AN ABSTRACT OF THE THESIS OF Philip C. Canfield for the degree of Doctor of Philosophy in Electrical and Computer Engineering Title: presented on August 2, 1990. A P-Well GaAs MESFET Technology /7l1 Abstract approved:. e"--A Redacted for Privacy UdV1U U. AlISLUL The semiconductor gallium arsenide (GaAs) has many potential advantages over the more widely used semiconductor silicon (Si). These include higher low field mobility, semi-insulating substrates, a direct band-gap, and greater radiation hardness. All these advantages offer distinct opportunities for implementation of new circuit functions or extension of the operating conditions of similar circuits in silicon based technology. these advantages has not been realized. However, full exploitation of This study examines the limitations imposed on conventional GaAs metal-semiconductor field effect transistor (MESFET) technology by deviations of the semiinsulating substrate material from ideal behavior. The interaction of the active device with defects in the semi-insulating GaAs substrate is examined and the resulting deviations in MESFET performance from ideal behavior are analyzed. A p-well MESFET technology is successfully implemented which acts to shield the active device from defects in the substrate. Improvements in the operating characteristics include elimination of drain current transients with long time constants, elimination of the frequency dependence of gds at low frequencies, and the elimination of sidegating. These results demonstrate that control of the channel to substrate junction results in a dramatic improvement in the functionality of the GaAs MESFET. The p-well MESFET RF characteristics are examined for different p-well doping levels. Performance comparable with the conventional GaAs MESFET technology is demonstrated. Results indicate that optimization of the p-well MESFET doping levels will result in devices with uniform characteristics from DC to the highest operating frequency. A P-Well GaAs MESFET Technology by Philip C. Canfield A THESIS submitted to Oregon State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy Completed August 2, 1990 Commencement June 1991 APPROVED: A Redacted for Privacy Professor of Electrical and Computer Engineering in charge of major Redacted for Privacy Head of departmenj of Electrical and Computer Engineering Redacted for Privacy Dean of Gra( Date thesis is presented August 2,1990 Typed by the author for Philip C. Canfield ® Copyright by Philip C. Canfield August 2, 1990 All Rights Reserved 1 ACKNOWLEDGMENTS Throughout my graduate studies many individuals at Oregon State University and in industry have contributed to my work and personal growth. Contributors in industry were primarily associated with TriQuint Semiconductor Inc. These include Wes Mickanin, Stu Taylor, Eric Finchem, and Bruce Odikirk who provided many long, engaging, and Rich thought provoking discussions about my graduate research. Koyama has shown a special interest in this work and contributed with many insightful questions and discussions as well as providing the opportunity to work two summers at TriQuint. I especially appreciate the time and effort Rich committed to serve on my graduate committee, and his carefully and thoughtful review of this thesis. Two individuals were especially inspirational early in my graduate studies and I continue to look up to them both. Angus McCamant and Reed Gleason (now with Cascade Microtech) have been exemplary role models. They have always shown a keen interest in, and excitement for my work. They were always eager to offer assistance when I was experiencing technical difficulties and provided countless hours of their time to discuss my progress. In addition, this work would not have been possible without the generous contribution of TriQuint Semiconductor and Tektronix in providing the means by which the ideas underlying this work could be brought to fruition. Tektronix provided the mask set which was used to investigate ideas which were developed during the summer of 1987 and TriQuint Semiconductor provided the processing and process design tools needed to implement the p-well MESFET technology. In 11 particular the work of Eric Finchem is especially appreciated. Eric devoted a significant amount of time selecting the appropriate implants, guiding the wafers successfully through the fabrication process, and personally conducting some of the process steps such as the processing required to make the quarter micron gates on some of the MESFET test structures. Tektronix solid state research labs also provided use of their facilities for deposition and anneal of the pohmic metal used on the test wafers. I also wish to acknowledge the support of the National Science foundation. Various NSF grants have provided my stipend and travel to several conferences which provided an immeasurable boost to my professional development. Other individuals who bear mentioning are Bill Vetanen of Tektronix (formally of TriQuint) who was always very interested in my work and helped develop my philosophy of semiconductor processing by his outstanding example and commitment to excellence. Norm Schienberg and especially Bob Bayruns of Anadigics have always shown a keen interest in this work. They have provided moral support with their enthusiastic interest and discussions of my work and have been long time advocates of a p-well MESFET technology in GaAs. Noel Fernandez of Hewlett-Packard helped with some test systems needed for last minute measurements and Tom Andrade of Gazelle Microcircuits provided many animated discussions about GaAs technology. I am deeply indebted to my major professor, Dave Allstot. has been instrumental in my progress and the quality of my work. Dave His enthusiastic participation and keen interest has been the highlight of my years in graduate school. Dave's ability work together with me 111 on this project as a college facilitated the development of my many weaknesses into strengths. closest friends. In addition Dave has become one of my It has been a rare privilege and honor to work with a college at this level who shares such a common philosophical outlook on life. I also wish to acknowledge the service of Prof. John Wager on my graduate committee. Many of John's questions have served as motivation to examine aspects of my work more carefully than I had The special assistance of Prof. John Owen during a been. particularly difficult time in my graduate career is also gratefully acknowledged. I want to express my deep gratitude to my family. inspired and nurtured my love of science. My mother has She has contributed greatly to the fire and enthusiasm I bring to my work. My father has shown a great degree of understanding and patience while teaching me He has provided me ample opportunities to provide for many things. myself and thus develop a high degree of self-esteem and confidence in my life. His unconditional love and support has been a constant comfort to me. My mother-in-law, Qienmei Liu, deserves special recognition for the care she has provided for my son the last two years. Her commitment has made the completion of this thesis immeasurably easier. My wife, Lingzhou, whom I met in a Quantum Mechanics class the first year of graduate school, has been a source of special comfort. Her undying support, devotion, dedication, and love have made my graduate experience rich and fulfilling. She has contributed many hours away from own graduate studies in Physics to iv assist in the preparation of some of the more tedious aspects of this document and many of the conference publications which resulted from this work. Her triumph and success in life against incredible odds and seemingly insurmountable obstacles has made any hardship I have faced seem trivial in comparison. V TABLE OF CONTENTS Page 1. INTRODUCTION 1 2. EXPERIMENTAL 10 3. 4. 2.1 Conventional Technology 10 2.2 P-Well Technology 11 2.2.1 Doping Profiles of the P-Well MESFETs 14 2.2.2 P-Well Test Mask Set 14 2.3 Drain Current Transient Measurements 20 2.4 Frequency Dependent Output Conductance Measurements 22 2.5 Sidegating Measurements 23 2.6 RF Characterization 26 DRAIN CURRENT TRANSIENTS 27 3.1 Introduction 27 3.2 Measurement of Conventional MESFETs 34 3.3 Analysis 39 3.4 Comparison of Analysis and Measurements 46 3.5 Measurement of Buried-Channel and P-Well MESFETs 50 3.6 Summary 56 FREQUENCY DEPENDENT OUTPUT CONDUCTANCE 57 4.1 Introduction 57 4.2 Measurement of Conventional MESFETs 60 vi 5. 6. 7. 8. 4.3 Analysis 62 4.4 Measurement of Buried-Channel and P-well MESFETs 64 4.5 Summary 67 SIDEGATING 68 5.1 Sidegating in Conventional Technology 68 5.2 Measurements of Sidegate Structures with Floating P-Type Implants 73 5.3 Measurements of Sidegate Structures with Contact to the P-Type Implants 77 5.4 Summary 82 RF CHARACTERISTICS 83 6.1 Equivalent Circuit Model 83 6.2 S-Parameters 85 6.3 Small-Signal Parameters 89 6.4 fT, f=x, and MAG 94 6.5 Summary 97 LIMITATIONS AND TRADEOFFS 99 99 7.1 Introduction 7.2 Drain Characteristics 101 7.3 Gate Characteristics 106 7.4 Summary 106 CONCLUSIONS 108 BIBLIOGRAPHY 111 vii APPENDIX A. Publications 119 A.1 Journal Publications 119 A.2 Conference Publications 121 viii LIST OF FIGURES Page Figure 6 1.1 Cross-sectional view of a conventional MESFET and the "self -backgating" effect. 2.1 Cross-sectional view of a conventional ion-implanted GaAs MESFET. 11 2.2 Cross-sectional view of the n-channel p-well GaAs MESFET. 12 2.3 Doping profiles for the different wafers fabricated in this study. 13 2.4 Layout of the mask set developed for the study of the p-well MESFET technology. 15 2.5 An example of the different p-well implant masks used to study the effect of the gate contacting the p-well. 16 2.6 Layout of the sidegating test cell showing the three different sidegate electrodes available. 18 2.7 Schematic diagram of the measurement system used to observe the drain current transients. 19 2.8 Illustration of the probe card used for the drain current transient measurements of GaAs MESFETs. 21 2.9 Schematic diagram of the measurement system for characterization of the frequency 22 dependent gds. 2.10 Cross-sectional view of the sidegate test structure implemented in the conventional technology. 24 2.11 Cross-sectional view of the sidegate structure implemented in the p-well technology. 24 2.12 Schematic representation of the sidegating measurement system. 25 ix 2.13 Schematic representation of the measurement configuration used for the RF characterization of the p-well technology. 26 3.1 Cross-sectional view of the various depletion regions which can contribute to drain current transients. 29 3.2 Input drain voltage waveform and corresponding drain current transient responses due to the three different regions of Fig. 3.1. 32 3.3 Measured drain current transient response of the conventional GaAs MESFET. A voltage square wave between 5.0 V and 1.0 V was applied to the 33 drain. 3.4 VGS = 0.0 V. Measured drain current transient response of the A voltage square wave conventional GaAs MESFET. between 5.0 V and 1.0 V was applied to the drain. 34 VGs = 0.0 V. 3.5 Measured drain current transient overshoot for an abrupt change in the drain voltage from 0.0 V to 5.0 V for a conventional GaAs MESFET. Each curve represents a different time scale ranging from 10 VGS = 0.0 V. mSec/Div to 10 nSec/Div. 35 3.6 Measured drain current transient response of the conventional GaAs MESFET. A voltage square wave between 5.0 V and 1.0 V was applied to the drain for three different pulse periods. 36 VGS = 0.0 V. 3.7 Measured drain current transient undershoot for an abrupt change in the drain voltage from 5.0 V to Each curve 1.5 V for a conventional GaAs MESFET. represents a different time scale ranging from 10 VGS = 0.0 V. mSec/Div to 10 gSec/Div. 37 3.8 Magnitude of the drain current transient overshoot for an abrupt increase in Vds from 0.0 V to 38 ads VGS 0.0 V. 3.9 Charge distribution along a path from the gate into the substrate of a conventional MESFET. The shaded regions signify the depletion regions which change due to the balancing of charge. 41 3.10 Cross-sectional view of a conventional MESFET showing the path used in Fig. 3.9. 42 x 3.11 Analytical results of drain current transient undershoot for the conventional MESFET results of Fig. 3.7. 46 3.12 Time dependence of the log of the magnitude of current overshoot normalized to t=0 for the measured current overshoot waveforms of Fig. 3.5. 47 3.13 Time dependence of the capture time constant for the transient current overshoot of Fig. 3.5. 49 3.14 Measured drain current transient overshoot response for a 0.0 V to 5.0 V drain voltage step on buriedchannel MESFETs for 1 mSec/Div, 10 ASec/div, and 100 nSec/Div scales. The peak p-type doping behind the channel is a) 2.0 x 10'6 cm-5, b) 4.0 x 1016 cm-3, 51 c) 8.0 x 1016 cm -5, and d) 16.0 x 10 3.15 cm-3. Measured drain current transient undershoot response for a 5.0 V to 1.5 V drain voltage step on buried channel MESFETs for 1 mSec/Div and 10 ASec/div The peak p-type doping behind the scales. 52 channel is a) 2.0 x 10'6 cm- 5, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 3.16 Measured drain current transient overshoot response for a 0.0 V to 5.0 V drain voltage step on p-well MESFETs for 1 mSec/Div, 10 ASec/div, and The peak p-type doping behind 100 nSec/Div scales. the channel is a) 2.0 x 10'6 cm-5, b) 4.0 x 1016 cm-3, 54 c) 8.0 x 1016 cm -5, and d) 16.0 x 1016 cm-3. 3.17 Measured drain current transient undershoot response for a 5.0 V to 1.5 V drain voltage step on p-well MESFETs for 1 mSec/Div and 10 ASec/div scales. The peak p-type doping behind the 55 channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 101 cm-3, and d) 16.0 x 1016 cm-3. 4.1 High frequency and low frequency small-signal current for a given applied small-signal vds 59 4.2 Measured frequency dependence of the output conductance for three different gate voltages. 60 VDS = 2.0 V. 4.3 Measured frequency dependence of the output conductance for different kips. 4.4 61 VGs = 0.0 V. Comparison of the frequency dependence of the output conductance model and measured data for three different temperatures. 64 xi 4.5 Measured frequency dependence of gds for the p-well MESFET at three different VGS The peak p-type doping behind the VDS = 2.0 V. channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, 65 c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 4.6 Measured frequency dependence of gds for the buried channel MESFET at three different VGS VDS = 2.0 V. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 66 5.1 Measured sidegating results for the conventional MESFET structure of Fig. 2.10 at room temperature without illumination. The drain current is shown in the upper trace and sidegate current is shown in the lower trace. Vps = 2.5 V and Ids(VsG =0.0 V) = 3.0 mA. 69 5.2 Measured drain current sidegating results with and without illumination for the conventional MESFET of Fig. 5.1. 70 5.3 Measured threshold voltage shift of a conventional MESFET for VsG = - 5.0 V. VGS = 0.0 V and Vps = 2.5 V. 71 5.4 Measured sidegating results of a conventional MESFET and an n* sidegate electrode surrounded by a p-type implant for Vps = 2.5 V and The equivalent peak p-type Ids (VsG=0.0 V) = 3.0 mA. doping levels were a) 2.0 x 1016 cm-3, 72 b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 5.5 Measured sidegating results of a buried channel MESFET and an n' sidegate electrode surrounded by a p-type implant for Vps = 2.5 V and The equivalent peak Ids( VsG=0.0 V) = 3.0 mA. p-type doping levels were a) 2.0 x 1016 cm-3, 73 b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cor3. 5.6 Measured sidegating results of a buried channel MESFET and an n+ sidegate electrode for Vps = 2.5 V The equivalent peak and Ids(VsG =0.0 V) = 3.0 mA. p-type doping levels were a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 74 xii 5.7 Measured sidegating results of a conventional MESFET and an n+ sidegate electrode in a p-well for Vin = 2.5 V and Ids(VsG=0.0 V) = 3.0 mA. The equivalent peak pp -type doping levels were 75 a) 2.0 x 10'6 cm- , b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 5.8 Measured forward biased p-i-n diode effects for a p-well MESFET and an e sidegate electrode. 76 5.9 Measured sidegating results of a buried channel MESFET and an n+ sidegate electrode in a p-well for Vin = 2.5 V and Ids(VsG=0.0 V) = 3.0 mA. The equivalent peak pp -type doping levels were 77 a) 2.0 x 106 cm- 3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 5.10 Measured sidegating results of a p-well MESFET and an n+ sidegate electrode in a p-well for Vps = 2.5 V and Ids(VsG=0.0 V) = 3.0 mA. equivalent peak pp -type doping levels were a) 2.0 x 10'6 cm- , b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 78 The 5.11 Measured sidegating results for various sidegate structures with and without illumination. a) Conventional MESFET and an n+ sidegate electrode surrounded by a p-type implant, b) buried channel MESFET and an n* sidegate electrode surrounded by a p-type implant, c) conventional MESFET and an n+ sidegate electrode in a p-type well, d) buried channel MESFET and an n+ sidegate electrode in a p-type well. 79 5.12 Measured sidegating result for the p-well MESFET technology with and without illumination. 80 6.1 Cross-sectional view of a p-well MESFET conceptually illustrating the depletion regions associated with the pn junctions and the conductive path to the source of the p-type region under the channel. 84 6.2 Small-signal equivalent-circuit model used for the p-well GaAs MESFET. 85 6.3 S21 responses for the p-well MESFETs with different n-channel and p-well doping levels. The sweeps are for bias voltages of Vos = 3.0 V and Vo = 0.0 V. 86 6.4 S22 responses for the p-well MESFET's with different p-well doping levels. The sweeps are for bias voltages of VDS = 3.0 V and 87 VGS = 0.0 V. 6.5 CGS and Cgd as a function of p-well doping levels for bias voltages of VDS = 3.0 V and 88 VGS = 0.0 V. 6.6 Small-signal gm values versus p-well doping levels at several different VGS bias voltages. VDS = 3.0 V. 89 6.7 Small-signal Rds values versus p-well doping levels at several different VGS bias voltages. 