in 0 a! 731...»: ~ aifierhfiltog 1nd...» ; ... Eras; sififi.‘ [.11'. t? :- ill-5' E)!( It. 6' 5.1.1:- . 1. ”win: I?! )l 14... If 2 a ti}... 37-1- '9' In... I . .3: rl J t .1 p.‘ 4. .1- ... 1:1: .V’IOL schiin .r..%h:iu rt!!! 1 .5101. I . I 3111200!!! :1 n {.3321- t .‘l . .1334 a II ,‘ x 2. .2 u..th.9.x.~iL h#1.. nil?! ” V c h. .s! . 3.5.9.1.:hnh r.» .5... ..£ a 0.. .ul.. 2!», r L I o t v \t 0?. A 9| mu." I #233732? . VDYI. I- vor.1l»rll.4.VV. Liv! . 1 A .Z . Y :1 ~ -I..~\ . . . 3 .5. .x...\v\ n..J.vI1v 3... .3 . 4 .nzutif :9... )5 A r a MICHIGM STATE III (II III II III III IIII IIIIIIIIIIIIII This is to certify that the dissertation entitled A Noise Study of Bipolar Junction Transistor Reliability presented by Chih-Chieh Jack Sun has been accepted towards fulfillment of the requirements for Ph.D. degree inElectrical Engineering Dr K. KM; at Major professor Dam/”limit @1993 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY MIChigan State University PLACE iN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE l MSU Is An Affirmative Action/Equal Opportunity Institution emmnt A NOISE STUDY OF BIPOLAR JUNCTION TRANSISTOR RELIABILITY BY Chih-Chieh Jack Sun A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Electrical Engineering 1993 ABSTRACT A NOISE STUDY OF BIPOLAR JUNCTION TRANSISTOR RELIABILITY BY Chih-Chieh Jack Sun Reliability issues, including hot-electron effects and electromigration effects, are investigated for bipolar junction transistors. The low-frequency l/f noise power spectral density and the common-emitter current gain (hFE) are the principal physical observables for this study. Bipolar n-p-n transistors manufactured with three different technologies are electrically stressed by reverse biasing or forward biasing the emitter-base junction at various temperatures (-75 C to 240 C) and times up to 2500 hours. After reverse-bias stress, an increase of the noise and a decrease of the gain were observed. The decrease of gain was caused by the increase of the base current with the collector current remaining unchanged. A linear relation between the noise increase and the base current increase is predicted theoretically and observed experimentally. The observed degradation is attributed to the hot-electron induced increase of interface states and oxide trapped electrons. Forward-bias induced post-stress recovery allows partial reversal of degradation and is believed to be due primarily to a reduction of the number of oxide trapped electrons. Thermal annealing, which is capable of removing interface states as well, produces a larger recovery of both gain and noise performance measures. With the final merit of the transistor being its lifetime, the forward-bias induced recovery results show that the lifetime of a transistor depends not only on the time over which a reverse bias is applied, but also on the history of biases applied between the reverse bias periods. For forward-bias stress, a combination of high currents and high temperatures produced gain degradations or failures of transistors for all three technologies as well as changes in transistor noise characteristics. Specifically, in all transistors stressed, the low- frequency noise initially decreased without a change in gain. For longer stress times, it was observed in two technologies that the noise eventually increased again concurrent with a gradual decrease in gain. Electromigration induced compressive stress is believed to play an important role in these phenomena. The third technology eventually showed failures of the transistors due to electromigration produced open circuits. A degradation model which explains the electromigration induced noise and gain changes on forward-bias stressed transistors is presented. Copyright © by CHIH-CHIEH JACK SUN 1993 To my parents and my wife ACKNOWLEDGMENTS The author wishes to express his appreciation to his academic advisor, Professor Donnie K. Reinhard, for his guidance and continuous encouragement during the course of this work. Appreciation is also extended to Dr. Timothy A. Grotjohn and his former Ph.D. student, Dr. Chi-Jung Huang, for their generous assistance and valuable suggestions during this research. This work was partially supported by IBM Corporation, Fishkill, NY. The author wishes to thank Dr. C.-C. W. Yu (IBM Cooperation, Kingston, NY) for his suggestions and comments. The author also thanks R. S. Hemmert for providing some of the devices, and Dr. Mohammad Aslam for packaging selected devices used in this investigation. vi TABLE OF CONTENTS LIST OF TABLES ......................................................................................................... ix LIST OF FIGURES ........................................................................................................ x Chapter 1 Introduction ................................................................................................... l 1.1 Motivation ..................................................................................................... 1 1.2 Objectives ...................................................................................................... 3 1.3 Preview of Thesis .......................................................................................... 4 Chapter 2 Background ................................................................................................... 6 2.1 Low-Frequency Noise Sources in BJT's ......................................................... 6 2.1.1 Physical Origins ............................................................................... 6 2.1.2 Identification of l/f Noise Sources in BJT's ................................... 10 2.2 A Review of BJT Degradation due to Electrical and Thermal Stress ............. 16 2.2.1 Hot-Carrier Induced Degradation .................................................. 16 2.2.2 Eiectromigration Induced Degradation .......................................... 22 Chapter 3 Reverse-Bias Stress Results ......................................................................... 27 3.1 Device Description ...................................................................................... 27 3.2 Device Characterization ............................................................................... 32 3.2.1 Low-Frequency Noise Measurement Method ................................ 32 3.2.2 I-V Characterization ...................................................................... 38 3.3 DC Reverse-Bias Stress and Post-Stress Recovery ....................................... 43 3.3.1 DC Reverse-Bias Degradation and Thermal Annealing .................. 46 vii viii 3.3.2 F orward-Bias Induced Recovery ................................................... 51 3.3.3 AC Stress ...................................................................................... 59 3.4 Voltage Dependence of Reverse-Bias Stress ................................................ 64 3.4.1 Experimental Methods ................................................................... 64 3.4.2 Results .......................................................................................... 66 3.4.3 Discussion ..................................................................................... 79 Chapter 4 Forward-Bias Stress Results ........................................................................ 84 4.1 Device Description ...................................................................................... 84 4.2 Experimental Methods ................................................................................. 90 4.3 Stress Effects on Gain .................................................................................. 92 4.4 Stress and Thermal Annealing Effects on Noise ............................................ 98 4.5 Discussion ................................................................................................. 114 Chapter 5 A Forward-Bias Stress Induced Degradation Model ................................... 116 5.1 Background ............................................................................................... 116 5.1.1 Effects of Emitter-Edge Dislocations on BJT Low-Frequency Noise ........................................................................................ 116 5.1.2 Dislocation Motion Under Forces ................................................ 120 5.1.3 Piezojunction Effect .................................................................... 123 5.2 Degradation Model .................................................................................... 128 Chapter 6 Summary and Conclusions ......................................................................... 142 BIBLIOGRAPHY ....................................................................................................... 146 LIST OF TABLES Table Page 3.1 The values of A13 and ASIbe used in Figure 3.36 and the corresponding stress history. ....................................................................................................... 83 4.1 The magnitudes of the current gain vs. stress time at VBE = 0.6 V for all seven transistors of the Hemmert test structure. ................................................. 100 5.1 The normalized hFE for transistor #2, #3 and #4 of the Hemmert's test structure. ........................................................................................................... 135 5.2 The normalized hFE decrease for transistor #2, #3 and #4 of the Hemmert's test structure. ..................................................................................................... 136 5.3 The electromigration-induced compressive stress gradient along the distance from transistor #1. ................................................................................ 139 ix LIST OF FIGURES Figure Page 2.1 Possible l/f noise sources inside a n-p-n transistor biased in forward- active mode. ........................................................................................................ 11 2.2 A circuit model for the rrns 1/f noise sources measured in a BJT. ......................... 13 2.3 The transistor bias circuits for low-frequency noise measurements. ...................... 14 2.4 An electromigraion—induced metal accumulation, with a passivation layer enclosing the metal, can result in compressive stress on the emitter. ..................... 24 3.1 Top view of the TIN3 device layout. .................................................................... 28 3.2 Top view of the TOUT device layout. .................................................................. 29 3.3 Doping profile for TIN 3 and TOUT devices. ........................................................ 30 3 .4 Pin out configuration of the transistors. ................................................................ 31 3.5 A set of three measurements of the emitter noise power spectral density on a TIN3 device under different bias conditions. ................................................. 35 3 .6 Two measurements of the emitter noise power spectral density on the same TIN3 device with its noise levels as shown in Figure 3.5. ............................. 36 3 .7 The quadratic current dependence of the noise power for a TIN3 device. ............. 37 3 .8 A set of three measurements of the emitter noise power spectral density on a TOUT device under different bias conditions. ............................................... 39 3.9 Two measurements of the emitter noise power spectral density on the same TOUT device with its noise levels as shown in Figure 3.8. ........................... 40 3.10 The quadratic current dependence of the noise power for a TOUT device. ........... 41 X 3.11 3.12 3.13 3.14 3.15 3.16 3.17 3.18 3.19 3.20 3.21 3.22 3.23 3.24 3.25 3.26 xi The noise measurement flow of the emitter-voltage power spectral density. .......... 42 Circuit configuration for the I-V measurements. ................................................... 44 Diagram for the I-V measurement setup. .............................................................. 45 Average Gummel plot for a population of 12 devices at various stress times (0 to 500 hours) for devices reverse stressed with VBE = -4 V and VBC=OVat23 C. .............................................................................................. 48 hp]; vs. VBE for the data shown in Figure 3.14. .................................................... 49 km; recovery of degraded transistors shown in Figure 3.14 as a result of post-stress thermal annealing at temperatures of 150, 175, and 240 C. .................. 50 The noise power spectral density measured at various stress times ( O to 100 hours) for a TIN 3 device reverse bias stressed with VBE = -4 V and V3c=0Vat23 C. .............................................................................................. 52 The noise was reduced almost to the pre-stress level after 240 C annealing for 60 hours. ......................................................................................... 53 DC reverse stress and post-stress thermal annealing effects on the normalized hp}; and normalized low-frequency noise levels 511,9. ........................... 54 F orward-bias induced hp}; recovery under various recovery conditions. ................ 56 The test sequence for an investigation of 1/f noise recovery due to post- stress forward bias. .............................................................................................. 58 Low-frequency noise reduction at various recovery time (0 to 60 minutes) due to forward bias applied to the emitter-base junction. ........................ 6O Forward-bias induced recovery effects on normalized by; and normalized low-frequency noise levels S1,”. .......................................................... 61 AC stress results with and without forward-bias (1.0V) cycles. ............................ 63 The voltage divider circuit used to perform reverse-bias stress under various stress voltages. ......................................................................................... 65 The experimental flow for the reverse-bias stress under various voltages. ............. 67 xii 3.27 Normalized hFE at VBE = 0.6 V for TOUT devices stressed under -3.5, 4.0 and -4.5 V. ..................................................................................................... 68 3.28 The emitter noise power spectral densities afier different times of -3.5 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.78 V and VBC = -1.5 V. ...................................................................................................... 70 3.29 The emitter noise power spectral densities afier different times of -3.5 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. ...................................................................................................... 71 3.30 The emitter noise power spectral densities after different times of -3.5 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. ...................................................................................................... 72 3.31 The emitter noise power spectral densities afier different times of -4.0 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.78 V and VBC = -1.5 V. ...................................................................................................... 73 3.32 The emitter noise power spectral densities after different times of -4.0 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. ...................................................................................................... 74 3.33 The emitter noise power spectral densities after different times of -4.0 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. ...................................................................................................... 75 3.34 The emitter noise power spectral densities afier different times of -4.5 V xiii stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.78 V and VBC = -1.5 V. ...................................................................................................... 76 3.35 The emitter noise power spectral densities after different times of -4.5 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. ...................................................................................................... 77 3.36 The emitter noise power spectral densities after different times of -4.5 V stress. The noise measurements are performed under the common- collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. ...................................................................................................... 78 3.37 ASIbe versus A13 for TOUT devices stressed under -4.0 and -4.5 V. ..................... 82 4.1 A schematic of the Hemmert test structure. .......................................................... 85 4.2 A set of three measurements of the emitter noise power spectral density on a TSEO3I device under different bias conditions. ............................................. 87 4.3 The quadratic current dependence of the noise power for a TSE03I device. ................................................................................................................. 88 4.4 Two measurements of the emitter noise power spectral density on the same TSE03I device with its noise levels as shown in Figure 4.2. ......................... 89 4.5 Current gain vs. VBE for TIN3 transistors at stress times of O, l, 4, 24, 100, 200, 500 hours. The forward stress was done at 240 C with a current density of 2 mA/umz. ............................................................................... 93 4.6 Normalized current gain of TIN3 transistors vs. stress time at various current levels and ambient temperatures. .............................................................. 94 4.7 Current gain of individual TSE03I transistors vs. stress time for a stress condition of 225 C and IE = 5 mA/umz. ............................................................... 96 4.8 Gummel plots for a TSE03I device showing the onset of high resistance xiv and eventually an open circuit during stress with an emitter current density of 5 mA/um2 at 225 C. ............................................................................. 97 4.9 Normalized current gain for transistors #1, 2, 3, and 6 of the Hemmert test structure vs. stress time. ................................................................................ 99 4.10 The noise power spectral density for transistor #1 of the Hemmert test structure. ........................................................................................................... 101 4.11 The noise power spectral density for transistor #2 of the Hemmert test structure. ........................................................................................................... 102 4.12 The noise power spectral density for transistor #3 of the Hemmert test structure. ........................................................................................................... 103 4.13 The noise power spectral density for transistor #4 of the Hemmert test structure. ........................................................................................................... 104 4.14 The noise power spectral density for transistor #5 of the Hemmert test structure. ........................................................................................................... 105 4.15 The noise power spectral density for transistor #6 of the Hemmert test structure. ........................................................................................................... 106 4.16 The noise power spectral density for transistor #7 of the Hemmert test structure. ........................................................................................................... 107 4.17 Normalized base current noise power at 1 Hz vs. stress time for the seven transistors of the Hemmert test structure. ................................................. 109 4.18 The noise power spectral density of a TIN3 transistor decreases afier a forward-bias stress, reaches a minimum and subsequently increases again. ................................................................................................................. l 10 4.19 The noise power spectral density of a TIN3 device decreases afier stressing with IE = 2 mA/um2 at 240 C for 66 hours. Also shown is the post-stress thermal annealing increase of the noise levels toward its original values. The annealing was done at 240 C for 66 hours. .......................... 111 XV 4.20 The noise power spectral density of a TIN3 device decreases afier stressing with JE = 2 mA/p.m2 at two different temperatures: first, 23 C for 100 hours, then, 240 C for 66 hours. ............................................................. 112 4.21 The noise power spectral density of a TSEO3I device decreases afier stressing at 225 C with IE = 5 mA/um2 for 90 hours. The noise level increases after a post-stress annealing at 225 C for 90 hours. .............................. 113 5.1 Forward-bias stress degradation model. .............................................................. 130 5.2 Normalized base-current noise power spectral density at 1 Hz vs. stress time for the seven transistors of Hemmert's test structure designed for investigating the electromigration-induced compressive stress. ........................... 133 5.3 The forward-bias degradation model for Hemmert's pressure sensors designed for testing the electromigration effect. .................................................. 137 CHAPTER 1 INTRODUCTION 1.1 Motivation Degradation of bipolar junction transistor (BJT) parameters, such as the current gain (km) and noise, is an important reliability consideration. Modern scaled-down bipolar junction transistors are widely used in integrated circuits, such as in BiCMOS gates, because of the advantages of high speed and low noise levels compared to MOSFET's. However, long-term degradation of BJT parameters may counter these advantages. The decrease of hpE in degraded BJT's can have a number of deleterious effects including bias level shifiing and degradation of speed performance of integrated circuits. The increase of the low fiequency noise levels in degraded BJT's may reduce the signal to noise ratio and may cause an increase of the sofi error rate. In order to determine proper corrective actions for the degradation effects, a detailed analysis and thorough understanding of the phenomena which affect reliability is required. The physical causes of device degradation can be introduced either during the manufacturing process or by externally applied electrical and thermal stress. An example of manufacturing induced degradation is the ionic contamination which can cause hFE degradation in BJT's. However, improvements in the quality of the passivation layers currently used in integrated circuit manufacturing produce transistors which are not expected to degrade by ionic contamination. This study is specifically concerned with BJT degradation due to externally applied electrical and thermal stress which includes hot 1 carrier effects and electromigration effects. In order to have a better understanding of these two different degradation phenomena, the low frequency noise level which is well known as a sensitive indicator is used as a physical observable in this study. In addition to noise measurements, device I-V characteristics are also measured. Hot carrier induced degradation of both hp}; and low frequency noise levels as a result of avalanching the emitter-base junction of a silicon transistor was observed by McDonald [2,14] in 1970. It was believed that the observation of an hFE decrease and l/f noise level increase came from the same physical phenomenon, which was an increase in the number of surface recombination states within the surface depletion region. Such an increase may be a result, for example, of broken Si-H bonds caused by hot carriers. In recent scaled-down BJT's, hot carrier induced hFE degradation has been also observed during reverse-bias stress with the bias below the avalanche voltage [4-9] and possibly during forward bias stress with very high current densities [1,12,13]. A below-avalanche reverse bias can cause hFE degradation if energetic carriers are created by a sufficiently high electric field obtained by reverse bias in the scaled transistors. As a result of reverse- bias stress, an increase in the effective number of surface recombination states will result in an hFE decrease and a noise level increase [14-16]. For the forward-bias stress, hot carriers in the base-emitter region are not likely to be generated by an electric field because the base-emitter region electric fields in the devices are much smaller during forward bias stress then during reverse bias stress. Watchnik et. a1. [1] proposed, however, that hot carriers created by the Auger recombination process during forward bias stress can also cause an increase of surface recombination states. Consequently, an hFE decrease during forward-bias stress would be observed. There is no report yet to describe the behavior of noise levels during the forward-bias stress although according to the above rationale the noise would be expected to increase because of the increase of surface recombination states. High forward-active current stress may also lead to BIT degradation as a result of electromigration induced compressive stress built-up above the emitter region as described by Hemmert et. al. [18,77]. It is noted that both hot carrier and electromigration effects can produce similar degradation effects, although the two effects are quite different physically. For example, both may cause hFE to decrease, both may cause an increase in the leakage current and in the emitter-base junction current in the low forward-bias voltage region, and both show annealing of degradation effects. Consequently, it is not sufficient to distinguish between these two phenomena only by monitoring hp}; changes. Since the two phenomena relate to different device design features, it is important to distinguish between them. Noise measurements are helpful in this regard. It is shown in this dissertation that after forward-bias stress on a BIT the noise levels actually decrease, which is opposite to the expectation of a noise increase caused by an increase of surface recombination states. This indicates that noise measurements can indeed provide information which enables one to distinguish between these two degradation phenomena. 