity, it can be calculated the a maximum drain current density well above 4 A/mm

is theoretically obtainable, yet no such current has been shown. The gradual chan-

nel approximation has been adapted for the insulated-gate AlN/GaN HEMT with

the inclusion of two practical mechanisms, that of contact resistance and electron

velocity saturation. It was shown that both mechanisms working together may

act to reduce drain current to the level achieved in the presented devices.

A simple lumped-element small signal model was developed based on con-

ventional FET circuit topology which allowed the modeling and de-embedding

of selected parameters from the measured S-parameters. It was found that the

removal of the pad impedance from the S - Z parameters yielded a 10 - 15 GHz

increase in unity current gain, ft. Accompanying the small signal modeling, de-

lay analysis was performed on the AlN/GaN HEMT based off Moll’s method. It

was found that the RC charging time of the gate-to-channel capacitances was the

dominant delay for AlN/GaN HEMTs. The results were compared to those of the

AlGaN/GaN HEMT. Furthermore, a conjoined method utilizing the Y-subtraction

de-embedding procedure to remove modeled extrinsic parameters allowed the ap-

proximation of effective electron velocity in the channel and was found to be ∼1.18

× 107 cm/s for the AlN/GaN HEMT in agreement with values reported in liter-

ature for its contemporary, the AlGaN/GaN HEMT. Points of improvement were

highlighted based on the given analysis.

145

A temporary digression from the AlN/GaN HEMT was made in Chapter 5 with

a specially designed test structure that was used to monitor trapped charge in a

SiN passivation layer for a conventional AlGaN/GaN HEMT. Negative threshold

voltage shifts of MIS gates within the gate-drain access region after bias-stressing

the HEMTs was observed and indicated the occurrence of either electron de-

trapping or hole trapping. A qualitative model was developed for the mechanism

involved which suggested hole trapping was cause based on gate current measure-

ments following Chynoweth’s model.

Finally, several multi-heterostructure devices were designed based off quali-

tative performance goals and corresponding proof-of-concept devices were fabri-

cated. A set of samples that explored an Al0.13GaN back barrier were fabricated

and analyzed. These device structures were the first of their kind. Additionally, a

multi-heterostructure dual-channel AlN/GaN HEMT was designed and fabricated

with the aim to exploit the ultra-thin barrier structure for increased drain current

density and mitigation of RF-DC dispersion.

The evolution from improved component design of the AlN/GaN HEMT to

the sophistication of multi-heterostructure based devices has been demonstrated.

From in-depth examinations of various aspects of the ultra-thin AlN/GaN HEMT

its potential has been emphasized and several records have been set in the pro-

cess. Moreover, several novel designs that utilize the AlN/GaN HEMT have been

proposed and realized. All of which have contributed to the advancement of III-

Nitride HEMT technology based on the extraordinary AlN/GaN HEMT structure.

146

7.2 Future work

The HEMT devices in this work show some of the best operation in terms

of standard performance metrics, or are the first of their kind. However, there

are many issues that burden the AlN/GaN HEMT from obtaining its theoretical

capabilities. Despite ohmic contact resistance as low as ∼0.2 - 0.5 Ωmm, achieved

through ultra-shallow pre-metallization etching, the formation of low resistance

ohmic contacts to AlN/GaN heterostructures remains difficult. As shown through

the GCA modeling of the AlN/GaN HEMT in Chapter 3, contact resistance is a

major factor in reducing transconductance and subsequently extrinsic frequency

performance. Recently, researchers at HRL demonstrated extremely low contact

resistance of 0.05 Ωmm to AlN/GaN heterostructures through an n+ regrowth

[129]. This method may be one of the better avenues for low contact resistance

but sacrifices processing simplicity. Ion implanting, advanced lithography and

etching techniques, as-grown n+ caps, and so on are all methods which have yet

to be tried and may offer other ways to reduce contact resistance for the AlN/GaN

HEMT.

The variety of options for dielectrics used for gate insulation is growing and

some have more favorable properties than others for the use in gate insulation

of the AlN/GaN HEMT. Atomic layer deposition of dielectrics has proven to be

convenient and produce high quality oxides. Very little has been explored in the

way of improvement of the oxide-semiconductor interface or the bulk oxide itself

by post-deposition thermal annealing. It is anticipated that vast improvements

may be made to the HEMT’s charge control and thereby transconductance and

frequency performance simply by removing or controlling the interfacial charge

states or mobile oxide charge in the layer structure.

147

Power performance of AlN/GaN HEMTs is as of yet unexplored. The pre-

dominant reason for this is the premature breakdown due to gate insulator failure

typically observed in these ultra-thin barrier HEMTs. However, there is ample

reason to pursue the advancement of power performance through the AlN/GaN

HEMT since the heterostructure has demonstrated that it can generate the high-

est mobile 2DEGs of any single heterostructure in any III-V system, and therefore,

potentially provide the highest current density of them as well. Several issues need

to be address in order to enable the AlN/GaN HEMT to be applicable for power

FET applications.

The application of field plated gates to the AlN/GaN HEMT are as of yet

unexplored. While some measure of frequency performance is given up when a

metallurgical gate extension is included, it is conceivable that for these ultra-thin

barrier AlN/GaN HEMTs there is ample frequency bandwidth to sacrifice as a

trade-off for improved power performance through increasing off-state breakdown

voltage by field plate design.

Terman’s method was developed for the oxide/heterojunction III-Nitride ca-

pacitor in Appendix A and applied to the oxide/AlN/GaN capacitor stack in

Chapter 2. The method allows for the extraction of the Dit spectrum provided

band discontinuities are well known and all other charge is accounted for. How-

ever, a number of caveats remain. Only a single sub-band was assumed in the

derivation which will lead to some error for high charge densities. Conduction

band non-parabolicity was not considered but is a likely occurrence for 2DEGs

since their associated energy level is further away from the conduction band edge.

Band non-parabolicity accounts for the softer “rolling off” of the C-V curve as

the 2DEG begins to deplete. Therefore, not accounting for band non-parabolicity

148

causes an increased ∆VGS and subsequently an over estimation of Dit near the

conduction band edge for sufficient non-parabolicity.

