An Ultra Low Power ADC for Wireless Micro-Sensor Applications by Naveen Verma B.A.Sc., Electrical Engineering and Computer Engineering, University of British Columbia Submitted to the Department of Electrical Engineering and Computer Science in partial fulfillment of the requirements for the degree of Master of Science in Electrical Engineering and Computer Scienco at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2005 © 2005 Massachusetts Institute of Technology. All rights reserved. ...... A uthor .......................... Department of Electrical Engineering and Computer Science May 16, 2005 C ertified by .................... ................ ..................... Anantha P. Chandrakasan Professor of Electrical Engineering Thesis Supervisor Accepted by..... .......... Arthur C. Smith Chairman, Departmental Committee on Graduate Students MASSACHUSETTS INS OF TECHNOOGY E BARKER MAR 0 3 2006 LIBRARIES 2 An Ultra Low Power ADC for Wireless Micro-Sensor Applications by Naveen Verma Submitted to the Department of Electrical Engineering and Computer Science on May 16, 2005, in partial fulfillment of the requirements for the degree of Master of Science in Electrical Engineering and Computer Science Abstract Autonomous micro-sensor nodes rely on low-power circuits to enable energy harvesting as a means of sustaining long-term, maintenance free operation. This work pursues the design of an ultra low-power analog-to-digital converter (ADC) whose sampling rate and resolution can be scaled to dynamically recover power savings. The proposed ADC has a sampling rate of 0-100 kS/s and a resolution of either 12 or 8 bits. The design is based on the successive approximation register architecture (SAR), which is suitable for scaleable, micro-power operation. Specifically, the number of active blocks has been minimized to allow efficient power-gating, which, inturn, has been leveraged to implement scalability features. Several new techniques to improve the efficiency of the ADC have been developed and employed. Analog offset calibration in the regenerative latch is used, to improve the power-delay product of the comparator; pre-amplifier cascade optimization is performed with consideration to thermal noise limitations; weak-inversion biasing is employed in the active amplifiers; passive switch-capacitors are used to generate the auto-zero reference voltage such the CMRR of the ADC is maximized; integrated capacitors are laid-out in a new common-centroid arrangement that minimizes edge effects; finally, the sub-DAC's transmission gain is adjusted to reduce non-linearities caused by the attenuating effects of parasitics. The ADC has been fabricated in a 0.18pLm CMOS technology. All circuits are powered using a IV supply, though bootstrapping is used internally. At a resolution of 12-bits, and a sampling rate of 100 kS/s, the power consumption of the entire ADC core is 26pW. The SNDR of the converter with a 48 kHz input tone is 65dB (10.55 ENOB) and the SFDR is 71dB. The power consumption decreases linearly with sampling rate, and is measured to be approximately 200nW at 500 S/s. Thesis Supervisor: Anantha P. Chandrakasan Title: Professor of Electrical Engineering 3 4 Acknowledgements This thesis has been a major accomplishment. It is not, however, my accomplishment alone. The immense support and friendship I've received from so many over the course of my degree and education so far, is perhaps the aspect most worth remembering. Let me start by thanking them here. Firstly, I must express my gratitude to my advisor and role model, Professor Anantha Chandrakasan. Anantha forces me to do my best; how is it that I can spend months analyzing a problem, and the first time I present it to him, I get a question I hadn't even thought of? But, most importantly, I thank Anantha for giving me something to strive towards. Of course, non-technically, the biggest impact in my life is courtesy of my mom and dad. Saying thank you to you guys is too trivial and inconsequential for everything you have done and meant. I only hope that you know, as you always do, how much your support means to me on a daily basis. My sisters, Angelee and Serena, have played a very important role, not only in my life, but specifically in the completion of this thesis. Thank you, both, for making it such a high priority to keep in tact my ability to write prose. Of course, your encouragement has meant even more. Engineering education, a microcosm of life, is an arduous process. I've gotten support, all the way, from my good friend, Sean. Neither of us quite fits into the EE scene, but we get each other through it, and, at the end of it, we have a good laugh. There is a group of masters who took it personally to make sure I understood what engineering really is: a tall order, considering, at the time, I barely understood V=IR. Thank you Dr. Lawrence, Dr. Jaeger, Dr. Pulfrey, and Dr. Salcudean. Here at MIT, I wish to thank everyone from Ananthagroup. I'm lucky to be associated with such a bright group, that continually inspires me to be more creative. Finally, I would like to acknowledge National Semiconductor for providing fabrication services for the prototype chip, as well as their development help and CAD support. 5 ~A- -Ka '-Or:No "-a --- -' -M -e''ESW:.r' 5-ilv--r--Niha us-a-t-Ms---ditie'-0.r R?-4.3W; rt-0---ishmafA'wMil r- ae:Oi- Contents 1 Introduction 1.1 1.2 2 3 19 Requirement Specifications . . . . . . . . . . . . . . . . . . . . . . . . 20 1.1.1 Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 1.1.2 Sampling Rate . . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.1.3 Scalability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.1.4 Input Interface . . . . . . . . . . . . . . . . . . . . . . . . . . 22 Architecture Selection . . . . . . . . . . . . . . . . . . . . . . . . . . 22 25 Precision Limitations 2.1 Low Overdrive of MOS Switches . . . . . . . . . . . . . . . . . . . . . 25 2.2 Charge Injection Errors . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.3 Mismatch of Passive Elements . . . . . . . . . . . . . . . . . . . . . . 27 2.4 Mismatch of Active Elements . . . . . . . . . . . . . . . . . . . . . . 27 2.5 Device Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.6 Threshold Voltage Hysteresis . . . . . . . . . . . . . . . . . . . . . . . 30 2.7 Other Error Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 Successive Approximation Conversion 33 3.1 Basic Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.2 Differential Successive Approximation ADC . . . . . . . . . . . . . . 35 3.3 Analysis of Error Sources . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.3.1 39 Capacitor Mismatch . . . . . . . . . . . . . . . . . . . . . . . 7 3.4 4 DAC Parasitics . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.3.3 Reference Voltage Error . . . . . . . . . . . . . . . . . . . . . 41 3.3.4 Comparator Offset . . . . . . . . . . . . . . . . . . . . . . . . 41 3.3.5 Sampling Switch Non-Linearity 3.3.6 Comparator Thermal Noise Summary . . . . . . . . . . . . . . . . . 42 . . . . . . . . . . . . . . . . . . . 42 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 Architecture Design and Theory 45 4.1 Global Architecture and Concepts . . . . . . . . . . . . . . . . . . . . 45 4.1.1 Conversion Plan . . . . . . . . . . . . . . . . . . . . . . . . . . 45 4.1.2 Sample Rate Scaling . . . . . . . . . . . . . . . . . . . . . . . 47 4.1.3 Resolution Modes . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.1.4 Self-Timed Bit Cycling . . . . . . . . . . . . . . . . . . . . . . 51 4.2 4.3 5 3.3.2 Block Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 4.2.1 Comparator Architecture . . . . . . . . . . . . . . . . . . . . . 54 4.2.2 DAC Architecture . . . . . . . . . . . . . . . . . . . . . . . . . 65 4.2.3 SAR Architecture . . . . . . . . . . . . . . . . . . . . . . . . . 72 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Circuit Design 5.1 5.2 73 75 DAC Circuit Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5.1.1 Switch Network . . . . . . . . . . . . . . . . . . . . . . . . . . 76 5.1.2 Capacitor array . . . . . . . . . . . . . . . . . . . . . . . . . . 83 Comparator Circuit Design . . . . . . . . . . . . . . . . . . . . . . . . 91 5.2.1 Preamplifers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5.2.2 Offset-Calibrating Latch . . . . . . . . . . . . . . . . . . . . . 96 5.2.3 Latch Level Restorer . . . . . . . . . . . . . . . . . . . . . . . 106 5.3 SAR Circuit Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 5.4 Clock Manager . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 5.5 Charge Pump and Voltage Multiplier . . . . . . . . . . . . . . . . . . 112 8 6 7 115 Testing and Characterization 6.1 Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 6.2 Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 6.2.1 Static Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . 118 6.2.2 Dynamic Noise and Linearity . . . . . . . . . . . . . . . . . . 120 6.3 Power Consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 6.4 Comparision Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 Discussions and Future Work 129 7.1 Effects of Applied optimizations . . . . . . . . . . . . . . . . . . . . . 130 7.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 7.2.1 Digital Optimization . . . . . . . . . . . . . . . . . . . . . . . 132 7.2.2 Programmable On-Chip Post-Filtering . . . . . . . . . . . . . 132 7.2.3 Resolution Scalable DAC . . . . . . . . . . . . . . . . . . . . . 132 7.2.4 Active Input Switch . . . . . . . . . . . . . . . . . . . . . . . 132 135 A ADC Fundamentals Linear Signal Processing . . . . . . . . . . . . . . . . . . . . . . . . . 135 A.1.1 Ideal ADC Model . . . . . . . . . . . . . . . . . . . . . . . . . 136 A.1.2 Quantization Noise . . . . . . . . . . . . . . . . . . . . . . . . 136 A.2 Non-Ideal ADC Model . . . . . . . . . . . . . . . . . . . . . . . . . . 137 A.2.1 Effective Resolution . . . . . . . . . . . . . . . . . . . . . . . . 138 A.2.2 Random Noise in SAR ADCs . . . . . . . . . . . . . . . . . . 138 . . . . . . . . . . . . . . . . . . . . . . . 139 A.1 A.3 Performance Normalization A.3.1 Sampling Rate . . . . . . . . . . . . . . . . . . . . . . . . . . 139 A.3.2 Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 A.3.3 Figure of Merit . . . . . . . . . . . . . . . . . . . . . . . . . . 141 9 List of Figures 1-1 Existing designs in a power-input frequency space (data courtesy B. G insburg, M IT). 1-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 Existing designs in a power-resolution space (data courtesy B. Gins- burg, M IT ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2-1 Effective resistance of a CMOS Transmission Gate. . . . . . . . . . . 26 2-2 Origin of charge injection error in a sampling circuit. . . . . . . . . . 26 2-3 Gain compression of a low voltage amplifier. . . . . . . . . . . . . . . 28 3-1 Logical block diagram of a successive approximation register ADC. . 34 3-2 SAR sampling phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3-3 SAR bit cycling of MSB. . . . . . . . . . . . . . . . . . . . . . . . . . 36 3-4 Fully Differential DAC during sampling. 37 3-5 Errors in 3 bit ADC transfer characteristics: error (c)Gain error (d)Linearity error [22]. 3-6 . . . . . . . . . . . . . . . . (a)Ideal case (b)Offset . . . . . . . . . . . . . . . 38 Effect of capacitor array top-plate parasitics (a)During sampling (b)During bit-cycling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 3-7 Effect of errors in Vregp. . . . . . . . . . . . . . . . . . . . . . . . . . 41 3-8 Equivalent sample-and-hold circuit. . . . . . . . . . . . . . . . . . . . 42 4-1 Block diagram of the final ADC. . . . . . . . . . . . . . . . . . . . . . 46 4-2 Final ADC conversion plan for (a)12-bit mode (b) 8-bit mode. .... 47 4-3 Waveforms showing power-gating control . . . . . . . . . . . . . . . . 50 4-4 Waveforms showing standard bit-cycling signals. . . . . . . . . . . . . 52 11 4-5 Waveforms showing self-timed bit-cycling signals. . . . . . . . . . . . 53 4-6 Delay for amplifiers with various output resistances. . . . . . . . . . . 56 4-7 Comparator gain structures (a)Linear amplifier cascade (b)Regenerative am plifier. 4-8 4-9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 Power-delay (normalized to CL) versus total gain for linear and regenerative am plifiers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 Auto-zeroed multi-stage preamplifiers. 60 . . . . . . . . . . . . . . . . . 4-10 Amplifier network (a)During auto-zeroing (b)During bit-cycling. . . 61 . . . . . . . . 63 4-12 Effect of changing GM at constant amplifier power-delay. . . . . . . . 64 4-13 Architecture of final comparator. . . . . . . . . . . . . . . . . . . . . 65 4-14 Architecture of final DAC. . . . . . . . . . . . . . . . . . . . . . . . . 66 4-15 Typical main-DAC/sub-DAC implementations . . . . . . . . . . . . . 68 4-16 Fully passive main-DAC and sub-DAC. . . . . . . . . . . . . . . . . . 68 4-17 Thevenin equivalent circuit for analyzing passive sub-DAC. . . . . . . 69 4-18 Differential-mode offset due to varying auto-zeroing voltage. . . . . . 70 . . . . . . . . . . . . . 71 4-11 Normalized total power-delay of auto-zeroed amplifier. 4-19 Common-mode independent charge sampling. . 4-20 Passive auto-zero reference voltage generation using capacitor arrays. 5-1 71 Implementation of switch sets in switch matrices (a)MSB switch set for positive array (b)Bits 10-0 switch set for positive array (c)MSB switch set for negative array (d)Bits 10-0 switch set for negative array. . . . 76 5-2 DAC network (a)During purging (b)During sampling. . . . . . . . . . 77 5-3 Distribution of injection error determined by poorly matched top-plate parasitic capacitances. 5-4 . . . . . . . . . . . . . . . . . . . . . . . . . . 78 Qualification of BCYCLEN and BCYCLEN signals with delayed sam pling signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5-5 Equivalent Resistance of input switch with respect to input voltage. . 80 5-6 Parasitic PN-junction on top-plate. . . . . . . . . . . . . . . . . . . . 81 12 5-7 Waveforms showing the origin of a transient spike during MSB bitcycling. 5-8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 Zoom-in of DAC differential output voltage as it recovers from a neg. . . . . . . . . . . . . . . 84 Implementation of purging switches . . . . . . . . . . . . . . . . . . . 85 5-10 Capacitor mismatch leading to largest DNL. . . . . . . . . . . . . . . 85 5-11 Effect of parasitic capacitance on sub-DAC top-plate. . . . . . . . . . 86 ative spike to the correct positive voltage. 5-9 5-12 Errors in ADC transfer characteristic due to compression in sub-DAC interpolation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 5-13 INL due to 200fF parasitic capacitance on top-plate of sub-DAC. . . . 88 5-14 INL after adjustment of coupling capacitance Cc. . . . . . . . . . . . 88 5-15 INL, after Cc adjustment, resulting from errors in estimation of parasitic capacitance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 5-16 Capacitor array layout (a)Conventional common-centroid (b)Equaledge ratio common-centroid. . . . . . . . . . . . . . . . . . . . . . . . 90 5-17 Bottom-plate to top-plate and dummy array coupling in poly-poly capacitor arrays. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5-18 Implementation of auto-zeroing switches (a)First and second stage of 12-bit cascade (b)Last stage (only stage for 8-bit path). . . . . . . . . 91 5-19 Pream plifier circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5-20 DAC and preamplifier waveforms during overdrive recovery . . . . . . 94 5-21 Change in cascade time-constant with respect to fractional changes in Gm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-22 Preamplifier biasing source. 96 . . . . . . . . . . . . . . . . . . . . . . . 97 . . . . . . . 98 5-24 Offset compensating latch during auto-zeroing. . . . . . . . . . . . . . 99 5-25 Block diagram of latch calibrating circuit . . . . . . . . . . . . . . . . 100 5-26 Latch voltage-follower calibration circuit. . . . . . . . . . . . . . . . . 100 5-23 Simplified schematic of complete offset-calibrating latch. 5-27 Latch voltage-follower equivalent circuit during first half of auto-zeroing phase. ....... .. ................................... 13 101 5-28 Latch circuit waveforms during the auto-zeroing phase. . . . . . . . . 102 5-29 Offset compensating latch during reset-resolve phase. . . . . . . . . . 103 5-30 Latch circuit waveforms during the reset-resolve phase. . . . . . . . . 104 5-31 Replica biasing circuit for latch. . . . . . . . . . . . . . . . . ... . . . . 105 5-32 Gain and phase of filtered and unfiltered auto-zeroing calibration circuit. 107 5-33 Latch level restoring circuit. . . . . . . . . . . . . . . . . . . . . . . . 108 5-34 Latch level restorer waveforms. . . . . . . . . . . . . . . . . . . . . . 108 5-35 SAR state machine circuit. . . . . . . . . . . . . . . . . . . . . . . . . 109 5-36 Metastability recovery circuits (a)Metastability detector (b)Metastbility reset......... ..................................... 110 5-37 Circuitry supporting resolution scaling. SLEEP signal is used to enforce power-gating. . . . . . . . . . . . . . . . . . . . . . . . . . . . .111 5-38 Internal clock gating circuitry. . . . . . . . . . . . . . . . . . . . . . .111 5-39 Charge pump circuit to generate voltages beyond VDD -........... 5-40 NMOS Dickson voltage multiplier [39]. .112 . . . . . . . . . . . . . . . . . 113 5-41 Voltage muliplier simulation. . . . . . . . . . . . . . . . . . . . . . . . 113 6-1 Micrograph of fabricated test chip. 116 6-2 Micrograph of full ADC with offset calibrating latch. 6-3 ADC Test setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 6-4 ADC Test PCB photograph. . . . . . . . . . . . . . . . . . . . . . . . 118 6-5 Code density histogram of ADC in 12 bit mode. . . . . . . . . . . . . 119 6-6 DNL and INL of ADC in 12 bit mode. . . . . . . . . . . . . . . . . . 120 6-7 Code density histogram of ADC in 8 bit mode . . . . . . . . . . . . . 121 6-8 DNL and INL of ADC in 8 bit mode. . . . . . . . . . . . . . . . . . . 121 6-9 ENOB versus input frequency for the ADC in 12 bit mode and 8 bit m ode. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 6-10 FFT of ADC output with 47.3kHz input tone. . . . . . . . . . . . . . 