1.2 Objectives The objective of this study is to provide an increased understanding of the degradation mechanisms, with particular emphasis on hot electron effects and electromigration effects, when a BJT is subjected to externally applied bias and temperature stress. Toward this objective, noise measurements are performed in addition to the traditional hp]; measurements in order to help obtain a more extensive view of the degradation phenomena. The measurements are designed to identify which physical noise sources are the principal contributors for any observed noise changes. Noise degradation models then can be built up, based on the experimental results, to correlate noise changes with the observation of has degradation. There are two main tasks addressed in this study. First, the low frequency noise measurement is used as a vehicle to distinguish the different physical causes of degradation, either hot-carrier induced degradation or electromigration effects, when BJT's are subject to forward bias stress. Explanations of the origins of the noise changes are evaluated. Second, for reverse-bias stress, a comprehensive study of device degradation and post-stress recovery on both gain and noise behavior is presented. A model which correlates the stress induced noise changes with the stress induced base current changes is presented. This model allows one to predict the noise changes, which requires a more complicated characterization process, by knowing the changes in base current. 1.3 Preview of Thesis Chapter 2 reviews previous studies of hot electron effects and electromigration effects on bipolar junction transistors. Chapter 3 presents the reverse-bias stress results of the present study. Various DC stress voltages and various AC stress signals are applied to stressed devices. Post-stress recovery mechanisms, including thermal annealing and forward-bias induced recovery, are also investigated. A model which correlates the stress induced noise changes with the stress induced base current changes is presented. Also included in Chapter 3 is the noise measurement method used in this study for identifying the dominant noise source within a given bipolar junction transistor. Chapter 4 presents the experimental results of the forward-bias stress on three kinds of bipolar junction transistors manufactured with different technologies, including a test structure design specially intended for investigating the effects of electromigration induced compressive stress. For all three different BJT technologies, a general trend in the noise analysis emerged: the low-frequency noise levels initially decreased after application of forward bias stress at high temperature. A model for interpreting the experimental data shown in Chapter 4 is given in Chapter 5. The background needed to build-up this model is first reviewed. Then, the model is presented which utilizes the concepts of both dislocation motion and band gap narrowing caused by electromigration induced compressive stress. This model presents qualitative explanations of both gain and noise changes on all three different BJT technologies. Finally, conclusions and results are summarized in Chapter 6. CHAPTER 2 BACKGROUND 2.1 Low-Frequency Noise Sources in BJT's 2.1.1 Physical Origins There are in general four kinds of low-frequency noise sources that may be observed when noise measurements are performed on bipolar junction transistors at room temperature, namely, (1) the surface recombination noise source, (2) the diffusion noise source, (3) the internal series resistance noise source, and (4) the burst noise source. In principle, there is additionally a fifih low-frequency noise source due to bulk recombination in the base region, but it has been shown [24] that the bulk recombination noise source is generally negligible compared with other low-frequency noise sources. The noise power spectra of the first three low-frequency noise sources are almost inversely proportional to the fi'equency; in other words, they have the frequency dependence of M“ , where a is around unity. Consequently they are ofien referred to as l/f noise sources. However, for burst noise, the noise power spectrum typically has a 1/f2 shape for frequencies lower than 104 Hz [25]. The physical origin of each of these noise sources is described as follows. The surface recombination l/f noise is associated with the fluctuation of the carrier density due to the recombination-generation process around the device surface [26-28]. The recombination at the surface proceeds via fast surface recombination states, but these surface recombination states are modulated by the trapping or detrapping of carriers at and near the surface. Consequently, the fluctuation in the surface recombination should be proportional to the fluctuation in the density of trapped centers, and hence to the density of surface recombination states. This recombination type of l/f noise source is sometimes also regarded as a "non-firndamental noise" because the magnitude of its spectrum can be strongly reduced by eliminating most of the surface recombination states. The recombination current may be only a small part of the total current in BJT's, but the recombination 1/f noise can be significant since the recombination is a quite noisy process. It is believed that the surface 1/f noise of a bipolar junction transistor comes from the recombination at the base surface or at the emitter-base space charge region [29]. Jantsch [30] obtained a expression for this surface noise current power spectral density as _ 0.1 (is)2 l - _ 9 rec f AN” (2.1) where f is the frequency, A is the surface area where the surface recombination takes place (e. g. the base surface area or the emitter-base junction surface area), Nit is the interface state density, and 73 is the surface recombination current. With an applied voltage V, is is given by [64] is =1—sO exp(qV/ZkT) (2.2) with isO cc AN“. (2.3). Since (792 is proportional to (ANi¢)2, Equation 2.1 predicts that the noise power spectrum increases linearly with Na. Since the total base current is 18 =180 exp(eV/ka) (2.4) where Ismsz, (2.5) by using Equations 2.1, 2.2 and 2.4, the noise intensity can also be expressed as , = _ [3. rec fA N it 130m (2.6) Consequently, the surface 1/f noise power spectrum can have a non-linear current dependence. It is noted that Equation 2.1 allows one to calculate the rrns noise current within a given frequency range as 1/2 fH (Irec)rms =|:fJ: Slmdf] « (2.7) 1. Note that since S1,“ goes as l/f, the noise power contained in a bandwidth between fL and f H is proportional to log( f” / f L)- Consequently, each decade of frequency contains the same noise power. For example, the noise power contained between 10 Hz and 100 Hz is the same as that contained between 0.1 Hz and 1 Hz. The diffusion 1/f noise [24,31,32] is caused by the mobility (u) fluctuation of charged carriers when current is flowing in the devices. Physically this fluctuation is due to the scattering of carriers. Because of the Einstein relation eD=kTu, the mobility fluctuation will lead to a fluctuation of the diffusion constant D. It has been shown both experimentally [33-36] and theoretically [24,37] that the diffusion 1/f noise can be one important noise source for solid-state devices governed by diffusion process, such as BJT's. The diffusion 1/f noise is also regarded as a "fundamental noise" because it should have the identical value from device to device for the same device structure. For a n-p-n bipolar junction transistor biased in the forward active mode, there are two diffusion current sources which can result in the diffusion l/f noise. One, S15", is due to the injection of electrons from the emitter to the base, diffusion in the base, and subsequent collection from the base by the collector. This noise current power spectral density can be expressed as [38] I D 0: N(0) S = ‘1 En n Hnl , 2.8 where f is the frequency, Dn is the diffusion constant for the electrons, W3 is the width of the base region, N(0) and N( W3) are the base-injected electron concentration at the emitter side and at the collector side, respectively, and 05”,, is the Hooge parameter [34] for electrons. A mobility fluctuation induced noise, SIB,” is also associated with the base current which results from the injection of holes from the base to the emitter and subsequent diffusion in the emitter. This noise current power spectral density can be expressed as [38] _ €113th O‘Hp P(0) SIB}, (f) _ WE2 f ln[P(WE)]a (29) where f is the frequency, DP is the diffusion constant for the holes, W5 is the width of the emitter region, P(0) and P(WE) are the electron concentration at the base side and at the emitter side, respectively, and 0:11,, is the Hooge parameter [34] for holes. Similar diffusion noise expressions can also be applied to a p-n-p transistor, however, with the opposite types of diffiision carriers. ' The diffiision 1/f noise sources can be distinguished from the surface recombination l/f noise source with the help of their current dependence. The power spectrum of diffusion 1/f noise is proportional to the magnitude of current (I). On the other hand, the power spectrum of surface recombination l/f noise is not linearly dependent on I and usually has a quadratic dependence on the current [29,3 5]. The l/f noise which originates from the internal series resistance in a BJT's was first proposed in 1980 [39] although there was a controversy [40] on the accuracy of the qualitative elaboration of this proposed idea. Recently a more definitive investigation of this noise source [41] was presented. In this study, the low-frequency noise measured in high bias current region (15 > IOOuA) was attributed to the base series resistance 10 fluctuation. Although the base series resistance (rb) consists of two parts: contact resistance (rbc) and bulk resistance (rbb), the results indicated that the observed l/f noise was mostly from fbb. According to the Hooge relation [33] Srbb(f) = “(2"be /be , (2.10) where f is the frequency, 05;, is the Hooge parameter for the base, and Nb is the effective number of charge carriers. Physically this noise is also due to mobility fluctuation caused by the carrier scattering. In general, the emitter series resistance re and the collector series resistance rc can also contribute to the l/f noise as Sr, and Src respectively. However, re is generally small, and in the common emitter configuration rc is in the output stage which will have less effect on device noise properties than rb, which is in the input stage. Consequently, Sr, and 5,, usually are neglected in comparison with Sn,- The mechanism that causes the burst noise in bipolar transistors is not well understood so far. However, it is generally accepted that the burst noise is a particular type of generation-recombination noise which is caused by the defects located near and inside the bulk emitter-base junction area. Nevertheless, opinions on the nature of the defects are divided. Hsu et al. [25] indicated metallic precipitates, Blasquez et al. [42] favored crystallographic defects, and in supporting this Mihaila et al. [43] presented results indicating that burst noise is due to the clusters of dislocations inside the emitter area. 2.1.2 Identification of l/f' Noise Sources in BJT's As mentioned in the previous sub—section, the possible low-frequency 1/f noise sources in an given n-p-n bipolar junction transistor are S1,“, S15", S13," and 5,1,, which are shown in Figure 2.1. The goal of this sub-section is to review some noise measurement techniques used for finding the location of 1/f noise generator(s) inside a BJT, and, ll h+ .."'.'fl"'-""‘O-""--"““" F /too;-nooa-cocoon-ov—cooooaooaoao I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I O O a ................................................................ Slm, = suface recombination l/f noise source S,“ = electron diffusion l/f noise source S,“ = hole diffusion l/f noise source 5'1, = base series resistance l/f noise source Figure 2.1 Possible 1/f noise sources inside a n-p-n transistor biased in forward-active mode. The arrows show the direction of charged carrier flow (not the current flow). 12 furthermore, deciding which of four above-mentioned l/f noise sources may be dominant in each generator(s). It is generally accepted that a circuit model for the l/f noise power spectra measured in bipolar junction transistors should include two independent noise current generators: Slbe located between the base and the emitter, and S1ce located between the collector and the emitter. There should, in principle, be a third noise current generator Slbc located between the base and the collector. However, the base-collector junction is usually reverse biased so that the base-collector current is slight and its noise can be neglected [35]. SIbe can be attributed to the combined effect of diffusion 1/f noise source (i.e., 813,, for a n-p-n transistor) and surface recombination 1/f noise source (81,“); however, usually one type of 1/f noise sources will be dominant. By checking the current (13) dependence, as described in Section 2.1. 1, one can determine the dominant contributor. On the other hand, Slce comes from the diffusion 1/f noise source (i.e., S15" for a n-p-n transistor) and should be proportional to the current 1c. Besides those two noise current generators, there is also a noise generator attributed from the base series resistance noise sources, Srb. Since there is a current 13 flowing through rb, S", can contribute to the low frequency noise in a transistor. Figure 2.2 shows the circuit model for the low-frequency l/f noise sources. Here ibe and ice are the rms noise currents of SIbe and SI“, respectively. It is noted that there also exits a rrns noise voltage of [32 Sn, across rb since 13 flows through rb. The rrns noise value of either a current power spectral density or a voltage power spectral density has a similar definition as shown in Equation 2.7. Kleinpenning [41] demonstrated a method to identify the 1/f noise sources by analyzing the results measured under both the common-emitter and common-collector configurations with different values of the external resistances. Figure 2.3 showed the circuits used for his noise measurement method. It is noted that the transistors have the noise circuit model shown in Figure 2.2. Both the collector-noise voltage spectral density (Sr/C) and the base-noise voltage spectral density (S V3) were l3 b rb b' C, ’c w r T W-—* lbe El gr“ CngVb’e, [3 ice e' re e Figure 2.2 A circuit model for the rms l/f noise sources measured in a BJT. l4 : VCC VCCTT_ _: VEE Figure 2.3 The transistor bias circuits under (a) common-emitter configuration and (b) common-collector configm'ation for low-frequency noise measurements. (After Kleinpenning [41]) 15 measured under the common-emitter configuration, and the emitter-noise voltage spectral density (S V5) was measured under the common-collector configurations. It may be shown that r- +R S +r+ +R 2S + +12128 SVE=[7: a”) 3)}2 [be [75 if B] Ice ([3 ) B’bXRlzf’ (2.11) ' Z [329), +RB +r +R ]2S +[r +0, +R +r +R ]25 “#32125 S = e E Ibe n B e E Ice 8 I'b X 2 VC 2 RC’ Z (2.12) r+ r+R 2S +r+R 2S +12S SVB=[7r me 175)] 1be22[9 E] [Le 13'be13, (2.13) with Z=RB+rb +rn+(l3+1)(RE +re)- (2'14) It is noted that a white noise term could be added to each of above noise equations. However for low frequencies, the white nose is dominated by l/f noise. By measuring SVE, SVC and S VB for known values of 11;, 1C, 113, RE, RC and RB, these three equations enable one to solve the three unknown noise generators, (1) S1ce, (2) Sibe, and (3) 54' Also, by modifying the bias point by varying the resistor values, it is also possible to check the current dependence of S1ce and Sjbe, and thus determine the physical origin for these l/f noise sources, as will be described in Chapter 3. An alternative noise measurement method has been developed by van der Ziel et al. [29]. The bias circuit for this method uses only the common-emitter configuration, however, with or without a R}; in the emitter lead. In this approach, basically four measurements are needed for a given bias point: (1) the collector noise spectrum with R5, (2) the collector noise spectrum without RE, (3) the base noise spectrum with R5, and (4) the base noise spectrum without RE. RE serves as a feedback resistor which allows the observation of the amplification or attenuation of noise generators, thus facilitating the l6 determination of noise generator. Also, in this approach, the 1/f noise caused by internal series resistance fluctuations was not considered. 2.2 A Review of BJT Degradation due to Electrical and Thermal Stress The physical causes of BJT degradation can be introduced either during the manufacturing process or by externally applied electrical and thermal stress. An example of manufacturing induced degradation is ionic contamination which can cause hFE degradation in BJT's. However, improvements in the quality of the passivation layers currently used in integrated circuit manufacturing produce transistors which are not expected to degrade as a result of ionic contamination. This section is specifically concerned with BJT degradation due to externally applied electrical and thermal stress, which includes hot-carrier induced degradation and electromigration induced degradation. 2.2.1 Hot-Carrier Induced Degradation It is believed that the physical cause of hot-carrier induced degradation is generally due to one of two mechanisms. The Si-Si02 interface may be damaged, resulting in an increase in the number of interface states. Secondly, carriers may be trapped in the oxide, either by surrnounting the Si-Si02 barrier, or by tunneling through the barrier to trap states in the energy gap of the oxide. In either case, the resulting change in trap density and band bending cause changes in BJT performance parameters. There are two distinctly different mechanisms by which hot-carriers may be created in a BJT. Hot-carriers are produced either by a large electric field or by Auger recombination [1]. The first mechanism certainly plays a dominant role in reverse-bias degradation of BJT's, and both mechanisms may play a role in forward-bias degradation of BJT's. 17 For reverse-bias degradation studies, McDonald [2] in 1970 showed that ha; degradation results when the emitter-base junction is reverse biased to the point of avalanche. He was able to explain the degradation induced leakage current by an increase in the number of surface recombination centers as a result of avalanche-induced surface- state formation within the surface depletion region. By using a gate-controlled transistor structure, he also determined that there was localized charge trapping within the oxide over the emitter-base junction as a result of avalanching. In the past twenty years, several improvements have bee; trade to avoid operating BJT's at avalanching biases. An example is the termination of wt lines by clamping diodes in integrated circuits. However, recent trends in bipolar transrstor scaling [3] have resulted in highly energetic carriers at the reverse biased emitter-base jL tion for below- avalanche voltages [4]. Hot carriers were reported to cause degradation ir. ivanced technology transistors not operated under avalanche bias conditions in sever:~ .;udies [4- 9]. Petersen et al. [5] described a hot-electron .iuced hFE degradation study in advanced self-aligned sidewall spacer n-p-n bipolar transistors. The transistors were operated with their emitter-base junctions reverse biased below the avalanche voltage and collectors opened. The carriers were injected by illuminating near the edge of emitter-base depletion region with a 5 mW He-Ne source. The measured characteristics taken before and after stress showed that the increase of the low level base current afier stress was accompanied by a shift in the ideality factor from near 1 to 2. This indicated that the probable cause for the higher base leakage current is an increase of interface state density in the sidewall oxide between the emitter and base. The work done by Joshi [6] showed hot-carrier induced degradation in a poly- emitter bipolar transistor fabricated by a BiCMOS process. Transistor hpg degradation was studied by reverse biasing the emitter-base junction under DC, pulse, and AC conditions. For the ceramic packaged devices, extended temperature (-55 to 175 C) tests 18 were performed. For a constant current stress, it was shown that hFE degradation depended on the total injected charge, which was equal to the stress current multiplied the stress duration, through the reverse biased emitter-base junction. A study of hot-carrier efi‘ects in self-aligned polysilicon emitter bipolar transistors was also presented by Burnett et al. [7] in 1988. The 17175 degradation under DC reverse- bias stress was described by a empirical model with fitting parameters. The model was able to predict the device lifetime over a wide range of reverse-bias stress currents for a given technology. Momose et al.'s paper [8] showed a study of temperature dependence of emitter- base reverse-bias stress degradation. It was found that the largest degradation occurred around 50 C. Their simulation work confirmed that both electron trapping in the oxide and interface state generation occurred during stress. However, the generation of interface states in the oxide near the emitter-base junction was the main cause for the degradation. In Burnett et al.'s later work [9], BJT's fabricated by BiCMOS technology were reverse biased at 110 K and 300 K respectively. Although the reverse current is smaller at 110 K than at 300 K for the same reverse voltage, the rate of degradation was observed to be four times larger at 110 K than at 300 K for the same reverse voltage and ten times larger for the same stress current. The increased degradation at low temperature is consistent with the increase severity of hot-carrier damage in MOSFET's [10]. In both cases, the carrier mean-free-path increases with decreasing temperature, allowing the carriers to achieve higher energy prior to scattering. Recently an investigation of process dependence of reverse-bias hot-carrier reliability in BiCMOS n-p-n transistors was performed [76]. As shown in the experimental results, the base doping concentration as well as the thermal history affect the hot-carrier induced hFE degradation. Rapid thermal annealing (RTA) was used in this study. The band-to-band tunneling current was identified as the dominant reverse-bias stress source. Consequently, any means which can reduce the band-to-band tunneling current should l9 improve the hot-carrier reliability. It was found that hot-carrier induced degradation was improved by reducing the base dopant concentration and by increasing the RTA time. For forward-bias induced hot-carrier degradation studies, Chen et al. [11] in 1986 reported the effects of forward active current stress on the electrical characteristics of self- align transistors. The transistors were mounted on high temperature ceramic packages and stressed in an ambient temperature ranging from 25 C to 280 C. During stress, the transistors were biased in the forward active mode by forcing a constant current density of 1 mA/um2 through the emitter. It was found that the transistors with higher initial non- ideality factor were more susceptible to degradation induced by high-current stress. The dominant changes in the electrical characteristics after stress were found to occur at the base-emitter junction with the collector current remaining unchanged. In the low current region, an increase of the non-ideality factor of the base current, and thus a decrease of hp}; was observed after stress, and was attributed to interface state generation underneath the sidewall oxide between the base and the emitter polysilicon. In the high current region, for transistors initially with nearly ideal base and collector current characteristics, hpg was found to increase afier stress. It was suggested that this was caused by an increase in the total Gummel number resulting fiom either electromigration of the dopant to the interface between the 11+ polysilicon and crystalline