90 VDS = 3.0 V. 6.8 Small-signal C, values versus p-well doping levels at several different VDS bias voltages. VGS = 6.9 91 0.0 V. Small-signal Rp values versus p-well doping levels for bias voltages of VDS = 3.0 V and 92 VGS = 0.0 V. 6.10 1H211 versus frequency for a conventional MESFET with bias voltages of Vps = 3.0 V and VGS = 0.0 V. 94 6.11 Small-signal fT values versus p-well doping levels with bias voltages VDS = 3.0 V and 95 VGS = 6.12 0.0 V. Maximum Available Gain (MAG) values versus p-well doping levels for bias voltages of 96 VDS = 3.0 V and VGS = 0.0 V. 6.13 Small-signal fmax values versus p-well doping levels for bias voltages of VDS = 3.0 V and VGS = 0.0 V. 7.1 Drain I-V characteristics for the p-well MESFET with the p-well tied to the source a) and the p-well tied to the drain b). The maximum VGS = 0.75 V and the step size 97 100 is 0.25 V. 7.2 Subthreshold characteristics for MESFET with the p-well extending The the channel under the gate. is VDS = 5.0 V and the step size the p-well outside upper curve is 2.0 V. 101 xiv 7.3 Subthreshold characteristics for the p-well MESFET with the p-well a) coincident with, b) pulled in 1/2 micron from, and c) pulled in one micron from the edge of the channel The upper curve is under the gate. VDS = 5.0 V and the step size is 2.0 V. 103 7.4 Gate I-V characteristics for the MESFET with the p-well a) pulled micron from, b) coincident with, c) pulled in one micron from the VDS the channel under the gate. 104 7.5 p-well out one and edge of = 2.5 V. Transconductance for the p-well MESFET with the p-well a) pulled out one micron from, b) coincident with, and c) pulled in one micron from the edge of the channel under the gate. VDS = 2.5 V. 105 XV LIST OF TABLES Page Table 2.1 Implant parameters for the experimental wafers. 13 2.2 MESFET structure identification for block A of Fig. 2.4. 17 2.3 Sidegate cell identification for the cells in block B of Fig. 2.4. 18 6.1 Equivalent-circuit small-signal element values for the different p-well doping The values are for MESFETs with levels. lAm gate lengths and 300 Am gate widths and bias voltages of Vos = 3.0 V and VGS = 0.0 V. 93 A P-WELL GaAs MESFET TECHNOLOGY 1. INTRODUCTION Interest in the use of gallium arsenide (GaAs) metal semiconductor field effect transistors (MESFETs) for integrated circuits (ICs) is based on several potential advantages this technology offers over silicon technologies. Among these are higher low-field electron mobility and peak velocity (high speed), greater radiation hardness (military and space applications), a direct bandgap (photonic applications), and the availability of semi-insulating substrates (good isolation and low interconnect and junction capacitances). Thus far, GaAs technology has been used successfully in the implementation of monolithic microwave integrated circuit (MMIC) applications, is showing increased strength in the area of high-speed digital applications, and only marginal success in the area of analog IC's and mixed-mode digital-analog circuits which require a high degree of precision (8 or more bits). While the success in the MMIC area has served to maintain a high level of activity, it has not been sufficient to provide adequate production volumes for most companies striving to survive commercially. The growth in demand for high speed-digital circuits of medium and large scale integration (MSI and LSI) has resulted in higher volumes and greater opportunities for many companies to become profitable. However, the inability to implement such ICs as 2 operational amplifiers which settle to more than eight bits of accuracy in less than a second [1], [2] has completely eliminated a large area of commercial interest for instrument vendors. Factors limiting the implementation of high-speed precision analog/digital circuits in GaAs technology are invariably related to These anomalies are a large assortment of GaAs device anomalies. principally due to deviations of the semi-insulating substrates from ideal behavior. Included in these anomalies are frequency-dependent small-signal parameters [1], [3]-[7], pattern and frequency-dependent propagation delays [8], drain current transients with long time constants [9]-[14], large absolute and matching tolerances between devices [15], [16], and sidegating or cross talk between devices [12], [16]-[20]. To reduce sidegating, large layout spacings have been used [21], particularly for MESFETs biased at high power levels as in many microwave applications. Another aid in reducing sidegating is the use of an isolation implant [22] or co-implantation of p-type dopants into the source-drain regions [23]. A p* blocking electrode [24], [25], and a Schottky shield [26] tied to the most negative potential has been proposed as well for the reduction in sidegating. For analog and digital circuits, frequency-dependent small-signal parameters result in frequency-dependent gain [1], [27] and modeling difficulties [1]-[3], [5], [27]. Drain current transients cause very long settling times in switches and amplifiers [1], [28] and hysteresis in differential amplifiers and comparators [2]. Pattern and frequency-dependent propagation delays result in timing nightmares in digital circuits [8]. An understanding of the 3 properties of the semi-insulating substrate is key to both understanding the source of these anomalies and to devising satisfactory solutions to these problems. GaAs semi-insulating substrates used in industry are almost exclusively grown by the liquid encapsulated Czochralski (LEC) technique. Residual impurities in the melt which are incorporated into the crystal as it is grown result in lightly doped p-type GaAs unless steps are taken to compensate the net concentration of acceptors [29], [30]. The most common residual shallow acceptor is carbon [31], [32]; however, trace amounts of other elemental donors and acceptors are also present. Some of the more common residual shallow impurities include the shallow donors Si, S, Se, and Te, and the shallow acceptors Mg, Mn, and Fe [33]. The net effective concentration of acceptors Mod is on the order of 1015 cm-3, and is determined by solving the charge balance equation in equilibrium; (1.1) Naeff=IiNai-IjNdi. The sum over i represents all the residual shallow acceptors and the sum over j accounts for all the residual shallow donors. In addition boron and nitrogen incorporated into the melt from the boron nitride crucible are present. However, their contribution to the compensation of the GaAs substrates appears to be of little consequence in currently available LEC semi-insulting material. The most common method of compensating the residual shallow acceptors, which also results in high resistivity material, is to grow the crystals slightly arsenic-rich [30]. Crystals grown under these conditions contain a large concentration of native defects 4 which are electrically active. The most abundant is the deep donor EL2 which is associated with an arsenic atom sitting on a gallium site in the crystal lattice [34]. When the concentration of EL2 is much greater than the concentration of residual acceptors, the Fermi-level is pinned near the EL2 energy level [35]. Since the energy level of EL2 in GaAs is near the middle of the band gap, the resistivity of the resulting material is high (p > 107 Q-cm). Typical concentrations of EL2 are greater than 1016 cm-3 in GaAs substrates [32], while the typical net concentration of shallow acceptors is on the order of 1015 cm-3 [32], [36]. Charge balance requires that the concentration of ionized deep donors equal Naeff [37]. This means there are about 1015 cm-3 ionized EL2. Hence, the majority of EL2 are filled and electrically neutral which implies that in equilibrium the Fermi-level is slightly above the EL2 level. Furthermore, since a relatively large percentage of the EL2 must change their charge state (- 1015 cm-3) for any significant movement of the Fermi-level, the Fermi-level is "pinned" near mid-gap. From a device standpoint, there are two factors of importance to consider here. The first is that the concentration of deep levels is large with a significant fraction of them ionized. Consequently, the Fermi level is near the energy level of the trap and any small movement of the quasi-Fermi levels can result in a significant change in the number of ionized EL2. The need to balance this charge, coupled with the long emission time constants associated with mid-gap traps, result in a transient time-dependent behavior which can take seconds to relax to steady state. When this relaxation process 5 occurs in the vicinity of an active device such as a MESFET, the time-dependent behavior will be reflected in the operating characteristics of the MESFET. The second consideration is related to the transition region from the active channel (formed by ion-implantation) to the semiinsulating substrate. This transition region is not abrupt. It typically extends about a micron into the substrate while the active channel is on the order of 0.2 Am thick [38]. Consequently the channel-substrate potential barrier is effectively graded and represents a region of varying conductivity. This facilitates the ability of the electric fields generated by the drain voltage of a MESFET, to wrap around the active channel and sweep charge down into the substrate. The resulting increase in the quasi-Fermi level can This charge change the charge state of deep levels in steady state. flowing through the substrate under the influence of the drain electric field is referred to as the sub-threshold current. The presence of residual acceptors actually acts to decrease the magnitude of the sub-threshold current by sharpening the electron tail and sharpening the barrier [38]. The modification of a MESFET's operating characteristics by the filling or emptying of traps in the vicinity of the MESFET can be discussed qualitatively with the aid of Fig. 1.1. With no electric field applied the to channel of the MESFET, electrons are continuously escaping over the potential barrier to the substrate. These electrons are continually trapped in ionized EL2 sites while others are emitted from neutral EL2 sites. In steady-state the 6 Drain Source N Channel Ionized Neutral EL2 EL2 Trap Trap Depletion Region LEC SI GaAs Substrate Figure 1.1. Cross-sectional view of a conventional MESFET and the "self-backgating" effect. concentrations of both trapped and free electrons are adjusted so that the average number of energetic electrons escaping over the potential barrier is matched by the average number of electrons emitted back into the channel. That is, the capture and emission rates are equal along the channel-substrate interface. In steady- state with a high electric field applied to the channel, both the rate of injection of electrons over the barrier, and the rate of emission of electrons back into the channel are increased, but must still be equal. 7 In a non-steady-state condition, as for example when the electric field between the drain and source is suddenly increased, more of the electrons in the channel gain enough kinetic energy to be scattered over the potential barriers. Because of the relatively short capture time constant, these electrons are quickly trapped in ionized EL2 sites. With the capture of these electrons the traps become electrically neutral, making the substrate adjacent to the channel more negatively charged due to the uncompensated residual shallow acceptors. Charge balance is achieved by nearby implanted shallow donors in the channel, with the accompanying increase of the depletion width in the channel. As a result of this pinching off of the channel, the drain current is decreased. referred to as "self-backgating" [37]. This effect is commonly Ultimately the problem lies with the unconstrained potential of the semi-insulating substrate which translates into an uncontrolled channel-substrate junction. In this work a systematic investigation of the interaction between the changing charge state of the deep-level defects and operation of the active device is conducted. Evidence is presented which supports the contention that an uncontrolled channel-substrate interface is responsible for the anomalies examined here. The use of p-type regions has been reported to help eliminate some of the MESFET anomalies. A systematic examination and comparison of the effect of various p-type layers is conducted and their relative impact on anomalous behavior is reported. It is demonstrated that the only technique which reliably eliminates these anomalous characteristics 8 is control of the channel-substrate junction by the introduction of a p-well which can be connected to a known potential. Chapter 2 details the experimental procedure and design used to study the characteristics of the anomalous behavior in MESFETs. The mask set and the various doping profiles used to fabricate the devices are described. Chapter 3 examines the nature of the drain current transients. In addition to the traditional examination of the transient response to a voltage step on the drain from a high to a low voltage, the response to a voltage step from a low to a high voltage is examined. The two responses are quite different and their behavior is shown to be consistent with modulation of the channel-substrate junction space charge regions due to filling and emptying of traps in the substrate. The transient behavior is analyzed using a simple square law relationship for the drain current and shown to agree with experimental data. The effect of p-layers and the p-well are examined and and the relative impact on the transient response of the drain current is discussed. Chapter 4 extends the results of the time-dependent behavior of the drain current to the frequency-dependent behavior of the output conductance. The expression derived for the drain current transients in chapter 3 is used to derive an expression for the frequencydependent output conductance expression. This expression is shown to be in close agreement with experimental results. The effect of p- layers and the p-well are examined and and the relative impact on the output conductance is discussed. 9 Chapter 5 examines sidegating in GaAs MESFETs. A review of the effect in conventional devices is presented and the effect of players is examined in detail. No report of a new technology in GaAs would be complete without an examination of reporting on its RF performance. Chapter 6 reports on the RF performance of the p-well MESFET for different p-well implant doses. An equivalent circuit model is implemented using TOUCHSTONE® (a commercially available software package from EESOF for RF modeling and measurements) and the FET parameters extracted. It is shown that the n-channel p-well MESFET implant parameters can be optimized to give RF performance which is comparable to MESFETs fabricated by a conventional technology, while maintaining the improved immunity to the low-frequency anomalies. Chapter 7 examines some of the limitations and trade-offs which arise in the device behavior due to the p-well structure and the ptype implants. Some of the issues addressed are pertinent to layout while others address design concerns. These results are summarized in Chapter 8 and an attempt is made to place these findings in perspective. future work are made. Several recommendations for 10 2. EXPERIMENTAL The experimental procedure and design used to evaluate the pwell MESFET technology are described in this chapter. Both the conventional and p-well MESFET technologies are presented. The implant matrix and the test structures on the mask set used for the different wafers in this study are shown in detail. Each of the measurement facilities and techniques used for the investigation of the drain current transient response, the small-signal equivalentcircuit output conductance (gds), sidegating, and the RF characteristics are described. 2.1 Conventional Technology A cross-sectional view of a conventional MESFET is shown in Fig. 2.1. Processing of the conventional MESFETs was provided by TriQuint Semiconductor, Inc. and reflects much of their common processing methodologies. The n-type channel region and the n' source and drain contacts are formed by ion-implantation of Si29" into the semiinsulating GaAs substrate. The implanted donors are activated with an 800 °C furnace anneal for 30 minutes. A gate recess is etched to set the desired saturation current (Id) and pinch-off voltage (Vp). The one micron gate is non-self-aligned within the three micron channel length. The drain and source contacts are made using Au/Ge/Ni ohmic metal, while Ti/Pd/Au metalization was used for the gates. 11 SEMIINSULATING GaAs SUBSTRATE L Figure 2.1 GaAs MESFET. 2.2 1 Cross-sectional view of a conventional ion-implanted P-Well Technology A cross-sectional view of the p-well MESFET is shown in Fig. 2.2. Processing of the p-well MESFETs was also provided by TriQuint Semiconductor, and thus the process technology is essentially the same as the conventional MESFET with additional p-type implants (Be') and a p-type ohmic contact. The additional p-type implants are composed of a bottom p-layer forming the p-well which completely surrounds the active channel region and the n' source and drain contact regions , a surface p-type implant, and a p' well contact implant provided by Tektronix solid-state research labs. The gate is non-self-aligned with a gate recess to remove the surface p-type layer under the gate and to adjust Vp and Idss The well is contacted 12 Well Drain Source PSurface N Burled Implant Channel LEC SI GaAs Substrate Figure 2.2 MESFET. Cross-sectional view of the n-channel p-well GaAs using a Au/Mn p-ohmic metal on the p' implant. In normal operation, the p-well (back-gate) terminal is connected to the source to guarantee that no pn junction ever becomes forward biased. By virtue of the well contact, the n-channel is isolated from the substrate by a reverse-biased pn junction, and is thus immune to perturbations in the charge distribution in the substrate and between devices due to electron capture and emission by defects. For purposes of comparison MESFETs were also fabricated with out a p' contact to the p-type well. These structures are also examined and will be referred to as buried-channel MESFETs. 