123 6-11 ADC power consumption with respect to sampling rate. . . . . . . . . 124 14 6-12 Figure-of-merit of this ADC compared with previous implementations (data courtesy B. Ginsburg, MIT). . . . . . . . . . . . . . . . . . . . 126 6-13 Power consumption of SAR ADCs with respect to input frequency (data courtesy B. Ginsburg, MIT). . . . . . . . . . . . . . . . . . . . 127 6-14 Power consumption of SAR ADCs with respect to resolution (data A-1 courtesy B. Ginsburg, MIT). . . . . . . . . . . . . . . . . . . . . . . . 127 . . . . . . . . . . . . . . . . . . . . . . . . . 136 Ideal model of an ADC. 15 List of Tables 6.1 Simulated and measured power consumption of ADC blocks. . . . . . 125 6.2 ADC performance summary. . . . . . . . . . . . . . . . . . . . . . . . 128 17 18 Chapter 1 Introduction Wireless sensor networks have received a lot of attention recently by various sectors of the research community. Although protocols and requirement specifications are still being defined at the communication, network architecture, node architecture, and circuit levels, design constraints and applications are already emerging. The vision of a micro-sensor network includes dense, intelligent nodes that are energy-autonomous and that operate and are deployed in an ad-hoc manner. Such networks have diverse applications ranging from military surveillance, reconnaissance, and damage assessment to environmental forest fire detection [1] and industrial process monitoring. The design of micro-sensor node hardware is constrained by several factors, many of which can be derived from collective considerations of the target applications. To be energy-autonomous, nodes must be powered entirely by an energy harvesting source. This places demanding, low-energy requirements on the constituent circuits. Ad-hoc deployment and operation requires that nodes be fault tolerant and able to adapt to unpredictable environments and network characteristics. Finally, high density and ubiquity places a cost constraint on nodes, reducing their acceptable price per unit to a few cents. Fundamentally, the architecture of an intelligent sensor network node must consist of an analog-to-digital conversion front-end (ADC), a digital signal processor (DSP), and a short range radio. Additionally, a power subsystem, which might be a simple as an energy scavenger regulator or as sophisticated as a dynamic voltage scaling DC- 19 DC converter, is a highly desirable peripheral [2][3]. This document will focus on the development of an ultra low power ADC suitable for sensor network nodes. Although design specifications and optimizations will be undertaken with this application in mind, the circuit techniques developed, and indeed the final ADC, will be useful for a variety of low power systems. The remainder of this chapter discusses the preliminary considerations associated with ADC design. Namely, Section 1.1 examines the target application and derives the performance requirements. Additionally, features beneficial, but not limited, to sensor applications are specified here. Section 1.2 surveys existing designs as a means of identifying ADC architectures that would be appropriate for this, low-power, implementation. 1.1 Requirement Specifications Within the vast application space encompassed by wireless sensor networks, acoustic monitoring of a battle field environment is an example that requires a sophisticated sensor node and, in particular, a non-trivial ADC. Consequently, it is a good vehicle for developing a system of general value. The performance requirements and features specified in the following subsections have largely been derived by anticipating possible functions and algorithms the sensor nodes might perform. In an effort to maintain broad usefulness, aggressive requirements, that also provide some academic gratification, have been specified. 1.1.1 Resolution In the case of an acoustic sensor front-end, the microphone, preamplifier (if it is present), and ADC all impose noise and distortion on the system. The microphone, can have a dynamic range in excess of 100dB [4][5]. The preamplifier, at the cost of power consumption, may be designed for virtually any dynamic range desired. In this case, the ADC limits the dynamic range of the front-end. As explained in Appendix A, all ADCs introduce quantization noise, limiting the sensor system's ability to 20 detect minute signals. The magnitude of acoustic activity is measured by sound-level or sound pressure level (SPL). The SPL associated with a quiet field is approximately 30dB, while that associated with a large (military) vehicle, one meter away, is approximately 100dB. Consequently, 70dB, which corresponds to just under 12 bits of quantization noise, is an upper limit of the dynamic range required of the ADC. Accordingly, in this design, a 12 bit ADC is pursued. 1.1.2 Sampling Rate Although the acoustic frequency band extends to 20kHz, military interest of the spectrum in a battle field is far more limited. Vehicular and personnel acoustic activity is typically within the kilo-Hertz range. As a result, a Nyquist rate converter with a sampling rate as low as 2-3kS/s may be sufficient. However, to avoid in-band signal-to-noise degradation due to aliasing, this imposes stringent requirements on the pre-filters. Specifically, very high-order anti-aliasing filters are required. Typically these are achieved through active circuits that may consume more power than the entire ADC. To allow exploration of a variety of anti-aliasing options, this ADC will be designed to have a sampling rate of 100kS/s. Of course oversampling, in this manner, would require post-ADC digital decimation filtering. 1.1.3 Scalability As mentioned previously, low-power operation is a paramount design objective for sensor subsystems. Since sensor nodes are reactive, and the ADC is the front-end unit in the detection path, conventional sleep-active state manipulations are of limited use. In particular, some form of ADC is responsible for detecting an event of interest and, accordingly, generating a wake-up signal. Of course, the level of sophistication required in that initial detection depends on the application, and may be met by a simple threshold detector. In any case, the ADC must remain "on" in some capacity. Where the power source is erratic and unpredictable, such as the case of an energy 21 scavenger, it is beneficial to scale the power consumption in response to the dynamically varying constraints. The two main dimensions, along which an ADC may be scaled in order to recover power savings are sampling-rate and resolution. In the extreme, the power consumption of the ADC should be vanishingly small. Highly reconfigurable solutions have been demonstrated [6] [7]. The characteristic of scalability is itself a feature that can impose considerable overhead, in the form of the reconfiguration logic/fabric as well as adaptive biasing circuitry. It is essential, therefore, to limit the performance space based on need, but also, based on ease of implementation. It, of course, follows that facilitating scalability should be considered early on, at the architecture selection phase. Precise specification of the scalability feature of this ADC will be deferred until the treatment of architecture and implementation issues. 1.1.4 Input Interface As discussed in Section 3.2, a differential ADC architecture is highly beneficial for robustness. This may, however, increase system complexity, as it requires a fully differential input signal path. Since the power consumption of the sensor system is liable to increase with the added complexity, this ADC will be designed to support single-ended inputs, though some degradation in precision is expected in this case. 1.2 Architecture Selection Architecture selection can begin with regards to the resolution, input bandwidth, and scalability requirements specified above. Since low-power operation is critical in this design, architecture selection will be driven by an examination of the power consumption of previous ADCs. Figure 1-1 plots the power consumption of existing designs against the maximum input frequency they can digitize. As shown, in the 1OOkS/s space, successive approximation register (SAR) ADCs and oversampling, delta-sigma ADCs achieve the lowest power. Pipelined and flash converters populate the high-speed, higher-power regime. Similarly, Figure 1-2 plots the power consump- 22 tion of existing designs against their resolution. Here, very low-power SAR converters are prominent at the 8 bit level. However, at the 12 bit level, the lowest power designs demonstrated are delta-sigma converters, though a couple of SAR converts also achieve reasonably low-power. Power VS Input Frequency A 10 0X 10 -2 10 -1 0 10~10' 0 1-5 0 106 O * Flash o SAR -O x Pipelined 10 102 104 10 108 Input Frequency (Hz) Figure 1-1: Existing designs in a power-input frequency space (data courtesy B. Ginsburg, MIT). With respect to the desired speed and resolution, SAR and oversampling ADC architectures are the primary candidates for this design. Delta-sigma modulation greatly eases the implementation requirements of analog blocks, and, for preserving accuracy, places an increased emphasis on digital circuits. This is amenable to modern technology scaling trends, and, as a result, efficient, low-voltage designs have been demonstrated in advanced technologies [8][9]. Delta-sigma converters benefit from the additional advantage that they fundamentally rely on oversampling, easing antialiasing requirements. Nonetheless, among the designs plotted in Figure 1-1 and Figure 1-2, the lowest power are SAR converters, achieving power levels of submicro23 Power VS Resolution Aj Xp 10 A AA A& 10-2 A A 10 * 0 10 0 A A A A A AA 10-5 - Flash SAR 0 x Pipelined A Al * 8 10-6 10-7 A 0~ 6 8 10 12 14 16 Resolution (Bits) Figure 1-2: Existing designs in a power-resolution space (data courtesy B. Ginsburg, MIT). Watts and micro-Watts [10][11]. This design will leverage the SAR architecture, which is suitable for micro-power operation, and apply new techniques to efficiently increase resolution capabilities to the 12 bit level. 24 Chapter 2 Precision Limitations Having identified the precision requirements for this ADC, we can begin to identify limitations in the implementation technology that must be addressed. This section surveys practicalities in circuits and devices that are relevant to the development of medium and high resolution ADCs. Often, these limitations will be overcome by appropriate circuit techniques, layout, and device sizing. Corresponding details are provided in Chapter 5. 2.1 Low Overdrive of MOS Switches The availability of ideal switches is extremely beneficial in analog circuit design. In this ADC, switches are used to control signal flow, enable charge redistribution on capacitors, and short nodes to ensure voltage equivalence. functions as a nearly ideal switch. At low gate overdrives (IcVGS The MOS transistor - VTN,P I), it provides very high "off" resistance, and at high gate overdrive, it provides relatively low "on" resistance. In low-voltage analog circuit design, however, reduced headroom limits the gate overdrive that can be applied. The effective resistance of a MOS transistor in deep triode is given by Equation 2.1. 1 (2.1) Reff,N,P AN,PCOX(1 25 IGS - VT N,P Here, PN,P is the device mobility, Cox is the oxide capacitance, and (1)are the device dimensions. The resulting range of voltages effectively passed by an NMOS is between 0 and VDD-VTN; VDD. The resistance of a CMOS transmission gate, such as the one shown in Figure the corresponding range for a PMOS is between VDD-IVTPI and 2-1, is the parallel combination, Reff,N Reff,P- In the limit, when VDD < VTN + IVTPI, a CMOS transmission gate approaches the "off" state for voltages near mid-supply, regardless of gate voltage. VIN R'L-RN-f Reff,N__ |0 Cefp VOUTtefP Figure 2-1: Effective resistance of a CMOS Transmission Gate. At low overdrive voltages, where the CMOS transmission gate resistance is high, we are faced with increased settling times due to the resulting RC time constant. Additionally, as this resistance becomes more dominate, its nonlinear nature (through the dependence on IVGS 2.2 - VTN,PI) becomes highly significant. Charge Injection Errors An additional limitation associated with MOS switches is the error in voltage observed on high-impedance nodes immediately following sampling. Consider the sampling circuit shown in Figure 2-2. VSAMP vF 2 r Rs TI CGSD-L COL Cs CGSD 'COL VOUr Cu Figure 2-2: Origin of charge injection error in a sampling circuit. 26 As VSAMP is reduced following sampling, the channel charge established in the NMOS escapes through its source and drain. The resulting voltage error depends on the interaction of the load impedances, CH, CS, and RS, with the MOS intrinsic capacitance, CGS,D, and extrinsic capacitance, COL. In strong inversion, the intrinsic capacitance is fixed at approximately 4 [12). In weak inversion, the intrinsic capac- itance is nonlinear and vanishes exponentially [13]. In both cases, circuit simulators, such as SPICE, represent charge injection behavior inaccurately. 2.3 Mismatch of Passive Elements Among the errors in an ADC's transfer characteristic, linearity errors are most difficult to correct. In a SAR ADC, the dominate source of nonlinearity is element mismatch in the feedback DAC (as explained in Section 3.3). To minimize this, highly linear and well matched passive elements are employed. Integrated poly-poly capacitors exhibit superior matching characteristics and are extremely stable, having low voltage and temperate coefficients (10-100 ppm/V and 20-30 ppm/SC respectively). Although native, untrimmed matching of +0.01% (suitable for up to 14 bits of precision) is achievable, this depends on proper sizing and layout techniques. 2.4 Mismatch of Active Elements Geometric tolerances as well as process dependant effects including diffusion, etch, thermal, and stress gradients all contribute to mismatch in MOS transistors. This mismatch results in variation that can be referenced to the physical MOS parameters VTO, -y (the body-effect coefficient), /t, Cox, W, and L [14]. All of these variations, except those associated with the one-dimensional quantities W and L, can be convolved with a double box function, defining the geometry of the devices, to quantify the overall variation. One dimensional analysis can be applied to W and L. If the proper layout practices of common-centroid arrangement and tight proximity are applied to matched devices, the residual spatial variations will have short 27 correlation distances with respect to the device dimensions and can be treated as random "white noise". In this case, the short range mismatch parameter associated with VTO dominates and can be used to estimate statistical variation between devices. At the circuit level, mismatch between devices manifests itself as an offset voltage which can be referred to the input of amplifiers. In a SAR ADC, offsets affecting the comparator result in overall offset of the ADC transfer characteristic. Though this can be calibrated using external hardware [15] or software [16], circuit support for offset cancellation can also be employed (as explained in Section 4.2). Note, comparator offset depends on temperature and is typically not static. This is particularly true in low-voltage implementations where the active amplifiers operate in weak inversion. An additional consequence of device mismatch is the resulting degradation in amplifier gain. Consider the simple amplifier in Figure 2-3. Assuming weak inversion operation of all devices, the available linear output range with a 1V VDD is approximately 700mV. Outside this range, all devices are no longer in saturation, and severe gain compression occurs. For devices with an intrinsic gain (gmro) of 20, the linear gain of the amplifier is approximately 10 (gm,N(roN 11 ro,P) where gm,N gm,P in weak inversion). The resulting input linear range is only 70mV. Any input referred offset of the same order will cause degradation in the amplifier's gain. Consequently, device offset must be considered when designing gain stages. 1V >10mV 0-] >10mV >10mV Figure 2-3: Gain compression of a low voltage amplifier. 28 2.5 Device Noise Noise is fundamental to circuits and cannot be avoided. Nonetheless, managing noise is critical particularly in the case of low-power, precision systems. Low-power implies either low voltage or low current. In both cases, the ability to derive signal power is reduced, and the limiting effect of noise power on achievable dynamic range is more significant. The noise inherent to electrical conduction, which comes about due to the discrete nature of charge, is known as shot noise. Assuming the arrival of charge carriers (i.e. electrons) is governed by Poisson statistics, and by treating their arrival as an impulse of current (causing a step change in charge), the variance in current, in a bandwidth Af, is given by the well known relationship of Equation 2.2[17]. 2 = 2qISIGAf (2-2) The signal power can be represented by ISIG and the noise power can be represented by a-. Accordingly the signal-to-noise ratio (SNR), can be represented by Equation 2.3. ISIG _ ISIG 2qAf 12- The important result of this relationship is that, SNR can be increased either by increasing the signal (resulting in increased power consumption) or by reducing the bandwidth. If this analysis is approached from a thermal noise point of view, an identical relationship is observed [18]. In a SAR ADC, this noise limits precision in at least three places. During initial sampling of the input signal (Section 5.1.2), we are limited by CSAMP noise due to the sampling switch, where CSAMP is the sampling capacitor. During auto-zeroing (Section 4.2.1), we are limited by BC Uk noise due to the active amplifier, where B is an implementation dependant parameter set by the amplifier and CAZERO is the auto-zeroing capacitor. Finally, during bit cycling, we are limited by noise in the 29 comparator (Section 5.2.1) which originates in the preamplifiers and latch. This last case can be extremely difficult to quantify, as comparator decisions do not depend on the average power of the input signal. Instead, a transient analysis that considers the time varying nature of the latching circuit must be applied. Since the arrival of electrons is treated as an impulse of current, the shot or thermal noise considered above has constant power over a broadband of frequencies. As a result, it is called "white noise". An additional type of noise, not inherent to conduction, but prominent in MOS devices, nonetheless, is -, or flicker, noise. Surface effects at the channel's silicon-oxide interface result in extraneous energy states (or "traps") for the carriers. Flicker noise is highest at low frequencies. Its power increases with device current (in a linear or quadratic manner depending on the degree of inversion) and decreases with channel area. Flicker noise affects comparator decisions if it is not cancelled appropriately. 