silicon, or the breaking of the interfacial oxide barrier between n+ polysilicon and crystalline silicon. However, no physical evidence was given to confirm these mechanisms. Auger recombination was identified by Wachnik et al. as a main source of forward- bias induced hot-electrons which produced BJT damage [1]. At room temperature, the forward bias was applied across the emitter-base junction with the base-collector junction shorted. Under this condition, electric fields in the BJT are relatively low since the base- emitter junction was forward biased, which reduced the electric field at that junction, and the collector-base junction had a zero bias applied. Also, the stress time and temperature ruled out the possibility of electromigration such as that reported by Hemmert et al. [18] 20 (see Section 2.2.2). They reported an observation of a fluctuating increase in the base current with the collector current remaining unchanged. Wachnik et al. proposed that the degradation was due to generation and diffusion of hot Auger electrons to the Si-SiOz interface at the sidewall oxide between the emitter polysilicon and base polysilicon. This was followed by a bond breaking event creating a dangling silicon bond, presumably a broken Si-H bond. The work also presented evidence for repassivation of that bond with continued stressing since fluctuation of the base current was observed. In 1990, Joshi [12] reported a study on degradation for self-aligned bipolar transistors operated under various forward-bias current densities (0.23 mA/um2 to 2.86 mA/umz) at various ambient temperatures (0 to 150 C). The most pronounced change in BJT characteristics was the base current and the hpg. The collector current remained unchanged during stress. hpE in the low current region (Ic < IOuA) was found to increase initially with stress time and then degrade very rapidly with subsequent stress, whereas by; in the high current region was found to increase monotonically with stress time. He also found the transistors with perimeter-to-area ratio larger than 1.5 were more susceptible to larger hFE changes. It was suggested that the major cause of [1,75 change was an increase of interface traps and/or positive charges due to Auger hot-electrons. A predictive model for the device degradation rate depending on the effective emitter stress current destiny and stress temperature was also developed. A investigation of very-high-current (up to 12.5 mA/umz) degradation on Si n-p-n transistors was done by Tang et al. [13]. The stressed devices were double-poly self- aligned n-p-n bipolar transistors with a emitter area of 0.8x2.0 umz. The emitter-base junction showed negligible degradation when the transistor was stressed in the normal mode such that the emitter-base junction was forward biased and the collector-base junction was reverse biased, so that electrons were injected fiom the emitter, diffused through the base and collected by the collector. However, the emitter-base junction showed degradation when stressed in the forward bias with the collector open. Also, the 21 emitter-base junction showed degradation when the transistor was stressed in the inverse mode such that the emitter-base junction was reverse biased and the collector-base junction was forward biased, so that electrons were injected from the collector and collected by the emitter. They observed a significant stress-induced leakage component in the base current in the later two stress modes. The cause of the junction degradation was believed to result from the degradation of the sidewall oxide damaged by the hot-carriers injected from the base-collector junction. Besides the above-mentioned reports, in which transistor DC current characteristics and hpE were primarily used as the physical observables, some other studies have also indicated that the low-frequency noise will be changed during reverse-bias hot- carrier degradation. McDonald [14] reported that I/f noise was increased significantly after avalanching the emitter-base junction of a bipolar junction transistor. By using the gate-controlled transistor structure, he showed that the mechanism that caused a l/f noise increase was identical to that responsible for hp}; degradation, which was attributed to an increase in surface recombination centers within the emitter-base space-charge region. Koolen et al. [15] also showed that an increase of low-fi'equency noise was correlated with the decrease of hFE of a bipolar transistor afier reverse-bias stress. The decrease of 11175 was identified as resulting from an increase of the non-ideal base current component with the collector current remaining unchanged. A modified SPICE noise model which described the noise behavior during the reverse-bias stress by relating the magnitude of the non-ideal base current component to the increase of the low-frequency noise was presented. Another model correlating the l/f noise magnitude with the hFE degradation of a polysilicon emitter n-p-n transistor fabricated by BiCMOS technology was proposed by Dreyer et al. [16]. At room temperature, a below-avalanche 5.2 V reverse bias was applied to the emitter-base junction. Consequently the noise voltage power spectral density was 22 related to by; degradation in the following form: 720,0...) = a x exp(—b - hFE" ), (2.15) where a and b were the fitting parameters, and km“ the normalized hFE defined as hp); after stress divided by original hFE. By checking the current dependence of the noise magnitude, they confirmed that surface recombination was the source of observed increased l/f noise. Furthermore, it was found that the slope of the noise spectrum was very sensitive to hpE degradation, with an increase in spectral slope from 1/f to l/f‘2 at the onset of degradation. However, the slope of the noise spectrum slowly dropped toward its initial value as the stress continued. There is no report found yet to describe the behavior of low-frequency noise levels as a result of the forward-biased induced hot-carrier degradation. However, one would expect to observe the same low-frequency noise increase as that in reverse-bias stress because the same physical degradation phenomenon, which is the increase of surface recombination centers, can be applied to both reverse-bias and forward—bias induced hot- carrier degradation cases. 2.2.2 Electromigration Induced Degradation Forward-bias stress with a very high current density can also cause electronrigration induced degradation in a BJT in several ways. Previous reports include the observation of an open circuit for the collector, a short circuit or a resistive shunt between the base and the emitter, a short circuit between the multiple-layer metalization, a change of contact resistance, the penetration of contact metal into Si substrate, and a change of Irma. Electromigration induced failure for silicon devices and integrated circuits utilizing Al metalization was presented by Black [17] in 1969. He described two kinds of failure 23 modes: (1) the formation of an electrically open circuit due to the occurrence of the metal voids, and (2) the penetration of aluminum into the Si emitter of a n-p-n bipolar transistor, by which the shorting of the underlying junction could occur. A theory which described the failure rate (open circuit) of metal conductor stripes was also presented. The theory expressed that the failure phenomena proceeded as the current density square multiplied by exp(-¢/kT), where d) is the activation energy, k is the Boltzmann's constant, and T is the absolute temperature. A study of electromigration induced hpg degradation attributed to electromigration induced compressive stress was proposed by Hemmert et al. [18] in 1982. Under forward-bias stress, electromigration will tend to cause an accumulation of metal above the emitter region. However, as shown in Figure 2.4, with a passivation layer enclosing the metal line the thickening of the metal will not be able to readily occur. Instead, there will be a build up of compressive stress normal to the substrate. The devices used for stress in their study were n-p-n transistors fabricated in a conventional manner on <100> p-type wafers. Aluminum penetration into the emitter was prevented by a chrome diffusion barrier. A Si02/Si3N4 passivation layer with different thicknesses was deposited above the devices. The experimental results showed that increasing the passivation thickness aggravated beta degradation. This phenomenon was explained by the thick wall pressure vessel problem [19], where equal volume of fluid accumulation(metal) resulted in higher pressure(stress) as the vessel wall(passivation layer) became thicker. According to the theory proposed by Kanda [20], when a n-p-n BJT built on a <100> wafer is subjected to a compressive stress, a net decrease in has should be observed for a long stress time. However, on a shorter time scale, a double minimum in hFE is predicted during the stress period. This double minimum in hFE is attributed to the stress dependence of the hole effective mass. Hemmert et al. did observe a slight double minimum in 111:]; degradation process. A post-stress annealing was also performed at various temperatures, and a complete recovery of 171:}; was achieved at 400 C in less than 15 minutes. The annealing PAS SIVATION (COMPRESSIVE STRESS) OXIDE L EMITI‘ ER j BASE k J OXIDE W COLLECTOR Figure 2.4 An electromigraion-induced metal accumulation, with a passivation layer enclosing the metal. can result in compressive stress on the emitter. (After Hemmert et al. [18]) 25 effect indicated that electromigration induced compressive stress is relieved by creep in the aluminum metalization. Wada et al. [21] performed electromigration tests on n-p-n bipolar transistors with three different metal layouts. Both transistor configuration 1 (Tr-1) and transistor configuration 2 (Tr-2) had a single emitter. However, Tr-l had single base and single collector contacts, but Tr-2 had double base and double collector contacts. Transistor configuration 3 (Tr-3) had a double emitter and triple base and collector contacts. Also, Tr-3 was the only transistor using double layer metalization. The stress condition was with Ic varying from 4 mA to 60 mA and with temperature varying from 150 C to 250 C. They observed that in all cases degradation was accelerated by higher temperature and higher current density. Also, the failure mode depended on the metal layout of the transistor. When the configurations of the emitter metalization and collector metalization were almost the same, the open circuit of the collector was the main failure mode. However, when the configurations of the collector metalization was larger than that of the emitter metalization, the resistive shunt of the emitter-base junction occurred. For Tr-l a gradual decrease of [mg was also observed before the open circuit of the collector was finally observed. For Tr-3, which had a collector metalization on top of the emitter metalization, emitter-collector short circuits were caused by hillocks breaking through the intermediate dielectric layer. Tang et al. [75] showed a metal migration into the polysilicon emitter afier stressing at an emitter current density of 1.63 mA/um2 for 47 hours at 90 C. According to post-stress TEM analysis, a metal filament was observed at the edge of the emitter polysilicon. However, it did not penetrate through the polysilicon. It is noted that the edge of the emitter polysilicon has the highest current density because of the emitter crowding effect. The metal filament lowered the emitter contact resistance. Also they suggested that a nrinute amount of metal, undetectable to the TEM analysis, could penetrate through the 26 polysilicon and reach the polysilicon/silicon interface. Under that circumstance, a lower hFE may be expected. Noise measurements have been used to predict the failure rate of metal film interconnectors for VLSI [22-23]. However, no report has been found using noise as an observable for the electromigration induced compressive stress effect on BJT's. In this dissertation, a noise study for such compressive stress on a BJT will be presented. Chapter 5 presents the experimental portion of this noise investigation, and a model for. interpreting the data is given in Chapter 6. CHAPTER 3 REVERSE-BIAS STRESS 3.1 Device Description There were two kinds of silicon n-p-n transistors which were manufactured by IBM used in this reverse-bias stress study, namely TIN3 and TOUT. These transistors were contained in silicon chips, which were bonded in 28 pin ceramic packages. Each chip had one TIN3 and one TOUT. The chips were traceable as to location on the wafer, wafer number and job number. Conventional oxide isolation methods [81] were used for fabricating TIN3 and TOUT. These devices did not include polysilicon emitter or polysilicon base structures. Instead, a first layer metal was used for device contacts. Figures 3.1 and 3.2 show the layouts for TIN3 and TOUT devices respectively. TIN3 and TOUT had the same emitter region layout with an area of 2.5x4.3 umz. The emitter contact cut is also the emitter ion-implant region. However, TOUT had a Schottky diode clamp and TIN3 did not. Both devices had the same doping profile as shown in Figure 3.3. The donor concentration on the emitter surfaces of both devices was 3.2x1020/cm3, and the acceptor concentration on the extrinsic base surface was 6.3x 1013/cm3. The intrinsic base width was 0.35 pm. Both devices were provided with double collector contacts, double emitter contacts, and a single base contact as shown in Fig 3.4. For TIN3, the pin- to-pin resistance between the two collector contacts were approximately 21 Q whereas for TOUT the resistance was approximately 5.2 (2. For the common emitter contacts shared 27 . V a M H—f—H \\ \' \\ Fgur31TpvrwthIN3dvrcely t.eTh dienm nisare units ofum.Metal region sear shownassolidbla ckareaso ontactc ansdascro sts-ha ched regions elsewhere. Actual metal regions exceed the plotted metal regions. \\\\\\\\\\\\\\\\\\\ \ \\ \ Base t \\ \\ \\\\\\\\\ \\\\\ :0a A JJALAAA 10' A AAAAAA‘ A xo|O A A AAA‘AA‘ xol. doping concentration A A A AAAAA. (917 A A ALAAA. 10" T r , 0.00 .230 .WOO .600 .000 1.00 1.20 1.W0 Figure 3.3: Doping profile for TINB and TOUT devices [81]. The impurity concentrations are in units of cm'3. 31 ll 12 13 17 I9 20 10: TOUT 141:1 TD4 18 TINS 15 16 Figure 3.4: Pin out configuration of the transistors. The three transistors have two common emitter pits. 32 by TIN3 and TOUT, the pin-to-pin resistance was approximately 8.5 Q. Both collector contacts and both emitter contacts were used during the stress conditions and during measurements to reduce parasitic resistance effects. 3.2 Device Characterization 3.2.1 Low-Frequency Noise Measurement Method As described in Section 2.1.2, the low-frequency noise sources inside a bipolar junction transistor can be identified by simple biasing techniques [41,29]. In this section, the noise measurement method used in this study is presented. This method used the noise equations obtained in [41]. As described in this section, it is found by this method that the dominant l/f noise source in both TIN3 and TOUT is the surface recombination type, and is located between the emitter and the base. By the same method, it will be shown later in Chapter 4 that the low-frequency noise source of transistors used for the forward-bias- stress study is also dominated by the surface recombination 1/f noise. The noise power spectral density was measured at room temperature by using a Hewlett-Packard 3561A Dynamic Signal Analyzer and averaging 200 sampling results. The frequency region measured was from 1 Hz to 100 KHz. The Hanning filter shape (window) of the HP 3561A was chosen in processing the data. The bandwidth was 375 ner when the noise measurement was performed from 1 Hz to 100 Hz, 37.5 Hz when the noise measurement was performed from 100 Hz to 10 K Hz, and 375 Hz when the noise measurement was performed from 250 Hz to 100 KHz. In order to reduce the effect of noise sources other than the test transistors, wire-wound resistors and battery power supplies were used in the noise measurement circuit. By using the common-collector bias circuit as described in Figure 2.3 (b), the emitter voltage-noise power spectral density (S V5) has the following form 33 S _ [6: 43(4) +RB)]251b, +16: “b +RB]ZSIce +(l3+l)21§Sq, V5 ' 2 Z x1233, (3.1) with Z=RB+rg +r,,+([3+1)(RE +re). (3.2) If a large RE is chosen in the bias circuit for this noise measurement, such that (B+1)RE >> R3 + r], + r), +re and RE >> re, , then SVE =1rn/13-(n, +RB)12S1,, urn/(3+0), +RB)/m251,, +4235... (3.3) Here, we assume [3 >> 1. For a given n-p-n transistor, SIbe can be attributed to the combined effect of base diffusion l/f noise and surface recombination 1/f noise. However, usually one type of 1/f noise sources will be dominant. By checking the current (13) dependence, one can determine the dominant contributor. Slbe has a dominant diffusion noise if SIbe cc [3. The surface recombination noise is dominant if SIbe oc (IB)7, with y z 2. On the other hand, S1“, comes from the emitter diffusion 1/f noise and should be proportional to the current IC. S", is attributed to the base series resistance noise and has a fixed magnitude. However, as shown in Equation 3.3, the third term in SVE which includes S", will show its current dependence as 132. The steps for checking the dominant noise source in SVE are presented as the following. By fixing the base resistance (RB) and changing the emitter resistance (RE), the bias conditions of the common-collector circuit as shown in Figure 2.3 (b) can be changed, and the current dependence of SVE thus can be checked. There are three bias conditions checked for a given device in our study. For example, the base-emitter bias (VBE) can vary as 0.78, 0.8 and 0.82V with a constant base-collector bias (VBC) of -1.5V. For both TIN3 and TOUT devices, these bias conditions correspond respectively to R5 = 510 KO, 200 K52, and 100 K!) with a constant R3 = 100 K!) used in the common-collector bias circuit with VEE = 36 V and VCC = 1.5 V. If SVE happens to have a current dependence of 34 132 (or 132), this means the base surface recombination noise, the base series resistance noise, or the combination of these two noise sources is dominant in SVE. In this case, further analysis has to be made. By examing Equation 3.3, one will find that for a fixed 13 value, the contribution of Sn, to S VE is a constant, but that of Slbe to 5V3 will depend on the value of R3. If R3 >> (I), —r,r/[3), SVE is approximately proportional to R32. Consequently a fourth noise measurement must be performed with one of the three bias conditions used before for checking the current dependence, however, with a different combination of R3 and R3. For example, for a TIN3 device the fourth measurement can be made with VBE = 0.82 V and VBC = -1.5 V by biasing with a smaller R3 (= 51 K0) but a higher R3 (z 100.5 KQ). It is noted that this R3 value (= 51 K52) should be much larger than (I), -r,,/fl). If S V5 from the fourth measurement is the same as that from the third measurement, then we know that S,,, has a dominant effect on SVE. If S VE from the fourth measurement is deceased and proportional to R32, then the surface recombination Sjbe is dominant. Otherwise, one would conclude that both Sn, and the surface recombination Slbe have appreciable effects on SVE, and their respective effect can be further decided by comparing the results from the third measurement with that from the fourth measurement. Figure 3.5 shows a set of measurements of the noise power spectral density on a TIN3 device under different bias conditions: VBE = 0.78, 0.80 and 0.82 V with a constant VBC = -1.5 V. Under the same bias conditions as the third measurement, VBE = 0.82 V and VBC = -1.5 V, Figure 3.6 also shows that the noise power spectral density is decreased to a quarter of the magnitude obtained in the third measurement when R3 is decrease to half of the value used in the third measurement. Figure 3.7 shows the quadratic current dependence of the noise power at 1 Hz. The results confirm that the surface recombination Slbe is the dominant noise source in TIN 3 devices. Figure 3.8 shows three measurements of the noise power spectral density on a TOUT device under the same bias conditions as Figure 3.5: VBE = 0.78, 0.80 and 0.82 V with a constant VBC = -1.5 V. Under the same bias conditions as the third measurement, Sve 35 TIN3 (80) 10-11 10‘12 10-13 10‘1Li 10-15 ' I T Y Y i r r l r v T v r l 100 101 10a 103 10‘+ 105 FREQUENCY Figure 3 .5: A set of three measurements of the emitter noise power spectral density on a TIN3 device under different bias conditions: VBE = 0.78, 0.80 and 0.82 V with aconstant VBC = -1.5 v. Sve 10-10 10-11 \ 10-12 10-13 10-1% 10‘15 36 TIN3 Ube=0 BEU r1*‘1*'1*fffir1 100 101 10a 103 10‘+ 105 FREQUENCY Figure 3.6: Two measurements of the emitter noise power spectral density on the same TIN3 device with its noise levels as shown in Figure 3.5. These two measurements are made under the same bias conditions, VBE = 0.82, and VBC = -l .5 V with a difi‘erent combination of R3 and R3. 37 10%0 TIN3<80> @ > 10”14 U) 10”? . *fi , 1 10‘5 10‘* 10‘3 Figure 3.7: The quadratic current dependence of the noise power at 1 Hz for a TIN3 device. 38 VBE = 0.82 V and VBC = -1.5 V, Figure 3.9 shows the noise power spectral density is decreased to a quarter of the magnitude obtained in the third measurement when R3 is decrease to half of the value used in the third measurement. Figure 3.10 shows the quadratic current dependence of the noise power at 1 Hz. The results also confirm that the surface recombination Slbe is the dominant noise source in TOUT devices. Although it does not apply to TIN3 or TOUT devices, it is noted for the sake of generality that if SVE happens to have a linear current dependence with [3 (or 13), this means the base diffusion 1/f noise of SIbe: the emitter diffusion 1/f noise (51cc), or the combination of these two noise sources is dominant in SVE. According to Equation 3.3, the contribution of SIbe to SVE can be minimized if R3 is equal to r" - rb. Consequently, a fourth measurement could be performed with the same bias conditions as the third measurement used before for checking the current dependence, but with R3 chosen to be equal to r” - rb. If the noise power measured in the fourth measurement is decreased dramatically, then the base diffusion l/f noise is dominant. If the noise power measured in the fourth measurement is almost unchanged, then the emitter diffusion l/f noise is dominant. However, as noted, the transistors used in this study did not have a dominant diffusion noise source. Instead, the quadratic current dependence indicates surface recombination noise. The noise measurement flow of the emitter-voltage power spectral density (S V5) used for identifying the noise sources inside a given n-p-n BJT is summarized in Figure 3.11. 3.2.2 I-V Characterization The I-V characteristics of the transistors being stressed were measured before any stressing began and at cumulative stress times of 1, 4, 24, 100, 200, 500, and 1000 hours Sve 39 TOUT (05) 10'101 10-15 1 T T v , I 1 w i v r I FREQUENCY Figure 3.8: A set of three measurements of the emitter noise power spectral density on a TOUT device under difl‘erent bias conditions: VBE = 0.78, 0.80 and 0.82 V with a constant VBC = -1.5 V. Sve TQUT Ube=0 BEU 10-11 10-12 10-13 10'1“+ 10-15 ' ' I ' ' r * 1 w 1 T ' 1 10° 101 102 103 10"+ 105 FREQUENCY Figure 3.9: Two measurements of the emitter noise power spectral density on the same TOUT device with its noise levels as shown in Figure 3.8. These two measurements are made under the same bias conditions, VBE = 0.82, and VBC = -1.5 V with a different combination of RB and RE- 41 fie TOUTCOS) 10 — , 01 m 1 ! 10‘13 l T 10‘5 10‘“ 3 4 19 Figure 3.10: The quadratic current dependence of the noise power at 1 Hz for a TOUT device. 42 Check 3v]; at three different bias conditions BI, 82 and B3 to determine the current dependence of SVE. (choose a large RE) st oer; / \ 5.,E 0:13 Repeat SvE at B3 with Repeat SvE at B3 with a . different combination of R3 R33?“ and a drfferent and RE. (R3 >> rb - r,,/ [3) RE. SvE changed ? SvI3 decrease dramatically? (unchanged) ’ Surface noise 51 be Figure 3.11: The noise measurement flow of the emitter-voltage power spectral density (SVE)~ 43 during the stress. The I-V characteristics measured consisted of the following set of six I- V measurements: (1)1EBO forward, 0.0 V s V33 3 0.8 V and I < 1 mA, (2) IEBO reverse, -2.0 V s V33 3 0.0 V, (3) ICBO forward, 0.0 Vs V30 3. 0.8 V and I < 1 mA, (4) 10130 reverse, -5.0V 3 V33 .<_ 0.0 V, (5) 13, 1C with VBC = 0.0 V and 0 S V33 5 0.8, and I < 1 mA (Gummel plot), (6) Family of curves, IC versus VC3 and 13. Figure 3.12 shows the circuit configuration for each of the tests. Double contacts were used for the emitter and the collector during device testing. Measurements were performed with a Hewlett-Packard 4145B Semiconductor Parameter Analyzer in the long integration time mode, except for the family of curves, for which the medium integration time was used. The measurements were performed at T = 23 i 1 C in a temperature-controlled environmental chamber as shown in Figure 3.13. The data collected was then transferred to a workstation where a database was created. The characteristics studied most extensively were the common-emitter current gain and Gummel plot curves. 