13 /- tr) I E 10 1 8 0 ....., 1017 1016 1--- w z 1 015 11111111 LILLL 0.0 0.2 WILWILLLL1-1 0.4 Implant parameters for the experimental wafers. Channel Si'29 Surface Beg' Backside Be9+ Dose Energy Dose Energy Dose Energy 1016 1013 (KeV) 1012 (KeV) 1012 (KeV) CM CM 3- 1.0 Doping profiles for the different wafers fabricated in Figure 2.3 this study. Peak P-Type Conc. 0.8 0.6 (microns) DEPTH Table 2.1 LU_L LLI --2CM CRIB 0.0 1.20 135 2.0 1.05 165 5.65 20 1.30 125 4.0 1.10 175 5.45 20 2.25 140 8.0 1.20 185 5.45 20 4.25 150 16.0 1.40 200 5.45 20 8.50 155 14 Doping Profiles of the P-Well MESFETs 2.2.1 The dependence of the MESFET characteristics on the p-well doping level was evaluated from wafers fabricated with peak p-type dopings of 2.0, 4.0, 8.0, and 16.0 x 1016 cm -3 behind the channel. Conventional control wafers with no p-type implants were also The theoretical doping profiles of the fabricated simultaneously. p-well MESFETs are illustrated in Fig. 2.3. These profiles were generated using TriQuint Semiconductor proprietary software which is used for process development and has been optimized for TriQuint's processing capabilities. To maintain constant Vp and Idss values, the surface p- and channel n-type implants were both adjusted so that the effective doping concentration in the channel regions are approximately equal for all samples. The implant doses and energies for the different wafers are given in Table 2.1. These doping profiles resulted in depletion-mode MESFETs with Vp = -1.5 V and Idss 25 2.2.2 200 mA/mm for all wafers. P-Well Test Mask Set A special mask set provided by Tektronix was developed to evaluate the dependence of the MESFET parameters on the presence of the various p-layers and the effect of tying the p-well to a known potential. The mask set is shown in Fig. 2.4 and consists of a large array of n-channel transistors, a few complementary transistors including vertical and lateral pnp bipolar transistors and p-channel MESFETs, special sidegating test cells, analog circuit building blocks, and analog circuits. Only the devices used in this study will be described in detail. For general DC characterization, drain 15 ... %\ IV Ve ki 111 il %x. Cc, I I! 8a is.ill''''. mo 110 id 4\ 00 0.0 6e1 in 1.10 5! ID fi 6 (4) 0 )1 0 a 0 59 0 0 44 iii 0 It 59 10 .6! 55! C4T5 1.11 61 65 is V; IX 59 0 'V XI Xi 6 0 0 X4 16 Ca . ve I cicc.'(I../Co o Y4 4! 8 'Ow .6 eti 0 5Y. 51 V 0 0 0 Vi iX Vilii0dage 0 0 60 X4 fu 74 m. Co o 55: 0 f2 X4 8 6 Co kt 41 0 ?X Xi 930 Ain, XibX Kt 0 .9 is 65 16 fti IV 91 Ci )4 0 ft 6 e6 6Y '5 II 00! 4 V X4 6 VV 6 !,9' 65 0 6365 fvf I 1 e;e: 1+.4 ti 7.5 W s cc' di 0 Co ) W. 01 et Ki et ! v o I!. Ii 11X 000 Ft,ra A 59 55! : 55! 0.4 iii F12-1 rkik pa. 4 it WIWA t re tix X4 .0 tf 0 .0 A Si 0 9; WV t teD33 MEM Clt C) : 5Y. 1)i. 14 C110.:0 ). h.. dwilt 5S vlo ;/: to: 1k /14 X; AC 601iN 14 it .1.1 Clark/ 011 tr;:k ;\ Figure 2.4 Layout of the mask set developed for the study of the p-well MESFET technology. 16 P; PLAYERS PULLED OUT 1 Z; PLAYERS PULLED OUT 0 p.m m PLAYERS PULLED IN 1 P-WELL M CHANNEL n n IMPLANTS An example of the different p-well implant masks used Figure 2.5 to study the effect of the gate contacting the p-well. current transient and output conductance measurements, the block of devices labeled A (Fig. 2.4) was used. Within this array are transistors with no p implants, top or bottom implants only, both top and bottom implants, and combinations of the above with and without a p-well contact. Because the gates are Schottky diodes, when the gate reverse biases the n-channel, it forward biases the p-well. In general, this is an undesirable effect which leads to high gate leakage currents and parasitic gating of the channel from the 17 Table 2.2 MESFET structure identification for block A of Fig. 2.4. Label; X X XX XX IL BM,BZ,BP1 TM,TZ,TP1 P,N2 Q,H,1,2,43 1. B and T correspond to the backside and surface p-type implants respectively. M, Z, and P correspond to the layout of the p-layers in the region of the gate metal outside the channel as illustrated in Fig. 2.5. These labels are not present if the corresponding p-type implant is not present. 2. P signifies a contact to the p-well and N signifies no contact. 3. backside. These characters represent the gate length in microns; Q and H are quarter and half respectively. So in addition to the above combination of devices, devices with the p-well implants outside the channel in the gate region, coincident with the gate region, and pulled inside the channel in the gate region were laid out as illustrated in Fig. 2.5. The available MESFETs are summarized in Table 2.2. The sidegate test structures are gathered together in the block of devices labeled B (Fig. 2.4). test cells on the mask. Fig. 2.6. There are 15 different sidegate The layout of the sidegate cell is shown in The test MESFET has a one micron gate length and a fifty micron gate width. The individual cells allow testing of sidegate sensitivity to electrodes on the source and drain sides of the transistor as well as the edge of the channel. The different cells 18 BGS5 S; SOURCE G; GATE D; DRAIN BG10; SIDEGATE 1 BG5; SIDEGATE 2 BGS5; SIDEGATE 3 BG5 Layout of the sidegating test cell showing the three Figure 2.6 different sidegate electrodes available. Sidegate cell identification for the cells in block B of Fig. 2.4. Table 2.3 Label; 1B XTB XTB Labels for the sidegate electrode. Labels for the MESFET. 1B 1 micron gate length MESFET for sidegating measurements. X X can be N for no contact or P for contact to the p-well connected to the source. TB T and B are the labels for the surface and the backside ptype implants respectively. These labels are only present if the implant is present. 19 HP 8112A PULSE GENERATOR 8 pH _93:00 50 0 9 KO TRIGGER TEK 111501 OSCILLOSCOPE TEK PS5010 DUAL POWER SUPPLY 1 0 K 10 n 4. A IEEE -488 BUS 10 KO Roans. 8 pH QQQ CONTROLLER Schematic diagram of the measurement system used to Figure 2.7 observe the drain current transients. allowed testing the impact of p-type implants and contacts to the pwell which is illustrated and described in section 2.5. The cells were laid out so that the n' sidegate electrode to n' source/drain spacing would be constant for BG10 and BG5 and the n" sidegate electrode to n-channel spacing for BGS5. The different cells are summarized in Table 2.3. The group of MESFETs in block C are for the RF characterization. Included in this array are MESFETs with different source-to-gate and drain-to-gate spacings as well as quarter, half, and one micron gate length devices. Only two of these was examined in this study: the p-well device and the MESFET with no p-implants on the control wafer, both with one micron gate lengths and one micron source-to-gate and 20 drain-to-gate spacings. The placement of the p+ well contact is different for these structures than the placement shown in Fig. 2.2. Since the interdigitated gate prevented the placement of the I)* contact adjacent and parallel to the source contact it was placed at the ends of the source n+ contacts. 2.3 Drain Current Transient Measurements A convenient means of detecting, comparing, and analyzing trap- related effects is to examine the drain current transient responses of conventional and p-well MESFETs when square wave voltages are applied to their drains. The experimental setup used to study drain current transients is shown in Fig. 2.7. The drain of the MESFET is terminated with a 50 ohm resistor to provide impedance matching with the pulse generator. The pulse generator was set to a 50 percent duty cycle to provide a square wave input and deliver independently adjustable high and low voltages. The total drain current was extracted from the voltage across a sense resistor (Rsense) on the source side of the transistor. The resistor value used in this study was 50 ohms, which when taken in parallel with the 50 ohm termination of the oscilloscope sampling head provided a 25 ohm sense resistor. The transient waveform was digitized by the oscilloscope and acquired by a controller over the IEEE-488 interface bus. With a resistor on the source side of the transistor it is difficult to maintain a constant VGS while the drain current is changing. assembled. To avoid this problem the gate biasing network was A dual tracking power supply was used to provide a voltage to a 10:1 resistive divider. The 8 a inductors were 21 PROBE CARD PULSE GENERATOR OSCILLOSCOPE it BIAS NETWORK z DUAL POWER SUPPLY -gt Illustration of the probe card used for the drain Figure 2.8 current transient measurements of GaAs MESFETs. provided to isolate the power supply from the fast rising and falling edges of the pulse generator. The capacitor between the gate and source of the MESFET is a speed up capacitor which was also needed for the falling and rising edges of the pulse generator waveforms. The complete bias network was assembled on a 1 inch square piece of circuit board and mounted on top of the gate probe to minimize stray lead inductance. Fig. 2.8. The probe card assembly is shown schematically in This probe and biasing configuration allowed 10 nSec rise and fall times with no ringing in the test circuit. 22 FUNCTION GENERATOR Roans., TRIGGER d DC POWER V SUPPLY LOCKIN F-7-1 'AMPLIFIER de A I DIGITAL VOLT METER IEEE-488 1\ BUS IEEE-488 BUS SYSTEM CONTROLLER Schematic diagram of the measurement system for Figure 2.9 characterization of the frequency dependent gds. 2.4 Frequency-Dependent Output Conductance Measurements The small-signal output conductance is defined as; ids (2.1) gds = Vds VGS = constant where ids is the small-signal drain-to-source current and vds is the small-signal drain-to-source voltage. The method by which gds is extracted in this study is illustrated in Fig. 2.9. A function generator is used to supply the small-signal voltage (vdd) to the drain of the MESFET, while the DC offset of the function generator was used to supply the the DC drain voltage. The small-signal 23 drain-to-source current is extracted from the small-signal voltage across Rsense on the drain side of the MESFET. This small-signal voltage is measured using a lockin amplifier which was triggered by the function generator. The constant VGs was supplied by a DC power vdd was adjusted until vds=50 mVpd was measured by the lockin supply. amplifier for all frequencies. The data was acquired over the IEEE-488 interface bus by the controller for several different frequencies and stored on disk. 2.5 Sidegating Measurements Sidegating is the decrease in the drain-to-source current of a MESFET due to a voltage applied to a nearby electrode. The nearby electrode can be an n+ contact to an implanted resistor, source/drain contact of a MESFET, or a piece of interconnect metal which is in contact with bare semi-insulating GaAs. A cross-sectional view of the conventional MESFET sidegate test structure is shown in Fig. 2.10. It consists of an n +- contact region in semi-insulating GaAs separated from the source of the transistor by a distance of 10 microns. The two quantities which will be examined most closely are the sidegate current (IsG) and the percent decrease in Ids. All the measurements are done with the MESFET reverse biased so that with zero sidegate voltage applied, Ids was 3.0 milliamps. For the p-well MESFET the sidegate electrode should simulate the source or drain of another MESFET which is in a p-well. The cross sectional view of the p-well MESFET sidegate structure is shown in Fig. 2.11. The other structures listed in Table 2.3 can be obtained by simply removing the respective implants from Fig. 2.11. 24 SOURCE SIDEGATE DRAIN SEMIINSULATING GaAs SUBSTRATE Cross-sectional view of the sidegate test structure Figure 2.10 implemented in the conventional technology. SIDEGATE SOURCE DRAIN PIBurled PWell Channel PSurface Implant SEMIINSULATING GaAs SUBSTRATE Cross-sectional view of the sidegate structure Figure 2.11 implemented in the p-well technology. 25 SMU 4 SMU 3 SOURCE HP 4145B SEMICONDUCTOR PARAMETER ANALYZER SIDEGATE SMU 1 DRAIN-1 GATE SMU 2 Figure 2.12 system. Schematic representation of the sidegating measurement Sidegating was measured using the experimental setup shown in Fig. 2.12. wafer. A probe station was used for doing the measurements on An HP4145B Semiconductor Parameter Analyzer was used to supply the bias voltages and to monitor the currents. The sidegate voltage was stepped to the most negative voltage and swept to zero volts at the slowest sweep rate to minimize the effects of hysteresis. The voltage step size was kept very small so that the sweep would be very slow. All measurements were done in a closed box and the influence of illumination was provided with the microscope light which was focused through the microscope lens onto the sample. The highest intensity available on the lamp was used. 26 DC SUPPLIES V ds gs 1-HP8510 WAFER 1111::11 COPLANAR WAVEGUIDE PROBES CONTROLLER Schematic representation of the measurement Figure 2.13 configuration used for the RF characterization of the p-well technology. 2.6 RF Characterization RF evaluation was done using the test setup shown schematically in Fig. 2.13. The S-parameters of the test samples were measured on wafer using an HP8510B network analyzer between 100 MHz and 26.1 GHz and coplanar waveguide probes from Cascade Microtech. supplied through bias tees in the network analyzer. 1 Am by 300 Am interdigitated gate MESFETs. DC biases were The samples were The S-parameters were then transferred to a personal computer for analysis. A physically based small-signal equivalent-circuit model was implemented using TOUCHSTONE®. The equivalent-circuit parameter values were obtained by an iterative process in TOUCHSTONE®. The iteration process adjusted the element values until the calculated values for the Sparameters of the equivalent-circuit were in good agreement with the measured S-parameter values. 27 3. DRAIN CURRENT TRANSIENTS The influence of deep level traps on the transient behavior of the drain current of GaAs MESFETs fabricated on semi-insulating substrates are examined in this chapter. The drain current transients are shown to arise from both the capture and emission of electrons by traps in the substrate near the channel. The transient behavior is described analytically using a square law model for the drain current which includes terms accounting for the modulation of the channel thickness due to changes in the width of the channelsubstrate junction depletion region. The drain current transient characteristics suggest that control of the floating channel-substrate junction is key to eliminating or reducing this anomalous behavior. The use of p-type implants to confine the channel-substrate junction is examined as a potential technique for accomplishing these goals. The use of a p-well technology is shown to eliminate the drain current transients with long time constants. 3.1 Introduction Understanding the dynamics of charge capture, as well as emission of charge by deep levels is crucial to understanding the transient behavior of GaAs MESFETs. Since n-channel MESFETs are a majority carrier device and the substrate is dominated by the large concentration of EL2, this discussion will be confined to electron traps. For an electron trap the emission and capture time constants are given by [34], [35], 28 ET) / kT e(Ec Te (emission) (3.1) (capture) (3.2) <v>n Nc ae 1 Tc <v>n n ac where, Ec ET is the trap energy below the conduction band minimum, k is Boltzman's constant, T is the absolute temperature, <v>n is the electron thermal velocity, Nc is the conduction band density of states, n is the electron concentration, and ae,c is the cross section for emission and capture, respectfully. The emission process depends on the material properties <v>n and Nc and the trap properties (Ec ET) and ae. Therefore, at a given temperature the emission time constant will be constant for a particular defect. For emission of an electron from EL2 at room temperature the time constant is approximately 17 mSec. different. The capture process is considerably It depends on the material properties through <v>n and n, and the defect properties through ac. The dependence of the capture time constant on n can result in behavior which is difficult to predict and model since capture can occur in regions of varying electron concentration. It is instructive to examine how the drain current responds to the capture and emission of electrons by traps in different regions of the MESFET. The MESFET can be divided into three different regions illustrated in Fig. 3.1: the gate-channel depletion region (I), the surface-channel depletion region outside the gate region (II), and the channel-substrate transition region (III). The transient behavior can be described in general terms by careful 29 SOURCE DRAIN Cross-sectional view of the various depletion regions Figure 3.1 which can contribute to drain current transients. consideration of the response of these three regions and the subsequent impact on the drain current after an abrupt change in the drain voltage. A step in Vds from an initial low voltage to a high voltage causes the gate depletion region to extend towards the drain. Subsequently, the subthreshold current flowing through the channel substrate transition region increases and extends further into the substrate in response to the electric field. As traps in region I emit their electrons due to the expansion of the depletion region, causing the depletion region to contract, the drain current 30 increases. The associated time constant is a direct reflection of the emission time constant of the trap in the gate depletion region. The transient response of the drain current due to modulation of the surface space charge region by variations in the potential along the surface has been examined in the past [14]. It has been suggested that a finite surface conductivity can lead to time dependent behavior in GaAs MESFETs [6]. The step from a low to high voltage on the drain of the MESFET results in surface leakage currents which can fill some of the surface states with negative charge, causing the surface depletion region to expand and Ids to decrease. As long as the surface conductivity is relatively constant, the time-dependent behavior will be a reflection of the distribution in capture cross sections present in the surface states. For the drain pulse this is dominated by the surface between the gate and the drain where the largest potential difference exists. In region III the electric field causes an increased rate of charge injection into the substrate region, extending the region through which conduction occurs. This is represented by an increase in the electron concentration deep into the substrate. Capture occurs in regions where there are ionized defects and an excess of electrons beyond the steady-state concentration prior to the drain voltage step. The excess electron concentration is a smoothly varying function of depth into the semiconductor. capture time also varies as a function