2.6 Threshold Voltage Hysteresis In a MOS device, inversion occurs as a result of energy band bending of the silicon in the channel region. This is induced by the electric field originating at the gate. At the on-set of strong-inversion, the concentration of channel forming carriers increases quickly with increasing gate voltage. Consequently, the strengthening electric field terminates at the silicon-oxide interface, and the additional voltage drop appears across the oxide. The ensuing reduction of oxide energy bands results in low-energy traps that can be filled by carriers via tunneling mechanisms [19]. Subsequently, when the gate voltage is reduced, carriers can take tens of milliseconds to leave the oxide traps and re-enter the silicon. The presence of the carriers in the oxide long after the stress voltage is removed leads to threshold voltage hysteresis in devices. This behavior is only observed for positive stress voltages and, as a result, affects differential inputs asymmetrically. The net effect manifests itself as a hysteresis in amplifier offset voltage. Threshold voltage hysteresis, on the order of millivolts, has been observed [19]. 30 The effects are far less pronounced in PMOS devices than NMOS devices, and can be managed further by minimizing signal induced band bending. 2.7 Other Error Sources In addtion to the error sources mentioned above, power supply noise and substrate noise can have a considerable effect on the precision of the ADC. Both of these noise sources may be correlated with signal transitions in the circuits. Particularly in the case where large swing digital signals couple to the sensitive analog blocks, like the comparator preamplifiers and latch, the ADC output can be subject to periodic error. For instance, since transitions in the SAR state machine occur in a regular fashion, spurious noise, associated with some fraction of the clock frequency, might degrade the ADC's precision. Lastly, kickback noise from the comparator, which comes about when the outputs start regenerating, is not particularly limiting in a SAR ADC. In a flash converter this noise appears at the inputs, which couple to other comparators through the resistive string. In the SAR ADC, only one critical decision is made at any time, and the comparator is reset between decisions, negating previous kickback effects. 2.8 Summary Details of how each of limitations above are managed are provided in the following chapters. Generally, both architecture and circuit techniques have been applied. For instance, low switch overdrive is overcome, in some cases, by bootstrapping, and, in other cases, by applying an architecture that avoids analog switches in the critical signal path. Similarly, charge injection errors are minimized through the use of counter-phased dummy switches as well as switching sequences that direct channel charge onto the appropriate nodes. Mismatch in passive and active elements are managed through proper sizing. Additionally, amplifier offset cancellation techniques are used. Finally device noise is limited in preamplifiers though gain, bandwidth, and 31 biasing considerations. 32 Chapter 3 Successive Approximation Conversion This chapter will describe the structure and operation of successive approximation (SAR) analog-to-digital converters. In Section 3.1, the basic operation of a SAR ADC will be considered. In Section 3.2, an enhancement to the architecture, enabling robust, differential operation, will be presented. Finally, in Section 3.3 constituent blocks of the SAR ADC will be analyzed for their contributions to errors in the overall conversion process. 3.1 Basic Operation Successive approximation register (SAR) ADCs are very popular for medium and high resolution applications. They achieve reasonably quick conversion times since they are based on an efficient convergence algorithm. A logical block diagram of a SAR converter is shown in Figure 3-1 below. Here the main blocks are a digital finite state machine (itself called a successive approximation register), a digital-to-analog converter (DAC), a sample-and-hold (S/H), and a comparator. The input signal offsets the DAC's output value, and the finite state machine SAR applies a binary search algorithm to the DAC to arrive at a digital representation of the analog input. In practice, the sample-and-hold is often combined with the DAC, 33 DAC + + VIN Comparator S/H Clock Figure 3-1: Logical block diagram of a successive approximation register ADC. and the entire structure is implemented using binary-weighted switch capacitors. Here, details of the conversion process for an N bit SAR ADC are described conceptually [20]. Operation, in the case of the practical implementation, employs additional enhancements. Basically, conversion requires two phases: sampling, during which the input voltage is stored, and bit-cycling, during which bits of the corresponding digital code are successively resolved. In particular, the digital code corresponds to where the analog input voltage lies within a range defined by the reference voltages, and Vref,p. Vref,N As shown in Figure 3-2, the input voltage is initially sampled onto the bottom plates of the DAC capacitor array. The resulting charge on the top plate of the capacitor array, QTOP, is equal to -(VIN - VrefN)2NCo. Next, bit-cycling starts by resolving the MSB. First, the bottom plate of the largest capacitor (2N- 1 C) is connected to the positive reference voltage, Vref,P, while the bottom plates of the remaining capacitors in the array are connected to the negative reference voltage, Vref,N, as shown in Figure 3-3. The resulting structure forms a capacitor divider with equal capacitance from the output to Vref,p and Vef,N- If we neglect the charge initially sampled on the top-plate, the output node will settle to 2refP refN. Instead, the sampled charge superimposes a voltage on the top node, 22 and the resulting output is VrefN.e, - 1 following charge conservation analysis. 34 IN + Vref,N. This can be verified by the Sample QTOP=-(VN-VrefN)2NCo STOP 2 VrefN VrefN 20o S1 SN-1 + SRES so VVref VIN To SAR Co O N SIN Ve VVrefN Figure 3-2: SAR sampling phase. QTOP = -(VIN - Vref,N)22NCON (VOUT - Vre)NCo fN)N1CO + (3.1) VOUT T = Vref,P + Vref,N 2 -VIN (3.2) -± Vref,N From this, we can see that for VIN equal to the mid-scale voltage VrefP+Vref,N ,VOUT 2 equals Vref,N, which is the comparator transition point. If the comparator resolves to "1", VIN is greater than the input mid-scale; if it resolves to "0", VIN is less than the input mid-scale. In the former case, the MSB capacitor bottom plate remains at Vref,p during subsequent bit-cycling. In the ladder case, the MSB capacitor is connected to Vref,N. The remaining bits are resolved in a similar manner, and the digital output code corresponds to the successive comparator decisions. 3.2 Differential Successive Approximation ADC The ADC discussed in Section 3.1 is not practical for medium or high resolution conversion. Specifically, it suffers from a high sensitivity to power supply noise. For instance, a 12 bit ADC, depending on Vref,p and Vref,N, will have an LSB voltage less 35 Bit Cycle QTOp=-(VIN-VrefN)2NCo STOP 20Co Vref,N 2 To SAR CO S- Vrefp VIN-N SIN Vrf, Vref,N Figure 3-3: SAR bit cycling of MSB. than 1mV. Supply noise will be attenuated through the capacitor divider and appear at the comparator input, degrading the resolvable signal. To mitigate this effect, a fully differential architecture can be employed. Figure 3-4 shows the resulting structure. Here, VCM is the common-mode voltage of the input signal. It ensures that equal and opposite charge is stored on each of the capacitor arrays. This, in turn, guarantees that the common-mode voltage of the DAC's output, during bit-cycling, does not change from conversion-to-conversion, regardless of the input common-mode voltage. ADC errors due to varying input common-mode are considered further in Section 4.2.2. In addition to improved power supply noise rejection, this approach has several other advantages. Firstly, the differential nature provides first-order cancellation of linear voltage coefficients in the DAC capacitors [21]. Non-zero capacitor voltage coefficients result in linearity errors, and as a result, the differential ADC improves linearity. Second, the thermal noise limited dynamic range of the entire ADC is doubled. This can be seen by considering that the input range has been extended from (Vref,P - Vref,N) to -(Vref,P - Vref,N). As a result the LSB voltage doubles, and the required minimum resolvable signal is larger. Note, however, if the ADC is 36 VIN+ Vref,N 2N-1C + _F eVrefP VrefN VIN- Figure 3-4: Fully Differential DAC during sampling. quantization noise limited, instead of thermal noise limited, the differential structure provides no improvement in dynamic range. Additionally, even in the thermal noise limited case, the extension in dynamic range comes at the cost of an increase in power consumption, since two capacitor arrays must be charged. The power increase in the DAC is, however, offset somewhat by a power decrease in the thermal noise limited comparator stages. 3.3 Analysis of Error Sources Any deviation from the ideal ADC transfer characteristic represents errors in the analog-to-digital conversion process. Figure 3-5 below enumerates static errors commonly observed in ADCs with respect to the ideal characteristics of a single-ended converter. A differential converter may be considered by means of a straight forward extension. 37 Any shift in the transfer curve is called offset error and results in a reduced conversion range. For instance, in Figure 3-5b, even if the input signal is shifted, the first -q16 of the analog input range is not properly converted, saturating the digital output at the lowest code. As shown in Figure 3-5c, equal deviations (from K/) in the width of code transition voltages results in gain error. Finally, as shown in Figure 3-5d, unequal deviations in the width of code transition voltages results in linearity errors. Digital Code Digital Code 111 - 111 - 110 110- 101 101- 100 100- 011 011 - 010 010- y Ideal Actual -. - ,v. 001 --doo 0001 / (a) ... '. Digital Code ref - Digital Code 111 - *Ideal 110101- VI/Vref % (b) Actual 111 - %/4 Ideal Actual 110- - -' ./ 101- 100- 100- 011- 011- 010- 010- 001- 001- 000 000 ref (c) - -j.... -y. I // r- ' %/ 3/ 1 VI/Vref (d) Figure 3-5: Errors in 3 bit ADC transfer characteristics: (a)Ideal case (b)Offset error (c)Gain error (d)Linearity error [22]. 38 In addition to the static errors considered in Figure 3-5, dynamic errors also influence the performance of ADCs. Non-linearities, for instance, may be dependant on the frequency of the input signal. Additionally, thermal noise in the converter is most clearly observed through dynamic analysis since static analysis typically applies averaging which attenuates zero-mean, random noise. Methods for precisely quantifying and correcting ADC errors are described in [22]. Among these, linearity errors are the most difficult to correct through post processing, and as a result, represent the most serious non-idealities in an ADC. The remainder of this subsection will examine the major sources of errors in a SAR ADC and how these manifest themselves with respect to the static and dynamic characteristics of the converter. 3.3.1 Capacitor Mismatch From the discussions in Section 3.1 it is clear that the DAC generates output voltages by means of capacitive voltage division. It follows, then, that ratiometric matching, and not absolute matching, of the capacitive elements is required. Since output values of the capacitor divider define the transition voltages, capacitor mismatch results in inconsistent deviations in the transition voltages, leading to linearity errors. This is a primary source of nonlinearity in the SAR converter. 3.3.2 DAC Parasitics During both sampling and bit-cycling the bottom-plates of all elements in the capacitor array are driven to either VIN, Vref,p, or Vref,N. The parasitics on the bottom plates are also driven to the corresponding voltage without affecting the sampled charge. Consequently, bottom-plate parasitic capacitance does not introduce errors. The top-plate of the capacitor array is a high impedance node during bit cycling and, as a result, the distribution of sampled charge is affected by associated parasitic capacitance. To understand this, consider the single-ended capacitor array of Figure 3-6a. The first thing to note is that, during sampling, the top-plate, and its parasitic capacitance, is charged to Vref,N 39 which is precisely the voltage towards which the DAC output converges during critical comparator decisions. This implies that, during bit-cycling, when the capacitor array is switched so as to output VrefN, the top-plate parasitic will hold exactly as much charge as it did during sampling. The remaining charge on the top-plate, which corresponds to the sampled voltage, will be appropriately distributed between the elements of the capacitor array. When the capacitor array is switched so as to output anything other than Vref,N, the result can be analyzed through the Thevenin circuit equivalent shown in Figure 3-6b. Here, the output voltage is attenuated through the capacitor divider composed of the DAC capacitor array, 2 NC, and the parasitic capacitance Cp. As a result, of this attenuation, the comparator must be able to resolve smaller signals. Vref,N S 2CO o p .--- VIN (a) Vref p Z bi2'Co Thevenin N Equivalent 2 N Cp VDAC,actual 2 -0 N-1 VDAC,ideal i--O .'P L, Vref,N (b) Figure 3-6: Effect of capacitor array top-plate parasitics (a)During sampling (b)During bit-cycling. 40 3.3.3 Reference Voltage Error While parasitic capacitance on the top-plate causes attenuation of the DAC output around the comparator trip point (Vref,N), staic reference voltage error on Vref,p causes attenuation (or amplification) with respect to ground. Namely, the DAC output voltage, referenced to Vef,N, is attenuated (or amplified). The resulting equiv- alent circuit is shown in Figure 3-7. referenced to Vef,N, Here, VDAC,ideal is the DAC output voltage rather than ground, and G represents the attenuation (or am- plification) as a result of Vef,p errors. The overall effect is a systematic compression (or expansion) of transition voltage widths leading to gain error in the ADC transfer characteristic. Static errors on Vref,N have a similar effect of compressing (or expand- ing) the DAC output voltage range, and also result in ADC gain error; in this case, the comparator input common-mode voltage is also shifted. GVDAC,ideadbN-1, -.. , 0 Vref,N + Vref,N > Figure 3-7: Effect of errors in Vref,p. From the argument above, it follows that non-static noise on the reference voltages can be highly problematic in the ADC. It is essential, therefore that the reference be properly de-coupled from other signals or power supplies. 3.3.4 Comparator Offset Figure 3-1 illustrates how the DAC output voltage is offset by the analog input signal before being resolved by the comparator. Comparator offset can be modeled as an additional summing operation with a static error signal at the comparator input. This can be combined with the ADC input signal, showing, explicitly, that comparator offset has the effect of adding offset to the entire ADC. 41 3.3.5 Sampling Switch Non-Linearity Since sample-and-hold (S/H) is the first operation and is applied directly to the input, any non-linearities in the S/H circuit contribute to overall ADC linearity errors. In SAR ADCs, sampling maybe a passive operation that occurs through MOS switches onto the capacitor array, as shown previously in Figure 3-2 and Figure 3-4. The capacitor array is typically quite linear. The sampling switches, however, are quite non-linear having an effective resistance dependant on the input signal voltage. This is not problematic at low frequencies, since the sampled voltage will track the input voltage, provided sufficient settling time. At high frequencies, however, the S/H circuit can be viewed as the voltage divider shown in Figure 3-8, where the sampled voltage is set by the non-linear MOS sampling resistance and the impedance of the capacitor array. R(VIN) VS/H VIN ZArray l/MWCArray Figure 3-8: Equivalent sample-and-hold circuit. Here, the sampled voltage, VS/H, and the input voltage, VIN, have the non-linear relationship, described in Equation 3.3, where R is a function of VIN. 1 VS1H = VIN 1(3.3) 1 + JWCArrayR While the error in magnitude observed due to this non-linearity might be small, the error in phase can be quite large, significantly contributing to distortion. 3.3.6 Comparator Thermal Noise In medium and high resolution SAR converters, the comparator must resolve signals of only tens or hundreds of micro-volts. Transient noise during the course of comparator settling, degrades the SNR of the ADC. Further, although SNR is effective 42 for characterizing noise power in the ADC, it does not fully elucidate the effect noise has on the resulting output code. In a SAR ADC, the code histogram resulting from noise can have a non-monotonic decay pattern from the mean code [23] as it depends on the joint probabilities of how individual bits are resolved (this is considered further in Appendix A). As a result, the effect of noise on the resulting output code, in addition to the dynamic range, should be considered. 