3.3 DC Reverse-Bias Stress and Post-Stress Recovery Previous work has shown that two critical bipolar junction transistors performance characteristics, current gain (1133) and low-frequency noise level, can be degraded by reverse biasing the base-emitter junction as described in Chapter 2. It has also been previously reported that the degradation in gain can be recovered at least in part by post- stress thermal annealing, and more recently, by application of a post-stress forward bias as well [6,8,78]. In this section, a similar recovery for transistor low-frequency noise is presented. 0 [lcso [ 11:30 (a) I 11:30 (h) 1 cao IC 4. Von (:1) Family of Curves Figure 3.12: Circuit configuration for the I-V measurements: (a) I530 forward and reverse, (b) ICBO forward and reverse, (c) Gummel plot, and (d) family of CUIVCS. 45 ‘I'hermon'on Environmental Chamber HP 41453 Semiconductor - Parameter Analyzer Test Device in Low Noise - IEEE 488 Fixture Bus Ethernet MSU Work Stations Figure 3.13: Diagram for the I-V measurement setup. TIN3 devices are used for this study. The reverse-bias-induced hot-carrier degradation on gain and noise and the post-stress thermal annealing will be shown in Section 3.3. 1. The forward-bias-induced post-stress recovery on both gain and noise levels on TIN3 devices will be shown in Section 3.3.2. By comparing the thermal annealing recovery of 1133 and noise with forward-bias-induced recovery of h33 and noise, estimates are provided for the relative role of interface state generation and trapped electrons in the degradation/annealing process. Finally, the gain degradation behaviors under AC stress will be provided in Section 3.3.3 which shows that the AC stress including forward-bias cycles produced less degradation than the AC stress without forward-bias cycles. 3.3.1 DC Reverse-Bias Degradation and Thermal Annealing TIN 3 transistors were divided into two groups for this study. In the first group, transistors were subjected to a reverse-bias stress for a longer stress time (up to 500 hours), and the post-stress thermal annealing was performed at various temperatures (fiom 150 C to 240 C). Only I233 changes were monitored for the devices in the first group. In the second group, transistors were subjected to the reverse-bias stress for a shorter stress time (up to 100 hours), and the post-stress thermal annealing was performed only at 240 C. Both h33 and low-frequency were observed for the devices in the second group. For both groups, the reverse-bias stress was applied to the base-emitter junction with the collector-base junction shorted. The stress voltage is -4 V which is below the breakdown voltage of approximately -5 V. The stress was interrupted for measurements at 1, 4, 24 and 100 hours for both groups and also at 200 and 500 hours for the first group. The Gummel plots and the h33 degradation of the transistors in the first group, as illustrated in Figure 3.14, 3.15 respectively, indicated that h33 degradation was due to an increase in base current with the collector current remaining unchanged. The increase in base current due to stress is particularly apparent at low values of V33. At low values of 47 junction voltage, the junction current due to diffusion and drifi is small compared to the stress-induced leakage current. At V33 = 0.6 V, 500 hours of stress causes a reduction in 1133 to 36% of its original value and at V33 = 0.8 V, a reduction in h33 to 82% to its original value. The post stress thermal annealing was performed at ambient temperatures of 150, 175, and 240 C after 500 hours of reverse-bias stress. The annealing experiments were interrupted at pre-selected times so that the I-V characteristics could be measured to determine the recovery rate of the current gain. The recovery of the normalized current gain (gain post-annealing / gain pre-annealing) at V33 = 0.6 V is shown in Figure 3.16. The recovery is due entirely to a reduction in base current since the collector current is unchanged by the annealing procedure. The thermal activation energy for this annealing process is in the range of 1.2 to 1.5 eV, which is comparable to the value of 1.6 eV reported in earlier work by McDonald [2]. For each annealing temperature, the largest recovery occurred during the earlier portion of the annealing process. Complete recovery of the gain was not observed at any of the time/temperature annealing combinations investigated here. An 85% recovery was achieved after annealing for 700 hours at 240 C. However, saturation of current gain recovery is not observed over the time scale of the study. For the second group, both the gain and the noise level were investigated. The collector-voltage noise power spectral density (S VC in Equation 2.12) was measured. The measurement circuit is under the common-emitter configuration, as shown in Figure 2.3 (a), with zero R3 and a large R3. Because the dominant noise source in a TIN3 device is the surface recombination type located between the base and the emitter (Slbe), according to Equation 2.12 the magnitude of S VC can be written approximately as R 2 2 VC =EZ—Lkg2—XSIb . [3.4] (RB +r1r) e 48 10'2 10" — ’01: (Amp) 10-10 _ 10‘14 —1 -' '\ o 1.2 Figure 3.14: Average Gummel plot for a population of 12 devices at various stress times (0 to 500 hours) for devices reverse stressed with V33 = -4 V and V3c = 0 V at 23 C. 49 80 hFE 0 0.2 0.4 0.6 0.8 l VBE(V0“) Figure 3.15: h33 vs. V33 for the data shown in Figure 3.14. 1.2 50 2.5 2 _ Normalized 24°C hrs v,£=o.6v 175 C 1.5 .— 150 c 1 ... l l 1 1 10° 101 102 103 10‘ 10’ Armani; Time (min) Figure 3.16: 1133 recovery of degraded transistors shown in Figure 3.14 as a result of post- stress thermal annealing at temperatures of 150, 175, and 240 C. Complete recovery would correspond to a normalized 1133 equal to 2.75. 51 Consequently, the magnitude of Sjbe can be carried out by measuring 51/0 In agreement with previous studies described in Chapter 2, a significant increase in l/f noise power spectral density after reverse-bias stress was observed, as shown in Figure 3.17. Figure 3.18 showed the noise was reduced almost to the pre-stress level after 240 C annealing for 60 hours. The stress/annealing results for both gain and noise level of the devices in the second group are summarized in Figure 3.19. The 1133 reduced monotonically with time to 60% of its initial value afier 100 hours of stress, and then, recovered to 89% of its original value after 60 hours of thermal annealing. Under the same conditions, the noise level increased to 4.6 times the original level after 100 hours of stress, and after 60 hours of thermal annealing, noise reduced to 1.63 times the original level. It is generally hypothesized that the physical cause of reverse bias stress induced degradation in modern BJT's can be due to two mechanisms. The Si-Si02 interface may be damaged, resulting in an increase in the number of interface states. Secondly, carriers may be trapped in the oxide, either by surmounting the Si-Si02 barrier, or by tunneling through the barrier to trap states in the energy gap of the oxide. From this view, the source of post-stress recovery may be due to (1) annihilation of interface states, (2) change in occupancy of interface states, and (3) detrapping of electrons from oxide traps. However, the thermal annealing results alone do not distinguish among these possibilities. In order to add fiirther insight into the degradation and recovery mechanisms, a means of recovery which does not involve thermal activation was investigated, as described in the next subsection. 3.3.2 Forward-Bias Induced Recovery Two groups of TIN3 devices were used in this study. For the first group, TIN3 devices were degraded by a -4 V base-emitter bias for 300 hours at 23 C without interruption. As for previous cases, the collector-base junction was short-circuited. Then 52 10'" . + : Ohours : . e- : 1hours ‘ ’ -v- : 4 hours _ 4..- : 24hours 10.20 REE“ . + :100hours 515: 21 i r 10' — ‘N. Wm) = . K. 10‘22 2 10‘23 1 10° 10‘ Frequemy (Hz) Figure 3.17: The noise power spectral density measured at various stress times ( o to 100 hours) for a TIN3 device reverse bias stressed with V33 = -4 V and V33 = 0 Vat 23 C. 53 10‘l9 — A 10’20 s 012711sz ‘ 10‘22 — 10‘23 I 10° 101 102 Beam (Hz) Figure 3.18: The noise was reduced almost to the pre-stress level afier 240 C annealing for 60 hours. 1 101 10 Stress Annealing <—-————> 3“ - Normalized SI“ Normalized hp; 8130112 1 q-a—l VBB=0°6V (Plotted as I) .............. (Plotted as U) B ....... —r —1 10 1 1 1 1 1° 0 20 40 60 80 100 160 Tune (hours) Figure 3.19: DC reverse stress and post-stress thermal annealing effects on the normalized 1133 and normalized low-frequency noise levels 51,, (normalized to the PRE- STRESS values). The ratio of the post-annealing 1133 (noise) to post-stress 1133 (noise) is 1.48 (0.35). 55 three post-stress recovery experiments induced by a 0.8 V forward bias were performed. Only 1133 was monitored for the devices in this group. For the second group, both the gain and the noise level were investigated. The stress was again made by a -4 V bias on the base-emitter junction for 100 hours at 23 C without interruption. Then the post-stress recovery by a 0.8 V forward bias on both gain and noise levels was investigated. Again, the collector-voltage noise power spectral density (Sr/C) was measured. In the first group of TIN3 devices, three forward-bias-induced recovery experiments were performed at 23 C. In the first, degraded devices had sequential Gummel plots measured with intervals between measurements of five minutes for a total of ten measurements. During the measurements, the base-emitter junction experienced a forward-bias increasing from 0 V to 0.8 V over a time period of 50 seconds. During the 5 minute time period between measurements the device was not biased. It is noted that the current at 0.8 V is sufficiently small to avoid appreciable self heating of the device since the power dissipated in the device at 0.8 V bias is 0.1 mW. For the emitter dimension for TIN3 devices, the thermal resistance can be calculated as 1.2 C/mW [80]. The second experiment was identical to the first except that the time interval between measurements was 8 hours. In the third experiment, a steady bias of 0.8 V was applied to the base- ernitter junction with the base-collector junction short circuited. In this experiment the bias was interrupted every 5 nrinutes and a Gummel plot was measured for a total of 10 measurements and 50 minutes of forward-bias time. The results of the three experiments are shown in Fig. 3.20. Significant recovery is observed in all 3 cases with no gain change in the control device which had received no previous reverse-bias stressing. Compared to the originally degraded value, the normalized 1133 at V33 = 0.6 V increases by approximately 15% after 10 measurements with a 5 minute interval between measurements. When the time interval is increased to 8 hours, essentially the same result is observed. This indicates that there is not appreciable spontaneous recovery between measurements on these time scales, but that the observed 56 1 .25 At = 5 minutes, bias = 0.8 V Normalized hrs VBE=O.6V Figure 3.20: Forward-bias induced 1133 recovery under various recovery conditions. The measured 1133 of a control device is also shown for comparison. 57 recovery in gain is due to the measurement process. In the third experiment, a larger gain recovery is observed with a 22% recovery. This may be interpreted as being a result of a longer exposure to forward-bias than in the first two experiments. It is also noted that, in contrast to the thermal annealing results, the gain recovery saturates with increasing time. The forward-bias gain recovery does not seem to involve interface states. If the gain recovery is due to a change in population of fast interface states as a result of simply shifting the Fermi level during forward-bias, then it should also be possible to return the states to their original post-stress populations by applying a reverse bias to shift the Fermi level in the opposite direction. Such a repopulation would be expected to occur on a similar time frame as that required for forward-bias gain recovery. However, this was not observed experimentally. Reversing the band bending by a subsequent reverse biasing of a device for 5 minutes does not undo the current gain increase achieved by forward bias gain recovery. Also, the forward-bias induced gain recovery is maintained over a long time period. For example, the 15% gain recovery shown in Fig. 3.20 drops to about 11% after 4 weeks of storage at room temperature. Therefore the forward-bias induced gain recovery is robust with regard to subsequent bias and changes in band-bending, as well as with time. This indicates that the recovery is more likely due to detrapping of oxide trapped carriers. Such detrapping could be caused by field assisted tunneling aided by the oxide electric fields present during forward-bias [79]. However, since only a partial gain recovery (approximately 20%) is achieved and because the gain recovery saturates, oxide trap states would appear to be the cause of only a fi'action of the current gain degradation. In order to investigate the effect of forward-bias on l/f noise recovery, a sequential two-step experiment was performed on TIN3 devices in the second group. In the first step, the 1/f noise of the transistor was measured with the transistor biased at 0.72 V. In the second step, the device is forward biased at 0.8 V for 20 minutes and the 1/f noise is re-measured. Gummel plots are also measured after each noise measurement as illustrated in the test sequence shown in Fig. 3.21. This two-step process was repeated several times 58 G=Gummelmeasurernent(v =OtoO.73Vforlrninute) N = Noise measurement ( V = 0.72 V for 20 minutes ) F = Forward-bias recovery ( VBE=0.8 V for 20 minutes ) V BE G N G F G N G F G N G F G N G Time [1 [2 l3 14 15 [6 1.7 18 Figure 3.21: The test sequence for an investigation of 1/f noise recovery due to post-stress forward bias. 59 for a given device. Ideally, the noise measurements would be made at as low a bias value as possible, so that bias induced recovery does not occur during the noise measurement itself. The value of 0.72 V represents a compromise between this concern and noise measurement sensitivity considerations. The value of forward-bias during the recovery part of the measurement should be higher than the one during the noise measurement, but not so high as to introduce thermal effects. Again, the value of 0.8 V was arrived at as a compromise between these two considerations. Figure 3.22 shows that there is in fact recovery of 1/f noise caused by the forward- bias. The l/f noise for an unstressed device when measured at V33 = 0.72 V is less than our measurement resolution 1023 AZ/Hz. Consequently, as was the case for current gain recovery, forward-bias alone produces only a partial recovery of noise degradation. The experiment illustrated in Fig. 3.21 produces information both on current gain recovery and noise reduction, and the effect of forward bias on both of these quantities is shown in Figure 3.23. It is noted that forward bias alone produces a much smaller gain recovery than thermal annealing. Also as shown in Figure 3.23, a saturated h33 recovery was observed. However, saturation of noise recovery is not observed over the time scale of this study. This indicates that a portion of the recoverable states affects noise properties but not gain. For control devices, which were not stressed, no change in h33 or noise was observed. 3.3.3 AC Stress Based on the forward-bias induced gain recovery observation, it would be expected that AC stress that includes forward-bias cycles should produced different results than AC stress without forward-bias cycles. To investigate the effect of non-DC stressing, transistors were stressed under two conditions. In the first group of six transistors, a 1 KHz alternating reverse-bias to forward-bias (-4 to +1 V) pulse train was applied to the 10'21 -I- : 0min €- : 20mm a. -- : 40min 4 a + : 60mm 1 . S». _22 . 5‘ (A2/Hz)10 - 10-23 I 10° 101 102 Frequency (Hz) Figure 3.22: Low-frequency noise reduction at various recovery time (0 to 60 minutes) due to forward bias applied to the emitter-base junction. 61 7.5 45 6.5 .— 5,“ 5.5 — at 10 Hz h” CKIAJ-zallz/IIZ) 3-4“) I/gg’==(1(5‘/ (Plotted as I) 4.5 —( (Plotted as CD 3.5 .4 2'5 1 1 1 1 1 1 1 1 35 I] (2 13 l4 (5 ‘6 l7 (3 Figure 3.23: F orward-bias induced recovery effects on normalized 1133 and normalized low-frequency noise levels 51,, (normalized to the POST-STRESS values). The ratio of the post-recovery h33 (noise) to post-stress 1133 (noise) is 1.24 (0.66). 62 base-emitter junction with the collector-base junction shorted. The duty cycle was 50% so that the devices were in forward-bias for half the time and reverse bias for half the time. In the second group, an alternating reverse-bias to zero (-4 to 0 V) pulse train was applied so that the devices were unbiased for half the time and reverse biased for half the time. Self heating of the device is in both cases not sufficient to cause thermal annealing effects. With 1 V forward bias on the base-emitter junction, the device power dissipation is 10 mW. The thermal resistance is 1.2 C/mW so the resulting temperature rise of 12 C is well below the temperature required for thermal annealing effects as reported in Section 3.2.1. Previous reports on AC stressing have indicated variable results. In one report, a significant difference in degradation between stress cycles which included forward-bias cycles and stress cycles without forward-bias was reported [6]. However, a second study did not observe any difference between the two stressing modes within the time of study, which was about one hour of reverse—bias-on time [78]. Our results for AC stressing for up to 2000 hours are shown in Fig. 3.24. For the first 100 hours, no apparent distinction is observed between the two stressing conditions. However, with longer stress times, a decrease in degradation rate is observed for the devices receiving forward-bias pulses. After 2000 hours, h33 has degraded to 26% of the original value for reverse bias only stressing, and has degraded to 45% of the original value for stressing which included forward-bias pulses. When comparing these results with those of Section 3.3.2, this would indicate that the difference is due to a lower trapped electron effect in those samples receiving forward- bias pulses. The degradation to 26% of the original value is interpreted as being due to the combined effect of an increase in interface states and an increase in trapped-electrons. The degradation to 45% of the original value is interpreted as being due to an increase in interface states, and fewer trapped electrons. The data also suggests that degradation due to interface states may accumulate more rapidly than does degradation due to trapped electrons. For short stress time, our results show that there is negligible difference 63 1_1 0.8 -2 Normalized hrs v” = 0.6 v 0.6 _ 0.4 — 0.2 l l l l 0 400 800 1200 1600 2000 Stressing Time (hours) Figure 3.24: AC stress results with and without forward-bias (1.0V) cycles. The AC signal was applied to the E-B junction with the B-C junction shorted. The fiequency of the AC signal is l KHz and the duty cycle is 50%. The power dissipation is 10 mW during the forward-bias cycles which will not cause thermal annealing since the device thermal resistance is 1.2 C/mW. between the two stress conditions, which is consistent with [78]. 3.4 Voltage Dependence of Reverse-Bias Stress In previous sections, the reverse-bias experiments were performed under the conditions of V33 = -4 V and V3c = 0 V. In this section, the results of reverse-bias stress under various stress voltages are presented. The stress devices used for this study are TOUT devices. 3.4.1 Experimental Methods The emitter-base junction of TOUT devices was reverse-bias stressed with the collector-base junction shorted at room temperature. A voltage divider circuit designed by [87] and shown in Figure 3.25 was used to perform stress under -3.5 V, -4.0 V and -4.5 V. For -3.5 V and -4 V stress, the total stress time was 200 hours, and for -4.5 V stress, the total stress time wais 24 hours. The stress was interrupted for both noise and I-V measurements at l, 4 and 24 hours for all three stress conditions, and also at 100 and 200 hours for -3.5 V and -4 V stress. A set of three emitter low-frequency noise power spectral densities from 1 Hz to 100 Hz was measured before and after stress. The noise was measured under the common-collector configuration as described in Section 3.2.1. This set of three noise measurements performed under three different bias points allows one to check the current dependence of the noise power (and consequently, the noise source inside the device). It is noted that the forward bias can induce post-stress recovery as described in Section 3.3.2. Consequently, the forward bias on the emitter-base junction during noise measurements should be kept as low as possible to minimize the forward-bias induced effect. However, the forward bias can not be arbitrarily low, or the 1/f noise will be too small to measure. 65 VCC=5V 0 3R1 4.5V 1R2 4.0V 2 = i- < R3 ? 35V _— _ _ fiN‘ Figure 3.25: The voltage divider circuit designed by [87] used to perform reverse-bias stress under various stress voltages. The values of resistors are : R1 = R2 = R3 = 47 Q, R4 = 317 Q, allothers are 100 K52. Under these concerns, the three measurement bias points were chosen as VBE = 0.74, 0.76 and 0.78 V with a constant VBC = -1.5 V. These bias conditions correspond respectively to varying RE = 2 MO, IMO and 510 K!) with a constant RB = 100 K!) used in the common-collector bias circuit with VEE = 36 V and VCC = -1.5 V, as shown in Figure 2.3 (b). The pre-stress noise measurements on TOUT devices under these three bias points show that the noise power has a quadratic current dependence. This indicates that the dominant noise source is the surface recombination type, which is consistent with the results described earlier in Section 3.2.1. Afier each noise measurements, a Gummel plot with VBE s 0.6 V was measured. A set of device I-V characteristics were also measured after the above-mentioned set of noise measurements. The device I-V characteristics are monitored by the following set of three measurements: (1) Gummel plot, [3 and 1C with VBC = 0.0 V and 0 _<_ VBE s 0.6 V, (2) 11380 reverse, stress voltage 3 VBE s 0.0 V, (3) 11580 forward, 0.0 V s VBE .<_ 0.6 V. It is noted that the forward bias on the emitter-base junction during I-V characterization is only 0.6 V, which is expected to cause a negligible effect on post-stress recovery. The experimental flow for this study is summarized in Figure 3.26. 3.4.2 Results The normalized hFE for devices under various stress voltages is shown in Figure 3.27, which indicates the hFE degradation caused by reverse bias stress and also the effects of by; recovery induced by the forward bias used for performing the post-stress noise measurements. It is clear to see that the forward-bias induced recovery plays an important role in this degradation study. For example, the hFE dropped to 83.9% of its initial value after 1 hour of -4 V stress, and recovered to 88.9% of its initial value after 15 minutes of 67 Initial noise measurements under three bias points and initial I-V characterization including a set of three measurements: Gummel plot, 1530 reverse and forward. 4 Reverse-bias stress on TOUT devices with stress voltages of -3.5, -4.0 and -4.5 V for pre-selected stress time interval. .1 i Gummel plot measurement Repcat (V35 5 0.6 V) till three noise measurements - are done. The noise measurement under one of the three bias points. Y I-V characterization including a set of three measurements: Gummel plot, 1530 reverse and forward. Figure 3.26: The experimental flow for the reverse-bias stress under various voltages. 0 SU NQRMQLIZED HFE at UBE 850 700 550 L+00 850 100 Figure 3.27: Normalized hFE at VBE = 0.6 V for TOUT devices stressed under -3.5, 4.0 REUERSE STRESS ON TOUT'S 4:99 :5 u _ 3 5 U 0 0 ~ i—% 00 $—% EU 1 1 1 * 1 10‘ 10° 101 108 103 TIME and -4.5 V. 69 0.78 V forward bias introduced by the first noise measurement, to 89.1% of its initial value after 15 minutes of 0.76 V forward bias introduced by the second noise measurement, and finally to 89.4% of it initial value afier 15 minutes of 0.74 V forward bias introduced by the third noise measurement. It is noted that the data presented in Figure 3.27 is measured at VBE = 0.6 V, which is different from the bias conditions of VBE = 0.74, 0.76 or 0.78 V used for the set of noise measurements. However, a similar hFE behavior will be expected to be observed for those bias conditions. The noise measurements of -3.5 V stressed TOUT device are shown in Figures 3.28 to 3.30. Figure 3.28 shows the noise results measured under the bias conditions of VBE = 0.78 V and VBC = -1.5 V. Figure 3.29 shows the noise results measured under the bias conditions of VBE = 0.76 V and VBC = -1.5 V. Figure 3.30 shows the noise results measured under the bias conditions of VBE = 0.74 V and VBC = -1.5 V. A fluctuation of the noise data is observed under -3.5 V stress. The noise magnitude initially decreases, then increases, then decreases again as stress time increases. The reasons for this behavior are unknown. However, because the degradation rate is much slower in -3.5 V stress than that in either of the other two stress voltages, the forward bias may have a more significant effect on these noise results. The noise measurements of a -4.0 V stressed TOUT device are shown in Figures 3.31, 3.32 and 3.33, which represent respectively the results measured under the bias conditions of VBE = 0.78, 0.76 and 0.74 V with VBC = -1.5 V. Also, the noise measurements of -4.5 V stressed TOUT device are shown in Figures 3.34, 3.35 and 3.36, which represent respectively the results measured under the bias conditions of VBE = 0.78, 0.76 and 0.74 V with VBC = -1.5 V. It is noted, as expected, that a larger noise increase will be observed if a device is stressed under a higher reverse bias. For example, under the measurement bias conditions of VBE = 0.74 and VBC = -1.5 V (Figures 3.32 and 3.35), the noise power at 1 Hz increases to 9.4 times the initial value after 24 hours of -4 V stressed, but to 124.2 times the initial value after 24 hours of -4.5 stress. Sve 7O TQUT(-3 5U STRESS),UBE=O 78U lnrtral 10‘13— 0 1 hr A L1 hrs 4 Y 8L1 hrs X 100 hrs 1 o 200 hrs Y 10‘“+ . . 