of position. Therefore the Neutralization of EL2 results in shallow acceptors which are now uncompensated. To achieve steady-state operation, this negative charge associated with 31 N.eff must be balanced. This is achieved by expansion of the channel-substrate depletion region on the channel side of the junction. The electrons which have been swept out of the depletion region leave behind the ionized shallow donors to balance the increase in uncompensated shallow acceptors. The drain current responds by initially increasing to a high value and then decreasing as the channel is pinched off due to the expansion of the channelsubstrate depletion region. The rate at which the depletion region expands is dependent on the rate of capture, and capture is dependent on the local concentration of electrons which is position dependent. Therefore, instead of a single discrete time constant, the time constant effectively changes with time as the ionized EL2 become neutral in regions of decreasing electron concentration. A step from a high to a low voltage on the drain results in considerably different behavior for all three regions. In region I the gate to drain depletion region collapses and the ionized EL2 captures electrons. Since the electron concentration is so high in the channel (z1017 cm-3), the capture time constant is much too fast (z10-12 sec) to observe. In region II the emission of charge from the surface states results in the depletion region relaxing, causing Ids to decay upward after the initial decrease. The decay is most likely to be a collection of time constants due to the distribution in energies of the surface states. In region III the transient response is dominated by the emission of charge from EL2. After the initial decrease in Ids due the abrupt change in Vds, Ids increases as the 32 INPUT V DS WAVE FORM REGION I TRANSIENT RESPONSE REGION II TRANSIENT RESPONSE REGION III TRANSIENT RESPONSE Figure 3.2 Input drain voltage waveform and corresponding drain current transient responses due to the three different regions of Fig. 3.1. electrons are emitted. The decay is therefore dominated by the time constant of EL2. The transient responses of the three regions are summarized schematically in Fig. 3.2. The response of region I is distinctly different from that of regions II and III which at least qualitatively appear to be similar. Differences are expected though based on the relative contributions each is expected to make for current transients resulting from abrupt changes in the drain voltage. When the drain is pulsed to a high voltage, region II contributes to the transient response primarily by increasing Rd, the 33 CONVENTIONAL MESFET TRANSIENTS TIME (50 mSec/DIV) Figure 3.3 Measured drain current transient response of the conventional GaAs MESFET. A voltage square wave between 5.0 V and 1.0 V was applied to the drain. VGS = 0.0 V. parasitic drain-to-gate resistance. Region III contributes to the transient response by increasing Rds, the drain-to-source channel resistance. Since in saturation Rd, > Rd, trapping in region III is expected to contribute much more significantly than trapping in region II. Also there exists the potential for a much broader spread in capture time constants in region III than in region II. distribution in the electron concentration from The large approximately 108 cm-3 in the substrate, to 1017 cm-3 in the channel of region III corresponds to approximately nine orders of magnitude variation in the capture time constant (10-3 to 10-12 seconds). 34 CONVENTIONAL MESFET TRANSIENTS 6 mA 0) a E 0 0 tr) 0 1 mA TIME (1 mSec/div) Measured drain current transient response of the Figure 3.4 conventional GaAs MESFET. A voltage square wave between 5.0 V and VGS = 0.0 V. 1.0 V was applied to the drain. 3.2 Measurement of Conventional MESFETs The transient response of the drain current to a square wave voltage input on the drain terminal with a period of 500 mSec is shown in Fig. 3.3. For a voltage step from an initial high value to a final low value, the drain current decreases to an initial low value and decays slowly upward to its steady-state value. This transient response will be referred to as current undershoot. The time required to reach steady-state is on the order of 200 mSec. 35 T=10-8 Q E (D 0 T=10-7 ...-..- ''...............................,Ibo T=10 -6 T=10-5 4 T=10 22.6 mA Time (T Sec/Div) Figure 3.5 Measured drain current transient overshoot for an abrupt change in the drain voltage from 0.0 V to 5.0 V for a conventional GaAs MESFET. Each curve represents a different time scale ranging from 10 mSec/Div to 10 nSec/Div. VGs = 0.0 V. However, the drain current response for a low to high voltage step does not display a transient on this time scale. When the time scale is expanded, as shown in Fig. 3.4 a transient is observed for this drain voltage transition. As the time scale is expanded further the transient becomes more dramatic as shown in Fig. 3.5. The drain 36 CONVENTIONAL MESFET 9.8 mA LOW . , . HIGH RESPONSE . . . . . 10 mSec 100 ASec 046411t1 tt14W-144401141111 44.01114441114114.01144411 1 p.Sec cn 8.8 mA . . . . . . . . TIME (100 nSec/Div) Measured drain current transient response of the Figure 3.6 conventional GaAs MESFET. A voltage square wave between 5.0 V and 1.0 V was applied to the drain for three different pulse periods. VGS 0.0 V. current increases to an initial high value and decays downward to its steady-state value. This transient response will be referred to as current overshoot. The considerable change in Ids on each scale indicates there are multiple time constants in the relaxation to steady-state. Another feature of the current transients is the dependence of the decay on the period of the applied square wave as illustrated in Fig. 3.6. As the period of the applied signal decreases, the time 37 19 mA Time (T Sec/Div) Measured drain current transient undershoot for an Figure 3.7 abrupt change in the drain voltage from 5.0 V to 1.5 V for a conventional GaAs MESFET. Each curve represents a different time scale ranging from 10 mSec/Div to 10 ilSec/Div. VGS = 0.0 V. constant and the magnitude of the transient overshoot change. This behavior arises because of the long time constant associated with emission. If the drain voltage switches low and traps do not have time to emit any electrons before the drain voltage swings high again, the fixed charge distribution will not have changed. Therefore, the initial current level will be the same as the final current level of the previous pulse. This behavior contributes to the frequency- and pattern-dependent propagation delays and result in considerable modeling difficulties. The drain current transients for the current undershoot are also shown on an expanded time scale in Fig. 3.7. Most of the change takes place on the two longest time scales which is more indicative of a single time constant. The changes occurring for much shorter 38 E 0 0 1 2 AV DS 3 4 5 (Volts) Magnitude of the drain current transient overshoot for Figure 3.8 an abrupt increase in VdS from 0.0 V to AVd$. VGs = 0.0 V. times may be the result of another defect level or may be thermal effects. Comparing the current transient behavior observed for these MESFETs with the discussion accompanying Figs. 3.1 and 3.2, it appears that trapping in region III is dominating the drain current transient behavior. This is not surprising since the channel substrate junction is floating, and is thus susceptible to modulation by charge accumulation in the substrate. The lack of any transient corresponding to region I might be surprising due to the large concentration of EL2 known to exist in semi-insulating GaAs. However, there is considerable evidence showing that EL2 is annihilated in the presence of high electron concentrations during 39 thermal annealing [41], [42]. These conditions are present in the n-channel regions during the implant activation anneal of an ion implanted process. Another result which supports this conclusion is shown in Fig. 3.8. This figure shows the magnitude of the current overshoot for a drain voltage step from zero volts to a high voltage. As the high value of VdS is increased, AIds increases. However, no current overshoot could be detected for drain voltage steps below 0.6 volts. In this range of voltage steps the MESFET remains in the linear region of operation and the electric fields are insufficient to sweep carriers over the channel-substrate barrier and into the substrate. Since transients corresponding to region I of Fig. 3.2 could not be detected, these results are direct evidence that EL2 can be annihilated by common processing steps such as the thermal anneals used for activating ion-implanted dopant atoms. 3.3 Analysis The channel substrate interaction can be modeled as a modulation of the active layer thickness by changes in the depletion layer thickness of the channel-substrate junction [12]. This interaction arises due to the capture and emission of electrons by EL2 and other electrically active defects in the substrate. For the active layer thickness and pinch-off voltage of the devices used in this study, a square law model of Ids is appropriate for analytical purposes. The 40 square law model most commonly used is the Curtice equation [43] given by, VT)2(1 + AVds)tanh(aVds), Ids = B(Vgs (3.3) where, VT is the threshold voltage, A is an empirical parameter related to the drain conductance, and B can be approximated as ides/2Kg. In this expression for B, A is the mobility, Z is the gate width, es is the dielectric constant, W is the active layer thickness, and Lg is the gate length. VT = Vbi We also have that, (3.4) Vpo, where Vbi is the built-in potential of the Schottky barrier and Vpo is the pinch-off voltage which is given by, qNdW2 (3.5) VP0 2Es In this analysis Nd is the doping in the active layer which is assumed to be constant. B can now be rewritten as; AZWqNd (3.6) B 4L9VP0 and Ids can be rewritten as; AZWqNd (Vgs Ids Vbi + VP0)2(1 + AVds)tanh(aVds). (3.7) 4LgVpo While the assumption that Nd is constant is in reality an over simplification, it can serve to ease the analysis considerably. Eventually the transient from of Ids will be expressed in terms of measurable parameters such as B and VP0 which depend on Nd, thereby removing the explicit dependence on Nd. 41 GATE METAL Old Waab ciNa I774 07. V7774 A Womb Charge distribution along a path from the gate into Figure 3.9 the substrate of a conventional MESFET. The shaded regions signify the depletion regions which change due to the balancing of charge. The modulation of Ids by changes in the channel-substrate transition region can be incorporated in Eqn. (3.7) by examining the charge density in Fig. 3.9. The figure represents the charge distribution along the X-X' path of the cross-sectional view of the MESFET in Fig. 3.10. In the neutral substrate, the acceptors are compensated by the ionized deep donors, i.e. Naeff = N. The shaded region represents the area where the above equation for compensation is changing i.e. Ncid+ <---> Ndd° by capturing or emitting electrons depending on how the drain voltage has changed. Using charge balance and the notation in Fig. 3.9, Nd(W Wh) = Na(Wsub AWsub). (3.8) 42 x x Cross-sectional view of a conventional MESFET showing Figure 3.10 the path used in Fig. 3.9. The amount of change that occurs due to a change in Vds is given by, NdAW = NaAWsub (3.9) where, AW = WI Wh. (3.10) If we let, ANdd41° = NaAWsub (3.11) be the total number of EL2 per cm2 which emit or capture electrons 43 (i.e. change charge state) then, AW can be expressed as, ANold+10 AW = (3.12) . Nd In Eqn. (3.12) ANdcr° is time dependent and is given by, ANdd+/°(t) = ANdd+/°(03)(1 ANdd+1°(t) = ANddllo (co) e-t/Tc etire) (emission) (3.13a) (capture) (3.13b) where the emission and capture time constants are given by Eqns. (3.1) and (3.2). The square law model for the drain current is an approximation which assumes that current saturation is due to pinch off of the channel. The expression for ANdd+/°(t) in Eqn. (3.11) is consistent with this approximation where X-X' of Fig. 3.10 passes through the point of pinch off in the channel. It is a measure of the sheet concentration of deep donors which change their charge state along a plane extending along AWsub of Fig. 3.9. Expressing the effects of capture and emission by traps in terms of their concentration is not particularly useful for modeling purposes since these quantities are rarely known with any accuracy. Furthermore, when the traps are native defects such as EL2, there always exists the possibility that their concentrations can change during processing [42]. A more tractable parameter is the channel thickness which is easier to extract from a simple electrical measurement of a MESFET. The change in the channel thickness can now be rewritten as a single expression for both capture and emission of 44 carriers, (3.14) Wc,e(t) = Wh,L ± AWettrc,e The c, h, and + are for emission, and e, 1, and are for capture. These terms are included in the Curtice equation (3.7) by replacing W with W(t). Therefore, Ids and Vpo become, K1 Wc e(t)[Vg + Ids(t) = Vp0(t)]2 (3.15) Vpo(t) where, AZqNd (3.16) (1 + AVds)tanh(aVds), xi 4Lg Vg = Vgs (3.17) Vbi, (3.18) Vpo(t) = K2 Wc,e(t)2, and qNd (3.19) K2 = 2es Expanding Eqn. (3.14) gives, KiWc,e(t) (3.20) [Vg2 + 2VgVpo(t) + Vp0(t)2] Ids(t) - VpoM Replacing V,(t) by Eqn. (3.18) gives, V g2 Ni + 2Vgx214c,e(t) + 22Wo,o(t)3 Ids(t) = K2 I . (3.21) t'ic,e(t) Inserting Wc,o(t) of Eqn. (3.14) and separating terms gives an equation for Ids of the following form; Ids(t) = Ids(m) ± AIds(t). (3.22) 45 The components of Ids(t) are given by, NiWn,L (3.23) [Vg + Voh,L]2, Ids(w) V Aids(t) = ic, f( )Vg2 Wh,LAWe" t/Tc,e -F 4- 11 1,1 poh,L "11,1. "Awe-tt rc,e ± AWetircie (2Vg + 3V poh,l) + 3K2wh,LAw2e-2ttrp,e ± K2Aw3e-3t/rp,e (3.24) 1, where, ciNdwh,i2 V (3.25) = 2es For the depletion mode devices examined in this report Wh,L AW. Therefore, the last two terms of Eqn. (3.25) which are higher order in AW can be ignored. Furthermore, the first term can be approximated as, V92 Wh,LAWCtirc,e t2 (3.26) V Wh,L ± AWE'tiTc,e VP°v h,l h,l tPrc'e. AIds(t) now becomes; AIds(t) = AIds(0)e-tmc,e (3.27) , KiAW AIds(0) = ± (3Vpoh,L2 + 2VtVpoh,i Vg2). (3.28) V Finally, the transient form of Ids can be expressed as, (3.29) Ids(t) = Ids(w) ± Aids(0)Cmc,e The expression in (3.29) is especially convenient for the characterization of drain current transients because it is expressed directly in terms of measurable parameters. For a substrate heavily 46 23 = 10 T = 10-3 T = 10-4 20 19 TIME (T Sec/Dev) Analytical results of drain current transient Figure 3.11 undershoot for the conventional MESFET results of Fig. 3.7. dominated by EL2, the characterization of the emission process is easily accomplished by measuring AIds(0) and Ids(w). The time constant and emission cross section of EL2 are well known and are easily to verified by measuring the transient at different temperatures. The capture process however, presents some difficulties in that the capture time depends on the concentration of excess electrons in the vicinity of ionized EL2. 3.4 Comparison of Analysis and Measurement The analytical expressions for the drain current can be compared with the experimental results of section 3.2. The most straight forward comparison of the analysis and experiment is a comparison of equation (3.29) to the transient data of Fig. 3.7. is shown in Fig 3.11. This comparison After adjustment of Te of EL2 for self-heating in the MESFET [5], the Ids curves compare nicely to the curves of Fig. 3.7. 47 0 10-9 1 10-8 10-7 10_6 10-5 10-4 10-3 10-2 10-1 Time (seconds) Time dependence of the log of the magnitude of current Figure 3.12 overshoot normalized to t=0 for the measured current overshoot waveforms of Fig. 3.5. As mentioned earlier, characterization of current overshoot is complicated by the absence of a distinct time constant. is an equation characteristic of exponential processes. Eqn. (3.29) The time constant can be extracted by manipulation of this equation to give; t ti Tc 9 (3.30) rlds(t1 In AIds(ti) where Tc is just the slope of the line generated from plotting the denominator of the above equation versus time; ti is some initial time, and t is some time later. A plot of the denominator in Eqn. (3.30) versus time is shown in Fig. 3.12 for the transient overshoot results of Fig. 3.5. For an exponential process with a single time 48 constant the curve should be a straight line when the horizontal axis is linear. However to fully examine the trends, the horizontal axis had to be logarithmic. These results demonstrate the absence of a simple exponential process. Clearly, if current overshoot and decay are the result of electron capture in the substrate, a single distinct time constant would not be expected. When the drain voltage is stepped from a low voltage to a high voltage, the electron tail in the channel-substrate junction shifts into the the substrate. This shift is in response to the changing electric field generated by the drain voltage. The constant potential lines wrap around the gate depletion region and the channel-substrate interface, pulling charge into the substrate along the electric field lines. in junctions decay exponentially. In general, carrier concentrations Likewise this new electron tail will decay exponentially into the substrate. Therefore, it is reasonable to expect the time constant to show a strong dependence on time in response to the exponential dependence of the electron concentration on position. Closer investigation of the time dependence of Tc can be accomplished by rewriting Eqn. (3.30) as the simple finite difference equation given below; ti _1 ti+1 (3.31) Tci rds(ti+1)] AIds(ti.1) This formulation of Tc relates the capture time to the changes in Ids created by capture of electrons in specific regions of the electron 49 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 10-9 10-8 10-7 10-6 10-8 10-4 10-3 10-2 10-1 Time (seconds) Figure 3.13 Time dependence of the capture time constant for the transient current overshoot of Fig. 3.5. tail as the MESFET reaches steady-state. The results are shown in Fig. 3.13 with a least squares fit to the data. The dotted line which has been fit to the data results in the following equation; r(t) = 1.21t°893. (3.32) This equation provides a good fit over a range of six orders of magnitude in time. The six orders of magnitude change in the time constant correspond to a six order change in magnitude in electron concentration which is easily transition region. achieved in the channel-substrate This equation can be used in Eqn. (3.29) to model the decay of the drain current in response to a drain voltage step from low to high. However, for this to be successful, the prefactor 50 to the exponential (AIds(0)) must be modified to represent the amount of change occurring while this time constant is valid. 