3.4 Summary All of the error sources mentioned in this chapter can lead to significant degradation in ADC performance. Consequently, they are addressed in the implementation of this design. For instance, although the reference voltage is provided from off-chip, it is heavily de-coupled on-chip to minimize transient errors. Comparator offset is cancelled using auto-zeroing and latch calibration. Input switches are bootstrapped to reduce the effect of their non-linear behavior. Finally, thermal noise is limited to an acceptable level in the comparator preamplifiers through gain, bandwidth, and biasing considerations. 43 44 Chapter 4 Architecture Design and Theory This chapter discusses the system architecture of the implemented ADC. A high-level description of the converter is provided in Section 4.1 with the purpose of explaining global features affecting the conversion process and its control mechanisms. Block level architectures are considered in Section 4.2 along with analysis supporting the applied optimizations. Implementation details, pertaining to design tradeoffs and circuit techniques, are presented in Chapter 5. 4.1 Global Architecture and Concepts Figure 4-1 shows a block diagram of the final ADC. Most of the blocks were introduced in Chapter 3. Here, however, the DAC has been separated into two sub-blocks: a main-DAC and a sub-DAC. The precise effect of this structure will be considered in Section 4.2.2. An additional block, the clock manager, is used to implement scalability in sampling rate, as described in Section 4.1.2. 4.1.1 Conversion Plan In the most straightforward case, a SAR conversion cycle consists of one phase of input sampling, typically combined with offset calibration (auto-zeroing), followed by one phase of bit-cycling for each bit that must be resolved. This implies that for a 45 Sub DAC 7 V,, Main DAC VIN+ Vref N VIN-* Comparator SAR VreVref,N Clock Manager, CLK Figure 4-1: Block diagram of the final ADC. 12-bit converter, at least 13 clock cycles are required; for an 8-bit converter, at least 9 clock cycles are required. The conversion plans for the 12-bit and 8-bit modes of this ADC are presented in Figure 4-2. Here, the conversion plan includes several phases in addition to those described above. Specifically, a phase has been added to purge the capacitor arrays; the sampling and auto-zeroing phases have been separated; and an extra half-clock cycle has been devoted to resolving the MSB. Clock cycles have been allocated to resolving the remaining bits in a standard manner. The purpose of the purge phase, as well as the rationale behind separating offset calibration and sampling, is explained in Section 4.2.2. Details pertaining to the auto-zeroing phase are provided in Section 4.2.1. Finally, the extra time spent resolving the MSB is justified in Section 4.1.4. In any case, a 17 clock cycle conversion period (in the case of the 12 bit mode) im46 4.,2 17 clock cycles Clockn---~-~I llFL 2 d- a, A 1jI EBit-Cycle -- L (10:0) (a) 12 clock cycles Clock 2 aN CU 0 Bit-Cycle (6:0) E U CO 421 (b) Figure 4-2: Final ADC conversion plan for (a)12-bit mode (b) 8-bit mode. plies a clock frequency of 1.7MHz to achieve a conversion rate of 1OOkS/s. This ADC will be designed to operate at a clock frequency of 2MHz, such that the maximum conversion rate is slightly higher that 1OOkS/s. 4.1.2 Sample Rate Scaling Section 1.1.3 examined the value of scaling the performance of the ADC to recover power savings in response to dynamically varying energy and performance constraints. Assuming a constant supply voltage, the power consumption of both digital circuits and analog circuits is directly proportional to their operating speed. In the case of digital circuits, this can be seen in Equation 4.1, which expressed 47 the power consumption for digital transitions (0 -+ 1). Here, a is the switching activity factor, CL is the load capacitance, VDD is the supply voltage, fcIk is the clock frequency, and d is the duty cycle (i.e. processing) [24]. ratio of time the circuit is actively Often, d is combined with a and may be considered a specific parameter associated with switching activity. Pdi = aCLVDDfclkd (4.1) Additional forms of power consumption, namely direct-path power and leakage power, are also present. However, these typically represent a smaller portion of the total power consumption. In the case of analog circuits, we can consider, as an example, a simple single stage amplifier. Here, the required gain, A, and bandwidth, f-dB, are given by the following expressions, where GM is the amplifier's transconductance, and Rut is the amplifier's output resistance. A = GMRUt 1 f-3B = 1 27r Rout CL (4.2) (4.3) For a simple differential pair implementation, the amplifier's GM coincides with the transconductance of the input devices, gm, and can be expressed in terms of the bias current, 'bias (note, this relationship is true in the weak-inversion regime [18] where efficiency is highest). gM = Ibiasq nkT Panaq nkTVDD Combining these results, the power consumption can be expressed as in Equation 4.5, where its proportional dependence on signal frequency is shown. 2 7TnkTVDDCLAf-3dB Pana = q (4.5) .5 Fundamentally, this result suggests that a linear power relationship is achievable with respect to processing frequency. With consideration to Equation 4.1, this rela48 tionship occurs explicitly in digital circuits through fclk, the clock frequency, and d, the duty-cycle ratio. In analog circuits, however, some flexibility exists in how this relationship can be leveraged. In a sampled system, which performs analog processing in the discrete-time domain, reduced signal frequency implies a reduced sampling rate requirement. With reduced sampling rate, the settling time of the analog circuits can be increased to allow lower power operation. This may be achieved, as suggested by Equation 4.2 and Equation 4.4, by increasing Rt and decreasing GM. Such an approach, however, would require dynamic adjustment of bias currents, which can be difficult to implement, particularly over a large range [7]. A more straightforward strategy is to apply power-gating to the analog circuits, such that they process at a constant rate, and then shut-off to enable scalability in performance and power consumption. One of the main advantages of the SAR architecture, for low-power implementations, is its limited number of constituent active components. Specifically, the comparator is the only active analog circuitry. The SAR (finite-state machine) is a digital block, and the DAC is based on the passive charge redistribution process (Section 3.1). Consequently, the SAR ADC architecture provides the opportunity for very efficient power-gating; static bias currents in the comparator, as well as digital clocking circuitry, can be shut off between conversions, minimizing power consumption to leakage levels. During this process, the only overhead is CL VD energy of charging and discharging load capacitors. The waveforms in Figure 4-3 show how power-gating is controlled in this design. Here, CNVRT is the input control signal to initiate a conversion, DOUT is the 12bit digital output code, and SLEEP is an internal control signal used to shut-off the clock and active circuits between conversions. Figure 4-3 shows the case where the sampling rate has been scaled to half the maximum rate. Chapter 5 explains how sleep modes are implemented at the circuit level, such that only leakage power is consumed during the sleep phases. 49 CLK CNVRT DOUT Data Valid Data Valid SLEEP Sleep Active Conversion Sleep Active Converion Figure 4-3: Waveforms showing power-gating control 4.1.3 Resolution Modes Along with sampling rate, Section 1.1.3 identified resolution as a performance dimension along which scalability can be implemented to achieve dynamic power savings. The inherent relationship between power consumption and resolution can be derived by considering analog and digital circuits separately. Resolution is a measure of the loss in precision, due to noise, imposed by a processing block. Dynamic range is defined as the ratio of maximum signal power to minimum signal power when the SNR has degraded to unity. Note that the signal power, at unity SNR, is equal to the accumulated power of all noise sources. add quantization noise. All ADCs However, their ability to resolve a higher number of bits is ultimately limited by intrinsic device noise. To characterize the power-resolution relationship in analog circuits, the effect of thermal noise, on limiting dynamic range, can, once again, be analyzed for a simple one stage amplifier. Equation 4.6 relates dynamic range, DR,to the power of a full-swing sine wave, 2 V 2 ^ m, and the noise power, 22 vno- DR= The value of v2 is given by -/, V2 Amp 2vno (4.6) where y is the equivalent number of noise sources in the amplifier and CL is the load capacitance. Then, by applying Equation 4.2, Equation 4.4, and Equation 4.3, we get the following expression. DR ~~ VmpPanaq 47rn(kT) 2 AVDDf- 3dB This results shows that power consumption in analog circuits is directly proportional 50 to dynamic range. Extrapolating to ADCs, where dynamic range is exponentially related to resolution (namely 2 N, where N is the number of bits), power consumption can be expected to vary exponentially with the number of bits to be resolved. This result, however, is not true in the case of digital circuits. The power consumption of digital circuits is simply linearly related to the number of bits. This can be seen, for example, in an adder circuit, where a 16-bit version requires twice the number of full-adder cells as an 8-bit version (assuming the critical path is not affected). Since ADCs are, generally speaking, mixed signal systems, the resulting powerresolution relationship expected in a scalable ADC, lies in between that for digital and analog circuits. A bias towards one or the other will depend on the exact distribution of power consumption between digital and analog blocks. The situation is complicated, however, by the ability to implement the feature of scalability. Although, in analog circuits, we expect exponential power savings with dynamically reduced resolution, in reality, this is limited by circuit practicalities, as described later in this chapter. In order to reconcile the difficulties associated with implementing power-resolution scaling, in this design, two discrete resolution modes have been chosen: a 12-bit mode, and an 8-bit mode. A feature of the SAR conversion algorithm, which resolves bits successively, starting with the MSB, is that the conversion process can be stopped at any point during bit-cycling, once the digital code has been determined to the desired resolution. Of course, this, alone, yields only linear power savings. In order to scale resolution more efficiently, the analog blocks, namely the comparator and DAC, should be retuned appropriately. Mechanisms to perform this adaptation in the comparator are described in Section 4.2.1. The difficulties associated with implementing scalability in the DAC are presented in Section 4.2.2. 4.1.4 Self-Timed Bit Cycling The most straight forward signaling to control bit-cycling in a SAR ADC is shown in Figure 4-4. Here, the SAR finite state machine generates the appropriate digital 51 control for the DAC at the start of the clock cycle (tek_ after the rising edge). During the first half of the clock period, the DAC and comparator preamplifiers (if present) settle to their correct values. Then, the comparator latch is triggered at the falling /edge of the clock. Subsequently, the comparator decision is used by the SAR logic to derive the next digital control for the DAC. In this scheme the preamplifier settling time, (plus the SAR clock-Q time, tsettle,PAMP, tck-q) must be less than half the clock cycle. CLOCK tck-q SAR DAC PRE-AMP \ seftle,PAMP LATCH T +-tiatch Figure 4-4: Waveforms showing standard bit-cycling signals. Alternatively, a self-timed scheme was suggested in [25], where the DAC and comparator preamplifiers borrow time from the latch, to ease their settling requirements. The signals associated with this scheme are shown in Figure 4-5. Here, the latch is once again triggered by the falling clock edge. However, now, following the latch resolution time, tlatch, the SAR immediately derives the next digital control word for the DAC. Subsequently, the DAC and preamplifiers start settling to their respective values while the clock is still low. As a result, the preamplifier settling time, can be longer by t + tck-q - tiatch - togic, where tdlk tsettle,PAMP, is the clock period. Specifically, if the preamplifiers dominate the bit-cycling time (i.e. the logic and latch delays are small), their settling time can be nearly doubled, reducing their power consumption by a factor of two. 52 CLOCK -tiogic SAR tsettle,DAC DAC PR E-AMP tsettePAMP LATCH -+ +tlatch Figure 4-5: Waveforms showing self-timed bit-cycling signals. Since the successive approximation algorithm works by dividing the conversion range in half during each phase of bit-cycling, the possible input swing to the comparator decreases by a factor of two each time. During the first few MSBs, however, large voltage swings are possible. The worst case sequence for the comparator to resolve involves a large input during one cycle, followed by a very small signal, of the opposite sign, during the next cycle. This condition is known as the overdrive case and is treated quantitatively in Section 5.2.1. The latch resolution time will, generally speaking, be dependant on the magnitude of the input signal. During overdrive conditions, tiatch will be very short during the first phase due to the large input. Subsequently, the preamplifiers will benefit from a lot of extra settling time, which they will need, as the weak input they are subsequently driven with limits the recovery time. In this manner, self-timed bit-cycling is particularly suitable for the SAR algorithm. The resolution process associated with each bit, except the MSB, has the opportunity to borrow time from the previous latching phase. Although, the MSB decision is not subject to overdrive, a condition nearly as limiting can be imposed by the preceding sampling phase. The corresponding details are described in Section 5.1.1. To ensure that the ADC timing is not limited by this effect, an extra half cycle has 53 been devoted to the MSB decision. 4.2 Block Architecture This section describes block level optimizations that have been implemented in this ADC in order to improve efficiency and facilitate scalability. Architectural concepts are presented here, and implementation details are provided in Chapter 5. 4.2.1 Comparator Architecture The comparator is responsible for resolving the sign of its input and generating a full-swing digital output based on its decision. An ideal comparator will be able to generate a decision despite an arbitrarily small input. In this sense, comparators must have nearly infinite gain. Fortunately, their non-linear and open-loop nature implies that cascading techniques and positive feedback can be used to increase efficiency without suffering instability. The next few subsections present optimizations and design approaches employed in the comparator. Specifically, trade-offs associated with linear and regenerative amplifier design, amplifier cascade optimization, auto-zeroing coordination, and performance scaling will be considered. Gain Optimization As mentioned, the open-loop nature of comparators presents the opportunity for employing cascaded amplifiers and positive feedback in order to achieve very high gain. The advantage of cascaded amplifiers is based, fundamentally, on the principle that the total gain is the product of the individual stages, while the time-constant is the sum of the individual stages. As a result, the constant gain-bandwidth constraint associated with a single stage amplifier is overcome. Positive feedback provides an even more efficient means of achieving high gain, since the signal continuously regenerates itself. This leads to a time-varying behavior, where the gain increases exponentially with respect to time. 54 To compare the efficiency of various amplification strategies, transconductor based gain blocks may be considered. For a given tranconductance and load capacitance, transconductance amplifiers, which have infinite output impedance, achieve a required output swing faster than voltage amplifiers, which have finite output impedance. This is shown in Figure 4-6, where the output voltage, normalized to the product of the ), is plotted with respect to time, normalized to the input voltage and Gm (i.e. GVou load capacitance (i.e. Here, the transconductor acts as an integrator, exhibiting i-). a voltage ramp at the output. The voltage amplifiers, having R"'t equal to 30 and 15 respectively, are subject to RC settling at their outputs. Assuming an output swing of 10 is required, the transconductor provides the least delay of the three options. Since, in the most efficient operating regime (i.e. weak inversion), amplifier GM is proportional to current, this plot shows that high-gain stages achieve a better powerdelay product than low-gain stages. Despite improved efficiency, high-gain stages suffer from several practical effects. These are considered later in this section. An analytical comparison of multi-stage and regenerative amplification was pursued in [26] where cascaded transconductors were assumed. Figure 4-7 illustrates the structures compared. In the case of the cascaded linear stages of Figure 4-7a, it was shown that the delay, tdly, to achieve a total gain of Atot is given by Equation 4.8, where GM is the transconductance of each stage, CL is the load capacitance of each stage, and N is the total number of stages. 