1 . , 1 10° 101 10a FREQUENCY Figure 3.28: The emitter noise power spectral densities afier different times of -3.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.78 V and VBC = -1.5 V. 71 1041_ TOUT(—3 50 STRESS),UBE=0 76U Sve D lh1t1a1 10_“+‘ 0 1 TWV A R hrs Y 2% hrs X 100 hrs 0 800 hrs 10‘1'5 . es 1 . m° 1m m3 FREQUENCY Figure 3 .29: The emitter noise power spectral densities after different times of -3.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. Sve w'“— 72 TQUTC-B 5U STRESS>,UBE=O 7%U D lhrtral 10”“h 0 1 hr A R hrs Y 8% hrs X 100 hrs 0 800 hrs 10-15 fl 1 fl 1 m° 1w m3 FREQUENCY Figure 3.30: The emitter noise power spectral densities after different times of -3.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. Sve 73 TQUT(-% 0U STRESS).UBE=O 78U 10'9 — 10‘10) A - ‘ e 10'114 ‘ "V \s; a , ‘ \. 6 10-18 _, I...\¢\ D lhltlal 10‘134 O 1 hr A R hrs Y 8% hrs X 100 hrs 0 800 hrs 10-1% ' ' 1 1 w 1 100 101 102 FREQUENCY Figure 3 .3 1: The emitter noise power spectral densities after different times of -4.0 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.78 V and VBC = -1.5 V. 74 TQUT(-% 0U STRESS).UBE=O 76U -10 10 7 Q ) 9 ;\ 10-11 ‘1 '9 -‘. A “ j T\ 9 \, ‘\ Sve 10.“+ ' ' 1 ' ' l FREQUENCY Figure 3 .32: The emitter noise power spectral densities afier different times of -4.0 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. Sve 10‘11 10-12 10-13 10'1'4 10‘15 75 TQUT(-% 0U STRESS),UBE=O 7%U ‘h ‘0 l 1 D Ihltlal 0 1 hr A R hrs Y 2% hrs X 100 hrs 0 800 hrs r r l ‘ FREQUENCY Figure 3 .33: The emitter noise power spectral densities afier different times of -4.0 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. Sve 10-10 10’11 10‘12 10‘13 lo'lLf l -GE-l 76 TQUTC-R 5U STRESS),UBE=O 78U 1h1tlal 1 hr L1 hrs 2L1 hrs l I ”‘1 FREQUENCY Figure 3.34: The emitter noise power spectral densities after different times of -4.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.7 8 V and VBC = -1.5 V. 809 77 TQUT(—L+ 5U STRESS>,UBE=O 76k) m‘m— C 4 6 7‘ é & a 3 10-11“ .9 I.) 1 1* e -12 ‘ \‘3‘ 10 E —‘ .'\ .5. 4‘5 m-m- D lhltral 0 1 hr 1 A L1 hrs Y 8% hrs l ._.1 10'H . . 1 - A A m° m1 m3 FREQUENCY Figure 3.35: The emitter noise power spectral densities after different times of -4.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.76 V and VBC = -1.5 V. Sve 78 TQUT(-% 5U STRESS),UBE=O 7%U 10-101 : 10_11.2. -"- g C. e T 7 ‘s‘ 's‘ e m‘m— ‘ _. 3" e e 0 10‘13— ‘ ~‘ 3 lhltlal 10‘1”“ 0 l tar A R hrs Y 8% hrs 10-15 .T I l r Y 1 m° 1w m3 FREQUENCY Figure 3.36: The emitter noise power spectral densities after different times of -4.5 V stress. The noise measurements are performed under the common-collector configuration with the bias conditions of VBE = 0.74 V and VBC = -1.5 V. 79 3.4.3 Discussion When the bipolar junction transistors in this study are subjected to a sufficient reverse-bias stress, hot-carrier induced degradation is clearly observed. It is generally believed that the physical cause for such degradation is due to one of two mechanisms. The Si-SiOz interface may be damaged, resulting in an increase in the number of interface states. Secondly, carriers may be trapped in the oxide, either by surrnounting the Si-Si02 barrier, or by tunneling through the barrier to trap states in the energy gap of the oxide. In the second case, the occupancy of previous existing interface states can be modulated by band bending and a shifted Fermi energy. In either case, the resulting change in the effective number of interface states causes a change of the surface recombination velocity, and consequently changes of the BJT characteristics. For example, both the base surface recombination current Ts and the surface recombination type 1/f noise S1,.“ can be increased after a hot-carrier stress, as shown in Section 3.3 for TIN3 devices and 3.4.2 for the TOUT devices. The surface recombination current is proportional to the surface recombination velocity and to the carrier concentration at the surface. The surface recombination velocity [64] is equal to ov,,,N,-,, where o is the carrier capture cross-section with a value around 10‘16 cm2 depending on the type of carriers [90], v”, is the thermal velocity which is equal to 107 cm/s at room temperature, and Ni, is the surface state density which can be increased by the hot-carrier stress as described earlier. Consequently, we can write the surface recombination current as the following fiom [89] 7. =qvm(a..An.+opAp.)AN.-.. (3.5) where Ans and Aps are the electron and hole concentrations at the surface respectively whose magnitude will depend on the bias condition, and 0,, and 0,, are the carrier capture cross-sections for electrons and holes respectively. It is noted that the surface 80 recombination current is linearly proportional to the surface state density N”. As described in Section 2.1.1, the surface recombination l/f noise has the following form =.01(1s)2 rec _f AN“ 1 (3.6) By substituting Equation 3. 5 into Equation 3. 6 and also by assuming 0,, ~ op= o, the surface recombination 1/f noise can be rewritten as 0.1 51,,- -—f-xq v,,,ol(An. Mp.) AN... (3.71 which also will be linearly proportional to Nit. Consequently, for a device subjected to a reverse-bias stress, the resulting increase of the surface recombination current (211;) and the resulting increase of the surface recombination 1/f noise (ASIrec) are both linearly proportional to the resulting increase of the surface state density (ANH). In the other words, the ratio between AS396 and ATS will have the following form ASI 0.1 AIS f xqvthop ( ) with 2 =(AnS1 +Aps1) (3 9) (A1132 +Aps2) and is independent of AN“. By plotting ATS versus AS1rec, a straight line with a slope equal to the right hand side of Equation 3.8 is expected. In general, the bias condition used for performing the noise measurement to obtain ASI,“ can be different from that used for performing the current measurement to obtain ATS. Consequently, in Equation 3.9 Ans 1 and Aps, correspond to the surface carrier concentrations under the bias condition used for the noise measurement, and Ans, and Ap52 correspond to the surface carrier concentrations under the bias condition used for the current measurement. 81 The above argument, which claims that the ratio between AS1rec and ATS is independent of AN,-,, can be verified by plotting the results of TOUT devices subjected to a reverse stress under various voltages as described in the previous section. Figure 3.37 plots the increased l/f noise at 1 Hz under the bias conditions of VBE = 0.74 V and VBC = -1.5 V, versus the increased 13 measured under the bias conditions of VBE = 0.6 V and VBC = 0 V (Gummel plot measurement) which is known as the surface recombination current (A75). The l/f noise shown here is the surface recombination 1/f noise located between the emitter and the base, SIbe, which is calculated by using Equation 3.3 from the measured results of the emitter noise power spectral density under the common-collector configuration (Figures 3.33 and 3.36). The stress conditions for obtaining these data are with VBE = -4 or -4.5 V and with VBC = 0 V. It is noted that, as shown in Figure 3.37, the increased base current (A13) and the increase surface recombination current (AS Ibe) can be expressed by a linear relation as LXI—1;" =a , (3.10) where a is obtained as 2.14x10"14 A/Hz by using the least-squares fitting the experimental data. If we substitute some numerical values into the left hand side of Equation 3.8, i.e. f = 1 Hz, q = 1.6x10'19 coul, v”, = 107 cm/s, o z 10'16cm2, and assume p is unknown, by using a = 2.14x10'14 A/Hz the magnitude ofp can be estimated as 1.34x1015 cm3. It is also worth noting that the above linear relation between A13 and ASIbe does not depend on the stress voltages. The values of A] B and AS139 and the corresponding stress voltages and times are shown in Table 3.1. delta (X10T81) 82 TOUT 8 00‘ 0 1 71'4 / / / 1a3~ // / / / / / 114— // / / /o / 857 — // m,’ / X o 5713 // / / l/ O -% 5U Stress em6~ m , m / / / / U / 000 E3 1 1 1 T r fi 000 l 33 8 67 S 00 5 33 6 67 8 00 delta <><10‘8> Figure 3.37: 4511).; versus A13 for TOUT devices stressed under -4.0 and -4.5 V. The base current and the noise power are both measured under the bias conditions of ‘VBE=WIWIV2md‘&Kf=-L5\L 83 AIB (nA) Asa. (Io-23A2/Hz) Stress History 2.4 2.41 -4 V for 1 hr 6.2 25.33 -4 V for 4 hrs 13.0 11.02 -4Vfor 24 hrs 25.8 24.55 -4 V for 100 hrs 32.9 76.50 -4 V for 200 hrs 34.5 64.18 -4.5 Vfor 1 hr 47.5 96.78 -4.5 V for 4 hrs 73.6 175.38 -4.5 V for 24 hrs Table 3.1: The values of A13 and “Mg used in Figure 3.36 and the corresponding stress history. CHAPTER 4 FORWARD-BIAS STRESS RESULTS 4.1 Device Description Three kinds of silicon n-p-n transistors fabricated with different technologies by IBM were used in this forward stress study, namely TIN3, TSEO3I and the Hemmert test structure. TINB's were fabricated by conventional oxide isolation technology [81], as described in Section 3.1, with emitter dimensions of 2.5x4.3 umz. TSEO3I's were advanced technology devices, fabricated with polysilicon base contacts and side-wall oxides, which had emitter dimension of 1.5x2.0 umz. Metal contacts to these devices were not by first level metallization, but rather by means of metal studs through oxide vias. Such studs serve to reduce electromigration induced stress directly over the transistor contacts. The Hemmert test structure is a test structure reported by Hemmert [77] which was designed particularly to investigate the electromigration induced compressive stress efl‘ects. As shown schematically in Figure 4.1, seven identical transistors with emitter dimensions of 2.5x4.0 umz have their emitter centered under and connected to a long conductor which is 4 pm wide and 2 mm long. Only one transistor on the end of the series of seven is electrically forward-bias stressed. As described in Chapter 2, a metal accumulation can occur above the emitter region of this electrically stressed transistor because of electromigration, and consequently a compressive stress can be built-up around the metal accumulation region. Since the forward-bias-stressing current does not flow through the conductor that contact the other six emitters, any change in the characteristics 84 85 BASE METAL 6 5 4 3 1 ENH 1 ”ER METAL EMITI'ER‘ CONT ACT 4...: 4.1. 1 - ELECTRON :f FLUX Figure 4.1: A schematic of the Hemmert test structure. Transistor #1 to #6 are separated equally by a distance of 10 um. Transistor #7 which is 100 um away from transistor #1 is not shown here. 86 of the other six transistors is attributed to electromigration induced compressive stress effects in the common conducting line over the emitters. These seven transistors essentially act as pressure sensors which sense the compressive stress along the long conductor at 0 (region of metal accumulation), 10, 20, 30, 40, 50, and 100 um from the region of the metal accumulation. All test dies from three kind of transistors, TIN3, TSE03I and the Hemmert test structure, were mounted on ceramic DIP packages. As described in Section 3.2. 1, the dominant l/f noise source of TIN3 devices was verified to be the surface recombination type located between the base and the emitter. The verification was done by checking the current dependence of the noise power spectral density measured under the common collector configuration. The same verification method was also performed on the new technology device TSEO3I. Figure 4.2 shows a set of three measurements of the noise power spectral density on a TSE03I device under different bias conditions: VBE = 0.77, 0.80 and 0.82 V with a constant VBC = -1.5 V. These bias conditions correspond respectively to R5 = 510 KO, 200 KS), and 100 K!) with a constant R3 = 100 K!) used in the common-collector bias circuit, referring to Figure 2.3 (b), with VEE = 36 V and VCC = 1.5 V. Figure 4.3 shows the quadratic current dependence of the noise power at 1 Hz. However, as described in Section 3.2.1, both the surface recombination 1/f noise (Slbe) and the base resistance fluctuation 1/f noise (Srb) have the quadratic current dependence. Consequently, in order to distinguish these two noise sources inside a TSE03I device, a fourth noise measurement has to be made, which will has the same bias conditions as the third noise measurement (i.e., VBE = 0.82 V and VBC = -1.5 V), however, with a different combination of RE and RB. If the result fi'om the fourth noise measurement is the same as that from the third noise measurement, then the base resistance fluctuation l/f noise (Srb) is dominant. However, if the result from the fourth measurement is changed and proportional to R32, then the surface recombination l/f noise (S1,,e) is dominant. Figure 4.4 shows that the noise power spectral density of the fourth noise measurement is decreased to a quarter of the magnitude obtained in the third 87 TSE031<#79) m‘m— A: Ube=0.82U 10*1— 0: Ube=0.80U m: Ube=0.77U Sve 10-13 10-1% '15 10 "T‘frl ‘1"V"1 Figure 4.2: A set of three measurements of the emitter noise power spectral density on a TSEOBI device under different bias conditions: VBE = 0.77, 0.80 and 0.82 V with a constant VBC = -1.5 V. TSEOBI<fi79> 10-11l 10'”31 m > on J 10‘V31 10.“+ ll 1 1 10‘5 10‘” *0‘3 Figure 4.3: The quadratic current dependence of the noise power at 1 Hz for a TSEO3I device. TSEO3I<fi79> Ube=0,880 H II t—AO KO E70 UU 10‘”- Y: Sve 100 101 103 :33 10‘+ 105 FREQUENCY Figure 4.4: Two measurements of the emitter noise power spectral density on the same TSEOBI device with its noise levels as shown in Figure 4.2. These two measurements are made under the same bias conditions, VBE = 0.82, and VBC = -l.5 V with a different combination of R3 and RE. 9O measurement when R3 is decrease to half of the value used in the third measurement. The results confirm that the surface recombination S Ibe is the dominant noise source in TSEO3I devices. 4.2 Experimental Methods Transistors from the above-mentioned three technologies were stressed with the base- collector junction short-circuited and the base-emitter junction forward biased. The stress was interrupted at the pre-selected times to measure the I-V characteristics and, for some devices, to measure the noise characteristics additionally. Device measurements were performed at a temperature of 23 C i 1 C in an environmental chamber. The method for I- V characterization was described earlier in Section 3.3.2. However, in this forward-bias stress study the I-V measurements with reverse-biased junctions (Icao reverse and 1530 reverse) were not performed in order to eliminate the possibility of degradation due to reverse-bias produced hot electrons. Consequently, the only potential source of hot electrons are those produced by Auger recombination during forward-bias stress as proposed in [1]. The noise measurements were performed under the common-emitter configuration or under the common-collector configuration to obtain the collector noise power spectral density (SVC) or the emitter noise power spectral density (S VE), respectively. Since the dominant noise source has been verified as the surface recombination type located between the emitter and the base (SIbe), the magnitude of Slbe can be carried out fi'om either SVE or SVC by using the equations described below 3V, =[rn/fl-(rb mars]... (4.1) 2 2 2 R VC =-———B B RC2 XSIb . (4.2) (R8 +r1t) e The above equations are also shown as Equations 3.3 and 3.4 in Chapter 3. 91 The stress conditions used for this forward-bias stress study varied a number of parameters including stress time, stress temperature and stress current. All devices were stressed with the base-collector junction short-circuited and the base-emitter junction forward biased. Post-stress thermal annealing was also performed on selected devices. These stress conditions and post-stress annealing conditions are listed separately below for the three technologies. It is noted that the symbol 15 represents the stress emitter current density. (I) TIN3 (A) Current gain degradation study: (1) IE = 105 A/cm2 at -75 C for 500 hours; (2) 15 = 105 A/cm2 at 175 C for 500 hours; (3) JE = 105 A/cm2 at 240 C for 750 hours; (4) JE = 2 x r05 A/cm2 at 240 C for 500 hours. (B) Noise characteristics study: (1) JE = 2 x 105 A/cm2 at 240 C for 100 hours; (2) JE = 2 x 105 A/cm2 at 23 C for 100 hours, and then IE = 2 x 105 A/cm2 at 240C for 66 hours; (3) JE = 2 x 105 A/cm2 at 240 C for 66 hours, followed by annealing at 240 C for 60 hours. (11) TSE03I (A) Current gain degradation study: (1) 15 = 2 x 105 A/cm2 at 225 C for 2400 hours; (2) 15 = 5 x 105 A/cm2 at 225 C for 500 hours; (3) 15 = 5 x 105 A/cm2 at 175 C for 2500 hours. (B) Noise characteristics study: (1) JE = 5 x 105 A/cm2 at 225 C for 250 hours; 92 (2) JE = 5 x 105 A/cm2 at 225 C for 90 hours, followed by annealing at 225 C for 90 hours. (111) The Hemmert test structure (A) Current gain degradation study: (1) JE = 5 x 105 A/cm2 at 200 C for 500 hours. (B) Noise characteristics study: ( 1) JE = S x 105 A/cm2 at 200 C for 500 hours. Control devices were also used which received temperature stress, but not electrical stress. In addition, desk control devices were used which received neither electrical or temperature stress, and were used to monitor the repeatability of the test procedures. The results of stress effects and post-stress thermal annealing effects on both gain and noise are shown in the following sections. 4.3 Stress Effects on Gain TIN3 devices showed a pronounced and systematic gradual degradation of gain with increasing stress time when subjected to a combination of high forward-bias current and temperature. Fig. 4.5 shows the ratio IC/IB, or hFE, versus VBE and stress time as obtained from the Gummel plots. The data represents an average of 12 devices which were stressed simultaneously. However the gradual decrease with stress time is representative of the degradation of the individual devices in the population as well. This degradation is a result of an increase in base current and not a change in collector current. For VBE = 0.6 volts, as shown in Fig. 4.6, the resulting gain degradation after 500 hours of stress at 240 C is approximately 20% when the stress current is 2 mA/um2 and approximately 10% after 750 hours when the stress current is 1 mA/umz. Stress was also performed at the latter current with ambient temperatures of -75 C and 175 C; however, 93 80 hrs 40- 20— iiififiil. a 5 0 0.2 0.4 0.6 0.8 l 1.2 VBE(V0”) Figure 4.5: Current gain vs. V33 for TIN3 transistors at stress times of 0, 1, 4, 24, 100, 200, 500 hours. The forward stress was done at 240 C with a current density of 2 mA/umz. 1.1 + l mA/tlm§ at -75 C G- : l mwz at 175 C -v- 2 1 "TA/«UN: at 7A0 C + :2m um at240C 1 Normalized « hrs VBE=O.6V 0.9 -— 0.8 l 1 10° 101 102 103 Stress Time (hours) Figure 4.6: Normalized current gain (average current gain after stress/average current gain before stress) of TIN3 transistors vs. stress time at various current levels and ambient temperatures. 95 no degradation in gain was observed during the 500 hour time period of the test. The gradual degradation and the temperature dependence is consistent both with electromigration induced compressive stress and with Auger hot electron induced damage. However, as discussed later, comparison with the Hemmert test structure indicates that electromigration is present and at least partially is responsible for the behavior shown in Figures 4.5 and 4.6. For TSEO3I devices, the degradation was quite different than for TIN3 devices. Figure 4.7 shows the result of forward-bias stress with 5 mA/um2 at a temperature of 225 C. No systematic change was observed up to 100 hours, but thereafter degradation in the population of 10 devices was observed. In contrast to the results for TSEO3I technology, the degradation for a given device was precipitous rather than gradual. Gummel plots of devices which had begun to degrade showed a increased series resistance such that more voltage was required to maintain a given current. Once degradation had begun, in about 30 to 40 hours the process culminated in an open-circuited transistor. This phenomenon is shown in Figure 4.8. The time required for a 50% drop out of the population was approximately 400 hours. The nature of the failure is indicative of electromigration, with openings in the line initially producing an increase in series resistance until a line is completely opened, finally producing an open circuit. Scanning electron microscopy (SEM) confirmed this conclusion. Further evidence of electromigration failure in TSE03I devices is found by considering the results when the devices were stressed at 225 C with a lower current density of 2 mA/umz. For this stress condition, the time required for 50% drop out was approximately 2400 hours. This indicates a degradation phenomena which proceeds as the current density squared, consistent with electromigration [17]. However for TSEO3I devices there is no evidence for gradual gain degradation due either to electromigration induced stress or to hot electrons. 200 L 150 ‘ hrs vggzoov ‘00 “ so _ '0 r r 10° 101 102 103 Stress Time (hours) Figure 4.7: Current gain of individual T813031 transistors vs. stress time for a stress condition of 225 C and 15 = 5 MW. 97 END STRESS on TSEOBI (T=885> 10-8 1 10'3 j 10“+ : 10‘S — i 10'6 j 10'? 9 m i T 10'8 j (.J -i 10-9 .4 10-10 3 ‘ [I]: 800 hr S 10%1e %“ O: 370 hVS . 43,21.“ , . A: 39Ht hr S rose. W5 hrs ~ ,, i .«0' ' «257’ X: L+13 hrs _ Knish.” rgi,..’:'.¢'“-' i, ": yr; 10 131‘”, or. Ioootlltmmmnflali; lib." "I”, I?“ i illlimiimwwmr flilli ...... -. ........ T l 00 0 800 400 600 1 00 1 20 UBE Figure 4.8: Gummel plots for a T813031 device (module #88) showing the onset of high resistance and eventually an open circuit during stress with an emitter current density of 5 mA/ltrn2 at 225 C. 98 Our gain degradation studies on the Hemmert test structure, as shown in Figure 4.9, confirm the results of Hemmert [77]. The first transistor which received the direct electrical stress shows the largest degradation in hFE after 500 hours of stressing with 5 mA/um2 at 200 C. For the other six transistors, the gain degradation decreases with increasing distance from the first transistor, consistent with expectations for electromigration induced compressive stress. It is noted that since these six transistors received no electrical stress, the possibility of hot electrons is ruled out. For all the transistors in the Hemmert test structure, the decrease in gain is due to an increase in base current. The magnitudes of the current gain vs. stress time at VBE = 0.6 V for all seven transistors of the Hemmert test structure are listed in Table 4.1. 4.4 Stress and Thermal Annealing Effects on Noise The low frequency noise characteristics were also measured for the three different bipolar transistor technologies. The stress condition for the Hemmert test structure was with a current density of S mA/um2 at T=200 C for up to 500 hours, and the noise levels were measured using the common emitter configuration. Transistor #1, through which the current flowed, showed a low frequency noise characteristic as shown in Figure 4.10. The noise initially dropped for the measurement taken after one hour of stress. Subsequent noise measurements taken at 4, 24, 100, 200 and 500 hours of cumulative stress showed a noise level that increased with stress time. The noise characteristics of transistor #2 which experiences a mechanical stress change only, i.e., no current flows through this transistor, showed a similar behavior with the noise decreasing during the first 24 hours of stress, as shown in Figure 4.11. The noise measurements at 100 and 500 hours of cumulative stress showed the noise level to be increasing with stress time. Transistor #3 showed the lowest noise level after 100 hours of stress, as shown in Figure 4.12. Figures 4.13 to 4.16 shows the noise behavior of transistors #4 to #7, respectively. The overall noise behavior of the 1 9 2 Normalized hrs v35=0.ov 0'5 " 1 '0 I I 100 101 102 103 Stress Time (hours) Figure 4.9: Normalized current gain (current gain after stress/current gain before stress) for transistors #1, 2, 3, and 6 of the Hemmert test structure vs. stress time. The stress was done at 200 C with 15 = 5 mA/umz. 100 Transistor Number #1 #2 #3 #4 #5 #6 #7 Initial 77.1 78.3 78.2 84.0 80.70 76.0 80.9 1 hr 77.9 81.5 81.5 83.8 82.4 78.6 82.5 4 hr's 67.4 83.2 81.1 84.2 81.6 80.2 83.1 24 hr's 61.2 80.9 81.5 83.8 81.5 81.3 81.1 100 hr's 53.9 78.3 77.3 81.8 82.0 80.8 83.6 200 hr's 57.6 76.8 76.9 80.3 82.2 80.0 84.3 500 hr's 4.88 69.8 75.0 76.4 79.4 77.1 83.4 Table 4.1: The magnitudes of the current gain vs. stress time at VBE = 0.6 V for all seven transistors of the Hemmert test structure. 