3.5 Transient Measurements of Buried-Channel and P-Well MESFETs The use of p-type implants to control the drain current transients have met with mixed results [9], [44], [45]. Much of the difficulty interpreting previously reported results has been due to the lack of direct comparison with control samples. Based on the transient results of the conventional devices, it is reasonable to postulate that control of the channel-substrate junction should result in improvements in the transient response of the MESFET. One possibility is to leave the p-type region floating as is typically done and hope that the increased barrier height and abruptness of the junction will significantly lower the rate of charge injection into the substrate and subsequent capture and emission [45]. Another possibility is to enclose the MESFET in a p-type well and confine the potential of the channel-substrate junction to a known voltage, thereby isolating the channel of the MESFET from fluctuations in the charge distribution in the substrate. The drain current transient response of samples with four different p-type implants behind the channel were examined and compared. The transient response for a low to high voltage step is shown in Fig. 3.14. Little improvement is seen in the samples with the lower p-type doping (2.0 and 4.0 x 1016 crif3 peak p concentration) behind the channel in comparison to the conventional device. The two samples with the higher doping (8.0 and 16.0 x 1016 cm-3 peak p concentration) behind the channel are significantly different. The 51 T=I0 -7 -5 T =10 T=10-3 Time (T Sec/Div) T=10-7 T=10-5 T=10-3 Time (T Sec/Div) c) E op T=10-7 T=10 T=1 Time (T Sec/Div) .4( E T=10 03 -7 T=10T=10-3 Time (T Sec/Div) Figure 3.14 Measured drain current transient overshoot response for a 0.0 V to 5.0 V drain voltage step on buried-channel MESFETs for 1 mSec/Div, 10 ASec/div, and 100 nSec/Div scales. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm- 3, c) 8.0 x 1016 cm -3, and d) 16.0 x 1016 cm-3. 52 a) T=10-3 ../1....... T:=10-5 Time (T Sec/Div) b) T=10-3 T=10-5 Time (T Sec/Div) Time (T Sec/Div) Time (T Sec/Div) Measured drain current transient undershoot response Figure 3.15 for a 5.0 V to 1.5 V drain voltage step on buried channel MESFETs for 1 mSec/Div and 10 ASec/div scales. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 53 time constant for the decay of the drain current increased The dramatically for the buried-channel MESFETs with higher doping. transient undershoot is shown in Fig. 3.15. As in the overshoot case, the two samples with lower p-type dopings behind the channel For showed little change when compared to the conventional device. the two samples with higher doping behind the channel, the magnitude of the transient increases. The time constant also changes for these two samples, becoming faster as the doping is increased. The drain current transients for the p-well MESFET with the same doping levels behind the channel have also been investigated. results of the current overshoot are shown in Fig. 3.16. The For the two samples with the lower peak p-type doping levels behind the channel, the results are similar to the buried channel structure. That is they show little improvement or change when compared to the conventional MESFET. The two samples with the higher peak p-type doping behind the channel are dramatically different. While the magnitude of the transients is slightly greater, the current has essentially decayed to its DC value in under 100 nSec. The drain current transient results for a high to low voltage step on the drain are shown in Fig. 3.17. The samples with the lower p-type doping levels showed little change in the response as opposed to the control samples. However, the sample with the peak p doping behind the channel of 8.0 x 1016 cm-3 showed no signs of any transient. 54 T=10 a E -7 T=10 -5 03 T=10-3 0 ,....., Time (T Sec/Div) E 03 Time E T=10 T Sec/Div) -7 OD T=10-5 T=10-3 V Time (T Sec/Div) T=10 -7 T=10-5 T=10 -3 Time (T Sec/Div) Measured drain current transient overshoot response Figure 3.16 for a 0.0 V to 5.0 V drain voltage step on p-well MESFETs for 1 mSec/Div, 10 ASec/div, and 100 nSec/Div scales. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 55 . . . . T=10-3 . 1.---... T=10-5 . . . . . Time (T Sec/Div) Time (T Sec/Div Time T Sec/Div Time (T Sec/Div) Measured drain current transient undershoot response Figure 3.17 for a 5.0 V to 1.5 V drain voltage step on p-well MESFETs for 1 mSec/Div and 10 gSec/div scales. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm -5, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 56 3.6 Summary The behavior of the drain current transient response to an abrupt change in the drain voltage of conventional MESFETs is strong evidence that this anomalous behavior is the result of charge capture and emission in the GaAs substrate. The ability to characterize this behavior using a square-law relationship for the drain current can be utilized in modeling these effects, although further parameterization of the current overshoot behavior must be accomplished. The concept of a floating channel-substrate junction modulating the MESFET current in conventional technology has been demonstrated from these results. The p-well MESFET was demonstrated to eliminate the drain current transients with long time constants and offers hope of a technology which is better suited for design of precision analog and digital circuits in GaAs. Significant reduction in the settling time of amplifiers and switches, reduction in hysteresis in differential circuits and comparators and reduction in frequency and pattern dependent propagation delay times are just some of the potential benefits possible with these improved device results. 57 4. FREQUENCY DEPENDENT OUTPUT CONDUCTANCE The characteristics of the frequency dependent output conductance are examined in this chapter. It is shown that gds(f) is related to the emission time constant of electrons from deep levels in the substrate near the channel. Using the expression for drain current transients in chapter 3, an analytical expression for the frequency dependent output conductance is derived. This analytical expression is compared with experimentally obtained results and shown to be in good agreement. The use of p-type implants to control the frequency dependence of gds is also examined. The p-well MESFET is shown to offer the greatest reduction in the frequency dependence of gds at low frequencies. 4.1 Introduction The output conductance is the small-signal response of the drain current to a small-signal voltage applied to the drain of a MESFET. Ideall y, gds is independent of frequency. However, in conventional GaAs MESFET technology gds is frequency dependent. Studies of the frequency dependence of gds have primarily focused on equivalent circuit modeling of the effect [1], [3], [4], [24], and very little on explaining its physical origins [46], [47]. The frequency response of gds is characterized by a low value at low frequencies, a high value at high frequencies, and a transition region which begins at about 10 Hz at room temperature. Just as the transient results of chapter 3 emphasized the deleterious role the floating channelsubstrate junction plays on the time dependent behavior of the 58 MESFET, it is also linked to the frequency dependence of gds at low frequencies. Conceptually the frequency dependent behavior of gds can be understood by carefully considering the interaction of defects in the substrate with the channel current when a small signal is applied to the drain. When the terminal voltages are applied to the transistor, it will reach steady state operation within a few hundred milliseconds as illustrated by the transient measurements of section 3.2. With the subsequent application of a small-signal voltage to the drain at a high frequency (f > 1/Te), the changes in Vds will occur too rapidly for any significant capture or emission to take place as illustrated by the results of Fig. 3.6. At a high frequency, the charge associated with the deep donors is fixed and the small-signal current is a reflection of this fixed charge distribution. At low frequencies, the channel substrate depletion region will respond to the slowly varying voltage on the drain. When the small-signal swings high, the rate of injection into the substrate and subsequent rate of capture in the substrate increases, resulting in a decrease of the drain current due to the expanding channel-substrate depletion region. When the small-signal voltage swings low the rate of injection into the substrate and subsequent rate of capture decreases and the current will increase due to the relatively higher rate of emission, which decreases the degree to which the channel is pinched off from the backside. That is, at low frequencies the traps can follow the applied signal and the magnitude of the small-signal drain current will be less than it is at high frequencies. Since gds is 59 Vds ds High frequency and low frequency small-signal current Figure 4.1 for a given applied small-signal vds. defined as ids/vds with VGs constant, and the peak-to-peak value of ids is smaller at low frequencies than at high frequencies we know that gds at low frequencies is smaller than gds at high frequencies (Fig. 4.1). It should be emphasized that emission is the rate limiting process. As observed from Fig. 3.5, most of the change in Ids occurs in the first couple of hundred microseconds for a capture process. Most of the change in Ids of Fig. 3.7. occurs between 100 microseconds and 100 milliseconds for the emission process. Before charge can be captured, there must be ionized deep levels to capture the charge. If the charge has not had sufficient time to be thermally emitted, 60 15 13 E 11 U) Vgs(volts) -o CD 0 0.75 7Q A 0 . .1 10 I . 100 . ....I 1000 0.0 0.5 . 1E4 1E5 FREQUENCY (Hz) Measured frequency dependence of the output Figure 4.2 conductance for three different gate voltages. Vps = 2.0 V. then no change in the charge distribution will occur before the small-signal swings high again. 4.2 Measurement of Conventional MESFETs Typical output conductance measurements versus frequency are shown in Fig. 4.2. At low frequencies gds is small, increasing between 10 Hz and 100 Hz and leveling off. The rise at 10 kHz may be due to another trap level or may be due to the thermal time constant of the GaAs. In Fig. 4.3, gds is shown as a function of frequency for different lips. As Vps increases, so does the magnitude of the change in gds from low to high frequency. For small Vps, the change in gds with frequency vanishes which supports the conclusions of the drain current transient results shown in Fig 3.8 suggesting that 61 2.5 2.0 0.5 10 0 1 10 10 2 3 10 10 4 105 Freq. (ME) Measured frequency dependence of the output Figure 4.3 conductance for different Vps. VGs = 0.0 V. there is no EL2 in the channel. That is, at low field conditions in the linear region of operation the small-signal on the drain does not modulate the charge states of defects in the substrate. The increase in gds as Vds increases is due to the increased penetration of the electrons into the substrate on the drain side of the gate. This allows for increased modulation of the channel due to the increased number of traps which can follow the signal. 62 4.3 Analysis The agreement of the above discussion with the experimental data reinforces the conclusion that emission of electrons from EL2 is the rate limiting process. Therefore only the emission part of Ids(t) need be examined in deriving the frequency dependent gds. Furthermore, from the drain current transient data, a single time constant is a satisfactory approximation for analytical purposes. This should allow the use of Eqn. (3.22) to calculate an analytical expression for gds. Rewriting Eqn. (3.22) in a more appropriate form gives, ic3 Ids ( t ) (1 + AVds)tanh(aVds) = V Pot x [Wt (V9 + VpoL)2 AWe-ttre(3Vpoh 1.2 + 2VtVp,,,h,1 (4.1) V92) ] where, AZqNd (4.2) K3 = 4Lg The output conductance is defined by, ales (4.3) gds avos VGS = const. Taking the derivative of (4.1) gives, K3 [Atanh(aVds) + (1 + AVds)asech2(aVas)] gds(t) = VPot x [WL(Vg + VpoL)2 - Awe-tpre(3vpohi2 + 2VtVpoilL V92)]. Since gds is a small-signal parameter, the time dependence is not particularly useful. However, the frequency dependence can be obtained by taking the Laplace transform of gds(t) in Eqn. (4.4). (4.4) 63 X3 [Atanh(aVdS) + (1 + AVds)asech2(aVds)] gds(S) V Poi (WI X ( + Vp01)2 S AW W) T. + s 2AW(Vp012 Vg2) (4.5) WT. + s which can be simplified to + (Wh/WL)Tes} 1 (4.6) gds(s) = gds(0) 1 + TeS S The term in brackets is the small-signal response, 1/s is the forcing function, and gds(0) is the low frequency value of gds(s) given by, X3 (3V92 gds(0) = 2VgVpo1 Vp012)W1 Vpot (4.7) x [Atanh(aVds) + (1 + AVds)asech2 (aVds)]. Finally, the small-signal response can be written as; 1 + [gds (m)/gds(0)1Tes (4.8) gds(s) = gds(0) 1 + TeS where gds(m) is identical to gds(0) in Eqn. (4.7) with WI replaced by Wh. This form of gds is particularly useful for modeling because it is expressed in terms of the measurable quantities, gds(m) and gds(0), and the time constant of EL2 which is well established. This equation is identical to a semi-empirical equation derived earlier [5]. However, here gds(s) is derived from the Curtice model and is dependent on the common set of modeling parameters B, A, a, and VP0. Implementation of a physically based model which accounts for the frequency dependence should therefore be much simpler. The robustness is illustrated in Fig. 4.4 which shows Eqn. (4.8) for gds data at three different temperatures. 64 20 16 275 °K X 325 °K,,.. .**-'375 ° K I 101 10 2 10 3 10 4 10 5 106. Freq. (1-1Z) Comparison of the frequency dependence of the output Figure 4.4 conductance model and measured data for three different temperatures. 4.4 Measurement of Buried-Channel and P-Well MESFETs Based on the improvements in transient response observed by use of the p-well MESFET, similar improvements in gds are expected. This is confirmed in Fig. 4.5 which show gds as a function of frequency for different Vgs. The two samples with the lower p-type doping behind the channel continue to show considerable variation with frequency. The two samples with the higher p-type well doping show that control of the channel-substrate junction is very effective in eliminating the frequency dependence of gds at frequencies below 100 kHz. 65 2G ) 15 i 43 0-0 10 0 V gs (..olls) O 0.75 A O 10 100 1000 0.0 0.5 1E4 1E5 1E4 1E5 FREQUENCY (Hz) 10 100 1000 FREQUENCY (Hz) 30.0 25.0 20.0 7 15.0 10.0 5.0 0.0 10 100 1000 1E4 1E5 1E4 1E5 FREQUENCY (Hz) 35.0 30.0 E 25.0 20.0 Nr 15.0 10.0 5.0 0.0 100 1COG FREQUENCY (11Z) Figure 4.5 Measured frequency dependence of gds for the p-well MESFET at three different VAS. Vps = 2.0 V. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm- 5, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 66 E E N Vgs(volts) ti O -0.75 0.0 O 0.5 10 1000 100 1E5 1E4 FREQUENCY (Hz) 10 100 1000 1E4 1E5 FREQUENCY (Hz) 100 1000 1E4 1E5 FREQUENCY (Hz) 10.0 d) 8.0 V 9S (volts) O -0.75 A 0.0 O 0.5 -6- 0 4,- 6' -6- -45- -A- - -It, -46- -6- 4.0 2.0 0.0 10 100 1000 1E4 1E5 IREOUEUCY (Hz) Figure 4.6 Measured frequency dependence of gds for the buried channel MESFET at three different VAS. V = 2.0 V. The peak p-type doping behind the channel is a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and d) 16.0 x 1016 cm-3. 67 The-buried channel samples on the other hand became much worse as the doping in the p-type layer is increased as shown in Fig. 4.6. For the two samples with the highest p-type doping the buried p-layer simply improves the efficiency of the capacitive coupling between the drain and the channel. 4.5 Summary The successful extension of the analysis of the drain current transients in chapter 3 to the frequency dependence of gds is convincing evidence that the two phenomena are linked. Basing the analysis on modulation of the drain current by varying the charge states of the deep levels in the substrate further substantiates the claim that the floating channel-substrate junction is responsible for this anomalous behavior. Finally, elimination of the frequency dependence at low frequencies by controlling the channel-substrate junction offers the opportunity to develop a technology which is minimally impacted by this anomalous behavior. 68 SIDEGATING 5. Chapters 3 and 4 explored many of the adverse consequences of an uncontrolled channel-substrate junction in GaAs MESFET technology which arise due to changes in the drain voltage. This floating junction can also be influenced by a voltage applied to a nearby electrode such as the drain or source of a nearby transistor, this effect is referred to as sidegating or backgating. The basic characteristics of sidegating in conventional technology are reviewed in this chapter. A substantial decrease in the anomalous behavior of gds and Ids were demonstrated in chapters 3 and 4 using the p-well MESFET. In this chapter, the influence of p-type implants on the sidegating behavior will be examined. Just as in the case of the transient and gds behavior, p-type implants have a strong influence on the sidegating behavior. The p-well MESFET is shown to offer the greatest immunity to sidegating. 