Termination of the AlN crystal leaves dangling surface bonds and a termination

of the assumed infinite periodic potential at the crystal boundary which has been

shown to cause increased surface state density near band edge [130] which can

amount to a band gap reduction. This mechanism has not been accounted for. A

reduction in barrier band gap, Eg, can occur for pseudomorphically-strained layer

structures [131] but is not included in the derivation given in Appendix A since

it amounts to a simple scaling of Eg and subsequently, ∆Ec. However, it is noted

that this is an important aspect to consider depending on the hetero-system under

analysis and by not taking energy barrier reduction into account, the result can

lead to misleading values in corresponding energy for the Dit spectrum. This is

particularly true for the III-Nitride system.

Regarding the noted improvements that may be had with the modification of

Terman’s method, it is noted that the development of the method does not ac-

count for the reality of the Fermi-Dirac (F-D) distribution of states to spread with

temperature. The development as given essentially “collapses” the F-D distribu-

tion to the Fermi level. Therefore, an artificial spreading of the trap distribution

occurs due to the mathematics in the method not accounting for this thermal dis-

tribution spreading of states. Therefore, this suggests that C-V measurements at

cryogenic temperatures be taken on the oxide/AlN/GaN heterostructure capaci-

tor to impose the condition on the system that more closely matches the assumed

mathematical distribution.

Surface state charging leading to current collapse and power slump is a partic-

ularly pressing issue with these ultra-shallow channels. Considering the feasibility

149

of fabricating AlN/GaN HEMTs with barriers as thin as 1.5 nm, and recalling the

propensity of the electron wave function to tunnel through thin barriers as well

as the proven generation of hot electrons in GaN channels, a question becomes

apparent of the possibility that these ultra-thin channel HEMTs may inherently

be inferior for power performance applications where high frequencies are also

required. The possibility that the high density 2DEG can efficiently communi-

cate with trap states at the AlN surface and act as a source of charge for those

traps may be to the HEMT’s own demise if the AlN surface can not be effectively

passivated. No theoretical consideration has been done along these lines of consid-

eration and may serve as part of the answer to this question. Although experiment

is the best final judge, to which the device structure proposed in Chapter 5 applied

to the AlN/GaN HEMT would be an ideal method for measuring such an effect.

Finally, regarding the development of the gradual channel approximation for

the AlN/GaN HEMT, the intrinsic region as it is analyzed in the given work is

taken only as a single region where the approximation is held valid. However,

this is certainly not a realistic picture and is likely in gross error for GaN-based

HEMTs, in particular. Since GaN may sustain such large electric fields before

critical breakdown, the intrinsic depletion region becomes very exaggerated near

the drain edge of the gate where the highest intensity fields exist. In this region,

all mobile charge is nearly depleted and fixed ionization centers are filled, thus

making the mathematics follow Laplace’s equation closer than Poisson’s equation.

Grebene and Ghandhi showed that the solution in this region is the linear super-

position of hyperbolic cosine functions when their fixed ionized donors were left

[132]. The picture for GaN channels differs in that the depletion region is the same

region responsible for conduction of current at the onset of the saturated current

150

characteristic (i.e. pinch-off) rather than the un-depleted portion of the channel

responsible in the doped-channel MESFET. Based on this consideration it would

be an interesting problem to see the effect of solving Laplace’s equation in the

depleted yet conducting region of a GaN channel. Furthermore, a question arises

based off the considerations of electron velocity saturation, that of the possibility

that channel pinch-off occurs before the onset of velocity saturation and not simul-

taneously as Shockley assumed. This may lead to a better understanding of the

I-V characteristic in the transition region between the low-field (linear) operation

regime and the high-field (saturated) regime.

7.3 Chapter 7 cited literature

129. A.L. Corrion, K. Shinohara, D. Regan, I. Milosavljevic, P. Hashimoto, PlJ.Willadsen, A. Schmitz, D.C. Wheeler, C.M. Butler, D. Brown, S.D. Burn-ham, M. Micovic, “Enhancement-Mode AlN/GaN/AlGaN DHFET with 700mS/mm gm and 112 GHz ft”, International Workshop on Nitride Semicon-ductors, 2010.

130. C. Kittel, Introduction to Solid State Physics ; 8th ed., John Wiley and Sons,New Jersey, 2005.

131. B.K. Ridley, W.J. Schaff, L.F. Eastman, J. Appl. Phys. 94, 3972, 2003.

132. A.B. Grebene, S.K. Ghandhi, “General theory for pinched operation of thejunction-gate FET”, Sol. Sta. Elec., vol. 12, pp. 573-589, 1969.

151

APPENDIX A

THE HETEROJUNCTION CAPACITOR;TERMAN’S METHOD

Terman’s method for the determination of interface-state trap density spec-

trum is adapted here from its original application (oxide/semiconductor junction)

[133] to one applicable to the polarization-doped insulator/AlN/GaN layer struc-

ture. A few points are worthy of mention as a preface for the derivation given

below. Terman’s method is a high-frequency C-V technique that hinges on the

trap time constant being too slow to respond to the ac voltage signal that rides on

the dc voltage sweep. Rather, the dc sweep voltage is adequately slow for the trap

to populate or de-populate. Therefore, the interfacial trap charge manifests itself

by a “smearing” of the C-V curve within the depletion zone as the trap states are

(de)populated when the fermi level is swept through the trap state energy (via

capacitor plate voltage, VGS). It is necessary, then, to have a foreknowledge of

the C-V operation of the structure if it were completely absent of any traps (such

a condition is considered “ideal”) in order to quantify the C-V smearing. Such is

the foundation of the method. This also imposes the constraint, that any trapped

charge that is extracted from the measurement is inherently below the ac signal

frequency. Therefore, traps that have a response time faster than this ac signal

frequency cannot be measured by Terman’s method. Interfacial trap states may

be donor or acceptor like. A donor-like trap is one in which the removal of an

electron from the state causes the state to become neutral or positively charged.

152

An acceptor-like trap is one in which the capture of an electron results in the state

becoming neutral or negatively charged.

A primary assumption is made in the following derivation, that all trapped

charge occurs from trap states at the insulator/AlN interface. This assumption

is fair considering the III-Nitride terminal surface is laden with dangling atomic

bonds as well as a surface polarization charge density. However, it is reasonable

that the insulator may also have trap states dispersed throughout it as well with

an unknown distribution. Terman’s method handles these traps by imaging any

such insulator trapped charge to the insulator/AlN interface. The same goes for

traps in the buffer layer.