101 Hemmert's device (#1) 10-18 w‘” D3 1h1t1al 01 1 hr A: S hrs 10_20 Y1 ELi ru's X= 100 hrs °= 800 hrs +3 500 hrs 1 0 -81 a; .D 10 -22 £0 10‘23 «- - ..«I‘ '53:»- 10 ‘8“ .— v vEg‘fifi‘ e;f' . st~ ‘ .151! 10‘85 r 3 10-26 1 ' I T I r I r I 100 101 108 103 10% 105 FREQUENCY Figure 4.10: The noise power spectral density for transistor #1 of the Hemmert test structure. 10”8— 1040; 10—21 I 10'22 — Slbe 1083— 10-29 _ 10-86 Hemmert's devlce (#8) 1h1t1a1 1 hr H hrs 8% hrs 100 hrs 200 hrs 500 hrs -9<>XOEl Y ' l ' ' l ' ' l ‘ ' l ‘ l 101 103 103 10‘+ :05 FREQUENCY Figure 4.11: The noise power spectral density for transistor #2 of the Hemmert test structure. 103 Hemmert's dev1ce (#3) m‘ms 10493 D: 1h1tlal 0: 1 hr A: % hrs _p_0 Y: 8% hrs 10 A X: 100 hrs l. 0: 800 hrs " i: 500 hrs 10-21 10-22 Slbe 10‘23 10-2'4 10-85 10‘26 I‘FII’I"I"1 100 101 108 103 1.0“+ 10S FREQUENCY Figure 4.12: The noise power spectral density for transistor #3 of the Hemmert test SENKnuflL 104 Hemmert's devrce (#H) 10*8— 10493 B: Ihltlal . 0= 1 hr J A: S hrs Y: 8% hrs 10720—4 . x. 100 hrs 0: 800 hrs 4‘ 500 hrs 10£1~ Q) ‘i 1: 10-22 fl m 4 0—23 _ i i 0’394 i J 0'35 l 13-86 ' ' I . * I I . I I ' I ' s I 100 101 103 103 10‘+ 105 FREQUENCY Figure 4.13: The noise power spectral density for transistor #4 of the Hemmert test structure. 10-18 10-19 10-80 10-81 10-88 Slbe 10-23 10‘29 10'85 10‘86 105 Hemmert's devlce (#5) — q q [I] ' I T 0 ° C -1 A ° 0 Y ' C —1 O X . .J ‘ o 9 oz" 9 - o . 3‘s. ’« a 0‘ u D was R'e“0.'\ 4‘ ° .‘e‘ “.93... o ., "‘8 \r.‘ -. s.“""“= "_ — "-‘-‘"“"‘~ We.‘ ’ fi \ D' \r ‘ é‘. ‘ ~ .4" ‘ \' " \~..;‘I ;. o- 2.5.: "K‘ I “ ’. “1‘ .‘ ‘.\\ ‘:_-;‘e‘ s ?--'. :0 ‘Q —‘ “AS-s. ". Q-Mh-n S; s“? u .\ _ "z\ x. ‘ ~ .‘ "e \“ . - -‘ 0‘ \ ‘ \ " 5 e‘ W ' «f: 0%.?! \{3‘ of“. \ '3 m —< . ~‘ .\~ ”Lu“... ‘. k‘t:‘\ ‘ b" .—_\‘:-. " ‘9:- :‘ xx '1 °' ‘3 v. “ ’. ‘. ’\ -8.."~.,"-.‘.‘-: ‘ \ ‘~. \'.( ".\ _. 9 -.\~-.. . ‘ .‘5. - V V v‘“‘ V '- —4 .J —4 1 4 I I I Initial 1 hr % hrs 8% hrs 100 hrs 800 hrs 500 hrs m0 Figure 4.14: The noise power spectral density for transistor #5 of the Hemmert test structure. 106 (#6) Hemmert's device .1 SSS a Srrr l Srhhh CVFh lhh 000 D. u-UQO lineal-9:3 DOAYXOA. .3 Av ._+ AU We S .16 EREQUENCY The noise power spectral density for transistor #6 of the Hemmert test Figure 4.15 SUUCIUI‘C. 10~18 10'19 10-20 10-21 10-88 Slbe 10'23 10‘2“} 10-85 10‘86 107 Hemmert s device (#7) Initial 1 hr % hrs 8% hrs 100 hrs 800 hrs 500 hrs $0X—(DOB O I \ a ‘ .: ‘ A 1,1 r .39. V (v..- ‘.:::3 ...1 V“_L‘.s‘~fif\‘ .- '3' -‘.-“- lay": .3 V $3.23,», "' ifs’cg‘r“ '3'... “ I ms, .53.»: ‘ 8 ° '1 19%;5. 25.9 \.- ~ _ \ V ‘ v V 4 ..._ g n". ,‘w‘ .. Vc‘- '.. t. —4 ~ . 4 oil . - \ 4 J - 4 a". q r 1 r T . I 100 101 ‘03 I03 I0“ 105 A A Figure 4.16: The noise power spectral density for transistor #7 of the Hemmert test structure. 108 Hemmert test structure, as shown in Figure 4.17, is an initial noise decrease and a subsequent noise increase as the transistor starts to show substantial gain degradation. These noise changes are produced in the case of transistors #2 and #3 exclusively by changes in the mechanical stress in the transistor structure since no electrical current flowed through either of these transistors. Experiments performed on both TIN3 and TSEO3I transistors also showed similar noise characteristic changes due to forward bias stress. TIN3 transistors showed an initial decrease in their low frequency noise characteristics when stressed at 240 C with a current density of 2 mA/umz. Figure 4.18 shows the noise power spectral density for pre-stress, 66 and 100 hours of cumulative stress time. It is significant that the observed noise decrease during the first portion of the stress (66 hours) required both the high temperature and the high current density conditions. This was confirmed by stressing the transistors with no current and just high temperature. This experiment did not reduce the noise significantly. In fact, if transistors with the noise previously reduced by applying both high current densities and high temperatures are subjected to just a high temperature of T = 240C (post-stress thermal annealing) the noise is seen to increase as seen in Fig. 4.19. Fig. 4.20 shows that the application of a current density of 2 mA/pm2 at 23 C also does not produce a significant noise decrease. Hence, it is the combination of both the high current density and high temperature that produces the noise reduction. For TIN3 devices the noise reduction occurs for the initial portion (66 hours) of the forward bias stress. As shown in Fig. 4.18, the noise increases after longer stress times, i.e. 100 hours cumulative stress. This same behavior was observed as presented above for the transistors of the Hemmert test structure. For both TIN3 devices and the transistors of the Hemmert test structure, the noise starts to increase, afier its initial decrease, at about the same time the gain begins to drop substantially. TSEO3I transistors also showed a drop in the low frequency noise characteristics as seen in Fig. 4.21. This technology did not show the later increase in the low fiequency noise as the other Normallzed N01se Power 109 m3 _ i m1 e / IN E Tr#1 O Tr#8 4 A Tr#3 10 Y Tr#% X Tr#5 0 Tr#6 e Tr#7 I 109 . . I . . I , . I . ..j (a) 10‘1 100 101 103 C) Stress Time Figure 4.17: Normalized base current noise power at 1 Hz vs. stress time for the seven transistors of the Hemmert test structure. Transistor #1 is the only transistor subjected to a electrical stress. 110 10-20 + : original -9- : 66hours 1041 7 + : 100 hours c 5": -22 (AZ/Hz)10 ‘ 10-23 _ 10‘“ I 10° 101 102 Frequency (Hz) Figure 4.18: The noise power spectral density of a TIN3 transistor decreases after a forward-bias stress, reaches a minimum and subsequently increases again. The stress was done at 240 C with IE = 2 mA/umz. 111 10-20_ +:original «948113811158ng -9-:afterannealing U : ‘ 104‘- . a A r 3 Slb - _ = ‘ _, (AZ/HZ) - - ’ _ _ . _ E ‘- 10‘22— ’ - - . _ _‘=\_‘ 10-23 I 1 10° 101 102 Frequency (Hz) Figure 4.19: The noise power spectral density of a TIN3 device decreases after stressing with .15 = 2 mil/um2 at 240 C for 66 hours. Also shown is the post-stress thermal annealing increase of the noise levels toward its original values. The annealingwasdoneat240Cfor66hours. 112 10-20 _ a- :original 9 :23 C, 100 hours -.- : 240 C, 1 hour ...- : 240 C, 65 hours 10-21 I 22 (A 2am ‘0’ ‘ 10-23 _ 10'“ I 10° 101 102 Frequency (Hz) Figure 4.20: The noise power spectral density of a TIN3 device decreases after stressing with 13 = 2 mA/trm2 at two difl’erent temperatures: first, 23 C for 100 hours, then, 240 C for 66 hours. 113 10'24 +: original +: after stressing ~94 after annealing 10-25 ‘ Slb (A 2IHZ) 10-26 __ 10'” I 10° 101 102 Frequency (Hz) Figure 4.21: The noise power spectral density of a TSEO3I device decreases after stressing at 225 C with JE = 5 mA/um2 for 90 hours. The noise level increases after a post-stress annealing at 225 C for 90 hours. 114 technologies. It is worth noting that TSEO3I devices did not show a gradual gain degradation as did the devices of the other two technologies, rather TSEO3I devices failed due to open circuits caused by electromigration. As also shown in Figure 4.21, a post- stress thermal annealing (225 C) caused the noise levels to increase toward the pre-stress levels, which is consistent with the annealing results obtained on TIN3 devices. 4.5 Discussion Gradual gain degradation due to base current increases was observed to occur as a result of forward bias stressing in two of the three technologies tested. T SEO3I devices did not show a gradual gain degradation, rather, a relatively abrupt failure occurred due to an open circuit produced by electromigration. The failures of TSEO3I transistors were verified as electromigration open-circuits using SEM examination. The TIN3 devices which had a gradual gain degradation were cross-sectioned for SEM examination. Any electromigration occurring in this technology was not large enough to be observed under SEM examination. As described in Chapter 2, previous studies of gain degradation due to base current increases under forward bias stress have attributed the effects to electromigration[18,77] or to Auger-recombination produced hot electrons [1]. Both of these phenomenon would show a gradual drop in the gain as the degradation proceeds. Also, both of these phenomenon would be expected to increase as the ambient temperature increases. The recent study of Hemmert [7 7] demonstrates that the gain degradation can occur due to electromigration as evidenced by the gain changes induced in devices receiving no electrical current but with forward bias current flowing through a neighboring device. The results of the present study on the Hemmert test structure, as shown in Section 4.4, indicate that electromigration also produces low frequency noise changes that can initially decrease the noise followed in some technologies by noise increases. This noise decrease is distinct from hot electron noise changes since hot 115 electrons produce noise increases[88]. The decrease in noise phenomenon was observed in all three technologies studied. This unique feature of the low frequency noise decreasing, at least initially, due to electromigration, provides an additional indicator as to whether the cause of gain degradation under forward bias stress in a particular technology is due to electromigration or hot electrons. A forward-bias degradation model which uses the concepts of dislocation movement and energy band narrowing both caused by electromigration induced compressive stress is presented in next chapter. This model gives a good explanation of the observed noise and gain changes after a forward-bias stress. CHAPTER 5 A FORWARD-BIAS STRESS INDUCED DEGRADATION MODEL As shown by the experimental results in Chapter 4, the forward bias combined with a high temperature can produce gradual changes on BJT's gain and noise characteristics. Namely, the low-fiequency noise initially decreases with the gain remaining unchanged; however, when the gain begins to decrease as the stress is continued, the low-fi'equency noise starts to increase. It is concluded that electromigration induced compressive stress plays an important role in these observed phenomena. The goal for this chapter is to propose a physical model, which can explain how such a compressive stress can cause the observed changes of the noise and gain characteristics. In Section 5.1, the background material to be used in developing the model is reviewed. There are two classes of physical phenomena which affect device performance that may result fi'om the application of compressive stress to a p-n junction. The stress may cause dislocations present in the crystal to move. Secondly, the application of mechanical stress can modify the energy band and change effective mass, lifetime, and mobility of carriers. Then, the model is presented in Section 5.2. 5.1 Background 5.1.1 Effects of Emitter-Edge Dislocations on BJT Low-Frequency Noise Dislocations in crystals refer to the line defects. For example, one kind of 116 117 dislocation can be considered as an extra plane of atoms partially crowded into the crystal. Another example of dislocations can be thought of as produced by cutting the crystal partway through and shearing it parallel to the edge of the cut by one atom spacing. The different kinds of dislocations can appear in the same crystal simultaneously. The dislocations can be easily observed by the etch pitting method. The etch pitting method requires a special ”dislocation" etchant, e. g. the Sirtl etchant [82], to reveal the locations of the dislocation lines. After a sample have been etched in such etchant, washed and dried, small pits can be observed on the surface of the sample by a microscope. These small pits indicate the points where dislocation lines cut the surface. Other methods are also used to observe dislocations, such as X-ray topography, scanning electron microscopy (SEM), or transmission electron microscopy (TEM). It is well known that dislocations can affect the electrical properties of semiconductors. For example, the low-frequency noise is one of the sensitive parameters to the effects of dislocations. In 1969, Green et al. [44] investigated the noise of planar p- n junctions affected by dislocations. In this study, dislocations were introduced by a bending procedure which caused mechanical stress, and the densities and the orientations of dislocations were monitored by both SEM and TEM. The investigated devices were divided into two groups. In the first group, the bending procedure was designed such that the dislocations emerged at the device surface near the junction. On the other hand, in the second group the bending procedure was altered such that the dislocations emerged at the device surface away fiom the junction. The densities of dislocations in both groups were varied. The low-frequency l/f noise spectra were then measured on both groups. The results from the second group showed that the 1/f noise did not change although the dislocation densities in the device were changed. However, the l/f noise of the first group was increased with the increase of dislocation densities. There is a very important conclusion drawn in this study: the low-fi'equency l/f noise increases with the increasing of the dislocation density ONLY when the dislocations emerge at the surface of the device 118 near the junction. In other words, only those dislocations at the surface of the device near the junction will affect the low-frequency 1/f noise properties. For a given silicon wafer, it was shown in another study by the etch-pit evaluation technique that phosphorus diffusion was also able to create dislocations [45]. It is believed that in this case an internal stress is induced by the high concentration of dopants, and that this internal stress is partially relieved by the generation of dislocations. For a n—p—n transistor, generally two kinds of dislocations can be generated during the emitter diffusion process. The first kind is the diffusion-induced dislocations [47], which appear inside the diffusion emitter area. The second kind is the emitter-edge dislocations [48], which appear around the edge of diffusion emitter area, usually after the emitter oxidation step. The density of each kind of dislocations depends on the emitter phosphorus-diffirsion conditions [49,54]. However, the densities of both kinds of dislocations will increase as the phosphorus doping concentration increases. These two kinds of dislocations can be observed by one of the above-mentioned methods, such as etching pitting, X-ray topography, SEM or TEM. Nishida [46] studied the effects of dislocations induced by impurity diffusion on gate-controlled n-p-n transistors, and he found that the low-frequency noise increased when more dislocations were created by a higher concentration of dopants. The dislocation density was counted by the etch-pit technique. However, Nishida did not identify whether the contribution to the 1/f noise changes came from the diffusion-induced dislocations or fi'om the emitter-edge dislocations. However, it was shown later that only emitter-edge dislocations affect the low- frequency l/f noise properties. Stojadinovic [49] investigated the low-frequency noise of silicon n-p-n transistors with different densities of emitter-edge dislocations and diffusion- induced dislocations. The etch pit was the observable of the dislocation density in this study. It was found that a transistor with a large number of diflirsion-induced dislocations had an almost equal low-frequency noise level as one without diffusion-induced 119 dislocations as long as the two transistors had a similar amount of emitter-edge dislocations. From Stojadinovic's observation and the conclusion [44] made by Green mentioned early, the indication is that only the emitter-edge dislocations emerging at the device surface will contribute to the low-frequency l/f noise. These emitter-edge dislocations are believed to induce surface states near the EB junction area of n-p-n transistors. Since the surface states accept and emit electrons, the l/f noise will be of the surface recombination type. It is noted that the low-fi'equency noise characteristics of the devices used in our study are in fact characteristic of surface l/f noise, as described in Chapter 3. The emitter-edge dislocations can also induce bulk generation-recombination centers in the base region of n-p-n transistors [43,46]. These generation-recombination centers give rise to generation-recombination noise which, however, is usually lower than the 1/f noise and the shot noise at room temperature [52]. Moreover, the surface states and bulk centers induced by emitter-edge dislocations have fluctuating charges, which can produce enhanced scattering fluctuations, and therefore can also give rise to mobility fluctuation 1/f noise. However, this mobility fluctuation type of l/f noise does not appear in our transistors. Mihaila et al. [43] also investigated the effect of emitter-edge dislocations on the low-frequency noise of bipolar transistors in relationship to the phosphorus surface concentration. An abrupt increase in both noise magnitude and dislocation density was observed for surface concentration greater than 4.3 x1020 cm'3. In their following work [53], an appearance of burst noise was sometimes observed, and they associated the existence of dislocation clusters inside the emitter area with the burst noise generation. It is noted that in some devices used for our forward-bias—stress study, the burst noise also appeared and disappeared in a random fashion. 120 5.1.2 Dislocation Motion Under Forces When an external stress is applied on dislocations, some reaction of the dislocations can be observed, such as the motion of dislocations and the change of dislocation shapes. The applied stress (a) required to cause a motion of the dislocations is called the critical stress (ac). The magnitude of critical stress depends on the strength of the lattice resistance. For example, dislocations in the heavily doped silicon crystals require a very high critical stress to start moving. This phenomenon is called dislocation locking. It is believed that an interaction between the impurities and the dislocations is the cause of the locking, and the locking strength will depend on the type and the concentration of the impurities. For example, a phosphorus doped silicon crystal has a dc value around 8.0x 107 dyn/cm2 with a doping concentration of 1.0x1014 cm'3 [86], but a larger cc of 1.1x108 dyn/cm2 is observed with a doping concentration of 1.5x 1019 cm'3 [55]. Some impurities have smaller ac values in silicon, e. g. a dc value less than 1.2x 107 dyn/cm2 is observed for a boron doped silicon crystal with a d0ping concentration of 2.0x1012 cm'3, and a similarly small 0c is found for an oxygen concentration of 1.5x10l7 cm'3 in silicon. Higher ac values are also found with values as high as 2.0x 108 dyn/cm2 for Sb doped silicon at a doping concentration of 8x 1018 cm'3 [50]. The basic hypothesis on dislocation motion was proposed long ago [56-58] in 1937. However, it was not until the late 1950's that experimental evidence for dislocation motion was obtained. The pioneering experimental study which demonstrated dislocation motion was performed by Johnson et al. [59] on LiF. They utilized etch pits to detect the positions of dislocations, and dislocation movements were traced by repeated etching. In this manner, the distance that a dislocation moved after an application of external force could be measured. Consequently, the average velocity of the dislocation versus the applied force was computed by knowing the duration of the external force. The results showed that the velocity was extremely sensitive to applied stress at low velocities: the 121 larger the stress was, the higher the velocity was obtained. However, a saturation of velocity increase was observed at high velocities. The maximum velocity is essentially the velocity of sound in the crystal. Following Johnson et al.'s work, Chaudhuri et al. [60] performed a similar study on four semiconductor crystals: silicon, germanium, gallium antimonide, and indium antimonide. The impurity concentration of Si was less than 1015 cm'3. The test conditions were with temperatures varying fiom 560 C to 940 C and with the applied stress varying from 1.1x103 dyn/cm2 to 3.5x1011 dyn/cmz. It is worth noting that the doping concentration is lower than the intrinsic carrier concentration under these test temperatures. The range of velocities measured was found to be between 10'6 and 2x 10'1 cm/sec. A relation between stress (a) and the dislocation velocity (v) was obtained in the following form v = v0(o/O'0)’" exp(-AH/kT), (5.1) with a stress exponent m = 1.4 :t 0.1, and an activation energy AH = 2.2 i: 0.1 eV. A constant stress value 00 is used for normalization, and v0 is a fitting parameter. For example, ”0(0/00) = 2x10‘2 cm/sec for an applied stress a = 6x108 dyn/cm2 at T = 560 C. As we can see in Equation 5.1, dislocations can move faster under conditions of higher temperatures and higher stress. It was also shown later [50,51,55,6l,62] by various methods that the dislocation velocity and its temperature dependence (activation energy) in semiconductors can be influenced by the impurity concentrations. However, the stress exponent m is not affected by the impurity carrier concentrations. Depending on their types, impurities can either decrease or increase the fitting parameter v0. For n-type silicon with doping concentrations larger than the intrinsic carrier concentration at the test temperatures, the velocity of the dislocation will increase with increasing doping. For example, in n-type silicon at 600 C, the dislocation velocity starts to increase with increasing doping when the 122 doping concentration is 1018 cm'3 or higher [61]. The intrinsic carrier concentration at 600 C is 3.4x10l7 cm'3. Kulkarni et al. [50] observed the dislocation velocity on n-type silicon by using Lang X—ray topography. They reported that the dislocation velocity varied linearly with the majority carrier concentration in a heavily doped silicon crystal. The activation energy AH is 1.4 eV for Sb doped silicon with a doping concentration of 8x1018 cm'3. Similar values of the activation energy were reported by other researchers. Patel [61] obtained 1.6 i 0.1 eV for a 5.5x1019 cm'3 As doped silicon. George [51] observed the activation energy as 1.41 eV to 1.5 eV for a P doped silicon with a concentration of 1.4x10l9 cm‘3. Imai [55] observed activation energy between 1.57 eV and 1.58 eV for a 1.5x1019 cm'3 P doped silicon. It has been shown [84,85] that the dislocation movement induced by an applied force in a silicon crystal is a reversible effect. In these studies, a line defect (dislocation) which has a half-loop shape with its two ends cutting the surface is subjected to a mechanical stress. The locations of the two ends of the half-loop can be revealed by the etch pitting method as described earlier. It has been shown that at a high temperature the size of the dislocation half-loop will expand under an applied force. This phenomenon can be observed by monitoring the movement of the two ends of the half loop on the surface represented by the etch pits, which move away from the original positions after the force is applied. However, if the sign of the applied force is reversed, at the same high temperature the etch pits can move back towards their original positions. Also, it is found that the dislocation velocity of backward movement is higher than that of forward movement. The physics behind this phenomenon of the observed velocity difference is not well understood. It is also indicated by another research group [83] that such a dislocation half- loop will also contract, i.e. the dislocation line moves backward towards its original position, if the sample is just put in a high temperature in absence of applied stress. Unfortunately, no detailed experimental data was given in [83]. One would expect that the 123 thermal-activated backward movement would be slower than that with a reverse applied force . 