5.1 Sidegating in Conventional Technology Sidegating is the result of a loss of isolation between two devices, which causes the behavior of one device to be modified by changes in the operating conditions of the other. In GaAs ICs, sidegating can occur in MESFETs, implanted resistors, current limiters and Schottky diodes. The electrode or device responsible for causing the loss of isolation and subsequent sidegating can be any one of the four devices listed above plus metal interconnect which lies on semi-insulating GaAs. Sidegating due to interconnect metal on semi-insulating GaAs can be eliminated by requiring all 69 CONVENTIONAL MESFET SIDEGATING 105 10-6 1 0-7 E 8 10 109 10-10 10 0 SIDEGATE VOLTAGE (1 V/div) Measured sidegating results for the conventional Figure 5.1 MESFET structure of Fig. 2.10 at room temperature and without The drain current is shown in the upper trace and illumination. sidegate current is shown in the lower trace. Vin = 2.5 V and Ids (VGS=°) = 3.0 mA. interconnect metal to lie on a passivating layer such as silicon nitride or silicon dioxide. Since current limiters, implanted resistors, and Schottky diodes will all be susceptible to the same sidegating mechanisms as MESFETs, only MESFETs will be examined in this chapter. Typical sidegating results are shown for the conventional MESFET technology in Fig. 5.1. A MESFET with a 50 gm gate width is first biased to three milliamps of drain current, followed by the application of a large negative sidegate voltage which is swept very 70 CONVENTIONAL MESFET SIDEGATING 100 W/0 ILLUMINATION WITH ILLUMINATION 0 . . 10 0 SIDEGATE VOLTAGE (1 V/div) Measured drain current sidegating results with and Figure 5.2 without illumination for the conventional MESFET of Fig. 5.1. slowly to zero volts. The drain current is shown as a percentage of Ids at a sidegate voltage of 0.0 V on the vertical axis on the left. The sidegate current is shown on the vertical axis on the right. The sidegate voltage which coincides with an abrupt decrease in the drain current and an abrupt increase in the sidegate current is referred to as the trap-fill-limit voltage (Vtfl) [48]. The Vtft corresponds to an applied voltage which results in a rate of injection of carriers into the substrate which exceeds the rate of emission of these carriers from the traps in a path between two electrodes. In this case the 71 CONVENTIONAL MESFET SIDEGATING 0.1 0.08 ca. 0.06 E 0 (f) 0.04 0.02 0.0 -2.5 V Vcs (0.25 V/div) 0.0 V Measured threshold voltage shift of a conventional Figure 5.3 Vps = 2.5 V. 5.0 V. MESFET for VsG = 0.0 V and VsG = two electrodes are represented by the sidegate n* electrode and the channel-substrate junction. The result is an unpinning of the Fermi-level and an increase in the conductivity of the material. The sidegating effect is further aggravated by illumination which acts to lower the substrate resistivity by generating free carriers. Fig. 5.2 shows measured sidegating results with and without illumination. With illumination, the loss of isolation causes the drain current of the given device to decrease with virtually zero volts applied to the sidegate electrode. In effect 72 (X ) (A ) 100.0 1E-05 10.00 (made /d1v 0+00 Cr 0000 (A ) SE-05 100.0 E.-00 Idly . LOG b) LOG a) I0 1E-12 -10.00 vSG 1.000/div .000 .0000 ( V) c) IS(A 1E-12 -30.0C VSG 1.000/01v ( V) d) 10 ) ) ecadu /01v 10.0 /01v (X ) 16G (A 100.0 AE-05 100.0 E+00 1E-05 10.00 decade /d1v 10.00 /01v ecade /01v 1E-12 -10.00 0000 E+00 /01vr 0000 .0000 vsG 1.000/01v .0000 ( V) 112 -10.00 VSd 1.000/01v ( V) Measured sidegating results of a conventional MESFET Figure 5.4 and an n+ sidegate electrode surrounded by a p-type implant for VDS = 2.5 V and Ids(VsG=0 V) = 3.0 mA. The equivalent peak p-type doping levels were a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cill3. then, the substrate now provides an effective electrical contact between the back of the channel and the offending sidegate electrode. In digital circuit applications, the shift in the threshold voltage due to a sidegate electrode is important [18]. Fig. 5.3 shows the square root of drain current versus Vgs for two different sidegate voltages. The threshold voltage shifts 350 mV for a sidegate voltage of 5.0 V. Not only does this result in the shifting of the turn on voltage of a digital gate, but it can cause a dramatic 73 ID IBG ID (A (X ) IBG (X ) ) (A I 100.0 1E-05 100.0 E+00 1E-05 10.00 cicada 1Idly .00 acetic E+00 Idly Idly 1.000/d1v VSG ( V) I0 I6G (A) ) 100.0 E+00 10.00° Idly VSG 1.000 /div 1.000/d1v ( V) MG ID (X (A) ) 1E-05 100.0 E+00 1E-05 'decade Idly 10.00 Idly scads Idly 1E-12 . .0000 1E-12 -10.00 .0000 -10.00 VSG .0000 0000 1E-12 .0000 .0000 (X Idly -10.00 ( V) .0000 .0000 VSG 1.000/01v lE-12 -10.00 ( V( Measured sidegating results of a buried channel MESFET Figure 5.5 and an n+ sidegate electrode surrounded by a p-type implant for The equivalent peak p-type 2.5 V and Ids(VsG=0 V) = 3.0 mA. VDS doping levels were a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. decrease in the speed of a digital circuit due to decreased current drive capability. 5.2 Measurements of Sidegate Structures with Floating P-Type Implants In order evaluate the relative performance of p-type implants and to optimize their usefulness, sidegate structures with different combinations of implants were tested. Results of samples with p-type implants around the sidegate electrode and no p-type implants around the MESFET are shown in Fig. 5.4. The drain current in the sample 74 ID 1SG ID (X ) (X 100.0 E+00 1E-05 100.0 E+00 10.00 Idly cicada 10.00' /div Idly .0000 .0000 1E-12 VSG ID (x 1.000 /div - 10.00 ( V) c) ) I8 (A I 100.0 E+00 10.00' Idly .0000 .0000 .0000 .0000 1.000/d1v (A ) 1E-05 decado /d1v 1E-12 . VSG ID (x ) - 10.00 1.000 /div ( V) rec d 1E-00 100.0 E+00 1E-05 decade /d1v 10.00 Idiv scads /d1v 1E-12 VSG 183 b) ) - 10.00 ( 0000 .0000 1E-12 VSG - 10.00 1.000/d1v ( Measured sidegating results of a buried channel MESFET Figure 5.6 and an n+ sidegate electrode for lips = 2.5 V and Ids(VsG=0 V) = 3.0 mA. The equivalent peak p-type doping levels were a) 2.0 x 101 cm-3, b) 4.0 x 10 16 cm-3, c) 8.0 x 1016 cm- 3, and 16.0 x 1016 cm-3. with the lightest doping around the sidegate electrode showed virtually no sidegating immunity. The three samples with the higher p-type doping showed no signs of sidegating for sidegate voltages as high as -10 volts, although the sidegate current steadily increased as the p-type doping increased. Similar results have been reported for sidegate electrodes co-implanted with p-type dopants [23]. The addition of the p-implants can act to increase the concentration of ionized deep levels, and consequently the concentration of traps which act to increase Vtft. 75 :asM ) 18G b) MI (A ) 100.0 1E-05 100.0 E+00 1E-OS 10.00 dey:te, 10.00 decade idly E+00 Idly .0000 .0000 Idly 1E12 -10.00 1.000/div VSG ( V) ID (X vSG 18( c) ) (A) 100.0 1E-05 10.00 d scads E+00 /01v Idly .0000 .0000 VSG 1.000/01v 1E-12 -10.00 ( V) .0000 .0000 1E-12 -11.00 1.000/01v ( V) 18G d) ID (Si (A ) 1E-05 100.0 E+00 10.00' /div, decade Idly .0000 .0000 VSG 1E-12 -10.00 1.000/01v ( V) Measured sidegating results of a conventional MESFET Figure 5.7 and an n* sidegate electrode in a p-well for Vin = 2.5 V and IdS(VsG =O V) = 3.0 mA. The equivalent peak p-type doping levels were a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 crif3. The sidegating results for an n+ sidegate electrode and a buried-channel MESFET are shown in Fig. 5.5. The sidegating threshold steadily decreases for this sidegate structure as the doping in the buried p-type layer behind the channel increases. The sharpness of the decrease in Ids also increases with the p-type doping level indicating that the p-type layer is forming a much more effective electrical contact between the sidegate electrode and the channel of the MESFET. 76 IBG ID (% (A ) iE E+00 \ decade /div 4.744/div 1E-11 .0001 VSG 1.000/div 10.00 Measured forward biased p-i-n diode effects for a pFigure 5.8 well MESFET and an n+ sidegate electrode. The sidegating results for a fully buried-channel technology are shown in Fig. 5.6. The sidegate structure consists of combinations of the previous two examples; a buried channel MESFET and a p-type implant around the n+ sidegate electrode. The sidegating results improve steadily as the p-type doping level increases. The sample with the highest level of p-type doping showed only a two percent decrease in drain current for a sidegate voltage of -10 V with no noticeable increase in the sidegate current. 77 ID o) (X VSG 1.000/dlY 113G (A iE-05 1E00.+00 0 acade Idly 10.00 Idly 1E-12 -10.00 SSG (A) 100.0 1E-05 10.00 d scads E+00 /d1v Idly .0000 .0000 VSG 1.000/d1v (X I 1E-12 -10.00 ( V) I(30 (A ) iE-05 cicada /01v 1E-12 .0(1800 VSG ( V) c) b) 10 ) /0 1.000 /div 10.00 ( V) d) (X) 100.0 E+00 10.00' idly ;decade /d1y .0000 .0000 VSG 1.000/d1v 1E-12 -10.00 ( V) Measured sidegating results of a buried channel MESFET Figure 5.9 and an 6+ sidegate electrode in a p-well for Vps = 2.5 V and The equivalent peak p-type doping levels were Ids(Vse0 V) = 3.0 mA. a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 5.3 Measurements of Sidegate Structures with Contact to the P-Type Implants The buried-channel MESFETs showed the ability to improve as well as degrade the sidegating of GaAs MESFET technology. However, similar or greater improvements should be observed for the p-well MESFETs. The results for a sidegate structure consisting of a conventional MESFET and an n+ electrode in a p-well are shown in Fig. 5.7. The results are similar to the results shown in Fig. 5.4 above, with only the sample with the lowest p-type doping showing any 78 IBG ID (X) (A ) a) 1E-05 100.0 E+00 decade /div 10.00/dim_ .0000 .0000 ID . ...._ VSG (X ) 1E-12 10.00 1.000 /div ( V) IBG (A ) b) 100.0 E+00 1E-05 10.00/d1v. .0000 .0000 ID 1E-12 VSG 2.000/div 20.00 ( V) (X ) IBG (A ) .0000 .0000 1E-12 VSG 2.000/d1v 20.00 ( V) Measured sidegating results of a p-well MESFET and an Figure 5.10 n+ sidegate electrode in a p-well for Vps = 2.5 V and The equivalent peak p-type doping levels were Ias(VsG=0 V) = 3.0 mA. a) 2.0 x 1016 cm-3, b) 4.0 x 1016 cm-3, c) 8.0 x 1016 cm-3, and 16.0 x 1016 cm-3. 79 18G a) IG (X ) 1E-05 100.0 E+00 150 /0 (A (A (x ) 1E-05 100.0 E+00 ccade /01v 1E-12 -20.00 .0033 .0000 2.000/U1v VSG 180 c) IO (X ) .0000 0000 ( V: CA ) 1E-12 -10.00 vSG 1U 1.000/01v ( v) ISO d) (% ) (A ; 100.0 E+00 1E-05 100.0 1E-05 10.00 decade 10.00 /d iv /d1v facade /0 iv E+00 .0000 .0000 VSG -- 1E-12 -20.00 2.000/01v ( V) -- 0000 .0000 VSG 1.000/01v 1E-12 -10.00 ( V) Measure sidegating results for various sidegate Figure 5.11 a) Conventional MESFET and structures with and without illumination. an n+ sidegate electrode surrounded by a p-type implant, b) buried channel MESFET and an n+ sidegate electrode surrounded by a p-type implant, c) conventional MESFET and an n+ sidegate electrode in a ptype well, d) buried channel MESFET and an n+ sidegate electrode in a p-type well. sidegating. This technique should be more resistant to sidegating than the structure of Fig. 5.4 since the zero biased pn junction insures that no electrons are injected into the substrate to fill the traps. In the case of the floating pn junction, the junction can forward bias to supply current to the substrate and fill the traps. One of the greatest dangers of using p* contacts is the potential to forward bias p-i-n diodes. This is illustrated by the results for a p-well MESFET and an n+ sidegate electrode shown in 80 PWELL MESFET SIDEGATING g 100 105 80 106 WITH ILLUMINATION W/0 ILLUMINATION 109 0 10-10 -20 0 SIDEGATE VOLTAGE (2 V/div) Measured sidegating result for the P-well MESFET Figure 5.12 technology with and without illumination. Fig. 5.8. Such a configuration can be disastrous and should be avoided at all costs. The sidegating results for a buried-channel MESFET and an n* electrode in a p-well are shown in Fig. 5.9. similar to the results of Fig. 5.6. These results are very The sidegate current remains low for all the samples with higher sidegate thresholds. The sample with the highest p-type doping decreased by approximately two percent at VSG = -10 V. The sidegating results which represent the complete p-well technology are shown in Fig. 5.10. The sample with the lowest p-type doping level is not shown since it exhibited the forward biased p-i-n diode behavior which is the result of the p-type doping being to low 81 to from a well defined p-type region around the n+ region. The sample with a peak p-type doping of 8.0 x 1016 cm-3 showed no measurable decrease in the drain voltage. Even for a sidegate voltage of -20 V and a 3 Am sidegate electrode to source gap. The sample with the highest p-type doping showed what appears to be punch-through effects. For the implant doses and the implant species (Be), this is not surprising. Be has a much greater implant side- straggle than Si, and for high concentrations it can diffuse quite readily. These conditions may be present for these doping levels leading to a much narrower effective sidegate-to-source electrode spacing. The best results were examined more closely to determine their resilience to adverse conditions. The effects of illumination were shown to adversely effect the sidegating results of conventional MESFET sidegate structures. Fig. 5.11 shows the sidegate results for the samples of some of the previous test structures under illumination. The samples with the conventional MESFET and sidegate electrode surrounded by a p-type implant showed the greatest immunity to sidegating. The sample with the p-well sidegate electrode showed no change even for a -20 V sidegate bias. However, such a configuration my not be practical since care would still have to be taken to insure that the p+ contact did not form a p-i-n diode with a nearby n+ contact. The two samples with the buried-channel MESFET essentially shorted out when the light was turned on. The p-layers behind the channel and the photogenerated carriers allowed the formation of an efficient backside contact to the channel of the 82 MESFET. The effectiveness of the p-well technology under illumination is illustrated in Fig. 5.12. Even though the sidegate current increased by two orders of magnitude, the drain current failed to decrease by any measurable amount. 5.4 Summary The sidegating behavior of a number of sidegate structures has been examined to determine the optimum combination of implants. The immunity of the p-well MESFET technology to sidegate voltages of -20 V with no decrease in the drain current even under illumination is further evidence that control of the channel-substrate interface is of the utmost importance in GaAs technology. In addition, the hazard of forward biased p-i-n diodes were illustrated, and degradation of sidegating in buried-channel MESFET structures was demonstrated. 83 6. RF CHARACTERISTICS The RF performance of p-well MESFETs and a suitable small-signal equivalent circuit model which accounts for the well capacitance and well contact resistance are examined in this chapter. The dependence of the equivalent circuit parameter values, unity current gain (fT), maximum frequency of oscillation (fmax), and maximum available gain (MAG) on the p-well doping is examined. When compared to control samples with no p-type implants, the p-well devices offer comparable RF performance. 6.1 Equivalent Circuit Model The small-signal equivalent-circuit modeling is an extension of existing, widely-used models with additional elements added to account for the constrained p-well. The number of additional elements was minimized to maintain reasonable accuracy, simplicity, and direct correspondence between the model and the physical device. This correspondence is illustrated in the cross-sectional view of the p-well MESFET shown in Fig. 6.1. In this cross-sectional view, the depletion regions between the n +- drain -to -p -well and n-channel-to-p- well is shown conceptually along with the associated equivalent circuit elements. The greatest reduction in the drain current transients, frequency dispersion in gds, and sidegating are realized when the p- well is doped sufficiently high so that the p-region under the channel is never completely depleted. This corresponds to a conductive p-type region constrained to the source potential via the 84 PWell LEC SI GaAs Substrate Cross-sectional view of a p-well MESFET conceptually Figure 6.1. illustrating the depletion regions associated with the pn junctions and the conductive path to the source of the p-type region under the channel. p' contact and which is modeled with the resistor R. The capacitances between the p-well and n"-drain and n-type channel regions are conveniently lumped into a single capacitor Cp. For the complete equivalent circuit model (Fig. 6.2), it is shown that C, and Rp in series form a branch in parallel with the Rd, and Cd, branches between the drain and source [4]. It is interesting to note that an RC branch is also commonly used to empirically model the lowfrequency g d, frequency dependence. The main difference here is that the RC branch in Fig. 6.2 is also in parallel with Rsc and Rdc, the parasitic source- and drain-channel resistances. Another difference 85 Small-signal equivalent-circuit model used for the pFigure 6.2. well GaAs MESFET. is that Rp and C, are process dependent because their values are controlled by implant doses. In this equivalent circuit model, Rp is determined by the implant dose and energy while C, is a junction capacitance which is easily calculated. By contrast, in modeling the frequency dependence of gds with an RC branch for a conventional MESFET, the values of R and C depend on the details of the defect compensation in the substrate material. This dependence on substrate defects makes it difficult to predict the circuit element values initially, and to account for process variations in their values. 