0 tdly,lin = m(At tNL!) GM (4.8) Then, multiplying by NGM gives an expression for the total power-delay product of the cascade. The delay associated with a regenerative amplifier was shown to be given by Equation 4.9. tdly,reg = CL Gm lrt(Atot) (4.9) Similarly, multiplying by GM (since only one transconductor is required) expresses the power-delay product of the regenerative stage. The results corresponding to the power-delay for a regenerative amplifier, as well as a 5-stage and 3-stage cascade, are 55 Output Voltage VS Time 20 out - 18 .R 16 4 =15 / 14 0~10 0 6 4 2 0 I 5 10 15 20 25 30 35 t/CL Figure 4-6: Delay for amplifiers with various output resistances. plotted in Figure 4-8. Here, it is shown that regenerative stages possess a superior power-delay figure of merit compared with linear stages. The difficulty with regenerative amplifiers is that they typically suffer from large input-referred offset. Typically offsets on the order of 50-100mV can be expected [27]. Accordingly, a common approach, is to use, less-efficient, offset-cancelled linear stages to achieve a signal swing larger than the offset floor of the regenerative latch, and then use the latch to generate full-rail, digital outputs. In this ADC, however, analog calibration will be applied to the regenerative latch to minimize its offset, thereby easing the gain requirements of the linear stages. Latch offset calibration has been explored previously [26][28], however, structures suitable for multi-step converters, such as SAR ADCs, have not been demonstrated. The primary challenge associated with multi-step converters is that the opportunity to calibrate does not occur prior to every decision. Calibration involves storing offset 56 cL CLCL! LL Gm fGm L- ICL CL Gm L (a) Gm (b) Figure 4-7: Comparator gain structures (a)Linear amplifier cascade (b)Regenerative amplifier. voltages on capacitors. However, latches typically need to be reset following every decision. During this process, the stored offsets are lost and must be regenerated by means of calibration. Since calibration requires application of a reference signal at the input, it is not a viable option during bit-cycling. Specifically, in a low- power, low-voltage design, the reference signal is difficult to generate, and difficult to apply, as analog switches, having a degraded gate-overdrive, should be avoided in the signal path. The circuit implementing latch calibration suitable for this SAR ADC is presented in Section 5.2.2. 57 Power-Delay VS Total Gain 120 Rgnrtv 5-stage -''' 3-stage - - 100 - 80 - -. 0 60 E 0Y) -lop -5 40 - - .. . 20 -- -'-- - - - - - -*** f'L 100 10 102 103 104 Figure 4-8: Power-delay (normalized to CL) versus total gain for linear and regenerative amplifiers. Preamplifier Chain Optimization Although regenerative amplifiers are most efficient from a power-delay point of view, achieving complete offset cancellation is difficult. Several techniques exist, however, to fully correct offsets in linear stages. Consequently, the common approach, of using linear stages preceding the regenerative stage, has been employed in this design. As suggested above, linear stages may be cascaded to improve their efficiency. Although it was shown that a desired output swing could be achieved fastest with high-gain or transconductance amplifiers, these require input offset cancellation techniques for auto-zeroing. As mentioned later, implementing this type offset cancellation effectively is difficult. Low-gain amplifiers do not suffer from the same effect, and consequently are used in this design. Specifically, a stage-gain of 3-4 is most practical to implement. Roughly speaking, amplifying 12-bit LSB voltages (58 100p V) to levels reasonable for calibrated latch offsets (~ 5mV), requires a cascaded gain between 30 and 60. This implies that three stages of preamplifiers should be used. Auto-Zeroing Optimization To fully correct for static (and low-frequency) errors in linear preamplifiers, techniques such as auto-zeroing and chopper-stabilization [29] may be employed. Chopper stabilization relies on modulating the input signal to a higher frequency, amplifying it in a low-noise frequency band, and then modulating it back to baseband. In low power implementations [30], modulation, at the input and output, is performed using passive, commutating switches. In the case of a SAR ADC, where the comparator is driven by a charge redistribution DAC, injection errors imposed by such switches on the critical nodes is prohibitive. Alternatively, auto-zeroing relies on sampling the amplifier offset, and noise, and subtracting this from subsequent outputs. Three common approaches to auto-zeroing are input offset storage (IOS), output offset storage (OOS), and auxiliary offset compensation [29]. Both IOS and auxiliary offset compensation involve sensing and storing the input referred offset, by means of feedback. For these techniques to be effective, high gain amplifiers are required, as the offset is attenuated by the factor of 1+Ao, where A 0 is the gain of the stage. Typically, the required gain for sufficient offset compensation is only achieved by cascoding or gain boosting techniques. Neither of these is amenable to low-voltage, low-power designs. Further, in IOS, charge injection errors, from auto-zeroing switches, appear at the amplifier input, and, as a result, get amplified by A 0 when reflected to the output. Output offset storage, however, does not require high gain amplifiers to achieve nearly perfect offset compensation and is not as sensitive to charge injection errors. Consequently, multi-stage, output offset compensation is employed in this ADC. The resulting structure is shown in Figure 4-9. Here, the charge on the auto-zeroing coupling capacitors is initially purged. This occurs at the start of the conversion as mentioned in Section 4.1.1. Then, the DAC generates a reference signal of zero differential amplitude (this is discussed in Section 4.2.2), and, simultaneously, the 59 three auto-zeroing switches are closed. The output-referred offset of each stage is stored on the auto-zeroing capacitors for 11 clock cycles. Then, the first auto-zeroing switch is opened, and the second stage is given another half clock cycle to cancel the ensuing charge injection error. This procedure is repeated in the next stage, such that the entire cascade is subject only to the charge injection error of switch AZ3, which is small when reflected back to the input [22]. PRG > PRG F I AZ3 AZ2 AZ PRG PRG Clock PRG L. PRG .J PRG AZ1 AZ2 AZ3 Figure 4-9: Auto-zeroed multi-stage preamplifiers. Typically, auto-zeroing is performed during sampling. However, in this design, the two operations are separated (as explained in Section 4.1.1 and rationalized in Section 4.2.2). Consequently, the auto-zeroing time is not constrained by the settling requirements of the sample-and-hold and can be optimized. The optimal time, determined to be 1- clock cycles in this design, can be estimated by considering the relative thermal noise contributions during auto-zeroing and bit-cycling. Figure 4-10 shows the networks relevant to noise analysis during the two phases, where it has been assumed that the load imposed by the subsequent amplifier stage is small. Here, the total capacitance seen during each phase should be optimized for minimum overall power-delay in the amplifier. Specifically, since the amplifier must settle only 60 once during auto-zeroing, but, once for every bit during bit-cycling, it is beneficial to decrease the noise contribution during auto-zeroing at the cost of increased settling time. CL CLI AZ AZ (b) (a) Figure 4-10: Amplifier network (a)During auto-zeroing (b)During bit-cycling. During bit-cycling, comparator decisions are subject to the noise sampled during auto-zeroing (on the CAZ capacitors), as well as the time-varying noise of the amplifier, which is always present. If the amplifier implementation is set (i.e. its gm, Rout, and equivalent number of noise sources does not change between auto-zeroing and bitcycling), the total noise variance, Vo,toit, is given by Equation 4.10, where CL is the load capacitance during bit-cycling, CAZ,tot is the total load capacitance during autozeroing (CAZ + CL), and a is an implementation-dependant constant. a 2 ono,tot = CL _+ a CAZ,tot (4.10) Normalizing this expression, such that the desired noise variance is equal to a, allows CL to be expressed in terms of CAz,tot. CL = CAZ,tot CAz,tot - (4.11) 1 The overall power-delay of the amplifier can be expressed as the sum of the GM-r products over the entire conversion. For a 12-bit ADC, this is given by Equation 4.12, where 12 clock cycles are required for bit-cycling, and the number of clock 61 cycles for auto-zeroing, K, must be determined for minimum power-delay. Note, it is assumed that neither GM, the amplifier transconductance, nor Rut, the output impedance, change during the conversion. K+12 GMTi = KGMRoutCAZ,tot + 12GMRoutCL (4.12) In the case that equally complete settling is desired during auto-zeroing and bitcycling, the ratio of CL to CAZ,tot is given by Equation 4.13, where the settling time during bit-cycling is set to 1 clock cycle. 1 K - Tbit-cycle RoutCL CL Tazero RoutCAz,tot CAz,tot (4.13) Now, using Equation 4.11 and Equation 4.13, CL and CAZ,tot can be expressed as follows. K±+1 CL = K K (4.14) CAz,tot = K + 1 (4.15) Finally, normalizing the expression for power-delay (Equation 4.12) by GMRout gives Equation 4.16. E K+12 T - K(K + 1) + R,,t 12 K+ I K (4.16) This function is plotted in Figure 4-11, where it is shown that minimum power-delay occurs for K ~~1.6. Based on this result 11 clock cycles have been used for autozeroing the amplifiers, and the load capacitances have been set appropriately. Recall, however, to improve the cancellation of charge injection errors, an additional clock cycle is used so that the auto-zeroing of each stage is extended with respect to the preceding stage. Resolution and Sample Rate Scaling As described in Section 4.1.2, linear power savings, with respect to reduced sampling rate, is straight forward to achieve. In this design, sample rate scaling has been 62 Power-Delay VS Auto-Zero Clock Cycles 28 27.527p-426.5 < 26 p 25.5_ 2524.52423.5 1 1.5 2 2.5 Auto-Zero Clock Cycles 3 Figure 4-11: Normalized total power-delay of auto-zeroed amplifier. implemented by power gating the preamplifiers and latch. This is controlled by the SLEEP signal, shown previously in Figure 4-3. Scaling the ADC's resolution means less overall gain is required, as the LSB voltage in 8 bit mode is 16 times larger than that in 12 bit mode. While the regenerative amplifier does not provide linear gain, and is therefore unaffected by this result, the preamplifiers can benefit from the reduced requirement. However, if the same threestage cascade is used, the power will only be reduced by a factor of 163. As mentioned in Section 5.2.1, the original preamplifier cascade is limited by the thermal noise performance requirement of the 12 bit converter. However, the reduced resolution greatly relaxes the capacitive load requirement of the first stage. Specifically, if the gain of each preamplifier in the three stage cascade is 4, the first two stages, providing a combined gain of 16, are not required. The power-delay product of the last stage, which is not thermal noise limited, coincides precisely with that achievable for an 8 bit converter. The Gm sets the power, and the Rt capacitance, which depends on parasitics, is fixed). 63 sets the delay (since the load If the power consumption and speed of the ADC is dominated by the preamplifier performance, adjusting these parameters proportionally would only change the maximum speed of the converter, and not its energy per conversion. This can be seen in Figure 4-12, where the effect of doubling GM, and accordingly halving R0 ,, is shown. Here, the power consumption doubles during the active conversion. However, the settling time, and therefore the clock period, is reduced by half. The net result is that the conversion completes in half the time, and the preamplifier shuts-off for the remainder of the sampling period. As a result, the average power consumed is unchanged. Clock tsette=NT=NRoUTCL PAMP 2 Clock L_ I Cq4 o t'see=NT'=NR'OUTCL =NROUTCL/2 PAMP Figure 4-12: Effect of changing Gm at constant amplifier power-delay. Of course, in reality, the relative power consumption depends on the portion of the total clock period consumed by preamplifier settling, and the portion of total power consumed in the preamplifiers. Nonetheless, to ease the implementation of the 8 bit mode, a single stage, identical to the last stage of the 12 bit cascade, can be used, and the clock rate can be set appropriately. As mentioned before, analog switches should not be used in the signal path as they are limited by low gate-overdrive in this low-voltage design. So, to avoid multiplexing the input signal of the third stage, an entirely separate preamplifier path is used. The final comparator structure is shown in Figure 4-13 below. The power gating switches are used to enforce sample-rate scaling, as well as resolution scaling, by shutting-off the inactive preamplifier path. 64 NSLP12 NSLP12 NSLP12 NSLP12 PRRP PRG PRG Ao=GmIRoUTI AZ3 AZ2 AZ1 Ao=GM2RouT2 PRG Ao=Gm3RouT3 NSLP8B NSLP8B PR 7AZ3 PR Ao=Gm3RouT3 Figure 4-13: Architecture of final comparator. 4.2.2 DAC Architecture The DAC is responsible for generating analog outputs at precisely the code transition voltages of the ADC. Figure 4-14 shows its complete architecture. Positive and negative outputs are generated using two sets of capacitor arrays that are coupled, through the top-plate sampling switch, only during sampling and auto-zeroing. During bit-cycling, the charge redistribution process is controlled by switches, in the positive and negative switch matrices, which drive the bottom-plates of the capacitors to the appropriate voltages. As shown, this ADC employs a fully differential DAC. Although the increased robustness is not crucial in 8 bit mode, and, as result, the extra overhead may be avoided, such configurability is not easily achieved in practice. In particular, switching from a differential DAC to a single-ended one would require shorting the top-plate of the positive array to Vef,p. However, now, the DAC outputs will swing around Vre,,p, which may be as high as VDD. Consequently, charge loss 65 from the critical node can occur through the top-plate sampling switch. Although this can be avoided, through proper bootstrapping, it results in a less desirable structure, which could degrade the ADC precision in 12 bit mode. Vref.P Vref.N VIN+ AZERO AZEROSwitch Matrix SAMPLE (Positive) BCYCL(11:0)--/-+ 12 -6 E2 x6 0_T I VOUT- Top-plate Sampling switch VOUT+ It CC. x6 Ma nDAC - 1 - - Su DAC 6 12 BCYCL(11:0)--\-c SAMPLE AZERO D1 TI' SAMPLE 6 A ZERO C, 7is x6 000 T T AC - ManDAC Swtc , Matrix itive) VINV Vref, Vref,N Figure 4-14: Architecture of final DAC. In a charge redistribution DAC, the ability to generate transition voltages accurately depends on the matching between constituent capacitive elements. In addition 66 to proper matching, a number of other considerations are critical during DAC design. For instance, in Section 1.1.4, the ability to convert both single-ended and differential signals was an identified requirement. Since, internally, this ADC uses a differential architecture, proper support of single-ended inputs can be interpreted as a commonmode signal rejection requirement. The associated high-level implementation details are presented in following subsections. Sub-DAC One of the limitations associated with medium and high resolution capacitive DACs is the binary relationship between elements. Specifically, in the case of a 12-bit converter, the largest capacitor is 2048 times bigger than the smallest capacitor. This binary relationship can be broken by using a sub-DAC [31][32]. Two structures, implementing M + N bit DACs, are shown in Figure 4-15. The basic principle here involves using the main-DAC to precisely generate transition voltages corresponding to the N most significant bits of the digital input code, and then interpolating between these using the M bit sub-DAC. The structure in Figure 4-15a uses capacitive charge redistribution to implement the interpolation, while that in Figure 4-15b uses a resistive string. Although the resistive implementation is inherently monotonic, it draws static current, and therefore has not been used in this low-power design. A major drawback of the structure shown in Figure 4-15a is that it requires an active, unity-gain buffer to drive the bottom plate of a main-DAC unit-capacitor. An alternative, fully passive implementation is shown in Figure 4-16 [33]. To analyze this structure, notice that the total capacitance of the main-DAC is ( 2 N - 1)CO. Also, the capacitance looking into the series combination of the coupling capacitor, Cc, and the sub-DAC, is equal to Co. However, unlike before, the bottom plate of the subDAC is involved in sampling the input. As a result, the main-DAC generates the MSB transition voltages by treating the series combination of the coupling capacitor and the sub-DAC as a single unit capacitor. To see how sub-DAC interpolation is performed, consider the Thevenin equivalent circuit shown in Figure 4-17. Here, voltages derived by the unloaded sub-DAC effect the output through the same capacitive divider 67 2N-ICO C "'0 C 2 4 2N "0 0 VIN V~f N Vref N - -Co CO Vref p VIN 2 C0 Co CO 0 Co- Vref p Vf Vf N N (a) (b) Figure 4-15: Typical main-DAC/sub-DAC implementations. (with ratio 1) as the previous implementations, thus 2 M+N transition voltages are generated. Vref,N--> C -- M Vref,P VIN VrefN 2lm1Co .- . 2Co Co Co V/N Vref,N Figure 4-16: Fully passive main-DAC and sub-DAC. Co essentially sets the transmission gain from the sub-DAC to the main-DAC such that proper interpolation is performed. Unlike the main-DAC, however, parasitic 68 Vref p N-1 i (2N C - Thevenin M-1 I , Z b2'C 1WI ~A 1 VrefN Cc qi-i'Co Equivalent C bi2=Co )CO N-I C0 Vref,P . 2v c (2"-)CO Co - Zqi2 I Vref,N VrefN 2mCo + VrefN Figure 4-17: Thevenin equivalent circuit for analyzing passive sub-DAC. capacitance on the top-plate of the sub-DAC causes linearity errors. Details of how these errors originate, and the technique used to manage them, is described in Section 5.1. Common-Mode Rejection Device mismatch in comparator preamplifiers leads to unbalanced currents in the two branches of the differential pair. Consequently, the gm's of the input devices are not equal. Then, due to the finite output resistance of the tail current source, commonmode input signals can lead to differential output voltages. Low-voltage designs, such as this one, are particularly susceptible to this effect since the reduced headroom leaves little margin for ensuring the tail device remains in deep saturation. Figure 4-18 demonstrates this in the case of 3o- mismatch applied to the preamplifiers discussed in Section 5.2.1. Here, the voltage at which the amplifier is auto-zeroed is varied, while the common-mode voltage, during active operation (i.e. bit-cycling), is fixed at 500mV. Although, perfect offset compensation is assumed, as shown, differentialmode offsets in the millivolts are observed even if the auto-zeroing voltage deviates by only 1OOmV. To avoid this, it is critical that the the preamplifiers be auto-zeroed with the appropriate reference signal. Specifically, the reference signal should be equal to the common-mode voltage seen during bit-cycling. Additionally, the DAC's output common-mode voltage should be well controlled to remain within the range amenable to the preamplifiers. 