The following development includes four parts. First, the ideal heterojunction

capacitance configuration is found (A.1.1) followed by the capacitance configura-

tion including trap states (A.1.2). The form for Dit is then given (A.2) and lastly,

a few remarks regarding the extent of the Terman methodology (A.3).

A.1 Capacitance-Voltage Relationship

The capacitance-voltage functionality of the dielectric/heterojunction capaci-

tor may be understood by employing the charge control model within the gated

region and considering the charge distribution shown in Fig. B.2. The voltage loop

equation (2.3) remains the same and the additional charge manifests itself through

the voltage-electric field relationship. By charge neutrality Qπnet = Qπ

AlN − QπGaN

and Qm = QπAlN + Qπ

net + Qit + Qw. Qπnet is the net polarization charge at the

AlN/GaN interface. QπAlN is the fixed polarization (spontaneous + piezoelectric)

charge of the AlN cap strained to the GaN lattice and QπGaN is the fixed polariza-

tion (spontaneous) charge of the GaN. The charge on the metal, Qm, is the metal’s

153

response to maintain charge neutrality and therefore is equal to the sum of charge

in the structure. Qw = −qns is the charge associated with the 2DEG and Qit is

the charge associated with the interfacial traps. Recalling the definition of capac-

itance (C = ∂Q/∂V ) it is noted that the total capacitance of the structure will be

the change in (image) charge on the metal gate, Qm, with respect to the change

in applied voltage, VGS. Given the above charge equations the magnitudes1 of

the local electric fields may be found by Gauss’ law and the associated boundary

conditions at the respective interfaces. The fields may be written in terms of the

charge on the metal following,

Fox =−Qm

εox, (A.1)

FA =QπAlN −Qit +Qm

εA(A.2)

The charge control equation, Eq. (2.3), may be written including terms of the

local electric fields in each layer

−VGS +

(ΦB + ΦoA −

∆Ecq

)+ kTηw −

Qm

εoxtox −

(QπAlN −Qit +Qm)

εAtAlN = 0,

(A.3)

where all quantities were previously defined and ηw = (EF − EC)/kT is the local

field normalized to kT that accounts for the conduction band bending which forms

the quantum well. The Fermi level deep in the GaN layer is taken to be held at

a voltage, VS. Grouping like capacitive terms and re-writing Eq. A.3,

−VGS +

(ΦB + ΦoA −

∆Ecq

)+ kTηw −

(QπAlN −Qit +Qm)

CA− Qm

Cox= 0. (A.4)

1The electric fields, F, are vector quantities but are written here simply as magnitudes sincethe vertical direction is inferred by the structure under analysis.

154

Figure A.1. Band diagram including interface trap states.

In this form the derivative with respect to VGS may be applied which allows the

determination of the capacitive relationship of the structure.

A.1.1 Ideal capacitive structure (Qit = 0)

Considering first, the ideal case where no trapped charge exists (Qit = 0),

taking the derivative of Eq. A.4 with respect to VGS gives

−1 +∂

∂VGS(kTηw) +

(1

CA+

1

Cox

)∂Qm

∂VGS= 0. (A.5)

Since Qm reflects the total charge involved with the heterostructure, it is reason-

able that the total capacitance of the structure is the derivative of Qm with respect

155

to VGS. The question follows, what does the ∂η/∂V term reflect in the structure?

CT =∂Qm

∂VGS(A.6)

∂VGS

kTηwq

=1

q

∂(kTηw)

∂Qm

∂Qm

∂VGS. (A.7)

Following Eq. A.7 the derivative may be separated as shown above. It may be

further simplified by recalling the charge equality of Qm,

q

kT

∂Qm

∂ηw=

q

kT

∂ηw(Qπ

G + qns) =q

kT

∂(qns)

∂ηw= CQ, (A.8)

where CQ is referred to as the “semiconductor” or “quantum” capacitance [134]

and is the response of the charge in the semiconductor due to a change in local

potential (ηw). Therefore, the outcome of the unknown capacitance of Eq. A.7 is

found to be

∂(kTηw)

∂VGS=CTCQ

, (A.9)

and finally an equation for the entire heterostructure without trapped charge may

be gained from the application of Eq. A.9 after some re-arranging and is found to

be

1

CT=

1

Cox+

1

CAlN+

1

CQ(A.10)

where CT is the total structure capacitance, Cox is the fixed oxide capacitance,

CAlN is the fixed AlN capacitance, and CQ is the quantum capacitance as de-

fined previously. It is apparent from Eq. A.10 that without trapped charge the

heterostructure capacitance is simply the series combination of capacitances asso-

ciated with each layer in the structure, which is to be expected.

156

A.1.2 Capacitive structure including interface traps (Qit 6= 0)

Considering the case where an interfacial trapped charge distribution exists at

the insulator/AlN interface. It is further assumed that no charge (mobile or fixed)

exists in the insulator. By imposing the derivative with respect to VGS as before,

the following capacitive relation results

−1 +kT

q

∂ηw∂VGS

+

(1

Cox+

1

CA

)CT +

1

CA

∂Qit

∂VGS, (A.11)

where again CT = ∂Qm/∂VGS is the total structure capacitance. Considering the

derivative ∂ηw/∂VGS and expanding,

∂Qm

∂VGS

∂(kT/q)ηw∂Qm

= CT∂(kT/q)ηw

Qm

. (A.12)

Recalling Qm = QπG − qns −Qit further expansion gives

∂Qm

(kT/q)ηw =∂(kT/q)ηw

∂QπG︸ ︷︷ ︸

=0

+∂(kT/q)ηw

∂qns︸ ︷︷ ︸=1/CQ

+∂(kT/q)ηw

∂Qit

, (A.13)

where the first bracketed term falls out and the second is equal to 1/CQ. This

gives the desired relationship

∂(kT/q)ηw∂VGS

= CT

(1

CQ− ∂(kT/q)ηw

∂Qit

), (A.14)

157

Similarly, considering the derivative ∂Qit/∂VGS by expanding it,

∂Qit

∂Qm

∂Qm

∂VGS=∂Qit

∂Qm

CT

→ ∂Qm

∂Qit

=∂Qπ

G

∂Qit

+∂(qns)