6.2 S-Parameters This expression for the output impedance of the p-well MESFET is obtained from the equivalent circuit of Fig. 6.2. Since Rds and Rp 86 P-WELL DOPING DEPENDENCE OF S21 20 0,2,4 X 10 16 cm -3 1166, 5 16 x 1016 1111 0 10-1 I C r11 I 3 114 8 x 1016cm -3 PI I 100 102 101 FREQUENCY (GHz) S21 responses for the p-well MESFETs with different nFigure 6.3. The sweeps are for bias voltages channel and p-well doping levels. of Vps = 3.0 V and VGs = 0.0 V. are much greater than Rs and Rd the output impedance can be approximated as the series RC branch in parallel with Rds. This introduces a pole-zero pair in the frequency response of the output impedance of the p-well MESFET given by, 1 + sR PC P Zds S (6.1) = Rds 1 + sCp(Rds + Rp) At low frequencies, the magnitude of Zds = Rds and starts to decrease at the pole frequency (1/27(Cp(Rds + Rp)). Zds begins to level off at the zero frequency, (1/2ffRpCp) and is just the parallel combination of 87 0,2,4 X 10 -3 16 cnn S22 responses for the p-well MESFET's with different Figure 6.4. The sweeps are for bias voltages of %fps = 3.0 V p-well doping levels. and VGs = 0.0 V. Rd, and Rp at high frequencies. In some cases the effect of this doublet can be observed in the S21 frequency response of the MESFET as shown in Fig. 6.3. For the MESFETs with a peak p-well doping of 8.0 and 16.0 x 1016 cm-3, the zero is clearly observable while the pole is at a lower frequency than could be measured accurately with the network analyzer. The effects of Rp and C, are also evident when S22 is plotted on a Smith chart in Fig. 6.4. As in Fig. 6.3, there is little effect for peak p-well dopings less than 8.0 x 1016 cm-3. For peak p-well dopings of 8.0 and 16.0 x 1016 cm-3 however, there is a considerable shift in S22 consistent with a decrease in the output 88 0.95 Cgs 0.90 0.85 0.80 0.75 0.10 0 ili 0.05 Cgd 0.00 0 2 4 6 1 I I 8 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) C gs and C gd as a function of p-well doping levels for Figure 6.5. bias voltages of Vos = 3.0 V and VGs = 0.0 V. impedance due to the parallel combination of Rds and Rp at high frequencies. The magnitudes of Rp and Cp depend primarily on the p-well doping. C, is also strongly dependent on the voltage applied to the drain while the applied voltages have only a minor effect on Rp. The p-well doping may also effect the relative magnitudes of some of the other parameter values. An evaluation of the equivalent circuit parameter values provides the information necessary for optimization of the p-well technology so that both satisfactory high frequency performance is achieved while maintaining the aforementioned benefits of junction isolation. 89 100 (1) E 800 < 0 60 400 0 cn 0 0 A A 0 0 O V V 0 V V V V V GS GS GS GS =0.5 V =0.0 V =-0.5 V =-1.0 V 20 F- 0 0 2 4 6 8 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) Small-signal gm values versus p-well doping levels at Figure 6.6. several different VGs bias voltages. lips = 3.0 V. Small-Signal Parameters 6.3 High frequency operation is most often characterized by fT, the frequency of unity-short circuit current gain. In general, fT can be expressed in terms of the small-signal parameter values as, gm (6.2) fT 27b.f(Cgs2 + 2CgsCgd) where, Cgs is the gate-source capacitance, C9d is the gate-drain feedback capacitance and gm is the small-signal transconductance. Cgs and C9d are shown as a function of peak p-well doping in Fig. 6.5. Because the pn junction capacitances are between the undepleted pwell and the n-channel and n' drain, no significant increase in Cgs or C9d is observed with doping. Cgs does increase slightly as the well 90 500 400 A ,O, A DI e--% C 3000 ..... A Q 1.0 V VGS VAS = -0.5V 3 0 03 Z 200 3 VGS = 0.0 V VGS = 0.5 V 100 0 ' 0 2 I I I I , I 4 6 8 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) Small-signal Rds values versus p-well doping levels at Figure 6.7. several different VGs bias voltages. Vps = 3.0 V. doping increases. However, this is due to the combined effects of slightly different channel profiles and different amounts of gate recess used to maintain equal Vp's for the different wafers. In Fig. 6.6, the RF gm is shown for different gate voltages as a function of p-well doping. It is interesting to note that as the p- well doping increases, gm actually increases slightly. This increased gm effect is beneficial to fT by acting to partially counter the effects of the increasing Cgs. Another parameter used for characterizing high frequency operation is fmax fmax is the frequency at which the maximum available gain becomes one and therefore, should have similar dependence on the equivalent-circuit element values. commonly used expression for MAG is given as [49]; A 91 DRAIN TO PWELL CAPACITANCE 2.5 O 2.0 A =lV V DS 2 V VDS V = 3 V DS 4 V VDS 4- 1.5 0 1 .0 V 0.5 8 a_ O V DS = 5 V 0.0 0 2 4 8 6 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) Small-signal C. values versus p-well doping levels at Figure 6.8. several different Vps bias voftages. VGS = 0.0 V. 1 (fr)2 (6.3) MAG(f)f (4/Zds+47rfTCgd)(Rg+Ri+R.+27rfTL.)+4fffTCgdRg where Zds is the expression given in (6.1) and replaces Rd, in the usually expression. For low p-well doping levels, Rds gradually increases as the p-well doping increases (Fig. 6.7) which is consistent with the tighter confinement of the channel carriers provided by the p-well. The apparent decrease in Rds at a well doping of 16.0 x 1016 cm-3 is apparently a limitation of the accuracy of the data. At this well doping, the pole is at such a low frequency that 92 105 104 a %Of 0C: 103 102 0 2 4 6 8 10 14 12 16 PEAK PWELL DOPING (1 016 cm-3) Small-signal Rp values versus p-well doping levels for Figure 6.9 bias voltages of Vps = 3.0 V and VGs = 0.0 V. it cannot be resolved using the network analyzer, and consequently the value of Rds cannot be accurately determined. The variation of the elements RP and Cp with well doping are shown in Figs. 6.8 and 6.9. In Fig. 6.8, Cp is shown for different Vps bias values. The voltage dependence of the junction capacitance is visible in the samples with the higher well dopings. In particular, the sample with 8.0 x 1016 cm-3 well doping shows a strong voltage dependence. For this sample, the depletion region has most likely reached through the p-well into the undoped substrate. R P in Fig. 6.9 decreases as the well doping increases. As expected, The values for Rp and Cp at the lower dopings should be taken as good 93 Table 6.1 Equivalent-circuit small-signal element values for the different p-well doping levels. The values are for MESFETs with lgm gate lengths and 300 gm gate widths and bias voltages of kips = 3.0 V and VGs = 0.0 V. EQUIV. CKT. PRAM. PEAK P-WELL DOPING LEVEL ( X 1016 cm-3) 0.0 2.0 4.0 8.0 gm (mS) 57.56 59.45 60.56 63.01 65.93 Tt (pS) 1.64 1.89 1.94 1.62 0.84 Cgs (fF) 772.12 797.45 818.91 871.30 887.28 Cgd (fF) 84.65 81.06 74.84 80.03 76.31 Cds (fF) 83.75 86.25 89.65 99.65 132.29 Cp (fF) 85.66 319.71 377.72 512.30 2356.42 Rd, (0) 280.76 298.64 301.72 190.45 109.92 Rp (0) 11.0 2.66 2.15 0.52 0.22 2.22 1.72 1.52 0.80 Rip (Q) Rgs (0) 2.55 101 101 100 100 16.0 104 Rsc (Q) 1.16 2.36 2.67 2.94 4.00 Rdc (Q) 0.75 0.63 0.01 1.13 5.19 Rs (Q) 0.01 0.02 0.02 0.03 0.02 Rd (Q) 0.01 0.00 0.05 0.13 0.25 Rg (Q) 0.01 0.01 0.01 0.11 0.01 Ls (pH) 1.25 1.03 0.94 0.96 0.71 Ld (pH) 1.10 1.09 0.97 1.03 0.92 Lg (pH) 26.25 16.06 14.23 18.78 14.85 94 50 40 m -o 30 .... 20 N = 10 0 109 1010 1011 FREQUENCY (GHz) 1H211 versus frequency for a conventional MESFET with Figure 6.10. bias voltages of lips = 3.0 V and VGs = 0.0 V. approximations which result in a satisfactory fit between the modeled and measured S-parameters. However, the trends do indicate the degree of control over these element values available with this technology. As indicated by the equation for MAG, other small-signal parameters are also important for high frequency performance. These small-signal parameter values are summarized in Table 6.1 for the five wafers tested. 6.4 fT, fm, and MAG The definition of fT given above is just the parameter values which result in 1H211 = 1 for the equivalent circuit model of the 95 14 13 10 9 2 0 4 6 8 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) Small-signal fT values versus p-well doping levels for Figure 6.11. bias voltages of lips = 3.0 V and VGs = 0.0 V. MESFET. However, H21 can be calculated directly from the S-parameters and fT determined for the frequency at which H21 = 1 as shown in Fig. 6.10. The slope of the line which has been drawn through H21 is fixed at 6 dB/octave and fT is determined from frequency at which the magnitude of H21 = 1. shown in Fig. 6.11. The dependence of fT on the well doping is As expected, based on variations in the small- signal parameter values, variation in fT is small. In fact, fT decreases less than 10 percent for the most heavily doped p-well sample as compared to the control sample. The maximum available gain of the MESFET was also calculated from the S-parameters at 10 GHz and is shown in Fig. 6.12 for 96 25 CONTROL WAFER CD 0 20- © o < 15 1 0 2 I I I I I I 4 6 8 10 12 14 16 PEAK PWELL DOPING (1016 cm-3) Maximum Available Gain (MAG) values versus p-well doping Figure 6.12. levels for bias voltages of %fps = 3.0 V and VGS = 0.0 V. different peak p-well dopings. A much stronger dependence on the well doping is apparent for MAG than for fr. This is expected based on the strong dependence of MAG on the output characteristics of the MESFET. However, the only sample with a marked decrease in the MAG is the sample with the highest p-type doping behind the channel as shown in Fig. 6.11. The sample with a well doping of 8.0 x 1016 cm-3 shows only a small decrease in MAG. fmax is the frequency at which MAG becomes one. extrapolating MAG to the frequency at which MAG = 1. fr. is obtained by Since finx is extracted from MAG it is expected to exhibit much the same dependence on the p-well doping as MAG. This is confirmed in Fig. 6.13 which 97 40 .0" N 0 x 30 4E 20 0 2 4 6 8 10 12 16 14 PEAK PWELL DOPING (1016 cm-3) Small-signal fr. values versus p-well doping levels for Figure 6.13. bias voltages of lips = 3.0 V and VGs = 0.0 V. shows that the samples with the more heavily doped p-well are adversely affected. 6.5 Summary The RF characteristics of a p-well GaAs MESFET technology have been presented and are shown to be comparable to conventional MESFETs. In spite of the heavily doped p-type region behind the channel and the large increase in capacitances on the drain, there is only minimal degradation in the performance of the devices. Furthermore since the additional pole and zero in the frequency response of the output characteristics are the result of an implanted pn junction capacitance 98 and a p-region conductance, the positions of the pole and zero can be set by appropriate modification of the implant parameters. 99 7. LIMITATIONS AND TRADEOFFS The results presented thus far offer compelling evidence that the p-well MESFET technology is superior to the conventional MESFET technology in many resects. However, as with many technological advances there are tradeoffs which must be considered. In this chapter some of the tradeoffs mentioned in the previous chapters are reviewed and some limitations which have not yet been examined are discussed. 7.1 Introduction Perhaps the most obvious concern which comes to mind is the potential for additional capacitance due to the undepleted p-type well. As was shown in chapter 6 the well doping can be adjusted to minimize parasitic capacitances while maintaining electrical control of the channel-substrate junction. Further reductions can be achieved by modifying the drain and source n+ implants. By lowering the implant energy, thereby decreasing the implant range and tail, it should be possible to maintain electrical contact with the channelsubstrate junction for lower peak p-type implant levels than was possible with the devices fabricated in this work. Another draw back examined in chapter 5 concerns the need to avoid independently biased p-type and n-type regions with semiinsulating GaAs in between. Forward biasing of the resulting p-i-n diode yields unacceptably large leakage currents. In addition to these concerns, there are other potential problems which deserve scrutiny. The effects examined in this 100 IDS CmA) 25. 00 2. 500 /div a) VOS . 5000/div ( V) VOS . 5000/d i v ( V) 5. 000 IDS CmA) 25. 00 2. 500 /div b) . 0000 5. 000 Drain I-V characteristics for the p-well MESFET with the Figure 7.1 p-well tied to the source a) and the p-well tied to the drain b). The maximum VGS = 0.75 V and the step size is 0.25 V. 101 IDS C A) 1E-01 decade /d iv 1E-09 -4.800 VGS .6000/div 0 .6000 C V) Subthreshold characteristics for the p-well MESFET with Figure 7.2 The upper the p-well extending outside the channel under the gate. curve is lips = 5.0 V and the step size is 2.0 V. chapter are illustrated with DC data, and include drain I-V as well as gate I-V and transconductance data. 7.2 Drain Characteristics Perhaps the most obvious concern with the p-well technology examined in this study is the need to insure that the drain and source do not become interchanged. If the source electrode becomes more positive than the drain then the drain-p-well pn junction will forward bias. This is illustrated in Fig. 7.1 where, Fig. 7.1 a) shows the normal drain I-V curves and Fig. 7.1 b) shows the result of switching the drain and source. For larger drain voltages the forward biased pn junction current dominates the channel current. 102 For operation in the linear regime and small drain bias in saturation this isn't a significant problem since the parasitic pn diode current is negligible compared to the channel current. Another important consideration is the subthreshold current, or the channel current when the MESFET is turned off. The subthreshold current for one of the p-well MESFET configurations is shown in Fig. 7.2. This device is completely enclosed in a p-type tub and shows good subthreshold characteristics with several orders of magnitude of exponential decay of the drain current before the gate-drain leakage However, as pointed out in section current begins to dominate. 2.2.2, the MESFET gate extends beyond the n-channel over the surrounding GaAs. For the results of Fig. 7.2, there is gate metal in contact with both n- and p-type GaAs. Therefore, when the gate voltage reverse biases the Schottky to n-channel region it is forward biasing the p-type region. This can result in large gate leakage currents which are discussed in the following section. One possible solution is to pull the p-type implant inside the n-type channel implant as shown in Fig. 2.5. The subthreshold characteristics are summarized in Fig. 7.3 for p-type implants coincident with the nchannel and pulled in one micron. In addition, the RF MESFETs which had the p-type implants pulled in 1/2 micron and are also shown. As the p-well is pulled inside the n-channel along the edges, the subthreshold characteristics steadily degrade. This degradation is simply related to the lack of any p-type implants along the outer edges of the channel, resulting in portions of the channel with a more negative pinch-off voltage than the central portion. 103 IDS ( A) 01 1 decode /d1v a) 1E-09 - 4. 800 0 .8000 VGS .8000/d1v ( V) VGS .8000/d1v ( V) VGS .8000/d1v ( V) IOS C 1E-01 decode /d1v b) 1E-09 - 4.800 O .8000 IOs ( 111 1E-01 decode /d1v c) 1E-09 O .8000 - 4.800 Subthreshold characteristics for the p-well MESFET with Figure 7.3 the p-well a) coincident with, b) pulled in 1/2 micron from, and c) pulled in one micron from the edge of the channel under the gate. The upper curve is Vps = 5.0 V and the step size is 2.0 V. 104 IG ) 1E-02. decode /dIv a) 1E -09- -9. 900 .0000 VGS 1.100/div V) VGS 1.100/dIv V) VGS 1.100/dIv IG ) 1E-02 decode /dIv b 1E-09 -9.900 .0000 IG CA ) 1E-02 decode /div c) 1E-09 -9.900 .0000 C V) Gate I-V characteristics for the p-well MESFET with the Figure 7.4 p-well a) pulled out one micron from, b) coincident with, and c) pulled in one micron from the edge of the channel under the gate. VDS = 2.5 V. 105 GM CMS) 15.00 E*00 1.500 /d1v a) 0000 -2. 450 VGS 0 .3500/d1v 0 . VGS .3500/d1v C V) . 7000 C V) GM CMS) 15.00 E+00 1.500 /d1v 0000 -2.450 7000 GM CMS) 15.00 E+00 1.500 /d1v CJ 0000 -2.450 VGS 0 .3500/d1v . 7000 V) Transconductance for the p-well MESFET with the p-well Figure 7.5 a) pulled out one micron from, b) coincident with, and c) pulled in Vos = 2.5 V. one micron from the edge of the channel under the gate. 106 7.3 Gate Characteristics As mentioned in the previous section, gate metal lying on the p-type GaAs can result in high leakage currents (Fig. 7.4). For the MESFET with the p-type region extending outside the n-channel the reverse gate leakage is very high. It should be pointed out that this is gate-source current via the p-well contact at the source and not gate-drain current which is very small as illustrated by the good subthreshold characteristics of Fig. 7.2. As the p-type material is pulled inside of the n-channel, the reverse leakage improves considerably. One other gate characteristic of importance is the transconductance which is shown in Fig. 7.5 for the three different p-well MESFET structures. The sample with the p-well extending outside the n-channel has a 'bump' in its shape near VGs=0.0 V. This is the result of the forward biased gate to p-type material providing a small amount of channel modulation through the pn channel-p-well junction. When the gate-channel Schottky barrier is reversed biased the gate-p-well Schottky barrier is forward biased. This allows the gate to reverse bias the channel from below the channel. While most of this action is shunted by the p-well contact to the source it does provide some parasitic gating of the p-well MESFET. As the p-well implant is pulled inside the channel the effect disappears. 7.4 Summary In addition to the tradeoffs illustrated in chapter 6 there are several others that must be considered in the p-well MESFET technology. These are primarily associated with the need to prevent 107 the gate Schottky metal from contacting the well. Pulling the p-well inside the channel along the edges as illustrated in Fig. 2.5 helps to reduce high gate leakage currents and parasitic gating of the channel by the p-well. However, it leaves a portion of the channel with a more negative pinch-off voltage which leads to poor subthreshold characteristics. 108 8. CONCLUSIONS Various anomalous effects in conventional GaAs MESFETs have been Failure shown to arise from the floating channel-substrate junction. to confine the potential of this junction allows variations in the charge state of defects in the substrate to modulate the behavior of the MESFET. This has been shown to lead to drain current transients with long time constants, frequency-dependent output conductance, and sidegating in GaAs MESFETs. Understanding the mechanisms responsible for these anomalous effects has motivated the development of a p-well MESFET technology which electrically confines the channel-substrate junction potential to the source potential. The p-well MESFET eliminates the drain current transients with long time constants and offers promise of a technology that is better suited for the design of precision analog and digital circuits in GaAs. Significant reduction in the settling time of amplifiers and switches, reduction in hysteresis in differential circuits and comparators, and reduction in frequency- and pattern-dependent propagation delay times are just some of the potential benefits possible with these improved device characteristics. That the p-well technology also demonstrated immunity to sidegate voltages in excess of -20 V with no decrease in the drain current even under illumination, is further evidence that control of the channel-substrate interface is of the utmost importance in GaAs technology. In spite of the heavily doped p-type region behind the channel and the large increase in capacitances on the drain, there is only minimal degradation in the RF performance of the devices. 