69 Input Referred Offset VS Auto-Zeroing Common-Mode 5 E co . 4 - ....0.. 1. ... . . . 1............... 0.3 0.35 - . . -.. ... 0.4 0.45 0.5 0.55 0.6 0.65 Auto-zeroing Common-Mode Voltage (V) 0.7 Figure 4-18: Differential-mode offset due to varying auto-zeroing voltage. As mentioned, proper conversion of single-ended inputs requires the ADC to have good common-mode rejection. By ensuring that the differential capacitor arrays sample equal and opposite charge, the comparator will not be subject to the varying common-mode of the input [21]. Additionally, the DAC outputs will be centered around vref,P+vef,N , which, by design, is desirable for the comparator preamplifiers. This can be seen in Figure 4-19, where the charge sampled on the top-plate is independent of the common-mode voltage, VCM. The only difference between this sampling network and that in Figure 3-4, is that no bias is actively applied to the top-plates of the arrays. If no charge is initially present on the capacitors, and the arrays are well matched, the top-plates will passively settle to the input common-mode voltage. One advantage of the purge phase, at the start of the conversion (Section 4.1.1), is that it eliminates the need for active VCM generation. Typically, auto-zeroing is performed during the sampling phase, using the topplate voltage as the reference. However, as shown above, during sampling, the topplate tracks the varying input common-mode voltage. 70 Further, the outputs of the + 1 N- C C . NCoV 1N + 2 + VCM VIN Q = 2NCoVN12 C Figure 4-19: Common-mode independent charge sampling. DAC during bit-cycling will not be centered around this voltage. Consequently, the top-plate voltage during sampling is not an appropriate auto-zeroing reference. This problem was solved in [25] by multiplexing a separately generated reference to the preamplifier inputs. In addition to imposing charge injection errors, this approach is not desirable in this low-voltage design, because it requires analog switches in the signal path. As a result, a separate phase has been devoted to auto-zeroing so that the capacitor arrays can generate an appropriate reference passively. In order to do this, the capacitor arrays are initially purged to remove any residual charge. Then, they are switched in a manner similar to MSB bit-cycling, which generates a transition voltage at mid-scale. To guarantee that the preamplifier inputs are zeroed, the top-plate sampling switch remains closed during this process. Note, this switch may experience low-overdrive, so it is bootstrapped, as described in Section 5.1. The resulting structure is shown in Figure 4-20. Vref P Vref P 2-1C' 2N-1Co (LSB Cap's) I (LSB Cap'si 2Co Vref N Vref N Figure 4-20: Passive auto-zero reference voltage generation using capacitor arrays. 71 Resolution and Sample Rate Scaling Since the DAC is fully passive, no special support is required to implement scalability in sampling rate. The array will only consume power when switched by a new digital input code. Ideally, scaling the ADC's resolution from 12 bits to 8 bits would allow removal of the 2 MSB capacitors in the main-DAC and the sub-DAC such that they don't participate in sampling or bit cycling. In practice, this is quite difficult to do. One option would be to disconnect the unused capacitors from the top-plate so that they are completely decoupled from the active array. This requires analog switches between critical nodes, and is, therefore, not straight forward. The use of bootstrapped switches, to resolve this problem, is considered in Section 7.2. Another option would be to disconnect the bottom plates of the MSB capacitors so that they do not impose any charge loading on the active arrays. Despite disconnecting their bottom-plates, however, the MSB capacitors continue to load the remaining arrays, as bottom-plate parasitic capacitances cause coupling to ground. Bottom plate parasitics are pro- portional the size of the active capacitance, and consequently, in the case of MSB capacitors, are quite large. Due to these practical limitations, no special technique has been used to facilitate power scaling in the DAC with respect to reduced resolution. In 8-bit mode, the last 4 capacitors in the sub-DAC are simply not bit-cycled. 4.2.3 SAR Architecture The SAR state machine is the digital block responsible for controlling the conversion process. Due to the self timed nature of this design, as well as its use of clock gating during sleep modes, asynchronous techniques are used extensively. During bit-cycling, successive resolution-of bits is triggered by completed comparator decisions. There is the possibility, however, that the latch does not resolve, remaining in a metastable state indefinitely. This could stall the entire bit-cycling process if it is not properly detected. Section 5.3 describes how latch metastability is sensed, and the manner in 72 which the SAR state machine recovers from this condition. 4.3 Summary A number of optimizations are applied to the global architecture of the ADC, as well as to specific blocks, in order to enable a low-power, scaleable design. Samplerate scaling is achieved through clock-gating, in digital circuits, and power-gating in analog circuits. Resolution scaling is achieved by activating a comparator path that provides the appropriate gain, bandwidth, and noise characteristics for the required precision. The efficiency of the bit-cycling process is improved by using a self-timed scheme that relaxes the settling time requirements of the DAC and preamplifiers. At the block level, a low-offset regenerative latch improves comparator performance by reducing the gain requirements of the less efficient linear amplifiers. Further, the common-mode rejection of the ADC is enhanced through proper sampling, as well as de-coupling the sampling and auto-zeroing phases. This allows an appropriate autozeroing reference to be generated passively using the charge redistribution DAC while avoiding analog switches in the signal path. Finally, since auto-zeroing is de-coupled from sampling, it has been optimized with consideration to the preamplifier noise contributions during the conversion phases. 73 Chapter 5 Circuit Design This chapter describes the circuit-level implementation of the ADC blocks. To achieve low-power operation the entire ADC has been designed to operate from a IV VDD. Both the digital state machine and charge redistribution DAC benefit from the reduced supply due to their quadratic power dependence on VDD. Analog blocks, however, only benefit in the case of non-noise limited stages. To maximize the noiselimited SNR, Vref,p has been designed to be as high as VDD, though it can be reduced slightly to meet the requirements of different applications. Similarly, VrefN is intended to be as low as OV (ground). These values allow a differential input signal swing of iV, and a single-ended input signal swing of IV. Further, in the differential case, it is assumed that the nominal common-mode voltage of the input is 500mV. Subsequent sections use these parameters to rationalize, and develop, the implemented circuit strategies. 5.1 DAC Circuit Design Architectural issues pertaining to the DAC were discussed in Chapter 4. Here, implementation details will be presented through consideration of the two main substructures: the first is the switch network, consisting of the switch matrices, as well as the sampling and purging switches; the second is the capacitor arrays of the differential main-DACs and sub-DACs. 75 5.1.1 Switch Network The DAC switch matrices consist of sets of switches, connected to the bottom plates of each capacitor. These switches control the charge redistribution and sampling processes. Figure 5-1 shows their actual implementation. The PRG, AZERO, and SMPL (SMPL) signals enforce the corresponding phases of the conversion cycle. The Cx (Cx) signals control successive bit-cycling for the 12 bits (or 8 bits), and the BCYCLEN (BCYCLEN) signals are qualifiers to globally enable bit-cycling. Their full significance is described later in this section. ~AIN+ V22 V () BTMPLT (1-0)+ BTMPLT_11+ Vre, N Vref,N (a) (b) BTMPLT_11- BTMPL T(1-0)- AIN- ~e 0 (1 T 0 Vr 60 C/ VrefP~ Vref N 0 Vref N 60M 'r- Vr,P AIN_ r Vref,P Vref N (d) Figure 5-1: Implementation of switch sets in switch matrices (a)MSB switch set for positive array (b)Bits 10-0 switch set for positive array (c)MSB switch set for negative array (d)Bits 10-0 switch set for negative array. The switch sets associated with the positive and negative arrays have inverted 76 functionality during the bit-cycling phase (i.e. the bottom-plate Vref,N and Vref,P connections are reversed), as shown. Further, during auto-zeroing, the capacitor arrays are switched in a manner similar to MSB bit-cycling, where the largest capacitor in the positive (negative) array is connected to capacitors are connected to Vref,N (Vref,P). Vref,P (Vref,N), and the remaining As a result, the switch sets associated with the MSB capacitors are slightly different than those for the other capacitors. The purge phase is shown in Figure 5-2a. Here, the bottom-plates of all capacitors are driven to Vref,N (which is nominally ground), and the purging switches are closed across the top and bottom plates. Following this, during the auto-zeroing and sampling phases, the top-plate sampling switch shorts the top-plates of the positive and negative arrays, equalizing any residual charge injection mismatch from the purging switches. The sampling phase is shown in Figure 5-2b. Physically, the top-plate switch is implemented as a single NMOS device. Both during auto-zeroing and sampling, it can be subject to low gate-overdrive, as its source and drain are at a voltage near mid-rail. As a result, its gate is bootstrapped using the charge-pump described in Section 5.5. The input switch, which is implemented as the CMOS transmission gate shown in Figure 5-1, can also be subject to low overdrive for input voltages near mid-rail. Consequently, the NMOS is bootstrapped in a similar manner. Input switch VIN. Vrf N VrefN VrefN (IV to 0%) Purging switch Top-plate switch 0.5V Top-plate switch Purging switch VrefN Vref,N VreftN Input Switch VIN- (OV to IV) (b) (a) Figure 5-2: DAC network (a)During purging (b)During sampling. Proper control of the top-plate sampling switch is critical to the precision of the sample-and-hold operation. If correct timing is not ensured, excessive injection er- 77 rors, and even loss of sampled charge can occur. Specifically, following the sampling phase, the top-plate sampling switch must turn off before the input switches turn off, and before the VrefP and Vref,N switches turn on to bit-cycle the MSB. Since these operations all occur, nominally, during the same phase, special circuitry is required to enforce their sequence. When it turns off, the top-plate sampling switch imposes charge injection. This directly affects the sampled value and introduces error if there is any mismatch between the charge injected on the positive and negative arrays. The precise proportion with which channel charge leaves the source and drain depends on the interaction between the device's capacitances with the load impedances on either side. To minimize residual mismatch, the impedance on either side of the sampling switch should be well matched. If the bottom-plate input switches turn off before the top-plate sampling switch, the capacitor arrays appear as a large impedance in parallel with main-DAC's top-plate parasitic capacitance. As a result, injection charge from the top-plate sampling switch will depend on the, relatively small, and poorly controlled parasitic capacitance, as shown in Figure 5-3. By ensuring that the bottom-plate input switches turn off after the top-plate sampling switch, the injected charge can be directed to the large, well-matched capacitor arrays. Consequently, the SMPL and SMPL signals driving the input switches are delayed using a series of weak inverters. The matching of the loads is further improved through input switch sizing as discussed later. No current path No current path TCP CPT Figure 5-3: Distribution of injection error determined by poorly matched top-plate parasitic capacitances. At the start of bit cycling, the charge redistribution process begins when the 78 bottom plates of the capacitor arrays are driven to Vrefp or Vref,N. At this time, differential outputs are generated by the DAC. Consequently, it is critical that the top-plate sampling switch be closed, or charge will be exchanged between the positive and negative arrays, corrupting the input sample. Since the input switches and the sampling switch are overdriven by relatively slow charge-pumps (described in Section 5.5), their turn-off time is difficult to predict. To ensure proper timing, the BCYCLEN and BCYCLEN signals are qualified, as shown in Figure 5-4, by buffered versions of the bootstrapped sample signals. This guarantees that bit-cycling will not begin until the sampling switch is completely off. BCYCL BCYCLEN SMPL_DLY---- SMPLDYBCYCLEN Figure 5-4: Qualification of BCYCLEN and BCYCLEN signals with delayed sampling signals. Input Switches As mentioned, the NMOS of each input switch in the switch matrices is overdriven to ensure the resistance of the CMOS transmission gate is not excessive. Specifically, sufficient settling time must be ensured, during the sampling phase, for analog inputs near mid-rail. The time-constant associated with the sample-and-hold can be estimated by considering the half circuit consisting of one CMOS input sampling switch, one of the differential capacitor arrays, and half of the top-plate sampling switch. For the used geometries, the worst case resistance of each input switch is 6.5kQ, while, at the expected common-voltage (500mV), the resistance of the top-plate sampling switch is 4.6kQ. Although, each capacitor in the array has an associated timeconstant, a conservative analysis assumes that the entire array (which imposes a load of 7pF) is charged by a single input switch. This results in a time-constant, Tsmgp, of approximately 62ns. For 12 bit accuracy, a settling of at least 9 time-constants is re- 79 quired. Consequently, 1-1 clock cycles of a 2MHz clock are sufficient for the sampling period. The entire array, with distributed time-constants, was simulated to confirm this result. In addition to settling time considerations, symmetry in the input switches is critical to minimizing charge injection errors. As mentioned above, keeping these switches "on" helps match the load impedance on either side of the top-plate sampling switch. This matching is further preserved by sizing the PMOS, and overdriven NMOS, such that the non-linear resistance of the input transmission gate is symmetric about the input common-mode voltage. The equivalent resistance is shown in Figure 5-5, for the used geometries and gate voltages; the typical process corner has been used for this simulation. Here, the difference in resistance is less than 300Q (approximately 10%) for differential input voltages centered around 500mV. Input Switch Resistance VS Input Voltage 4000 3500- 3000 .0 (n 2500 - - 2000- 15001 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Input Voltage (V) Figure 5-5: Equivalent Resistance of input switch with respect to input voltage. Finally, the input switches introduce non-linearity to the sample-and-hold process due to their dependence on input voltage. The effect they have on degrading SNDR depends on their precise voltage dependence. This is considered further in Chapter 80 6. Reference Switches The source-drain regions of the top-plate sampling switch, which is implemented using an NMOS device, introduce parasitic PN-junctions on the critical top-plate nodes, as shown in Figure 5-6. VIN- N+ VrN Vre P Vep VreN 0- Co Co Co Co Figure 5-6: Parasitic PN-junction on top-plate. During sampling, a voltage is generated across the array capacitors. Subsequently, after the sampling switch is opened, the impedance on the top-plate node is high, and the bottom plate is driven to Vref,N or Vref,p. As a result, top-plate voltages beyond the rails are possible. Specifically, when the MSB is bit-cycled, following the sampling phase, the voltage on one of the top-plate nodes can undershoot below ground, forward-biasing a PN-junction. As a result, top-plate charge can be lost, corrupting the sampled input. To avoid this, the PMOS devices, which connect the capacitor bottom-plates to Vref,p, have been sized much stronger than the NMOS devices, which connect the bottom-plates to Vref,N. As a result, the top-plate voltage never undershoots below ground, and no charge is lost. A drawback to this approach is that the DAC outputs can spike in the wrong direction during MSB bit-cycling before settling to the correct value. As shown in Figure 5-6, all capacitors in the negative array (right hand side), except the largest one, connect to Vrej,p, whereas in the positive array (left hand side), only the largest 81 capacitor connects to Vef,p. Due to the strong PMOS devices used to make the Vref,p connections, the top-plate of the negative array initially approaches Vej,p faster than the top-plate of the positive array, even if a negative input voltage is sampled. This transient impulse is shown in Figure 5-7 below. Here, during the sampling phase a voltage near 1LSB is applied to the array. Then, at the start of MSB bit-cycling (BCYCL1 1), the DAC's negative output (VoUT_) is observed at the differential outputs ( 3 ,d rises sharply, and a negative pulse plot). Figure 5-8 shows the slow settling during this period as the weak NMOS device, which is connected to Vef,N, finally redistributes the charge, generating an output with the correct sign. Due to bandwidth limitations, the impulse seen here is smeared out in time through the comparator's preamplifier chain. As a result, the comparator is subject to a near overdrive condition. Since self-timed bit-cycling (explained in Section 4.1.4) leverages latch time borrowing, which benefits all bits except the MSB, an extra half clock cycle is given to MSB bit-cycling so that it does not limit the ADC's performance. Purging Switches The purging switches short the common top-plate node to the six, separate bottomplate nodes of each capacitor array. Their implementation is shown in Figure 5-9 below. Here, a half-sized dummy switch, MN2, is used to minimize charge injection error. It turns off slightly after the main switch so that its injected charge is not lost, through the switches, to the bottom plate. Although, it is not strictly required, this device does help control the common-mode output voltage of the DAC, by reducing extraneous top-plate charge. In the structure shown, there is a single NMOS associated with the top-plate, MN1, and one for each of the bottom plates, MN3 - MN8. This arrangement minimizes the top-plate parasitics during the conversion process, by decoupling MN3 - MN8 from the top-plate when MN1 turns off. 82 Conversion Phase - - -'-'-----'-- - SAMPLE BCYCL_11 -' 00.5 0 - - 0 - - - - - - 4.2 4 4.4 5.2 5 4.8 4.6 DAC Ouput Voltages 1 S-- CD I VOUT+ ' I OUT- 05 0 4.2 4 4.8 4.6 4.4 DAC Differential Output Voltage (VOUT+ 5 - 5.2 VOUT) 0.5 CD 0) 0 -II - 0 -0.5 4.2 4 4.4 4.6 4.8 5 5.2 Time (g s) Figure 5-7: Waveforms showing the origin of a transient spike during MSB bit-cycling. 