5.1.3 Piezojunction Effect When a piece of n- or p-type semiconductor is subjected to a mechanical strain, a change of its resistance can be observed. This is called the piezoresistive eflect, which can be explained by the mobility change caused by the strain-dependent effective mass of the semiconductor. However, when p-n junction devices are subjected to a large compressive stress, 109 dyn/cm2 or larger, the I-V characteristics can also be changed. This phenomenon is called the piezojunction eflect, which can be correlated to the piezoresistive effect, but can not be adequately explained only by the mobility change. It is noted that the electromigration-induced compressive stress is evaluated to be as high as 1010 dyn/cm2 [7 7]. Consequently, the piezojunction effect should be considered in our forward-bias-induced degradation model. The piezojunction effect is manifested in various ways depending on the junction configuration, and the stressing technique. In this sub-section, the experimental results performed by other researchers and the proposed models are reviewed, with special focus on the change of the forward-bias current and the mag. The total forward-bias current (11:) of a p-n junction can be divided into two parts: the ideal current (Id) and the recombination current (1,). The ideal current is attributed to the minority carrier diffusion and has the following form: Id =WAInpmn/rnm +pnmp/rp)“21xlexp(qV/kr) -11, (5.2) where 71,, and pn are the minority carrier concentrations of electrons in the p side and the minority carrier concentration of holes in the 11 side respectively, 1.1,, and up are the electron and hole mobilities respectively, In and IP are electron and hole lifetimes 124 respectively, A is the junction area, and V is the applied voltage. Assuming that the stress has a negligible effect on A, then the stress-dependent parameters which need to be considered in Id will be the minority carrier concentrations, the mobilities, and the lifetimes. The recombination current sometimes is dominant in the low voltage range and can be expressed as [64] Ir = (qW/ZM ovrthnr exp(qV/ZK T ) ~ Nr no (53) where o is the carrier capture cross section, v”, is the carrier thermal velocity, N, is the trap density, and n,- is the intrinsic carrier concentration. Again, if we assume that the stress has a negligible effect on W and A, then the dominant changes in I, can come from N, and ni. It is also noted that the minority carrier concentrations and intrinsic carrier concentration can be related to the energy gap (Eg) and the effective mass (m‘) as [64] nppp =p,,n,, ="i2 “(may2 Hug/2 )l/2(m;m:2 )l/4 exp(—Eg/kT), (5.4) where n" and pp are the electron concentration in the n side and hole concentration in the p side respectively, "’1; .is the light-hole effective mass, mzh is the heavy-hole effective * . . . . . mass, m, rs the longrtudrnal electron effective mass, and m: transverse electron effective mass. As described below, research has correlated the observed changes of IF and 1,. with the stress related behavior of the parameters like the mobilities, the lifetimes, the effective masses, and the energy gap. Early work in the piezojunction effect investigated the effect of hydrostatic pressure upon diodes [65]. A small change in the electrical properties was observed which could mainly be attributed to the change of energy gap. It has been shown [66] that under hydrostatic pressure (P) the energy gap change (8Eg) could be described as 6Eg =axP, (5.5) where a = -2x10'12 eV-cm2/dyn for silicon. In 1962, Rinder reported an anisotropic stress effect (ASE) on the characteristics of p-n junctions [67]. The stress was applied perpendicularly to the device surface, by 125 means of a diamond stylus with a radius of curvature of 17 um. It is worth noting that in the ASE study the junction area is much larger than the contact area of the stressing element. A large stress, in the range of 1010 dyn/cm2 or larger, could be easily obtained. A significant increase of both forward-bias and reverse-bias current was observed, and these effects increased with increasing temperature. It was noted by the author that these effects could not be interpreted as being simply due to the piezoresistive effect, which would show an opposite temperature dependence. In his later work [68], the ASE was found to increase strongly with decreasing junction depth. In the same study a tentative model was also proposed to explain qualitatively the observed piezojunction effect caused by the ASE. This model assumes that dislocations can be created during stress, and these dislocations can be elastically "indented" by the stress to a depth which was comparable to that of very shallow junctions. Under this circumstance, new generation-recombination centers which are induced by dislocations become electrically effective in or above the space charge region and result in the increase of the recombination current. When the stress is removed, it is hypothesized that dislocations are able to move back elastically. Consequently a reversible effect can be observed in the ASE. On the other hand, if the stress is large enough to produce permanent dislocations with electrical consequences as previously described, the partially irreversible effects of the ASE can be observed. This model gives a good explanation on the irreversible effects observed in the ASE. However, this model is not adequate to explain some other observations. For example, the current of a device under the ASE test sometimes increases at first and then decreases. Consequently, other models were also proposed later to explain the ASE piezojunction effect [20,69-74]. As indicated earlier in Equation 5.2, the minority carrier lifetime is one of the parameters that should be considered when the device current is changed by a mechanical stress. Sloan et al. [72] investigated experimentally the compressive stress effect on the minority carrier lifetimes on doped silicon wafers. The lifetime was determined by the 126 method of photoconductive decay. The excess minority carriers were created in a piece of semiconductor by optical excitement. The rate of decrease of minority carriers was monitored by the change in conductivity with time measured immediately after ceasing optical generation. Consequently the minority carrier lifetime could be determined. In this study, the hydrostatic pressure was used as the means to apply mechanical stress. It was found that the minority carrier lifetime of electrons (tn) decreased and the minority carrier lifetime of holes (tp) increased after applying compressive stress. The physical cause of these lifetime changes was not fully understood. This lifetime data was taken into account by Kanda [73,74] who also developed a piezojunction model, but he showed that the effect of the lifetime is negligible compared with other variables. Using BJT's, Edwards [71] observed both collector current and base current increases afier a local stress was applied. He also investigated the influence of the strain distribution on the changes of 11175. It was found that when the stress was applied over a relatively large area using a blunt stressing element, the hp]; increased for the low values of current; however, when the stress was applied by a sharper stressing element, a decrease of has is observed over a wide current range. The changes of the emitter efficiency caused by the changes of the energy gap (Eg) alone under compressive stress were used to interpret the experimental results. A more sophisticated model was developed by Wortman et al. to explain the increase of the ideal part [69] and the recombination part [70] of the forward-bias current under a compressive stress. This model was based on the stress-induced variations in energy band structure and, consequently, their effect on minority carrier concentrations. From Equation (5.4), the minority carrier concentration changes, np/npo and pn/pno, with 0 denoting zero applied stress, is equal to the square of intrinsic carrier concentration change, (ni/n;0)2. Instead of only considering the energy gap change as proposed by Edwards [71], the stress-induced changes in all conduction band minima (AEcl, AECZ, ---) as well as two valance band maxima (AEVI, AEVZ) for both heavy and light holes were 127 taken into account. For example, the piezojunction model for silicon devices, given that silicon has six conduction band minima, could be presented as 1 (n,- /n,.O )2 = gleXPMEvr /kT) +exp(AEv2 /kT)] x[exp(—AEcl /kT) + +exp(-AECG /kT)] 2 w. (5.6) In this model, the effective mass was assumed to be constant during the stressing. The changes of mobilities and lifetimes were also neglected. Since there was only a part (A S) of total junction area (A ,) that was subjected to stress, Equations 5.2 and 5.3 are modified to include the stress effect on Id and 1, resulting from the stress-induced changes of np, p", and (In-)2. These modified equations are ’_d tags = . 5.7 [do A, ’4th ( ) I A -A A 2 —'—=———! S+—Syf,’ . (5.8) 1’0 Al At Following the band structure model proposed by Wortman et al., Kanda [20,73,74] also investigated the stress dependence of the ideal forward-bias current under ASE. However, Kanda included the stress-induced effects on mobilities, lifetimes, and hole effective masses. It was concluded that the changes of mobility and lifetimes were small in comparison with the changes in minority carrier concentrations. On the other hand, because of the stress-induced effect on the hole effective mass, it was predicted that for the <100> orientation of silicon the diode current had a double peak at a stress around 109 dyn/cmz. Also, the hp}; would go through a minimum at around a similar stress magnitude. In the work done in 1978, the noise behavior of a Si piezo-transistor was studied. It was found experimentally that the noise behavior did strongly depend on the exact location on the emitter where the mechanical stress was applied. If the stress was applied close to, or right over, the emitter-base junction edge, the noise increased with 128 stress. Otherwise, the noise decreased with stress. No further explanation for this observation was given. Hemmert et al. [77] used bipolar junction transistors as pressure sensors to monitor the electromigration-induced compressive stress in conductors of integrated circuits. The common emitter current gain (hFE) of the transistors subjected to such compressive stress decreased because of the increase of the base recombination current with the collector current unchanged. From Equation 5.3, the increase of the recombination current can be described as due to the increase of intrinsic carrier concentration, which was caused primarily by compressive stress induced band gap (Eg) narrowing as described in Equation 5.4. If hFEn represents the ratio of hp}; before stress to hFE after stress, then hFEn = exp(—AEg / 2kT). (5.9) Also from Equation 5.5, the magnitude of the compressive stress can be evaluated by knowing the amount that Eg changes, thus 2kT1fl(hFE,,) = -AEg = -axP. (5.10) From Hemmert's results, a large compressive stress (up to 1010 dyn/cmz) was deduced to be present on the stressed transistors. It is noted that the gradual degradation of hp); observed in our forward-bias stress devices has the same qualitative degradation behavior, namely only the base recombination current increases with the collector current remaining unchanged. 5.2 Degradation Model As shown in Figure 2.3 in Chapter 2, under forward-bias stress, electromigration will tend to cause an accumulation of metal above the emitter region. However, with a passivation layer enclosing the metal line the thickening of the metal will not be able to occur. Instead, there will be a build up of compressive stress normal to the substrate. In 129 our forward-bias stress study, the electromigration-induced compressive stress is claimed as the main origin of the observed changes of gain and the noise characteristics. It is worth noting that there are some distinct properties of the electromigration- induced compressive stress which are different than the anisotropic stress used in the ASE studies or a hydrostatic pressure. The electromigration-induced compressive stress offers a non-uniform stress over the device surface around the emitter region and the emitter-base junction beneath the field oxide. The magnitude of the compressive stress will depend on the device configurations. For example, it has been shown that a higher compressive stress can be achieved with a thicker passivation layer [18]. Also, the stud contacts used in TSEO3I devices will reduce the electromigration-induced compressive stress on transistors. Besides the dependence of device configurations, the magnitude of the compressive stress will also depend on the forward-bias stress conditions. The devices will be subjected to a larger compressive stress with higher stress current densities, higher stress temperatures and longer stress times. As shown in our experimental results, the low-frequency noise initially decreases with the gain remaining unchanged; however, when the gain is ready to decrease as the stress is continued, the low-frequency noise starts to increase. The decrease of the gain is caused by the increase of the base recombination current with the collector current remaining unchanged. Also as described in Chapter 4, the low-fi'equency noise is identified as the surface l/f noise located between the base and the emitter. Figure 5.1 shows a specific degradation model for these forward-bias stress research. The stressed devices are scaled-down transistors, which have very high emitter doping concentrations (usually larger than 1020 cm'3). Because of the high doping concentration, emitter-edge dislocations are presented which can contribute to a part of the surface l/f noise as described in Section 5.1.1. When the forward-bias stress starts, a small compressive stress is initially built-up upon the emitter edge. It is proposed that when the compressive stress becomes large enough (>1. 1x108 dyn/cmz) to move the COMPRESSIVE STRESS A : EMITI'ER-EDGE DISLOCATIONS 131 locked dislocations in heavily doped n-type silicon, the emitter-edge dislocations will move in the direction of the force away from the surface, as described in Section 5.1.2. Therefore, the surface recombination center induced by the dislocations will decrease, and a drop of 1/f noise power spectra thus can be observed. Equation 5.1 enables us to estimate the dislocation velocities in our devices under the forward-bias stress conditions. It is also noted that the dislocation velocity is linearly proportional to the doping concentration [50]. Consequently, one can use the following relation m l’rzfl 91 exp A”. 1.-.]- (5.11) V2 "2 02 k 71 T2 to calculate the dislocation velocity v 1 of the device with a doping concentration of n, under compressive stress of a, at temperature T 1 if the values of v2, 112, 02, T 2 and the activation energy AH are known. Kulkarni et al. [50] measured a dislocation velocity of 3.6x 10'5 cm/sec at 650 C with the compressive stress of 2.2x 108 dyn/cmz. In their study, the doping concentration was 8x1018 cm’3. Also a activation energy of 1.4 eV was obtained. Our devices have the doping concentration about 20 times larger then the samples used in Kulkarni et al.'s study. The stress temperatures in our forward-bias stress experiments are in the range of 200 C to 240 C. With the activation energy of 1.4 eV, the dislocation velocities in our devices will be 3.85x10'll cm/sec at 200 C or 5.74x10'10 cm/sec at 240 C under the same compressive stress of 2.2x108 dyn/cmz. This means that the dislocations can move away from the surface to a distance of 13 A at 200 C or of 206 A at 240 C after an hour of compressive stress of 2.2x108 dyn/cmz. It is worthy noting that under the same compressive stress condition, the dislocation velocity at room temperature is only 3.7x 10'3 A/hour, which is too small to affect the dislocation density at the device surface. This explains why the noise drop is not observed during a room- temperature forward-bias stress. If we also consider the force dependence for the dislocation velocity, i.e. (01/02)“ term in Equation 5.11 with m = 1.4, a higher velocity 132 can be obtained under a larger compressive stress. For example, under the compressive stress of 7.0x 108 dyrI/cm2 the dislocation can move a distance of 66 A per hour at 200 C. It is noted that the compressive stress can also act on the field oxide and can possibly affect the base-emitter junction beneath the oxide. However, according to Equation 5.10 the hFE change at room temperature is less than one percent under the stress magnitude of 2.2x108 dyn/cmz, and is 2.7 % under 7.0x108 dyn/cmz. A movement of dislocations which would result in a decrease of surface state density would in principle cause the base surface recombination current to decrease. In fact, in this study an initial slight increase in gain due to a reduction in base current was observed on TIN3 and Hemmert devices. However, this effect cannot be unambiguously attributed to a movement of dislocations. Other possible explanations exist, such as temperature induced relief of pre-existing stress [18]. In some cases, control devices that received only the temperature stress also showed this slight gain increase. As the forward-bias stress continues, a larger compressive stress will be established over the stress transistor resulting in a non-negligible band gap narrowing. Hence an increase of the base recombination current can be observed. Assume a substantial decrease of 11175, for example, is 4%. By Equation 5.11, a compressive stress larger than 1.1x109 dyn/cm2 is required to achieve a 4% hFE decrease. Also, as shown in Equation 2.1, the surface l/f noise is proportional to the recombination current squared. Under this condition, the surface l/f noise will start to increase because of the increase of the base recombination current. It is noted that the movement of dislocations is still in action at this stage, but its effect is compensated by the effect of the base recombination current increase. The experimental results performed on TIN 3 devices can be well explained by the above model. For TSEO3I devices, only the initial decrease of 1/f noise was observed, and a subsequent increase of the noise was not shown. The l/f noise reached a minimum value, 133 102 l .l L 101 — . (b . 3 ,1 o 0. cu Ul . O U OJ . N E: 13 Tr#l g 0 11138 5 A Tr#3 Z: 10-17 '5 Y Tr#L+ >< Tr#5 Ii 0 “”116 4‘ Tr#7 . 10‘8 4 . l . . r . . I s 10"L 100 101 10a 10‘ Stress Time Figure 5.2 Normalized base-current noise power spectral density at 1 Hz vs. stress time for the seven transistors of Hemmert's test structure designed for investigating the electromigration-induced compressive stress. Transistor #1 is the only transistor subjected to a electrical stress. The stress was done under the conditions of J = 5 mA/pm2 at 200 C. 134 and then stayed in that low value. The gain remained approximately at a constant during the forward-bias stress till the fatal failure of the devices finally occurred. In this case, the l/f noise decrease can be explained by the motion of emitter-edge dislocations cause by the compressive stress (~ 108 dyn/cmz) built-up on the emitter edge, as described earlier. Because the dislocations contribute to only a part of the surface l/f noise, the continuous decrease of the 1/f noise will not occur. However, because TSEO3I devices have the stud contacts, which is used to minimize the effect of electromigration-induced compressive stress, a large compressive stress (> 109 dyn/cmz) which can cause the increase of the base recombination current will not be obtained. Under this condition, a later noise increase and gradual gain decrease will not be observed. For Hemmert's BJT pressure sensors designed for investigating the electromigration-induced compressive stress, as described in Chapter 4, only the first transistor was subjected to the electrical stress. Therefore, any observed changes on the rest of the transistors will be attributed to electromigration effects. As shown in the experimental data of Chapter 4 and repeated here in Figure 5.2, the overall noise behavior is still that of an initial noise decrease and a subsequent noise increase as the transistor starts to show substantial gain degradation. However, the noise minimum occurs at a later time with the increasing distance from the transistor subjected to the electrical stress. For example, the noise levels of transistor #2 reached a minimum after 24 hours of stress. The noise levels of transistor #3 varied less but seemed to reach a slight minimum after 100 hours of stress. Aspects of the observation of the gain and the noise changes can also be qualitatively explained by the above-mentioned degradation model by considering the varying magnitude of compressive stress acting over each transistor. Table 5.1 shows the normalized 11175 for transistor #2, #3 and #4. The normalized by; is defined as the hp]; after stress divided by the initial has. It is noted that there is a small increase in the by: before the hat; decrease occurs. Hemmert et al. attributed this increase to the hypothesis that the electromigration produced compressive stress 135 stress time hFEn of hFEn of hFE" of (hours) transistor #2 transistor #3 transistor #4 0 1.000 1.000 1.000 1 1.041 1.042 0.998 4 1.062 1.037 1.003 24 1.033 1.042 0.999 100 1.000 0.989 0.975 200 0.981 0.984 0.957 Table 5.1 The normalized hpg (hFE before stress / hp]; after stress) for transistor #2, #3 and #4 of the Hemmert's test structure. 136 stress time h'FEn of h'FE, of h'FEn of (hours) transistor #2 transistor #3 transistor #4 24 0.972 1.000 0.996 100 0.942 0.949 0.972 200 0.923 0.944 0.954 #3 and #4 of the Hemmert's test structure. Table 5.2 The normalized hp}; decrease (hFE / the highest value of hp];) for transistor #2, 137 COMPRESSIVE STRESS OXIDE --------------------------- BASE 2 EB JUNCTIONS 8 10 18 20 (um) A :EMITER-EDGE DISLOCATIONS Figure 5.3. The forward-bias degradation model for Hemmert's pressure sensors designed for testing the electromigration efl‘ect. Only the second and the third transistors are shown. The electrical stress is applied only on the first transistor. Transistors are separated equally by 10 um measured from the center of each transistor. The emitters have a length of 4 um, as shown in the cross section, and a width of 2.5m. 138 compensated some initial built-in stress from the fabrication process [18]. Additional compressive stress builds as the degradation process is continued and will result in hpg decrease. It is then convenient to define another quantity to monitor hFE decrease. The normalized hp]; decrease (h'FEn) is defined as the hp]; divided by the highest has, and the values of h'FEn is shown in Table 5.2. A schematic of the second and the third transistors in Hemmert's test structure are shown in Figure 5.3. The center of transistor #1 is located at x = 0 urn which is not shown in Figure 5.3. The center of transistor #2 is located at x = 10 um, and that of transistor #3 is located at x = 20 pm. For each transistor, the largest compressive stress is obtained on the E-B junction located on the left side of the emitter contact. The emitter size is 4 um long (as shown in Figure 5.3) and 2.5 pm wide. Consequently the EB junction subjected to the largest compressive stress for transistor #2 is located at x= 8 pm, and that for transistor #3 is located at x = 18 um. The compressive stress at the location of the E-B junction can then be calculated by monitoring the gain decrease of the respective transistor. Table 5.3 shows that calculated pressure by using the data provided in Table 5.2. With the help of Equation 5.10, the pressure at x = 8 um and that at x = 18 um can be calculated by using the data of the normalized by; decrease (h'FEn) for transistor #2 and #3.. It has been also shown [77] that the electromigration-induced compressive stress has an exponential decay form of exp(- 0), where L0 is called the characteristic length. It has also been shown [77] that L0 will be increased as the stress time increases. The values of L0 versus the stress times in our study was evaluated as 10.25 pm after 24 hours, 26.2 um after 100 hours and 38 um after 200 hours of forward-bias stress. It is noted that the pressure at x = 18 um after 24 hours of stress is carried out by interpolating with the help of the data fi'om transistor #4. As described earlier, the compressive stress acting on the emitter edge will cause the dislocation to move if the compressive stress is large enough to overcome the locking 139 Pressure (dyn/cmz) stress time x = 8 pm = 18 um (hours) 24 7.30x108 2.75x108 100 1.56x109 1.07x109 200 2.06x109 1.59x109 Table 5.3 The electromigration-induced compressive stress gradient along the distance (x) from transistor #1. 