109 Furthermore, since the additional pole and zero in the frequency response of the output characteristics are the result of an implanted pn junction capacitance and a p region conductance, the positions of the pole and zero can be set by appropriate modification to the implant parameters. The results presented in this thesis open up broad avenues of potential research opportunities. One of the more obvious would be the implementation of analog and digital building blocks to begin quantifying improvements in circuit performance. The presence of independent p-type implants and p-ohmics should allow for the development of a technology with a complementary devices as well. Complementary devices may be particularly useful in level shifting applications, replacing large stacks of diodes. In the general area of technology development, there are some potential advantages to be gained from an epitaxial based equivalent to the p-well MESFET. As demonstrated in chapter 7, care must be taken to avoid contacting the p-well with the gate metal. While this reduces the gate leakage currents and parasitic gating of the channel by the p-well, this results in poor pinch-off and subthreshold characteristics due to the channel n-type implant not being coimplanted with p-type material at the edges of the channel. The mask set used for this study is ideal for the investigation of many device phenomena which might be impacted by the material. Stress-induced piezoelectric charges, 1/f noise, gate transients, frequency-dependent transconductance, and low frequency oscillations 110 are all properties which could potentially be impacted by the presence of p-type implants and bare further scrutiny. Extension of the device modeling outlined in chapters 3 and 4 could lead to a combined large-signal, small-signal, and DC model capable of modeling frequency- and pattern-dependent delay times, provided the transient overshoot term can be better quantified. 111 BIBLIOGRAPHY 1. N. Scheinberg, R. Bayruns, and R. Goyal, "A low frequency GaAs MESFET circuit model," IEEE J. of Solid-State Circuits, vol. 23, no. 2, pp. 605-608, April 1988. 2. L.E. Larson, J.F. Jenson, H.M. Levey, P.T. Greiling and G.C. Temes, "GaAs differential amplifiers," IEEE GaAs IC Symp. Tech. Dig., pp. 19-22, 1985. 3. J.M. Golio, M.G. Miller, G.N. Maracas, and D.A. Johnson, "Frequency-dependent electrical characteristics of GaAs MESFET's," IEEE Tran. on Electron Devices, vol. ED-37, no. 5, pp. 1217-1227, May 1990. 4. C. Camacho-Penalosa and C.S. Aitchison, "Modeling frequency dependence of output impedance of a microwave MESFET at low frequencies," Electon. Lett., vol. 21, no. 12, pp.528-529, June 6th 1985. 5. P.C. Canfield, S.C.F. Lam, and D.J. Allstot, "Modeling of frequency and temperature effects in GaAs MESFETs." IEEE J. of Solid-State Circuits, vol. 25, no. 1, pp. 299-306, Feb. 1990. 6. P.H. Ladbrooke and S.R. Blight, "Low-field low-frequency dispersion of transconductance in GaAs MESFET's with implications for other rate-dependent anomalies," IEEE Trans. on Electron Devices, vol. ED-35, no. 3, pp. 257-267, Mar. 1988. 112 7. J. Graffeuil, Z. Hadjoub, J.P. Fortea, and M. Pouysegur, "Analysis of capacitance and transconductance frequency dispersions in MESFETs for surface characterization," SolidState Electon., vol. 29, no. 10, pp. 1087-1097, Oct. 1986. 8. S. Barton, "Characterization of high-speed (above 500 MHz) devices using advanced ATE techniques, results, and device problems," IEEE International Test Conf. Tech. Dig., 1989. 9. T. Itoh and H. Yanai, " Long-term drift of GaAs MESFET characteristics and its dependence on substrate with buffer layer," Inst. Phys. Conf. Ser. 45, pp. 326-334, 1979. 10. S. Makram-Ebeid and P. Minondo, "The roles of the surface and bulk of the semi-insulating substrate in low-frequency anomalies of GaAs integrated circuits," IEEE Tran. Electron Devices, vol. ED-32, no. 3, pp. 632-642, Mar. 1985. 11. T.W. Hickmott, "Temperature dependence of FET properties for Cr-doped and LEC semi-insulating GaAs Substrates," IEEE Trans. Electron Devices, vol. ED-31, no. 1, pp. 54-62, Jan. 1984. 12. C. Kocot and C.A. Stolte, "Backgating in GaAs MESFET's," IEEE Trans. on Electron Devices, vol. ED-29, no. 7, pp. 1059-1064, July 1982. 13. R. Yeats, D.C. D'Avanzo, K. Chan, N. Fernandez, T.W. Taylor, and C. Vogel, "Gate slow transients in GaAs MESFETs causes, cures 113 and impact on circuits," IEEE Inter. Electon Device Meeting Technical Digest, pp. 842-845, 1988. 14. S.R. Blight, R. H. Wallis, and H. Thomas, " Surface influence on the conductance DLTS spectra of GaAs MESFET's," IEEE Trans. on Electron Devices, vol. ED-33, no. 10, pp. 1447-1453, Oct. 1986. 15. L. G. Salmon, "GaAs substrates: Issues for VLSI," Proc. 6th Semi-Insulating III-V Materials Conf.: Toronto, to be published, 1990. 16. R.Y. Koyama, B. Odekirk, W.A. Vetanen, E.P. Finchem, and I.G. Beers, "'Parasitic' effects and their impact on gallium arsenide integrated circuits," Proc. 5th Semi-Insulating III-V Materials Conf.:Malmo, G. Grossmann and L. Ledebo; eds., Bristol: Adam Hilger, pp. 203-212, 1988. 17. C.P. Lee, S.J. Lee, and B.M. Welch, "Carrier injection and backgating effect in GaAs MESFET's," IEEE Electon Device Letters, vol. EDL-3, no. 4, pp. 97-98, Apr. 1982 18. S. Subramanian, P.K. Bhattacharya, K.J. Staker, C.L. Ghosh, and M.H. Badawi, "Geometrical and light-induced effects on backgating in ion-implanted GaAs MESFET's," IEEE Trans. on Electon Devices, vol. ED-32, no. 1, pp. 28-33, Jan. 1985. 19. C.P. Lee, "Influence of substrates on the electrical properties of GaAs FET devices and integrated circuits," Proc. 2nd Semi- 114 Insulating III-V Materials Conf.:Evian, S. Makram-Ebeid and B. Tuck; eds., Natwich, UK: Shiva Pub., pp. 324-335, 1982. 20. M.S. Birrittella, W.C. Seelbach, and H. Goronkin, "The effect of backgating on the design and performance of GaAs digital integrated circuits," IEEE Trans. Electron Devices, vol. ED-29, no. 7, pp. 1135-1142, July 1982. 21. H. Goronkin, "Sidegating and backgating in GaAs ICs," Proc. IEEE/Cornell Conf. on High-Speed Semiconductor Devices and Circuits, pp. 26-37, 1983. 22. D.C. D'Avanzo, "Proton isolation for GaAs integrated circuits," IEEE Trans. Electron Devices, vol. ED-29, no. 7, pp. 1051-1059, July 1982. 23. E.P. Finchem, W.A. Vetanen, B. Odekirk, and P.C. Canfield, "Reduction of the backgating effect in GaAs MESFET ICs by charge trapping at the backgate electrode," IEEE GaAs IC Symp. Tech. Digest, pp. 231-234, 1988. 24. K. Onodera, "A novel GaAs MESFET for suppression of low temperature sidegating," IEEE GaAs IC Symp. Tech. Digest, pp. 215-218, 1989. 25. A.S. Blum and L.D. Flesner, "Use of a surrounding p-type ring to decrease backgate biasing in GaAs MESFET's," IEEE Electon Device Letters, vol. EDL-6, no. 2, pp. 97-99, Feb. 1985. 115 26. C.P. Lee and M.F. Chang, "Shielding of backgating effects in GaAs integrated circuits," IEEE Electon Device Letters, vol. EDL-6, no. 4, pp. 169-171, Feb. 1985. 27. L.E. Larson, "An improved GaAs MESFET equivalent circuit model for analog integrated circuit applications," IEEE J. Solid-State Circuits, vol. 22, no. 4, pp. 567-574, Aug. 1987. 28. L.E. Larson, C.S. Chou, and M.J. Delany, "An ultrahigh-speed GaAs MESFET operational amplifier," IEEE J. Solid-State Circuits, vol. 24, pp. 1523-1528, Dec. 1989. 29. G.M. Martin, J.P. Farges, G. Jacob, and J.P. Hallais, "Compensation mechanisms in GaAs," J. Appl. Phys., vol. 51, no. 5, pp. 2840-2852, May 1980. 30. D.E. Holmes, R.T. Chen, K.R. Elliott, C.G. Kirkpatrick, and P.W. Yu, "Compensation mechanism in liquid encapsulated Czochralski GaAs: importance of melt stoichiometry," IEEE Trans. Electron Devices, vol. 29, no. 7, pp. 1045-1051, Jul. 1982. 31. P.J. Pearah, R. Tobin, J.P. Tower, and R.M. Ware, "The role of carbon in the compensation of semi-insulating LEC GaAs," 5th Semi-Insulating III-V Materials Conf.:Malmo, G. Grossmann and L. Ledebo; eds., Bristol: Adam Hilger, pp. 195-200, 1988. 32. R.K. Boncek and D.L. Rode, "Characterization of residual carbon in semi-insulating GaAs," J. Appl. Phys., vol. 64, no. 11, pp. 6315-6321, Dec. 1988. 116 33. C.G. Kirkpatrick, R.T. Chen, D.E. Holmes, and K.R. Elliott, "Growth of Bulk GaAs," Gallium Arsenide Materials, Devices, and Circuits, M.J. Howes and D.V. Morgan; eds., New York: John Wiley and Sons, pp.39-117, 1986. 34. J. Lagowski, J.M. Parsey, M. Kaminska, K. Wada, and H.C. Gatos, "On the behaviour and origin of the major deep level (EL2) in GaAs," Proc. 2nd Semi-Insulating III-V Materials Conf.:Evian, S. Makram-Ebeid and B. Tuck; eds., Natwich, UK: Shiva Pub., pp. 154-159, 1982. 35. R.W. Tang, L. Sargent, and J.S. Blakemore, "Compensation assessment in 'undoped' high-resistivity GaAs," J. Appl. Phys., vol. 66, no. 1, pp. 256-261, Jul. 1989. 36. Y. Nakamura, Y. Ohtsuki, Y. Itoh, J. Kikawa, and Y. Kashiwayanagi, "Role of carbon in LEC SI-GaAs," Inst. Phys. Conf. Ser. No. 79, pp. 49-54, 1985. 37. J.F. Wager and A.J. McCamant, "GaAs MESFET interface Considerations," IEEE Trans. Electron Devices, vol. 34, no. 5, pp. 1001-1007, May 1987. 38. R. Anholt and T.W. Sigmon, "Substrate-impurities effects on GaAs MESFETs," J. of Electronic Materials, vol. 17, no. 1, pp. 5-10, Jan. 1988. 39. W.B. Leigh, J.S. Blakemore, and R.Y. Koyama, "Interfacial effects related to backgating in ion-implanted GaAs MESFET's," 117 IEEE Trans. Electron Devices, vol. 32, no. 9, pp. 1835-1841, Sept. 1985. 40. J.S. Blakemore, "Use of detailed balance with known attributes of GaAs midgap levels, in deriving new insights into their energies and properties," Proc. 4th Semi-Insulating III-V Materials Conf.:Hakone, H. Kukimoto and S. Miyazawa; eds., Tokoyo: Ohmsha, Ltd., pp. 389-394, 1986. 41. J. Lagowski, H.C. Gatos, J.M. Parsey, K. Wada, M. Kaminska, and W. Walukiewicz, "Origin of the 0.82 eV electron trap in GaAs and its annihilation by shallow donors," Appl. Phys. Lett., vol. 40, no. 4, pp. 342-344, 1982. 42. D.C. Look, "Annealing and thermal cycling effects in semiinsulator GaAs," 5th Semi-Insulating III-V Materials Conf.:Malmo, G. Grossmann and L. Ledebo; eds., Bristol: Adam Hilger, pp. 1-10, 1988. 43. W.R. Curtice, "A MESFET model for use in the design of GaAs integrated circuits," IEEE Trans. on Microwave Theory and Techniques, vol. 28, no. 5, pp. 448-456, May 1980. 44. W. Mickanin, P.C. Canfield, E.P. Finchem, and B. Odekirk, "Frequency-dependent transients in GaAs MESFETS: Process, geometery, and material effects," IEEE GaAs IC Symposium Tech. Dig., pp. 211-214, 1989. 118 45. P.C. Canfield, L. Forbes, R. Gleason, and A. McCamant, "Drain current transient suppression in buried channel GaAs MESFETs," Proc. 4th Semi-Insulating III-V Materials Conf.:Hakone, H. Kukimoto and S. Miyazawa; eds., Tokoyo: Ohmsha, Ltd., pp. 392-396, 1986. 46. S. Choi and M.B. Das, "Determination of and significance of frequency dependent output conductance in ion-implanted microwave MESFETs on SI GaAs," Proc. 6th Semi-Insulating III-V Materials Conf.:Toronto, to be published, 1990. 47. D. Gitlin, C.R, Viswanathan and A.A. Abidi, "Output impedance frequency dispersion and low frequency noise in GaAs MESFETs," Proc. 18th European Solid-State Device Research Conf.: Montpellier, 1988. 48. M.A. Lampert, "Injection currents in insulators," IEEE Proceedings, vol. 50, no. 8, pp. 1781-1796, Aug. 1962. 49. P.H. Ladbrooke, MMIC Desigin: GaAs FETs and HEMTs. Boston: Artech House, pp. 256, 1989. APPENDIX 119 A. PUBLICATIONS The work contained in this thesis has resulted directly in a number of publications. Additionally my graduate studies have resulted in several publications which are only peripherally related to the central topic of this thesis; many of them arising from collaboration with other students and professionals in industry. They are compiled below for those interested in a full accounting of my contribution to the general body of GaAs technology knowledge during my years in graduate school. A.1 Journal Publications 1. P. Canfield and L. Forbes, "Gate bias dependent low frequency oscillations in GaAs MISFETs," IEEE Electron Device Lett., vol. EDL-6, no. 5, pp. 227-228, May, 1985. 2. P. Canfield and L. Forbes, "Suppression of drain conductance transients, drain current oscillations and low-frequency generation-recombination noise in GaAs FET's using buried channels," IEEE Trans. on Electron Devices, vol. ED-33, no. 7, pp. 925-928, July, 1986. 3. P. Canfield, J. Medinger, and L. Forbes, "Buried channel GaAs MESFETs with frequency independent output conductance," IEEE Electron Device Lett., vol. EDL-8, no.3, pp. 88-89, Mar., 1987. 120 4. P. Canfield and L. Forbes, "Buried channel GaAs MESFETs with immunity to ionizing optical radiation effects," IEEE Electron Device Lett., vol. EDL-8, no.3, pp. 113-115, Mar., 1987. 5. P. Canfield and L. Forbes, "Lateral n-p-n bipolar transistors by ion-implantation into semi-insulating GaAs," Solid State Electronics, vol. 31, no. 1, pp. 123-125, Jan., 1988. 6. H.C. Yang, P.C. Canfield, and D.J. Allstot, "A one transistor GaAs voltage reference circuit," Electronics Letters, vol. 25, no. 7, pp. 464-465, 1989. 7. P.C. Canfield, S.C.F. Lam, and D.J. Allstot, "Modeling of frequency and temperature effects in GaAs MESFETs," IEEE Journal of Solid-State Circuits, vol. 25, no. 1, pp. 299-306, Feb., 1990. 8. W. MicKanin, P.C. Canfield, E.P. Finchem, and B. Odekirk, "Transients in GaAs MESFETS: Process, geometry, and material effects," EURO III-Vs Review, vol. 3, no. 1, pp. 22-25, Feb., 1990. 9. P.C. Canfield and D.J. Allstot, "A p-well GaAs MESFET technology for mixed-mode applications," IEEE Journal of Solid-State Circuits, accepted for publication, vol. 25, no. 6, Dec. 1990. 10. L.L. Peng, P.C. Canfield, J.R. Arthur, and D.J. Allstot, "Trap effects on p-channel GaAs MESFETs: A temperature dependent drain 121 current transient study," Submitted to the IEEE Trans. Electron Devices. 11. P.C. Canfield and D.J. Allstot, "A p-well MESFET Technology for multifunction MMIC's," Submitted to the IEEE Trans. Electron Devices. 12. P.C. Canfield and D.J. Allstot, "Sidegating in GaAs MESFETS with p-type implants and its elimination using a p-well technology," IEEE Trans. Electron Devices, to be submitted. 13. P.C. Canfield and D.J. Allstot, "Substrate contribution to drain current transients and frequency-dependent output conductance on GaAs MESFETs," IEEE Trans. Electron Devices, to be submitted. 14. P.C. Canfield and D.J. Allstot, "The impact of p-type implants on the drain current transients and output conductance in GaAs MESFETs," IEEE Trans. Electron Devices, to be submitted. A.2 Conference Publications 1. L. Forbes, P. Canfield, R. Gleason and A. McCamant, "Lowfrequency noise in GaAs FETs," Proc. 3rd Semi-Insulting III-V Materials Conf.: Kah-Nee-Ta, D. C. Look and J. S. Blakemore; eds., Natwich, UK: Shiva Pub., pp. 392-396, 1984. 2. L. Forbes, P. Canfield, R. Gleason, and A. McCamant, "Seperation of generation-recombination noise and 1/f noise components on GaAs FETs," Abstracts of the 1984 Device Research Conf., IEEE 122 Trans. Electron Devices, vol. ED-31, no. 12, pp. 1986, Dec., 1984. 3. P. Canfield, L. Forbes, R. Gleason, and A. McCamant, "Drain current transient suppression in buried channel GaAs MESFETs," Proc. 4th Semi-Insulating III-V Materials Conf.: Hakone, H. Kukimoto and S. Miyazawa; eds., Tokyo: Ohmsha, LTD. and Amsterdam: North-Holland Publ. Co., pp. 573-578, 1986. 4. P. Canfield, L. Forbes, R. Gleason, and A. McCamant, "Suppression of deep level trapping related effects in GaAs MESFETs using a buried channel structure," Abstracts of the 1986 Device Research Conf., IEEE Trans. Electron Devices, vol. ED-33, no. 11, pp. 5. 1851, Nov., 1986. L. Forbes and P. Canfield, "Suppression of deep level trapping related effects in GaAs MESFETs using a buried channel structure," GaAs REL WORKSHOP abstracts, paper 1-2, 1986. 6. P.C. Canfield, J. Medinger, D.J. Allstot, L. Forbes, A.J. McCamant, W.A. Vetanen, B. Odekirk, E.P. Finchem, and K.R. Gleason, "High speed quarter micron buried channel MESFETs with improved output characteristics for analog applications," IEEE/Cornell Conference Proceedings, pp. 247-254, 1987. 7. P. C. Canfield, D. J. Allstot, J. Medinger, L. Forbes, A. J. McCamant, W. A. Vetanen, B. Odekirk, E.P. Finchem, and K. R. Gleason, "Buried channel GaAs MESFETs with improved small-signal 123 characteristics," IEEE GaAs IC Symposium Technical Digest, pp. 163-166, 1987. 8. H.C. Yang, P.C. Canfield, and D.J. Allstot, "GaAs buried channel MESFET analog integrated circuits," Proc. of the IEEE Inter. Symp. on Circuits and Systems, pp. 1607-1610, 1988. 9. E.P. Finchem, W.A. Vetanen, B. Odekirk, and P.C. Canfield, "Reduction of the backgating effect in GaAs MESFET ICs by charge trapping at the backgate electrode," IEEE GaAs IC Symposium Technical Digest, pp. 231-234, 1988. 10. S.C.F. Lam, P.C. Canfield, A.J. McCamant, and D.J. Allstot, "Analytical model of GaAs MESFET output conductance," IEEE GaAs IC Symposium Technical Digest, pp. 203-206, 1988. 11. D.J. Allstot, P.C. Canfield, P.K. OR, and H.C. Yang, "A GaAs MESFET voltage reference," IEEE Inter. Electon Device Meeting Technical Digest, pp. 774-777, 1988. 12. P.C. Canfield, A.J. McCamant, and D.J. Allstot, "Gate and drain transient measurements of conventional and buried-channel MESFETs," Workshop on Instabilities in III-V Devices, Sedona, Arizona, Ap. 24-26, 1989. 13. W. Mickanin, P.C. Canfield, E.P. Finchem, and B. Odekirk, "Frequency-dependent transients in GaAs MESFETS: Process, geometry, and material effects," IEEE GaAs IC Symposium Technical Digest, pp. 211-214, 1989. 124 14. P.C. Canfield and D.J. Allstot, "A P-well GaAs MESFET technology," IEEE Inter. Solid State Circuits Conf. Technical Digest, pp. 242-243, 1990. 15. P.C. Canfield and D.J. Allstot, "The potential of p-well GaAs MESFET technology for precision integrated circuits," Proc. of the IEEE Inter. Symp. on Circuits and Systems, pp. 3065-3068, May, 1990. 16. P.C. Canfield and D.J. Allstot, "RF Characteristics of p-well GaAs MESFETs," Proc. of the IEEE MTT-S Inter. Microwave Symp. Digest, pp. 1077-1080, May, 1990. 17. P.C. Canfield and D.J. Allstot, "A p-well GaAs MESFET technology for precision integrated circuits," Proc. 6th SemiInsulting III-V Materials Conf.: Toronto, accepted for publication, May, 1990.