5.1.2 Capacitor array As mentioned in Section 3.3, mismatch in the capacitor array is the dominant source of low-frequency non-linearity in the ADC. Since this ADC does not employ capacitor calibration [34], the elements must be sized and laid-out appropriately to manage mismatch. Additionally, during the sampling process, thermal noise can limit the SNR of the ADC. During the sampling phase, the sub-DAC appears as a unit element. The total capacitance seen by the input, is the series combination of two arrays, each having a total value of 26Co, where Co is the unit capacitance. The resulting sampled noise 83 DAC Differential Output Voltage (VOUT+ - VOUT-) 1 0.8 0.610.4 E 0.2 0 CZ 0 -0.2 -0.4 -0.6 F -0.8 -1 4.5 4.55 4.6 4.65 4.7 Time (p. s) Figure 5-8: Zoom-in of DAC differential output voltage as it recovers from a negative spike to the correct positive voltage. power is given by Equation 5.1. 2 Vno,smpl _kT 32Co (5.1) This relation suggests that, with a signal of 1V amplitude, a sampling-noise limited SNR of 74dB (12 bits) is achievable with a unit capacitance of 6.5fF. The analysis below shows that capacitor sizing is instead limited by matching requirements. Poly-poly capacitors have been used for the DAC arrays due to their superior matching characteristics and stability. To determine the sizing required for sufficient linearity, consider the static transfer characteristic of the ADC. DNL is a measure of the deviation of code transition voltages from the ideal LSB voltage (after correcting for gain error). In the case of a 12 bit SAR ADC, the transition voltage associated with digital code 2047 is derived by the voltage divider shown in Figure 5-10a, where 84 TPPLT MN2 (Half-size dummy switch) MNI DDi~2 r~- 0011- itI rI-I 1 I-O k 0~ k t- U- -J Figure 5-9: Implementation of purging switches. the MSB capacitor and one unit capacitor are connected to Vref,N, and the remaining capacitors are connected to Ve,,p. The next transition voltage, associated with digital code 2048, is derived by the voltage divider shown in Figure 5-10b, where all the switchable capacitors are reversed. The worst mismatch scenario is where the MSB capacitor has a positive or negative error bias with respect to the remaining capacitors. Applying these considerations to the matching characteristics of the technology (provided by the vendor), a unit capacitance of 100fF was determined to be sufficient. Vref P 2N-2CO . . 2-1c0±0 . T -C: 2C0 Vref P C0O -2 T 2N2 OT ** . 2C 0 oT COT COT C0 Vref,N Vref,N Code 2047 (a) Code 2048 (b) C0 Figure 5-10: Capacitor mismatch leading to largest DNL. In addition to sizing, proper layout techniques are essential to ensuring the best matching of capacitive elements. Common-centroid arrangement has been used to 85 minimize the effects of linear process gradients, and dummy devices have been employed to ensure all elements are subject to similar etching environments. Further, a metal, electrostatic shield is used above the poly-poly capacitors to isolate their topplate from the bottom plate routing wires [35]. Coupling between the top-plate node, which is common to all capacitors in the array, and individual bottom plate wires adds to the net capacitance of the associated elements. If this is not well controlled, it can be a significant source of element mismatch. Note, the solid metal shield does increase the parasitic capacitance to ground associated with the top-plate. However, this does not affect the linearity of the ADC, at least in the case of the main-DAC. Parasitics associated with the top-plate of the sub-DAC are discussed below. Sub-DAC Transmission Compensation Parasitic capacitance on the top-plate of the sub-DAC leads to linearity errors in the ADC. This can be seen in Figure 5-11, where the Thevenin equivalent of the DAC network is considered. As mentioned in Section 4.2.2, voltages generated by the subDAC, VubDAC, are reflected to the output through the capacitive division between the net unit capacitor, Co, and the total main-DAC capacitance, ( 2 N _ 1)CO. In this way, the sub-DAC interpolates between transition voltages generated by the main-DAC array. Vref p VreP 1 i0 N 1) CO (2 Cc C0 The venin M-1 q2'C0 Equivalent M AI b,2'C 0 1 Vref,N I2.o (2N-1)C 0 VmainDAC -1 CC 2 MC CP' VsubDAC -LT jgo- VrLfN Figure 5-11: Effect of parasitic capacitance on sub-DAC top-plate. However, as shown in Figure 5-11, parasitic capacitance on the top-plate of the sub-DAC, has the effect of attenuating the sub-DAC output through the resulting parasitic capacitor divider (formed by 2 mCO and Cp). The sub-DAC is, thus, subject 86 to gain error, which manifests itself as a constant compression in code transitions within the interpolation range. However, this appears as non-constant deviations at the main-DAC transistion voltages, and thus contributes non-linearity. A sample transfer charcteristic, with a 2 bit sub-DAC, is shown in Figure 5-12 below. - ...... - MainDAC transitions Ideal subDAC transitions Compressed subDAC transitions Max. INL-+ - Figure 5-12: Errors in ADC transfer characteristic due to compression in sub-DAC interpolation. In this design, the resulting non-linearity has been minimized by adjusting the transmission gain of the sub-DAC, which is set by the coupling capacitor, Cc. By increasing Cc, the effective weight of the sub-DAC interpolation increases, compensating for the attenuating effects of the parasitic capacitance. The top-plate parasitics were estimated to be approximately 200fF based on RC extraction of the layout. A zoom-in of the resulting INL for the first 320 codes of this 12 bit converter, with no adjustment to Cc, is plotted in Figure 5-13. Here, as expected, a saw-tooth pattern is observed with a period equal to the sub-DAC interpolation range (i.e. 26). The worst case INL is -1.9LSB. Figure 5-14 shows the maximum INL expected for various values of Cc. The uncompensated value is 101.59fF. However, as shown, INL is minimized for a Cc of 104.76fF. The value used in this ADC was set accordingly. Note, adjusting Cc in this manner introduces gain error since the mainDAC no-longer sees a unit capacitance looking into the coupling capacitor. Figure 5-14 shows INL after proper gain-error correction. Since the method used here requires estimating the total parasitic capacitance on 87 INL VS Code 0 -0.5 Cl) -1 z -1.5 -2 0 50 100 150 Code 200 250 300 Figure 5-13: INL due to 200fF parasitic capacitance on top-plate of sub-DAC. INL with subDAC Parasitics VS Coupling Capacitance 41 1 - 1 _1 106 108 110 3 Cl) 2 -J z 1 E E 0 -J X -1 -2 -o30 100 102 104 Coupling Capacitance (fF) Figure 5-14: INL after adjustment of coupling capacitance Cc. the top-plate of the sub-DAC, it is worth evaluating the sensitivity of the INL to that parameter. Figure 5-15 shows, the resulting maximum INL if the actual parasitic capacitance differs by 100fF (50%) in either direction. In this range, a maximum INL of less than 0.8LSB is observed, so, some error in the estimated parasitic capacitance can be tolerated. 88 INL with predicted subDAC Parasitics VS Actual parasitics 1 M 0.5C,) z E 0 - E -0.5 -1 100 150 200 250 Parasitc Capacitance Capacitance (f F) 300 Figure 5-15: INL, after Cc adjustment, resulting from errors in estimation of parasitic capacitance. Edge Effect Minimization As mentioned, an array of dummy capacitors are used around the active elements to ensure a uniform etch environment. The top and bottom plates of the dummy elements are grounded. A conventional layout of this sort would look like the array shown in Figure 5-16a [27][35]. Here, the array has a total of six binary weighted capacitors, Co - C 5 (where C 5 = 2 5CO), in an interleaved and common-centroid arrangement. However, the conventional common-centroid array suffers from non-linearities due to the parasitic coupling between the bottom-plates of the poly-poly capacitors and the top-plates of neighboring elements. This is shown in Figure 5-17, where CP 1 to Cp3 all cause coupling to the common top-plate of the array, increasing the net capacitance of the actual element. CP4 , however, causes coupling to the top-plate of a grounded dummy device, and does not effect the net capacitance of the element. Thus, since active elements on the outside edge of the array do not experience the additional, parasitic increase in their capacitance, they are subject to a source of mismatch. Figure 5-16b shows our layout approach which was used in this design. It attempts to equalize the ratio of outside edges to capacitance for the largest capacitors. It is 89 Dlimnty A ray D mrhy A ray: C5 04 C5 C4 : C5 C4 C5 : , 4 1 d52 ---- q3 02 1 0 1 2 Ci C3J [iC . .. 1C11C4Cd I li 5 05 04 05 1 1: Dummoy Array: j iC5 : IC4 C4 C5 _5 Dumm y Array (a) (b) Figure 5-16: Capacitor array layout (a)Conventional common-centroid (b)Equal-edge ratio common-centroid. critical to focus on the largest capacitors, since their errors have the most impact on the linearity of the ADC. In this case, C5, C4 , and C3 are subject to 18, 10, and 4 outside edges, respectively; the ratios of 1, and 7 outside edges to capacitance are 18, 32' 16' . The corresponding ratios in the conventional case (24, 4, 32t16d and 1) are not as well 8 matched. Dummy Elements~Gun .P . . -- Ground r-Top-Plate Active Elements Figure 5-17: Bottom-plate to top-plate and dummy array coupling in poly-poly capacitor arrays. 90 5.2 Comparator Circuit Design The comparator auto-zeroing strategy was introduced in Section 4.2.1. As shown in Figure 5-18a, the first two stages of the 12-bit preamplifier cascade are auto-zeroed at a high impedance node. By design, the optimal input, and output, common-mode voltages are equal to mid-rail. Consequently, instead of active biasing, an initial phase to purge the charge on the capacitors is all that is required. Note, however, sufficient settling, requires that the auto-zeroing switches, which are implemented using NMOS devices, be overdriven using the charge-pump circuit. The final preamplifier, stage, which drives the latch, must have an output common-mode voltage slightly less than mid-rail. Consequently, it is auto-zeroed at the required bias voltage, as shown in Figure 5-18b. Generation of this bias point is considered in Section 5.2.2. Here, the auto-zeroing switch is delayed with respect to the biasing switches to equalize their charge injection error. Additionally, half-sized dummy switches not shown are used. PRG PRG IVCMLe!!h I -r -7_ PRG PRG VOUr- VIN+ V"N+ A JAZEROAZERO J VOUT+ V/&.' PRG -EROVou-r- (boosted and delayed) AF VIN- FAZERO AZEROA > VOUT+ --- PRG AZEROj 71 -7- FAZERO VCMW~th PRG PRG (b) (a) Figure 5-18: Implementation of auto-zeroing switches (a)First and second stage of 12-bit cascade (b)Last stage (only stage for 8-bit path). The remainder of this section discusses the design details associated with the preamplifiers and offset calibrating latch. 91 5.2.1 Preamplifers The preamplifier circuit is shown in Figure 5-19. This topology allows for a compact layout compared to a resistively loaded differential amplifier, and it achieves a good balance of offset and gain characteristics. Specifically, the VGS of the diode connected loads limits the output swing, easing overdrive recovery. Despite the reduced output range, however, offsets (due to 3- device mismatch) do not degrade the gain severely, as in the case of a current source loaded differential amplifier. VDD SLEEP P3 MP1 MP2 DVOUT+ VOUT-VIN+ =_ SLEEP MN4 MN2 MN1 VtaiioJ MN3 VIN- SLEEP] MN5 Figure 5-19: Preamplifier circuit. In Figure 5-19, the input devices, MN1-2, are biased in weak-inversion to maximize gm at the given current level. The load devices are biased in strong-inversion such that the gain of the stage is between 3 and 4. The devices MP3 and MN4-5 facilitate sample-rate scaling, by performing power gating, such that the preamplifiers shut off between conversions. Additional design details are considered below. Noise Considerations The preamplifiers are central to determining the actual precision achieved by the ADC. Excessive flicker and thermal noise can exceed the quantization noise, limiting the overall dynamic range. Auto-zeroing acts as a high-pass filter [29], and is effective at managing flicker noise. Here, design constraints imposed by thermal noise are 92 considered. The noise contributions of the diode connected loads, MP1-2, appear directly at the outputs, and are attenuated by the square of the amplifier's gain, when input referred. To simplify this analysis, the total noise will be estimated by considering the weak-inversion input devices, MN1-2. Their combined input referred noise power is given by Equation 5.2. v2 = 2 (f2nkT (5.2) gml,2 The integrated noise depends on the amplifier's bandwidth, which is set by the output resistance, Rose, and load capacitance, CL1. Alternatively, Rest can be expressed in terms of the amplifier's gain, A0 , giving the following expression. V2. -ni nkT AOCLl (5.3) From Equation 5.3, the load capacitance required to achieve 74dB of SNR (i.e. 12 bits), is 104fF. The actual value used in this design has been set more conservatively to account for the accumulation of other errors. Settling Time The time-constant associated with the first preamplifier is set by its R,t and the load capacitance, CL1, determined from the noise analysis above. Since subsequent stages are not noise limited, their load capacitance is determined by the fixed parasitics associated with their outputs. Accordingly, RC extraction of the layout provides a reasonable estimate for these output capacitance, CL2 and CL3, associated with each of the next two stages. The output resistance of the amplifiers is, then, constrained by the settling-time required of the cascade. Specifically, the overdrive case, where the preamplifiers are driven strongly in one direction, and then weakly in the other direction is considered. The corresponding waveforms are shown in Figure 5-20. Although the full-rail swing of VDD is not possible due to saturation in the amplifier, it can be used to analyze the worst-case overdrive, where the output of the cascade must recover from +1V. Assuming a 93 subsequent output swing of 5mV is required to ensure the latch resolves correctly, the settling requirements of the cascade can be derived. If each of the three preamplifiers has a gain of 3, a -L-2 input voltage (-240pV) will ultimately settle to -6.75mV at the output. The number of time-constants, N, required to exceed the required 5mV swing is given by Equation 5.4. : AC a Pre-amp output DAC output IV OV -- mlf ------to---- c t -240piV u -5 V p fl------o -6.75mV during-bit- Figure 5-20: DAC and preamplifier waveforms during overdrive recovery. N =-ln 1 - 1.00675) =.0 -6.35 (5.4) The preamplifiers are allowed to consume most of the clock period during bitcycling, since the digital logic and DAC impose very little delay and settling time. As mentioned in Section 4.1.4, the latch resolves quickly during overdrive conditions, allowing the preamplifiers to borrow a significant amount of time to recover. In this design, to ensure some timing margin, approximately 350ns, of the 500ns clock period, have been allocated to the preamplifiers to recover from overdrive. Assuming the preamplifier stages are identical, their output resistance, R., is constrained by the overdrive recovery time and the settling requirements. This is shown in Equation 5.5, where the method of zero-value time-constants has been applied. NRot(CL +0 CL2 + CL3) = 350ns (5.5) The maximum output resistance is derived using CL1, CL2, CL3, and N from above. Accordingly, for a stage gain of 3-4, the minimum GM of each amplifier can be derived, establishing the power consumption of the cascade. 