140 strength on dislocations. This usually requires a compressive stress larger than 103 dyn/cmz. The stress temperature for these devices is at 200 C. It has been estimated in Section 5.1.2 that the dislocation can move 13 A per hour at 200 C under a pressure of 2.2x108 dyn/cm2 or 66 A per hour at 200 C under a pressure of 7.0x108 dyn/cmz. On the other hand, the compressive stress acting on the E-B junction will cause the narrowing of the energy gap and consequently will result in a decrease of the gain. We also estimated that it requires 1.1x 109 dyn/cm2 to observe a substantial 4% gain decrease. From the data in Table 5.3, after 24 hours the compressive stress built-up on the emitter edge of transistor #2 (x z 8 pm) is around 7 .30x 103 dyn/cmz, which enables dislocations to move at a velocity of around 66 A/hour at 200 C. Also the compressive stress on the E-B junction (x = 8pm) is 7.3x103 dyn/cmz, which is too small to cause an appreciable increase of the recombination current. Consequently we will observe a noise decrease with the gain remaining almost unchanged at this stage. However, after 100 hours of forward- bias stress the compressive stress built-up on the E-B junction of transistor #2 is large than 1.1x109 dyn/cm2 which can result in an appreciable increase of the base recombination current. That is why the decrease of the gain and the increase of the noise level will be observed after 100 hours on transistor #2. In between 0 hours and 100 hours, there is a noise minimum. On the resolution of the time scale used here, the minimum noise was measured at 24 hours. However, the actual minimum may be somewhat earlier or latter than this. Consequently the true minimum may be deeper than shown. Additionally it is noted that for transistor #1 , the stress covers the entire perimeter with relatively equal magnitude because the current flows vertically down into the device. However, for transistor #2, the edge of the perimeter close to transistor #1 is subjected to a larger stress prior to the far edge. Consequently, the minimum would be reduced and spread out with time relative to transistor #1. For transistor #3, after 24 hours of forward-bias stress at 200 C the compressive stress built-up on the emitter edge (x z 18 um) is around 2.75x108 dyn/cmz, which is less than that on transistor #2. Under this compressive stress, the 141 o dislocation is just about to move and moves slowly in a velocity of around 13 A/hour at 200 C. Consequently a smaller noise drop will be observed on transistor #3 afier 24 hours. Also at this moment the compressive stress on the EB junction of transistor #3 is not large enough to produce the gain changes. After 100 hours, a larger compressive stress is built on the emitter edge. However, the compressive stress on the EB junction at this time is also large enough (1.07x109 dyn/cmz) to produce the increase of the base recombination current, so the noise decrease caused by the dislocation movement was compensated by the noise increase caused by the base recombination current increase. Since there was no measurement between 24 hours and 100 hours, any noise minimum that occurred was not well measured in this experiment. Also, the initial noise for transistor #3 was different (smaller) than transistors #1 and #2, so some differences in noise variation may be expected. It is also noted that as shown by the results in Chapter 4, the post-stress noise levels will return towards the initial higher values after thermal annealing. However, total recovery is not observed on the devices after around 60 hours of annealing at T = 225 to 240 C. It is known that the application of high temperature annealing afier the electromigration-induced compressive stress has built-up due to forward-bias stress would allow the metal to creep reducing the pressure. Hemmert et al. [18] showed that such compressive stress could be totally removed after 15 minutes of annealing at 400 C. Also, as described in Section 5.1.2, the dislocation movement induced by an external compressive stress can be reversed by thermal annealing. The post-annealing noise behavior suggests that a part of the dislocations which initially moved away fi'om the surface due to the compressive stress can move back to the surface after annealing. The thermal annealing is believed both to remove the electromigration-induced compressive stress and to activate the backward movement of dislocations. Consequently, a increase of noise levels can be observed after annealing. CHAPTER 6 SUMMARY AND CONCLUSIONS In this study, reliability issues resulting from hot-carrier effects and electromigration effects in bipolar junction transistor structures were investigated. The physical observables for this study included device I-V characteristics and low-fi'equency noise properties. Four kinds of bipolar junction transistors with different manufacturing technologies were used: namely, TIN3, TOUT, TSE031 and the Hemmert test structure. TIN3 and TOUT devices were manufactured by conventional oxide isolation methods. However, TOUT had a Schottky diode clamp but TIN3 did not. TSE031 devices were advanced technology transistors, fabricated with polysilicon base contacts and side-wall oxides. The Hemmert test structure, which contained a set of seven transistors, was a test structure designed particularly to investigate the electromigration induced compressive stress effects. A detailed description of these devices is presented in Section 3.1 and Section 4.1. The noise measurements for this study were performed by biasing the transistor in a forward active bias condition under both the common-collector and common-emitter configurations. The dominant low frequency l/f noise source of transistors was identified by determining the current dependence of their noise power spectral densities. A quadratic current dependence of the noise power was observed; consequently, it was concluded that the dominant pre-stress and post-stress low frequency 1/f noise source for the transistors under study is due to surface recombination states located between the base and the emitter. 142 143 Reverse-Bias Stress For the reverse-bias stress study, both TIN3 and TOUT devices were investigated. Transistors were electrically stressed at 23 C by applying three reverse DC bias voltages (- 3.5 V, -4.0 V, and -4.5 V) to the base-emitter junction and by applying two AC conditions (0 to -4 V and 1 to -4 V) to the base-emitter junction. The base-collector junction was short-circuited. Stress times varied up to 2000 hours. Two post-stress recovery mechanisms, namely thermal annealing and forward-bias induced recovery, were also studied. Observed degradation phenomena included a decrease of hFE and an increase of low-frequency noise levels after reverse bias stressing of the base-emitter junction of bipolar junction transistors. The observed gain and noise changes are attributed to both interface states and oxide trap states produced by hot electrons. Forward-bias induced recovery allows a partial reversal of degradation and is believed to be due primarily to a reduction in the number of electrons trapped in the oxide. Thermal annealing, which is capable of removing of interface states as well, produces a larger recovery. Furthermore, because of the effect of forward-bias induced recovery, AC stress that included forward- bias cycles produced less degradation than AC stress without forward-bias cycles. Consequently, the degradation of the transistor not only depends on the time over which a reverse bias is applied, but also on the history of biases applied between the reverse-bias periods. Also, it was found experimentally and theoretically that after DC reverse-bias stress the increased base current was linearly proportional to the increase surface recombination surface noise power. The slope of this linear relation did not depend on the stress voltages. This linear relation allows one to predict the noise changes, whose characterization requires a more complicated experimental process, by knowing the base current changes after stress. Forward-Bias Stress For the forward-bias stress study, TIN3, TSE031 and the Hemmert test structure were the devices under investigation. The devices were electrically stressed at various 144 temperatures (from -75 C to 240 C) with large stressing emitter current densities (fiom 1 mA/um2 to 5 mA/umz). Stress times varied up to 2500 hours. It was shown that the combination of high currents and high temperatures produced gradual gain degradations on both TIN 3 transistors and the transistors on the Hemmert test structure, and produced precipitous failures of TSE031 transistors. Similarly the application of high currents and high temperatures produced changes in transistor noise characteristics for all three technologies. Specifically, in all transistors stressed, the low-frequency noise initially decreased. It was also observed in TIN3 transistors and in the transistors on the Hemmert test structure that the noise eventually started to increase again at the point in time when an appreciable gain decrease was observed as the stress continued. For TSE031 transistors, the initially decreased noise stayed in the decreased levels till the precipitously failures finally occurred. The failures were due to electromigration produced open circuits. Degradation Mechanisms Reverse-bias degradation of gain and noise attributes is consistent with hot electron degradation. However, for the forward-bias induced changes in gain and noise, it is believed that electromigration induced compressive Stress played an important role in the phenomena of noise decrease followed by noise increase observed during forward-bias stress. As mentioned earlier, the dominant l/f noise source inside the stressed transistors is due to surface phenomena. Consequently, the result of forward-bias stress induced reduction of 1/f noise indicates that the stress reduces noise producing states at the surfaces. Such surface states may be due, for example, to crystal dislocations, which have been reported to be caused, for example, by high doping or by process induced stress and have previously been shown to be related to l/f noise. As described in Chapter 5, the magnitudes of electromigration-induced compressive stress in this study are sufficient (> 108 dyn/cmz) to cause appreciable dislocation motion, and the direction of the force is away from the device surface. Consequently, from this viewpoint, electromigration induced compressive stress tends to drive surface dislocations beneath the surface where 145 they are no longer contribute to surface 1/f noise. However, as the forward bias stress continues to long stress times, the compressive stress eventually increases to a point (z 109 dyn/cmz) sufficient to cause appreciable band-gap narrowing. At this point, there is a resulting increase in base recombination current causing gain degradation. Also, since 1/f noise is proportional to the recombination current squared, the l/f noise at a given base- emitter voltage increases accordingly with the gain decrease. The forward bias induced noise decrease, at least initially, is opposite to expectations for hot electron effects, by which a monotonic noise increase with stress times will be observed. Consequently, the low-frequency noise behavior provides an indication as to whether the cause of gain degradation under forward bias stress in a particular technology is due to electromigration or hot electrons. Future Studies There are three suggestions for the future studies related to the current work. First, a simulation of hot electron induced noise changes after a reverse-bias stress would be useful in order to compare the simulation results with the experimental results presented in this dissertation. Secondly, the forward-bias stress effect should be further studied on advanced poly-emitter devices, and the results should be compared with the current results performed on conventional and advanced poly-base transistors. Consequently, the physical causes, either hot electron effects or the electromigration effects, of forward-bias induced degradation on advanced poly-emitter transistors also can be identified. Finally, the direct experimental evidence of electromigration effects on near- surface dislocations is necessary to confirm the hypothesis developed in Chapter 5. For example, methods of producing etch pits may be used to observe the movement of dislocations after an electromigration induced compressive stress is built up. [1] [2] [3] [4] l5] [6] BIBLIOGRAPHY R. A. Wachnik, T. J. Bucelot, and G. P. Li, "Degradation of bipolar transistors under high current stress at 300 K," J. Appl. Phys, vol. 63, pp. 4734-4740, 1988. B. A. McDonald, "Avalanche degradation of hFE," IEEE Trans. Electron Devices, vol. ED-17, pp. 871-878, 1970. P. M. Solomon and D. D. Tang, "Bipolar circuit scaling," in ISSCC Dig. Tech. Paper, pp. 86-87, 1979. D. D.-L. Tang, and E. I-Iackbarth, "Junction degradation in bipolar transistors and the reliability imposed constraints to scaling and design," IEEE Trans. Electron Devices, vol. ED-35, pp. 2101-2107, 1988. S. A. Petersen, and G. P. Li, "Hot carrier effects in advanced self-aligned bipolar transistors, " IEEE International Electron Device Meeting Technical Digest, pp. 22- 25, 1985. S. P. Joshi, R. Lahri, and C. Lage, "Poly emitter bipolar hot carrier effects in advanced BiCMOS technology," IEEE International Electron Device Meeting Technical Digest, pp. 182-185, 1987. 146 [7] [3] [9] [10] [11] [12] [13] 147 J. D. Burnett, and C. Hu, "Modeling hot-carrier effects in polysilicon emitter bipolar transistors," IEEE Trans. Electron Devices, vol. ED-3 5, pp. 2238-2244, 1988. H. S. Momose, Y. Niitsu, H. Iwai, and K. Maeguchi, "Temperature dependence of emitter-base reverse stress degradation and its mechanism analyzed by MOS structures," Proc. IEEE 1989 Bipolar Circuits and Technology Meeting, pp. 140- 143, 1989. J. D. Burnett and C. Hu, "Hot-carrier degradation in bipolar transistors at 300 and 110 K -- effect of BiCMOS inverter performance," IEEE Trans. Electron Devices, vol. ED-37, pp. 1171-1173, 1990. T.-C. Ong, P. K. Ko, and C. Hu, "SO-A gate-oxide MOSFET's at 77 K," IEEE Trans. Electron Devices, vol. ED-34, pp. 2129-2135, 1987. T. C. Chen, C. Kaya, M. B. Ketchen, and T. H. Ning, "Reliability analysis of self- aligned bipolar transistor under forward active current stress," IEEE International Electron Device Meeting Technical Digest, pp. 650-653, 1986. S. P. Joshi, "Reliability imposed constraints on current density of high performance poly-emitter bipolar transistors," Proc. IEEE 1990 Bipolar Circuits and Technology Meeting, pp. 180-183, I990. D. D.-L. Tang, E. Hackbarth, and T. Chen, "On the very-high current degradations on Si n-p-n transistors," IEEE Trans. Electron Devices, vol. ED-37, pp. 1698-1706, 1990. [14] [15] [16] [17] [13] [19] 1201 148 B. A. McDonald, "Avalanche-induced l/f noise in bipolar transistors," IEEE Trans. Electron Devices, vol. ED-17, pp. 134-136, 1970. M. C. A. M. Koolen, and J. C. J. Aerts, The influence of non-ideal base current on l/f noise behaviour of bipolar transistors," Proc. IEEE 1990 Bipolar Circuits and Technology Meeting, pp. 23 2-235, 1990. M. L. Dreyer, and J. Durec, "Low frequency l/f noise and current gain degradation in BiCMOS n-p-n transistors," Proc. 30th IEEE Rel. Phys. Symp, pp. 95-101, 1992. J. R. Black, "Electromigration failure modes in aluminum metallization for semiconductor devices," Proc. IEEE, vol. 57, pp. 1587-1594, 1969. R. S. Hemmert, G. S. Prokop, J. R. Lloyd, P. M. Smith, and G. M. Calabress, "The relationship among electromigration, passivation thickness, and common-emitter current gain degradation within shallow junction NPN bipolar transistors," J. Appl. Phys, vol. 53, pp. 4456-4462, 1982. S. Timoshenko, Strength of Materials, Part II, 2nd ed., New York: Van Norstrand, 1941,pp.236. Y. Kanda, "Effect of compressive stress on silicon bipolar devices," .1. Appl. Phys, vol. 44, pp. 389-394, 1973. [21] 149 T. Wada, M. Sugimoto, and T. Ajiki, "Electromigration in bipolar NPN transistors," Solid-State Electron, pp. 1409-1412, 1988. [22] J. G. Cottle, T. M. Chen, and K. P. Rodbell, "A comparison between noise measurements and conventional electromigration reliability testing," Proc. 261h IEEE Rel. Phys. Symp, pp. 203-208, 1988. [23] A. Diligenti, P. E. Bagnoli, B. Neri, S. Bea, and L. Mantellassi, "A study of [24] [25] [26] [27] PM electromigration in aluminum and aluminum-silicon thin film resistors using noise technique," Solid-State Electron, vol. 32, pp. 11-16, 1989. A. van der Ziel, "Unified presentation of l/f noise in electronic devices: fundamental 1/f noise sources," Proc. IEEE, pp. 233-258, 1988. S. T. Hsu, and R. J. Whittier, "Characterization of burst noise in silicon devices," Solid-State Electron, pp. 867-878, 1969. S. T. Hsu, J. T. Fitzgerald, and A. S. Grove, "Surface-state related l/f noise in p-n junctions and MOS transistors," Appl. Phys. Lett., vol. 12, pp. 287-289, 1968. A. van der Ziel, "The surface recombination model of p-n diode flicker noise," Physica, vol. 48, pp. 242-246, 1970. A. L. McWhorter, " 1/f noise and germanium surface properties," in Semiconductor Surface Physics, R. H. Kingston, Ed. Philadelphia, PA: Univ. Pennsylvania Press, 1957, pp. 107-208. [29] [30] [31] [32] [33] [34] [35] [36] 150 A. van der Ziel, X. Zhang, and A. H. Pawlikiewicz, "Location of l/f noise sources in BJT's and HBJT's - 1. Theory," IEEE Trans. Electron Devices, vol. ED-33, pp. 1371-1376, 1986. O. Jantsch, "Flicker ( 1/f) noise generated by a random walk of electrons in interfaces, " IEEE Trans. Electron Devices, vol. ED-34, pp. 1100-1115, 1987. F. N. Hooge, "l/fnoise," Physica B, vol. 83, pp. 14, 1976. T. G. M. Kleinpenning, "l/f noise i p-n diode," Physica B, vol. 98, pp. 289, 1980. F. N. Hooge, T. G. M. Kleinpenning, and L. K. J. Vandamme, "Experimental studies on l/f noise," Rep. Prog. Phys, vol. 44, pp. 479-532, 1981. X. Zhu, and A. van der Ziel, "The Hooge parameters of n+-p-n and p+-n-p silicon bipolar transistors," IEEE Trans. Electron Devices, vol. ED-32, pp. 658-661, 1985. A. H. Pawlikiewicz, and A. van der Ziel, "Location of 1/f noise sources in BJT's and HBJT's - ll. Experiment," IEEE Trans. Electron Devices, pp. 2009-2012, 1987. A. H. Pawlikiewicz, A. van der Ziel, G. S. Kousik, and C. M. Van Vliet, "Fundamental l/f noise in silicon bipolar transistors," Solid-State Electron, vol. 31, pp. 831-834, 1988. [37] [33] [39] [40] [411 [42] [43] [44] 151 A. van der Ziel, P. H. Handel, X. Zhu, and K. H. Duh, "A theory of the Hooge parameters of solid-state devices," IEEE Trans. Electron Devices, vvol. ED-32, pp. 667-671, 1985. A. van der Ziel, "Proposed discrimination between l/f noise source in transistors," Solid-State Electron, vol. 25, pp. 141-143, 1982. M. Stoisiek, and D. Wolf, "Origin of 1/f noise in bipolar transistors," IEEE Trans. Electron Devices, vol. ED-27, pp. 1753-1757, 1980. T. G. M. Kleinpenning, "On l/f mobility fluctuations in bipolar transistors," Physica B, vol. 138, pp. 244-252, 1986. T. G. M. Kleinpenning, "Location of low-frequency noise sources in submicrometer bipolar transistors, " IEEE Trans. Electron Devices, vol. ED-39, pp. 1501-1506, 1992. G. Blasquez and J. Caminade, in Noise in Physical System, D. Wolf, Ed. Berlin: Spinger, 1978, pp. 60-63. M. Mihaila, and K. Amben'adis, "Noise phenomena associated with dislocations in bipolar transistors," Solid-State Electron, pp. 109-113, 1983. D. Green and A. G. Jordan, "Effects of dislocations on the noise of planar p-n junctions," Int. J. Electron, vol. 27, pp. 159-176, 1069. [45] [46] [47] [43] [49] [50] [51] [52] 152 S. Prussin, "Generation of Distribution of Dislocations by Solute Diffusion, " J. Appl. Physics, vol. 32, pp. 1876-1881, 1961. M. Nishida, "Effects of diffirsion-induced dislocations on excess low-frequency noise," IEEE Trans. Electron Devices, vol. ED-20, pp. 221-236, 1973. R. A. McDonald, G. G. Ehlengberger, and T. R. Huffman, "Control of difl’usion induced dislocations in phosphorus diffirsed silicon, " Solid-State Electron, vol. 9, pp.807,l966. J. M. Fairfield, and G. H. Schwuttke, "Strain efi‘ects around planar diffused structures," J. E. Lawrence, J. Electrochem. Soc., vol. 115, pp. 416-422, 1968. N. Stojadinovic, "Effects of emitter edge dislocations on the low-fi'equency noise of silicon planar n-p-n transistors," Elect. Lett., vol. 15, pp. 340-342, 1979. S. B. Kulkami and W. S. Williams, "Dislocation velocities and electronic doping in silicon," J. Appl. Phys, vol. 47, pp. 4318-4324, 1976. A. George and G. Champier, "Velocities of screw and 60° dislocations in n- and p- type silicon," Phys. Stat. Sol, part A, vol. 53, pp. 529-540, 1979. T. Wade and A. van der Ziel, "Noise associated with recombination in the emitter space charge region of transistors" Solid-State Electron, vol. 19, pp. 909-910, 1976. [53] [54] [55] [56] 1571 [58] [59] [60] 153 M. Mihaila, K. Amberiadis, and A. van der Ziel, "l/f, g-r and burst noise induced by emitter-edge dislocations in bipolar transistors," Solid-State Electron, vol. 27, pp. 675-676, 1984. R. S. Popovic, and N. D. Stojadinovic, "Dependence of dislocations on emitter phosphorus diflusion conditions and their effects on electrical characteristics of silicon planar NPN transistors," Phys. Stat. Sol, part A, vol. 52, pp. 433-440, 1079. M. Imai and K. Sumino, "In situ X-ray topographic study of the dislocation mobility in high-purity and impurity-doped silicon crystals," Phil. Mag, part A, vol. 47, pp. 599-621, 1983. E. Orowan, Z. Phys, vol. 89, pp. 634, 1934. M. Polanyi, Z. Phys, vol. 89, pp. 660, 1934. G. 1. Taylor, Proc. R Soc., vol. A145, pp. 362, 1934. W. G. Johnson and J. J. Gilman, "Dislocation velocities, dislocation densities, and plastic flow in lithium fluoride crystals," J. Appl. Phys, vol. 30, pp. 129, 1959. A. R. Chaudhuri, J. R. Patel and L. G. Rubin, "Velocities and densities of dislocations in germanium and other semiconductor crystals, " J. Appl. Phys, vol. 33, pp. 2736-2746, 1962. [61] [62] [63] [64] [65] [66] [67] [63] [69] 154 J. R. Petal, L. R. Testardi and P. E. Freeland, "Electronic effects on dislocation velocities in heavily doped silicon," Phys. Rev., vol. 13B, pp. 3548, 1976. V. N. Erofeev, V. I. Nikitenko and V. B. Osvenskii, Phys. Status Solidi, vol. 35, pp. 79, 1969. C. Kittel, Introduction to Solid State Physics, 5th Edition, John Wiley & Sons, Inc., 1976,pp.570. S. M. Sze, Physics of Semiconductor Devices, 2nd Edition, John Wiley & Sons, Inc., 1981. H, Hall, J. Bardeen and G. Pearson, "The effects of pressure and temperature on the resistance of PN junctions in germanium," Phys. Rev, vol. 84, pp. 129-132, 1951. W. Paul and D. M. Warchauer, "Optical properties of semiconductors under hydrostatic pressures—II silicon," J. Phys. Chem. Solids, vol. 5, pp. 102-106, 1958. W. Rindner, "Resistance of elastically deformed shallow p-n junctions, " J. Appl. Phys, vol. 33, pp. 2479-2480, 1962. W. Rindner, and 1. Braun, "Resistance of elastically deformed shallow p-n junctions. II," J. Appl. Phys, vol. 34, pp. 1958-11970, 1963. J. J. Wortman, J. R. Hauser and R. M Burger, "Effect of mechanical stress on p-n junction device characteristics, " J. Appl. Phys, vol. 35, pp. 2122-2131, 1964. [70] [71] [72] [73] [74] 175] [76] 155 J. J. Wortman and J. R. Hauser, "Effect of mechanical stress on p-n junction device characteristics. 11. Generation-recombination current," .1. Appl. Phys, vol. 37, pp. 3527-3530, 1966. R. Edwards, "Some effects of localized stress on silicon planar transistors," IEEE Trans. Electron Devices, vol. ED-l 1, pp. 286-294, 1964. B. J. Sloan and J. R. Hauser, "Effects of uniaxial compressive stress on minority- carrier lifetime in silicon and germanium," J. Appl. Phys, vol. 41, pp. 3504-3508, 1970. Y. Kanda, "Effect of stress on germanium and silicon p-n junctions," Japan. J. Appl. Phys, vol. 6, pp. 475-486, 1967. Y. Kanda, Y. Kanazawa, T. Terada, and M. Maki, "Improved Si piezo-transistors for mechano-electn'cal transducers," IEEE Trans. Electron Devices, pp. 813-817, 1978. D. D. Tang, C. Wong and P. A. McFarland, "Metal migration into polysilicon emitter afler very hight current stress," IEEE Elec. Dev. Lett., vol. 13, pp. 265-266, 1992. C. J. Varker, D. Pettengill, W.-T. Shiau and B. Reuss, "Hot carrier induced hFE degradation in BiCMOS Transistors," Proc. 30th IEEE Rel. Phys. Symp, pp. 58- 62, 1992. [77] [78] [79] [80] [31] [82] [83] [34] 156 R. S. Hemmert, and M. Costa, "Electromigration-induced compress stresses in encapsulated thin-film conductors, " Proc. 29th IEEE Rel. Phys. Symp, pp. 64-69, 1991. J. D. Burnett, Ph.D. dissertation, University of California/Berkeley, 1990. G. Blasqez, G. Barbottin and V. Boisson, "Instabilities in bipolar devices," in Instabilities in Silicon Devices: vol. 2, ed., G. Barbottin and A. Vapaille, North- Holland, 1989. R. W. Keyes, "Thermal problems of the pulsed injection laser," IBM Res. Develop. , vol. 14, pp. 154-167, 1970. I. Antipov, "Proposed process modifications for dynamic bipolar memory to reduce emitter-base leakage current," IEEE Trans. Electron Devices, vol. ED-27, pp. 1649-1654, 1980. E. Sirtl and A. Adler, Z. MetaIlK., vol. 52, pp. 527, 1961. A. George, C Escaravage, G. Champier and W. Schroter, "Velocities of screw and 60° dislocations in silicon," Phys. Stat. Sol., part B, vol. 53, pp. 483-496, 1972. V. I. Nikitenko, B. Ya. Farber and E. B. Yakimov, "Asymmetry of dislocation mobility in semiconductors," JET? letter, vol. 34, pp. 233-236, 1981. [35] [36] [87] [33] [89] [90] 157 V. I. Nikitenko, B. Ya. Farber and E. B. Yakimov, "Asymmetry of isolated dislocation mobility in silicon single crystals," Crystal Res. T echnol., vol. 19, pp. 295-302, 1984. 1. E. Bondarenko, V. G. Eremenko, V. I. Nikitenko and E. B. Yakimov, Phys. Stat. Sol., part A, vol. 60, pp. 341, 1980. C.-J. Huang, Ph.D dissertation, Michigan State University, 1992, pp.39. C. J. Sun, D. K. Reinhard, C.-J. Huang, T. A. Grotjohn and C.-C. W. Yu, "Hot- carn'er-induced degradation and post-stress recovery of bipolar transistor gain and noise characteristics," IEEE Trans. Electron Devices, vol. ED-39, pp. 2178-2180, 1992. 7 A. S. Grove, Physics and Technology of Semiconductor Devices, John Wiley, Chichester (1971). I E. Dubois, " Surface recombination in bipolar transistors accurate determination of carriers capture cross-sections of Si/Si02 states by means of process/device simulations," Proc. IEEE 1991 Bipolar Circuits and Technology Meeting, pp. 178- 181, 1991. ”lllllll'llllllll“