∂Qit

− ∂Qit

∂Qit

=∂qns∂Qit

− 1

=∂(qns)

∂(kT/q)ηw

∂(kT/q)ηw∂Qit

− 1 (A.15)

where ∂(qns)/∂(kT/q)ηw = CQ. Therefore, the resulting capacitive relationship

results

∂Qit

∂VGS= CT

(CQ

(kT/q)ηw∂Qit

− 1

)−1. (A.16)

The unknown term links the local potential in the GaN well to the population of

trapped charge at the insulator/AlN interface. In other words, it is the 2DEG’s

response to the addition of the trapped charge density, Qit, at the insulator/AlN

interface. It may be understood by reconsidering the voltage loop between the

energy level of the trap state and the Fermi level within the well. As can be seen

in Fig. B.2 the voltage loop equation follows,

(Eit − EF,i)− qEAlN tAlN −∆Ecq

+ (EF,w − EC,w) = 0

→ ηit =EAtAkT

+∆EcqkT

− ηw + Eo (A.17)

where ηit = (Eit − EF )/kT is the local change in potential normalized to kT at

the insulator/AlN interface, EF,i is the Fermi level at the insulator/AlN interface,

Eit is the interface reference energy level, EF,w is the Fermi level in the quantum

well, and Ec,w is the conduction band energy in the GaN quantum well. By setting

Eit = EC,A then EF will be allowed to sweep through the top of the AlN conduction

band through any trap states at and below EC,A. Although in practical terms, the

158

range in which EF is swept through the barrier specifically will be only a fraction

of the full band gap. This is due to Fermi level pinning deep within the buffer

layer, limitations of EF,i sweep when ns → 0 for wide Eg semiconductors, as well

as electric field limitations the barrier layer(s) are able to sustain. Including the

trapped charge the field in the AlN, EAlN tAlN , is given by (QπA − Qπ

G − qns)/εA

and the derivative may be taken,

∂ηit∂ηw

= − 1

CA

∂qns∂kTηw

− 1 = −(

1 +CQCA

)(A.18)

which leads to the simplification of Eq. A.16

∂Qit

∂(kT/q)ηw= Cit

∂ηit∂ηw

= −Cit(

1 +CQCA

)→ ∂Qit

∂VGS= CT

(1

CQ( 1−Cit(1+CQ/CA)

)− 1

), (A.19)

where Cit is the capacitance associated with the (dis)charging of the interface

traps. The other term in Eq. A.11 that is under question may similarly be

considered,

∂(kT/q)ηw∂VGS

= CT

(1

CQ− ∂(kT/q)ηw

∂Qit

)= CT

1/CitCQ1/Cit + (1/CA + 1/CQ)

. (A.20)

159

Finally, Eq. A.11 describing the total capacitance may be re-written with the

above results included to yield

− 1 + CT1/CitCQ

1/Cit + (1/CA + 1/CQ)+

(1

Cox+

1

CA

)CT

− CTCA

Cit

Cit + (CQCA

CQ+CA)

= 0 (A.21)

Solving for the total capacitance of the system, CT , including trapped charge, and

after some algebraic manipulation CT may finally be written in a form that may

be easily interpreted,

1

CT=

(1

Cox+

1

Cit + ( 1CA

+ 1CQ

)−1

)(A.22)

This form shows the capacitance configuration of the system including trapped

charge at the insulator/AlN interface. It can be seen that CA and CQ are se-

ries connected and subsequently in parallel with the trap capacitance, Cit. This

capacitance block is finally in series with the oxide capacitance, Cox. This configu-

ration is reasonable in that if no traps were present (Cit → 0) the series connected

capacitance configuration of Eq. A.10 would result.

A.1.3 Quantum capacitance

The capacitances associated with the thin films (Cox and CA) are fixed since

their electronic boundaries are fixed for all VGS. However, the “quantum” or

“semiconductor” capacitance of the system is a strong function of VGS through

the normalized potential of the GaN QW, ηw. The functional form of the quantum

capacitance may be found by considering the quasi-2D charge distribution within

160

the GaN quantum well as given by Eq. 2.2 and taking the derivative, ∂ns/∂ηw as

shown below

∂ns∂ηw

=qNeff

kT

∂ηw[ln(1 + eηw)]→ CQ(ηw) =

qNeff

kT

(1

1 + e−ηw

)(A.23)

whereNeff = m∗kT/π~ is the effective conduction band density of states assuming

single subband population in the quasi-2D system. This is the functional form

with respect to ηw which is used in combination with the derivation above. If

desired, the quantum capacitance may be re-casted as a function of VGS through

the VGS − ηw relationship given by the CCE in Eq. A.4.

A.2 Terman’s method for Dit determination

The method Terman proposed was to compare an “ideal” capacitor (i.e. band

discontinuities known, no extraneous fixed charge, no traps, etc.) which is achieved

through modeling the structure, to the measured equivalent capacitor laden with

interfacial traps. The comparison is made through the difference in the measured

and ideal C-V curves. This can be seen by taking the derivative of Eq. A.4 with

respect the normalized energy, ηit,

∂VGS∂ηit

− kT

q

∂ηw∂ηit− 1

CA

∂Qit

∂ηit−(

1

CA+

1

Cox

)∂Qm

∂ηit= 0 (A.24)

where several of the quantities have already been found and are stated again for

161

convenience,

∂ηw∂ηit

=−1

1 + CQ/CA(A.25)

∂Qit

∂ηit=kT

qCit (A.26)

and the quantity, ∂Qm/∂ηit can be found by the derivative separation,

∂Qm

∂ηit=∂qns∂ηit

− ∂Qit

∂ηit= −∂Qit

∂ηit+∂qns∂ηw

∂ηw∂ηit

= −kTqCit +

kT

qCQ

(−1

1 + CQ/CA

). (A.27)

Putting these equalities together, Eq. A.24 can be re-written in terms of the

constituent capacitances and re-arranged to yield a relationship for the capacitance

associated with the interfacial trapped charge, Cit,

∂VGS∂ηit

=kT

q

(CitCox

+1

1/CQ + 1/CA(

1

Cox+

1

CA+

1

CQ)

)→ Cit =

q

kTCox

∂VGS∂ηit

− Cox1/CQ + 1/CA

(1

Cox+

1

CA+

1

CQ

)(A.28)

The capacitance difference can be written,

Cit − Cit|ideal =q

kTCox

(∂VGS,meas∂ηit

− ∂VGS,ideal∂ηit

), (A.29)

where it is assumed that ηit in both the measured and ideal systems are identical.

The relationship between capacitance and density of charge states (D = C/q2)

162

can be derived from the Cit difference and is written as

Dit =CoxqkT

∂∆VG∂ηit

∣∣∣Cmeas=Cideal

(A.30)

where ∆VG = VGS,meas − VGS,ideal taken at a common capacitance. Since ηit is

not a measurable quantity, it may be found through the relationship given by Eq.

A.4 which is a function of the measurable quantity, VGS. The form of Dit in Eq.

A.30 has a distinct likeness to that of the SiO2/Si system [135] with the exception

that the derivative is with respect to the normalized local potential, qηit, which is

spatially displaced from the mobile 2DEG.

A.3 Discussion of Terman methodology

Terman’s method to measure interface trap density (Dit) of the SiO2/Si sys-

tem has been adapted for a multilayer heterojunction capacitor as given above.

The keystone principle behind the method is to compare the measured C-V char-

acteristic of the real structure to the C-V character of an electronically ideal (no

unknown or unintentional charge) structure with the same layer properties. The

comparison is made through the derivative of the difference in voltage values that

correspond to the same capacitance, with respect to the local potential, η.

A few comments are worthy of mention regarding the intricacies of Terman’s

method. It is noted the energy scale that corresponds with Dit depends on barrier

height and band offsets and thus, accurate values of these material parameters

are necessary for the estimation of ηit with respect to VGS. Inaccuracies in these

quantities result in errors in the corresponding energy of the interface states. If

an inaccuracy in band alignment occurs or fixed charge exists in the oxide the

measured C-V will deviate by a simple shift in voltage. This will have no effect

163

on the calculation of Dit since the derivative of the difference will differentiate out

any simple voltage shift. However, if there are trap states distributed throughout

the oxide layer, Terman’s method accounts for this charge by “imaging” it to the

interface. This will lead to an over estimation of Dit. For a “pure” oxide without

imbedded trap states, this issue does not exist. It is assumed in the derivation that

the oxide/AlN interface is dominantly responsible for charge trapping given the

density of inhomogeneous atomic bonding of the crystalline surface with the oxide.

If there is mobile oxide charge present, this will also cause a skewing of the C-V

which will be interpreted as interfacial traps, further causing an overestimation

of Dit. It has also been argued at length that Terman’s method becomes too

inaccurate for Dit below 1010 cm−2eV −1 dictated by resolution limitations of the

test equipment [135]. While this may be true, the use of deposited, non-annealed

oxides in this work has led to Dit values far in excess of the claimed lower limitation

and therefore are believed to be true.

Considering the current theory that III-Nitride crystals have the property of a

dominant surface donor state [136] there exists the propensity of the Terman anal-

ysis to track the donor states as one of the interface states. A typical occurrence

when extracting Dit is an abrupt increase in the trap density near the band edges.

This character arises from two potentially indistinguishable mechanisms that are

both ultimately due to charge that is not accounted for in the ideal structure but

manifested in the C-V curve. The first being the sensing of the thermal popula-

tion of charge in the conduction band at the oxide/semiconductor interface that

occurs due to thermal spreading. Secondly, an increase in surface state density

occurs within the band gap due to the termination of the periodic crystal at the

surface [137]. Such states can have higher densities near the band edges and will

164

add to the apparent Dit.

Finally, Energy band non-parabolicity is not taken into account in the deriva-

tion given since it results in a loss in simplicity with only marginal improvements

in accuracy of Dit for some systems but warrants a comment on its effects on

sub-band population, C-V curvature, and Dit. Altschul et al. developed a simple

approximate energy band model for the deviation from parabolic band strucuture

[138]. In his work, the population of multiple 2DEG sub-bands was calculated

that accounted for conduction band non-parabolicity. This circumstance is ex-

acerbated for quantum confined 2DEGs where sub-band energies occur at levels

above the conduction band edge where non-parabolicity is pronounced. A greater

degree of non-parabolicity manifests in a softer C-V roll-off at the onset of deple-

tion. Thus, not accounting for such an effect may lead to slight over-estimation in

Dit ad energy levels closer to the conduction band in heterosystems that possess

very non-parabolic band structure.

A.4 Appendix A cited literature

133. L. M. Terman, “An investigation of surface states at a silicon/silicon oxideinterface employing metal-oxide-silicon diodes”, Sol. Stat. Elec., Vol. 5,285, 1962.

134. S. Luryi, “Quantum capacitance device”, Appl. Phys. Lett., 52, 6, 1988.

135. D.K. Schroder, “Semiconductor Material and Device Characterization”, sec-ond ed., Wiley Publishers, New York, 1998.

136. J.P. ibbetson, P.T. Fini, K.D. Ness, S.P. DenBaars, J.S. Speck, U.K. Mishra,“Polarization effects, surface states, and the sorce of electrons in AlGaN/GaNheterostructure field effect transistors”, Appl. Phys. Lett., 77, 2, 2000.

137. C. Kittel, “Introduction to Semiconductor Physics”, 8th Ed., Wiley publish-ers, New Jersey, 2005.

138. V.A. Altschul, A. Fraenkel, E. Finkman, J. Appl. Phys. 71, 9, 1992.

165

APPENDIX B

DERIVATION OF CURRENT-VOLTAGE CHARACTERISTICS OF THEAlN/GaN HEMT BY GRADUAL CHANNEL APPROXIMATION

B.1 Linear I-V functionality

Considering the 3 dimensional block of unintentionally doped GaN that com-

prises the channel immediately below the gate region in Fig. B.1, Ohm’s law may

be employed as the foundation for current flow functionality. Knowing electric

field is equal to the gradient of potential, Ohm’s law1 (J = σF ) can be integrated

along the channel

JLg = −qµ∫ VD

VS

ns dV, (B.1)

where J is the electron current flux (per unit gate width) passing through the plane

of width, Wg, Lg is the length of the channel immediately below the gate (or gate

length) and ns is the 2 dimensional current carrying electron charge density within

the channel. The mobility, µ, is assumed to be constant with depth (y-direction)

although in reality the scattering mechanisms involved in determining the effec-

tive mobility differ with depth and cause the mobility to be depth-dependent.

Nonetheless, the assumption of a constant mobility greatly reduces complexity in

the derivation with little loss of accuracy in the overall product. The electron flux

is a vector quantity as is the the differential, dV. For simplicity, these quantities

will from here on assume no special vector designation since the directions are

1Some clarity may be gained in the following derivation through the expansion of Ohm’s lawby the Drude model: J = (q2nτ/m∗)F .

166

Figure B.1. Illustration of the active region in a HEMT where the GCAapplies. External resistances are included to depict the quasi-circuit

under consideration.

understood. Furthermore, the current density is taken as the opposite direction

as the electron flux and therefore Eq. B.1 will written as a positive quantity.

The charge in the channel is a function of the voltage applied to the gate

metal above. The charge-potential relationship may be found through Poisson’s

equation, 52V = ρ/ε. The equation contains 3D dependency of potential on

charge but may be reduced to a 1D dependency by imposing two conditions,

1. charge is constant along the width of the gate (z-direction).

2. the change in electric field vertically (y-direction), which is responsible for

the confinement of the 2D channel charge, is much larger than along the

channel (∂F/∂y >> ∂F/∂x) and thus,

∂2V

∂x2+∂2V

∂y2∼ ∂2V

∂y2

167

The second condition is well known as the gradual channel approximation [139],[140].

Imposing these conditions, Poisson’s equation reduces to

∂2V

∂y2=ρ

ε. (B.2)

From this form of Poisson’s equation the energy profile of the conduction band

taken vertically below the gate can be traced to yield a relationship between the

electric potential and charge that is given in section 2.1 by

−VG + ΦB + Vox + ΦoA − VAlN −∆ECq

+Eoq

+(EF − Eo)

q+ VS = 0, (B.3)

which may be rewritten in terms of the fields present in the heterostructure

−VGS +

(ΦB + ΦoA −

∆ECq

)+ Foxtox − FAtA +

Eoq

+(EF − Eo)

q= 0, (B.4)

where Eo is the energy of the lowest occupied sub-band and (EF − Eo) is the

difference in energy between the Fermi level and the lowest sub-band. Note that

the conduction band is taken as the zero-energy reference in this derivation. These

quantities may be found by an approximate solution to Airy’s equation for the

triangular well [142],[143],

Eo '(

9π~q2ns8εGεo

√8m∗

)2/3

, (B.5a)

EF − Eo =π~2

m∗ns. (B.5b)

The integration of the RHS of Ohm’s law requires the 2D charge density, n, to

168

Figure B.2. Band diagram of the insulated gate AlN/GaN HEMTshowing pertinent parameters used for the GCA.

169

be a function of the changing potential along the channel (V(x) due to VG, VS,

and VD). Equation B.4 can again be rewritten in terms of the charge involved in

creating the fields (Gauss’s law)

− VGS + V (x) +

(ΦB + ΦoA −

∆Ecq

)+QπAlN

Cox−Qnet

(1

Cox+

1

CAlN

)+ qns

(1

CAlN+

1

Cox+

π~qm∗

)+ (qns)

2/3

(9π~q

8εGεo√

8m∗

)2/3

= 0, (B.6)

where V(x) is the channel voltage that varies along the gate length, QG is the GaN

spontaneous polarization charge, and Qnet is the net polarization charge between

the strained AlxGaN layer and GaN where x = 1 for AlN. These charge densities

are given below, respectively

QπAlN = q(Psp + Ppz)AlN , (B.7a)

Qnet = QπAlxGaN = ∆Psp + 2

(e31(x)− e33(x)

c13(x)

c33(x)

)×(a(x)− aGaN

aGaN

). (B.7b)

For convenience, equation B.6, can be written −V (x) + VGS = α + βns + γn2/3s

so that the V(x)-ns functionality can be seen. Referring back to equation (B.1),

the current integral must be solved by integrating the charge in the channel by

the the differential voltage slice, dV , from the voltage at the source to the voltage

at the drain. This operation may be performed by first writing dV in terms of a

differential charge slice, dn. Therefore, differentiating equation B.6 with respect

170

to ns allows the desired transformation of the element, dV ,

d

dns[−V (x) + VGS = α + βns + γn2/3

s ]

−→ dV (x) = −dns(β + 2/3γn−1/3s ), (B.8)

which allows the re-writing of equation B.1

I =qµWg

Lg

∫nsdns +

2

∫n2/3s dns

](B.9)

where β = q(1/Cox + 1/CA +π~2/m∗) and γ = (9π~q2/8εrεo√

8m∗)2/3. The upper

integration bound, which is given by the charge at the drain side of the gate set

up by VD, is given by a root of the third order polynomial (in ns) of equation B.6.

Performing this integration and re-inserting values for the place holders β and γ

the final form for the drain current can be written

IDS =qµWg

Lg

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)n2s +

2

5

(9π~q2

8εG√

8m∗

)2/3

n5/3s

]ns,D

ns,S

.

(B.10)

The voltage functionality is contained in the ns terms through Eq. B.6. This I(ns)

relationship is for the linear region of the HEMT characteristic where channel

pinch-off has yet to occur and the gradual channel approximation remains valid.

It is assumed that at a specific voltage, Vknee, the lateral field below the gate

is sufficient to cause scattering-limited velocity saturation. In this condition any

additional voltage beyond Vknee, current is limited from increasing. In Silicon

technology the knee voltage is called the pinch-off voltage since drain depletion is

significant enough to “pinch off” the conduction channel. This voltage is defined

as VGS for which the charge density below the gate, ns = 0. Through equation

171

B.6 the knee voltage may be found,

Vknee = Vpo =

(ΦB + ΦoA −

∆ECq

)−Qnet

(1

Cox+

1

CAlN

)+QπAlN

CAlN. (B.11)

It is assumed that there is no parallel conduction (from residual buffer charge or

poor 2D confinement) or self-heating effects that occur such that there is neither

positive or negative drain conductance within the saturation regime. This assump-

tion forces a non-changing drain current with any additional drain voltage thereby

making IDS within the saturation regime constant. This a poor assumption that

forces an ideality on the saturated drain characteristics but allows the simplifica-

tion of the I-V functionality for ease of understanding fundamental physics behind

the HEMT.

B.2 Transconductance

Transconductance is defined by gm = ∂IDS/∂VGS at a fixed value of VDS usually

taken in the saturated current regime. Considering the scenario above where IDS

is a function of ns the derivative may be expanded as follows;

gm =∂IDS∂qns

∂qns∂VGS

=CGq

∂IDS∂ns

, (B.12)

where CG = ∂qns/∂VGS and was given by equation A.10 and takes the form,

CG = (1/Cox + 1/CA + 1/CQ)−1. The derivative of the current is easily found;

∂IDS∂ns

=qµWG

LG

(βns +

2

3γn2/3

s

), (B.13)

172

thus, making the transconductance

gm =µWG

LGCG

(βns +

2

3γn2/3

s

). (B.14)

Recalling β and γ the full form of the transconductance follows,

gm =µWG

LGCG

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)ns +

2

3

(9π~q2

8εG√

8m∗

)2/3

n2/3s

]. (B.15)

The typical form the transconductance takes for HFETs and HEMTs as given in

Chapter 2 follows, gm = qvsatCG. Taking this in mind in combination with the

definition of current, J = qnsve, it can be seen that a charge-dependent velocity

term applicable in the linear regime of operation is present in the equation for

transconductance

ve =µWG

LG(βns +

2

3γn2/3

s ), (B.16)

which can be verified by considering the current density J = qnve where ∂J/∂qns =

ve which leads to

ve =∂J

∂qns=µWG

LG

(βns +

2

3γn2/3

s

), (B.17)

which shows agreement. Up until now it has been assumed that the velocity hits

a maximum (saturation velocity) without proof or reason. The following section

discusses the nature and implications of the occurrence of velocity saturation.

173

B.3 Current saturation

The scattering-limited velocity saturation approximation (SLVA) is given by

ve =µE

1 + µEvsat

, (B.18)

where vsat is the saturated electron drift velocity. The approximation seeks to

give a phenomenological explanation to the occurrence of electron velocity satura-

tion. Taking into account the SLV approximation in combination with the GCA

development given above, the current density may be written in compact form

J =qµns

∂V∂x

1 + µ∂V/∂xvsat

. (B.19)

By working out the integration for the current density by the same method as

shown above (Eq. B.10), the modified current density is found to be

JDS =qµ

Lg + µVDS

vsat

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)n2s +

2

5

(9π~q2

8εG√

8m∗

)2/3

n5/3s

].

(B.20)

The inclusion of the SLVA introduces a scaling term (Lg + µVDS/vsat) which is

taken as a VDS-dependent “effective gate length” since it always adds to Lg an

amount determined from the electric field in the channel. Therefore, a saturated

electron velocity serves to reduce the drain current density by a proportion of VDS

and vsat while maintaining equivalent ns functionality as in the ideal form of Eq.

B.10.

The transconductance may be found in the same was as the development of

Eq. B.15. Again, using the definition of transconductance (gm = CG

q× ∂IDS/∂ns)

174

the SLVA-modified transconductance is written

gm =qµCg

Lg + µVDS

vsat

[q

2

(1

Cox+

1

CAlN+

π~qm∗

)ns +

2

3

(9π~q2

8εG√

8m∗

)2/3

n2/3s

].

(B.21)

The SLVA reduces the transconductance in the same way as it does the current

density JDS, by the scaling of the vsat-dependent effective gate length.

B.4 External resistance

External resistance acts to reduce the intrinsic voltage drop in the channel, thus

causing a reduced drain current as well as shifting the I-V characteristics so that

they occur at higher applied voltages (due to the imposed voltage division). The

internal voltages in the resistor-HEMT circuit: V intDS = V ext

DS−IDS(RS,a+RD,a) must

be found numerically. Since IDS is found by applying V intDS , the I-V relationship

cannot be written in closed form. Therefore, the values for V intD and V int

S are found

through an iterative calculation using the equations:

V intD = V ext

D −RD,aIDS, (B.22a)

V intS = V ext

S −RS,aIDS. (B.22b)

In this work, the Jacobi method for numerical calculation is used for a faster

convergence of intrinsic voltage values. The equations above are modified to take

the form

V intD = (1− α)V ext

D − αRD,aIDS, (B.23a)

175

V intS = (1− α)V ext

S − αRS,aIDS. (B.23b)

The damping term, α, is used to prevent the iterative calculation from swinging

to the extrinsic rail values and also forces a faster convergence on the intrinsic

voltage values. Once the internal source and drain voltages are known for given

external resistances, the three-terminal transistor functionality is obtained (along

with Vg).

The common lumped-element approach to extract intrinsic transconductance

functionality is through considering the HEMT as a two-element resistance/conductance

circuit (1/gext = RS + 1/gint) where g = 1/R such that the intrinsic/extrinsic re-

lationship may be written

gm,int =gm,ext

1 +RSgm,ext, (B.24)

where RS = RC + Ra,S, RC is the contact resistance of the source contact, and

Ra,S is the source access resistance due to the biased gate-to-source channel re-

gion. Typically, for GaN-based HEMTs RC >> Ra,S such that RS ' RC and is

especially true for the AlN/GaN HEMT given its very low sheet resistance and

high contact resistance.

B.5 Validity of the GCA

The condition that defines the GCA is that the change in potential energy along

the channel is much less than taken vertically through the structure (∂V/∂y >>

∂V/∂x) such that Poisson’s equation can be accurately approximated to a single

dimension in y. Therefore, the GCA remains valid along the gate up to the drain

176

depletion region under certain bias conditions. Under certain conditions (pinch-off

in current saturation) the requirement that a constant current flows from source

to drain (under constant on-state bias) requires that the electric field in the drain

region under the gate increases (laterally and vertically) such that the lateral

change in field is comparable to the vertical chance forcing the GCA to no longer

remain valid. Grebene argued that an infinitesimal thin conducting “filament”

was left responsible to maintain current flow in the saturated portion of field effect

transistor I-V characteristics and used superposition in the x and y directions to

solve Laplace’s equation for the saturated characteristic (see appendix in Ref.

[140]). In this work it is assumed, as it is in Shockley’s work, that once pinch-off

conditions are reached, drift velocity saturation simultaneously occurs and pins

the drain current at a constant value.

B.6 Appendix B cited literature

139. W. Shockley, “A Unipolar “Field-Effect” Transistor”, Proc. of I.R.E., pp.1365-1376, 1952.

140. A.B. Grebene, S.K. Ghandhi, “General theory for pinched operation of thejunction-gate FET”, Sol. Sta. Elec., vol. 12, pp. 573-589, 1969.

141. S.M. Sze, in Physics of Semiconductor Devices, 2nd ed., Wiley Publishers,1981.

142. J.H. Davies, “The physics of low-dimensional semiconductors”, Cambridgepress, 1998.

143. D. Jena,“Polarization induced electron populations in III-V nitride semi-conductors; Transport, growth, and device applications” Ph.D. dissertation,University of California, Santa Barbara, 2003.

177

APPENDIX C

GENERAL AlN/GaN HEMT PROCESS

I. Ohmic level

• Ohmic lithography:

- Sample clean: ACE, METH, IPA, DI rinse and N2 blow dry

- Dehydration bake at 105 oC for 30 sec

- Spin LOR-5, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 180 oC for 5 min

- Spin 1805, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 105 oC for 60 sec

- Alignment and exposure (intensity = 7 mW/cm2 for 5 sec)

- Develop in 353 for 90 sec

- DI rinse and N2 dry

• Ohmic metal deposition and lift-off:

- O2 plasma descum for 1 min (Pressure = 900 mT, Power = 10 W, T= 25 oC)

- Ohmic pre-metallization ICP etch according to target depth (leave ¡2 nm AlN remaining), (Pressure = 5 mT, ICP/RIE Power = 100/25 W, T= 25 oC, Ar/BCl3/Cl2 flow rates = 10/10/20 sccm) NOTE: rate in OxfordICP is ∼ 0.53 A/sec

- NH4OH:H2O (1:10) dip for 15 sec, DI rinse

- E-beam deposit ohmic metal stack: Ti/Al/Ni/Au (300/2000/400/200A)

- Lift off metal by spraying with ACE followed by IPA, DI rinse, N2 blowdry. NOTE: verify metal thickness by profilometry.

Comment 1: Minimize time between pre-metallization etch, NH4OH dip,and insertion into evaporation chamber (no more than 15 min. maximum).

178

Comment 2: Measure as-deposited metal 2-point I-V on 4 µm separationCTLM for evolutionary data of contacts. Result will further dictate neces-sary anneal temperature.

• Ohmic contact anneal:

- Anneal sample at 800 oC for 30 sec in AET RTA using recipe: 80030.

II. Mesa level

• Mesa isolation lithography:

- Sample clean: ACE, METH, IPA, DI rinse and N2 blow dry

- Dehydration bake at 105 oC for 30 sec

- Spin 1811, ramp 5 sec at 500 RPM + 30 sec at 4000 RPM

- Bake at 105 oC for 60 sec

- Alignment and exposure (intensity = 7 mW/cm2 for 20 sec)

- Develop in 353 for 45 sec

- DI rinse and N2 dry

- Hard bake at 120 oC for 60 sec

Comment: Measure device characteristics: 2-point I-V on 4 µm separationCTLM, buffer isolation I-V, Hall effect, TLM, open channel I-V.

• Mesa isolation etch:

- Etch 500 Awith Trion ICP (Pressure = 5 mT, ICP/RF Power =100/25W, T = 25 oC, Ar/BCl3/Cl2 flow rates = 10/10/20 sccm)

- Remove PR in solvents: ACE, Meth, IPA, DI rinse and N2 blow dry

- Measure step height by profilometry

III. Gate level

• Oxide deposition:

- NH4OH:H2O (1:10) dip for 15 sec, DI rinse

- Blanket deposit oxide of choice (optimized deposition conditions varywith oxide variety)

• Gate lithography (Lg = 1 µm)

- Sample clean: ACE, METH, IPA, DI rinse and N2 blow dry

- Dehydration bake at 105 oC for 30 sec

179

- Spin LOR-5, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 180 oC for 5 min

- Spin 1805, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 105 oC for 60 sec

- Alignment and exposure (intensity = 7 mW/cm2 for 5 sec)

- Develop in 353 for 90 sec

- DI rinse and N2 dry

Comment: Check for over exposure/development of PR in microscope.

• Gate metal deposition and lift-off:

- O2 plasma descum for 1 min (Pressure = 900 mT, Power = 10 W, T= 25 oC)

- NH4OH:H2O (1:10) dip for 15 sec, DI rinse

- E-beam deposit ohmic metal stack: Ni/Au (200/3000 A)

- Lift off metal by spraying with ACE followed by IPA, 1165 (LORremoval), IPA, DI rinse, N2 blow dry. NOTE: verify metal thickness byprofilometry.

IV. Overlay level

• Overlay lithography:

- Sample clean: ACE, METH, IPA, DI rinse and N2 blow dry

- Dehydration bake at 105 oC for 30 sec

- Spin LOR-5, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 180 oC for 5 min

- Spin 1805, ramp 5 sec at 500 RPM + 30 sec at 3000 RPM

- Bake at 105 oC for 60 sec

- Alignment and exposure (intensity = 7 mW/cm2 for 5 sec)

- Develop in 353 for 90 sec

- DI rinse and N2 dry

• Overlay etch:

- O2 plasma descum for 1 min (Pressure = 900 mT, Power = 10 W, T= 25 oC)

180

- Over-etch oxide in Trion ICP using SiN recipe (Pressure = 50 mT, ICPPower = 40 W, T = 25 oC, Bias = 0 V, SF6 flow rate = 26 sccm)

- Remove PR in solvents: ACE, Meth, IPA, 1165 (LOR removal), IPA,DI rinse and N2 blow dry

Comment 1: Measure device characteristics: buffer isolation I-V, Hall effect,transfer and drain characteristics, off-state breakdown, gate lag, C-V.

Comment 2: Measure device characteristics for sub-micron gate length de-vices: transfer and drain characteristics, off-state breakdown, small signalS-parameters, gate lag, large signal (if GLR warrants).

181

This document was prepared & typeset with pdfLATEX, and formatted withnddiss2ε classfile (v3.0[2005/07/27]) provided by Sameer Vijay.

182