94 Bias Current Optimization The above result was derived under the assumption that the GM's and R0 st's of all stages are equal. Since the time-constant of the first stage, which is loaded by the largest noise limiting capacitor, likely dominates the delay of the cascade, reducing its Rout can benefit overall performance. Specifically, reducing the overall time-constant, Ttot, while keeping the total power constant requires increasing the GM of an earlier stage, so that its Rout may be reduced, while decreasing the GM of a later stage. Essentially, this corresponds to overall power-delay optimization, where the power is constant, an the delay is minimized. In the case of the three-stage cascade, consider increasing the transconductance of the first stage to Gm + aGM, and reducing the transconductance of the third stage to Gm - aGM. The resulting change in timeconstant is given by Equation 5.6. A-tot = CLRot(G Gm +aGm 1) + CL3 Rout ( G - aGM - 1) (5.6) The percentage change in the overall time-constant is plotted in Figure 5-21 for the CL1, CL3, and Rout values of this design. As shown, the time-constant is minimized for a equal to 0.4. Although this optimization does not give the global minimum, it is a tractable approach and gives a result that is practically as close. The preamplifier bias currents are, thus, set accordingly. Biasing Circuit Figure 5-22 shows the biasing source employed to generate the preamplifier tail currents [36]. Here, MN1-2 are in weak-inversion, and MP1-2 are in strong-inversion to improve the matching of the mirror. The current setting resistor is off-chip so that the bias currents may be manually adjusted during testing. However, the required resistance is approximately 40kQ, which is amenable to integration if desired. Like the preamplifiers, the biasing source is power gated so that it may be turned-off between conversions. Additionally, during testing, the entire source can be disabled, to allow an external biasing option through MP4. Finally, although it is likely not needed, a 95 Percentage Change in Time-Constant VS Changes in II :n II 1510- 0 -~0 0 o -5-10-15-20 0 0.2 0.4 0.6 0.8 1 Figure 5-21: Change in cascade time-constant with respect to fractional changes in Gm. start-up device, MN3, has been used in this design. Since the non-zero, equilibrium current is quite low (-1pA), the difference in the leakage currents of MN1-2 (due to their sizing ratio) should force the circuit to start up by itself. 5.2.2 Offset-Calibrating Latch A simplified schematic of the offset calibrating latch is shown in Figure 5-23 below. Here, MP1-MP2 form a regenerative load, MN1-MN2 form the input pair, and MN3MN4 form tail biasing sources, whose drains can be isolated. As described below, this allows the differential structure (composed of MN1-MN4) to be configured as two separate voltage-follower amplifiers. The remaining devices facilitate the calibration operations. Complete operation of this circuit occurs over two phases: an auto-zeroing phase and a reset-resolve phase. The auto-zeroing phase coincides with auto-zeroing of the preamplifiers and is used to calibrate the relative offsets of the voltage-followers. The reset-resolve phase coincides with bit-cycling and allows successive latch decisions to 96 Off-chip bias VDD E ENABLE MP4 SLCPMP3 MP1 MP2 N3 MN1 -Tail bias MN2 Off-chip resistor (40kn) Figure 5-22: Preamplifier biasing source. be made while enforcing offset cancelation. Auto-Zero Phase Here, operation of the latch during the auto-zeroing phase is considered. As mentioned, MN1-MN4 can be configured as voltage followers, thereby reflecting incremental differences in the input voltages to the source nodes of MN1-2. The purpose of the auto-zeroing phase is to eliminate the differential offset in these source voltages. Figure 5-24 shows the relevant circuit elements. The gray devices are turned off by driving the appropriate values onto their gates (through switches not shown). During this time, it assumed that zero differential voltage, with the proper common-mode value, is applied to the inputs. This is achieved using the biasing circuit described later. Initially, switches Scscd, SdiffMd, and Sfb are all closed. This biases MP7-8 and MN3-4 while forcing the source voltages of MN1-MN2 (i.e. the follower outputs) to be equal. If SdiffMd were open, the difference in the source voltages would be determined by the Vcs's of MN 1-2, which depend on the devices' drain currents. Offsets in MN3-4 and MP7-8 would cause a difference in the drain currents, and offsets in MN1-2 would 97 V8 MP7 MPMP mplk! MP O - VOUr- MP2 VOUr.C MP4 MPS MP RSLV VIN+1-> T ! -- MN1 MN2 1MN3 MN4 [-<VIN- T Figure 5-23: Simplified schematic of complete offset-calibrating latch. cause VGS shifts. In this case, however, current flows through SdiffMd to equalize the source voltages. During the second part of the auto-zero phase, Scd and SdiffMd are opened. Firstly, if the feedback connection through Sfb is ignored, this dramatically increases the impedance at the drain of MN1-2 due to the resulting NMOS and PMOS cascodes. Secondly, the current path through SdiffMd is eliminated, exposing all offsets by changing the branch currents. A block diagram of the resulting structure is shown in Figure 5-25. The feedback network is composed of a transconductor, gmn3,4, driving the high impedance node at the drain of MN1-2, Vhg. This drain impedance is the equivalent resistance, Re.. , of the parallel NMOS and PMOS cascodes. Here, the transconductor current, ifb, is combined with the incremental offset current, i0 s, which is caused by opening SdiffMd. The resulting current, Zer,, is multiplied by the output resistance of the voltage-follower, Rfiiw, "mp1,2on1, . which is approximately equal to ro3 ,4 The feedback loop attempts to minimize ierr and, therefore, attenuate incremental changes in the voltage-follower outputs, Voutflwr. 98 _177 MP8 MPP7MP 7, Vc, F--OVd -3r SOc RSLV VN1 MN1 MN3 MN4 SdMid | VIN- sd I soSL Clock __ [N2 | | Auto-Zero, S Sowd, SaMW RSLV Figure 5-24: Offset compensating latch during auto-zeroing. Intuitively, the feedback loop can be thought of as means of enforcing drain currents in MN1-2 that are equal to those before opening SdiffMd. Since the drain currents determine the corresponding VGs's, the source voltages will return to their original values. Specifically, opening SdiffMd causes incremental changes in the branch currents. When these appear at the high-impedance drains of MN1-2, they result in large voltage swings which are fed-back to the gates of MN3-4. Consequently, MN3-4 are rebiased so that the incremental voltage is eliminated. This requires canceling the incremental current, returning the source voltages to their original values. The calibration loop may also be analyzed using a small-signal circuit model. For simplicity, the impedance of the PMOS cascode, composed of MP5-MP8, is assumed to be infinite. The resulting right-half calibration circuit is shown in Figure 5-26, where i0, represents the change in current caused by opening SdiffMd, and vca is the incremental change in MN4's gate voltage. Once again, resulting changes in vout,flLwr 99 iOs + -- voutfll, Rscd Vhg 'err + If_ gmn3,4 Rmw Figure 5-25: Block diagram of latch calibrating circuit reflect errors in the desired calibration. r.2 4 g2 i 0~ Vin-Vout OW 0 r04 9" V.8, Figure 5-26: Latch voltage-follower calibration circuit. The following equations are derived from KCL and KVL analysis. = VOS of' + gm4vcal VcaI = Vout,fllwr + (5.7) (5.8) Voutfllwrgm2ro2 From these, the error in the follower output voltage can be determined with respect to iOs. Vout,f Iwi _= r_4_________(__.9) __ 1 + gm4ro4 (1 + gm 2 ro 2 ) gm4 9m 2 ro 2 (5.9) Now, to estimate i0.,, a Thevenin equivalent of the voltage follower outputs can be considered. Figure 5-27 shows the resulting structure, where the output impedance is approximately equal to 1 , if ro, 2 are large and the drain impedances are low. Note, the drain impedances before opening Sc8 cd should be considered here. Then, i,, can be approximated by Equation 5.10, where v, 8 is the uncalibrated, half-circuit voltage-follower offset voltage. 100 1/gm1 Vfilwrjcd/ + 1/gm2 1OS _ VfIrwrid+Vos Figure 5-27: Latch voltage-follower equivalent circuit during first half of auto-zeroing phase. ios = gm 2 vos (5.10) Using Equation 5.9, the follower offset voltage, after calibration, is given by Equation 5.11, where, as shown, the uncalibrated offset is attenuated by an amount near the intrinsic gain. Voutfl = vOS (5.11) gm4ro2 Figure 5-28 shows the simulated waveforms observed during the auto-zeroing phase. Here, 3a- offset voltages have been applied to each pair of devices in the latch circuit by placing an ideal voltage source in series with one of the gates. The voltage sources have been applied so as to give the worst-case, overall offset. As shown, during the second part of this phase, when Sc.cd and Sdiff Md open, the gate voltages of MN3-4 get re-biased such that the outputs of the voltage-followers converge. Note, at the end of this phase, the voltage-follower outputs are not exactly equal, differing by approximately 1mV. Reset-Resolve Phase Here, operation of the latch during the reset-resolve phase is considered. Figure 5-29 shows the relevant circuit elements. Once again, the gray devices are turned off by driving the appropriate values onto their gates (through switches not shown). Additionally, during this phase, RSLV is high, and the parallel NMOS and PMOS devices, MN5-6 and MP5-6, effectively short their source-drain connections. While Clock is high, the sources of MN1 and MN2 are de-coupled, and their 101 Auto-Zeroing Switches ) 0)0.5 0 > 0 3 .4 S osod' e Sdif S Md - b 3.6 3.8 4 4.2 I MN3-4 Gate Voltages , 4.4 4. 6 0.44 a) 0)0.42 0 > 0.4 - 4 -oi-oe- 3 .4 3.6 -Follower -- 3.8 4 4.2 Ouput Voltages, - - 4.4 4.6 ,0.155 a) 0.15 - - t 0.145 - I I I a 0 0.14 3 .4 - ... 3.6 3.8 4 time ([t s) * I 4.2 4.4 4.6 Figure 5-28: Latch circuit waveforms during the auto-zeroing phase. voltages are determined by VIN+ - VGS,MN1 and VIN- - VGS,MN2, respectively. Of course, VGS,MN1 and VGS,MN2 are set by the bias currents, which were calibrated to result in equal source voltages for zero differential inputs. To understand the behavior of the regenerative loads, MP3-4 should initially be ignored. Since Saux and closed, positive feedback is disabled, and the voltages generated across Srgnrt Crgnrtl,2 are hold MP1-2 at a stable bias condition for the given branch currents and device offsets. However, at the falling clock edge, Sa and Srgnrt open, and SdiffMd closes. If the source voltages of MN1-2 were initially equal (which only occurs for the zerodifferential input condition), the branch currents will remain unchanged, preserving metastability in MP1-2. However, any difference in the source voltages will cause a perturbation in the branch currents and trigger regeneration in MP1-2. Note, Sa,, is an auxiliary switch [37], which is delayed with respect to Srgnrt, to prevent positive feedback until the charge injection errors from the Srnt switches are equalized. A practical problem in the circuit described above originates from the finite r's 102 VB Crggvn Cg1sC tii~ Sc, Vour+C< ->VoUTr LiI' Vc$~(:Q.. TN5 V N+ MP4 MP2 MP MP3 .... .... MW RLV MNI MN2 [-cVIN- SA~ MN4 MN3 Clock~ Sx, Srgnrt | Saux SdfflMd Vo, Vout. RSLV Figure 5-29: Offset compensating latch during reset-resolve phase. of MN1-2. Proper operation of the voltage-followers, requires that the drains of MN1-2 be subject to a low impedance. Voltage swings on these nodes will cause incremental changes in the portion of the branch currents passing through the devices' r0 's as opposed to their gm generators. Consequently, their VGS's will deviate form the calibrated values. MP3-4, which are initially diode-connected, provide a low impedance path while the circuit stabilizes. Then, during regeneration, the diodeconnections are broken, enabling strong positive feedback. Figure 5-30 shows the simulated waveforms observed during the reset-resolve phase. Once again, 3a- offset voltages have been applied to each pair of devices, in the directions resulting in worst-case overall offset. An input differential signal of i1.5mV is successfully resolved. Note, while clock is high, the outputs are not 103 exactly equal. This is due to the offsets in MP1-4, which cause different voltages to be generated across Crgnrt,2- Clock 00.5 0 >0 7.8 8 8.2 ~0.414 8.4 8.6 8.8 9 Diffqrantial lnput 9.2 9.4 9.6 I -~~ IN+ -- 0)0.412 0 > 0.41 -,-,-,-,LatchOutptI -I- 7.8 V 8 1 8.2 I I 8.4 II I - a) 8.6 8.8 Latch Output f 9 M 9.2 a 7.8 8 8.2 9.6 I iOUT+ -~~ 00.5 0 > 0 9.4 " - 8.4 8.6 8.8 time (g s) 9 9.2 OUT- 9.4 9.6 Figure 5-30: Latch circuit waveforms during the reset-resolve phase. Biasing Two bias voltages are needed within the latch circuit. The first, Vscd, is simply at ground. MP5-6 are biased in moderate-inversion, leaving over 500mV for the VDS's of the load devices, MP1-4 and MP7-8. Additionally, MN1-4 are all in weak-inversion and have a VDS,sat of slightly more than 1O0mV. The branch currents (as well as the nominal drain voltages of MN1-2 during auto-zeroing) are set by the relative sizing of MN3-4 and MP7-8. In this implementation, branch currents of approximately 1.5pA are used. The VGS of MN3-4 is approximately 400mV, so that MP7-8 are in stronginversion making the biasing less sensitive to supply variations. During reset-resolve, the nominal drain voltages of MN1-2 are set by the diode loads, MP3-4. The second bias voltage, VB is generated using the replica circuit shown in Figure 104 5-31. Here, MP9-10, which replicate the regenerative and diode loads, are active during the reset-resolve phase, and MP12 is active during the auto-zeroing phase. In addition, to VB, this circuit generates the bias VCM-IN which is an appropriate common-mode voltage for the inputs. Consequently, it is used to set the DC bias of the last, auto-zeroed preamplifier stage. MP10 MP9 MP12 RSLV VCM-IN 7 MN7 MN8 Figure 5-31: Replica biasing circuit for latch. Both the latch currents, and the biasing circuit currents, are turned off between conversions to yield linear power savings with respect to reduced sampling rate. Although the branch currents of the replica biasing structure can be scaled back compared to those used in the active latch, there is practical limitation associated with this approach. Reducing the sizes of devices in the biasing circuit degrades their matching characteristics. Since the current levels are quite low and the devices are in weak-inversion, additional offsets can lead to complete inactivation. Consequently, current levels have not been further reduced in the biasing circuit. Implementation Analysis Several implementation issues associated with the latch circuit are worth noting. First, the only switches that introduce charge injection error are ccd and Sfb. However, since they are used to bias static nodes, large gate capacitors (Ccscd, CA 105 ), as well as dummy switches, can be used to minimize this effect. The injection errors from Srgnrt and S,, are equalized by the auxiliary switch Sa.. Finally, Sdiff Md, which actually triggers the regeneration, does so by shorting two nodes. Consequently, the possibility of false regeneration due to residual mismatch in charge injection is naturally avoided. This is a particularly beneficial feature of the latch circuit. During the reset-resolve phase, the regenerative load is initially stabilized by breaking the positive feedback. However, as mentioned in [28], the finite "on" resistance of the switches may not allow the positive feedback to be fully broken, resulting in instability. Care, in this regard, is particularly required for low-voltage implementations. Finally, the feedback structure used during auto-zeroing can also be subject to instability if parasitics are not properly managed. After the Scscd switches are open, each branch of the latch may be thought of as a high-gain cascode amplifier whose output is fed-back to its input through Sfb. Although the resulting high-gain amplifier is a single-stage structure, the branch currents are quite low, resulting in low gm's for MN1-2. Consequently, their sources present a relatively high impedance. Excessive parasitics on these nodes, or additional capacitances, which may be added to smoothout transient spikes, can degrade the phase margin of the amplifier. Figure 5-32 plots the gain and phase of the amplifier for the case where no filtering capacitance has been added and the case where 400fF of filtering capacitance has been added. As shown, the phase margin degrades considerably in the filtered case. Note, Cfb is the compensation capacitor in this loop. Its value has also been reduced in the filtered case shown in Figure 5-32. 5.2.3 Latch Level Restorer The latch circuit described above does not achieve full-swing outputs. Although, it is possible to reach VDD, output voltages below 250mV can not be generated since MP3-4 remain on despite regeneration. So, the level-restorer circuit, shown in Figure 5-33, is used to generate full-swing digital outputs. Here, depending on the latch outputs, one of the PMOS input devices, MP1 or 106 Gain 20 - M.:.............~....... 0-- No Filtering Filtering . . 0. .. .. ......................................... -10 103 10 Xi 10 10 10 108 Phase 0 No Filtering Filtering -- -50. -.- -100.- .- - .- .- phas -150..Degraded :margin: -200. 103 104 10 10 Frequency (Hz) 10 10 Figure 5-32: Gain and phase of filtered and unfiltered auto-zeroing calibration circuit. MP4, will turn on when Clock goes low. The corresponding NMOS, MN2 or MN4, whose gate was initially discharged, will turn on. The other NMOS will, dynamically, remain off. This circuit is robust, as it requires VLATCH+ and VLATCH- to only swing slightly more than VTp1,4, in order to generate full-swing digital outputs. The signals corresponding to the 8 bit and 12 bit paths are ANDed together at this point, to drive the control signal of the SAR state machine. Waveforms showing the operation of the level restorer circuits are given in Figure 5-34. As shown, the full rail digital signals are derived for latch outputs of less than 400mV. 5.3 SAR Circuit Design The SAR state machine is the digital block that controls the conversion cycle by generating purging, auto-zeroing, sampling, and bit-cycling signals at the appropriate times. In this, low-speed, design, the SAR imposes minimal delay with respect to the target clock frequency. As a result, a standard cell implementation has been leveraged. 107 VL AT CH+- =-I VDD MP Clock VLATCH- m--- 4 VL AT CH- MP3 VDD Cockf VLATCH+ MP2 6 -UTMP5 C7ok?:DOUT DOUT C1ok MNI MN2 MN3 MN4 Figure 5-33: Latch level restoring circuit. Clock a> >0.5 0 8 8.5 -- 9 9.5 Latch Outputs ' --- % - 0.5 0 LATCH+ LATCH- t - 9.5 10 0 8 8.5 9 Level Restorer Outputs -OUT - =- a 0. > 0 8 8.5 time ( t s) 9 9.5 10 Figure 5-34: Latch level restorer waveforms. The core state logic, in this SAR, is a simple shift register, which makes up the top row of flip-flops shown in Figure 5-35. The conversion is initiated by an activehigh pulse, CNVRT, which sets the first flip-flop in the shift register and resets the remaining flip-flops (the reset of lower flip-flops is not shown). In all cases, the flipflops are set or reset asynchronously, as the clock is internally gated (as described 108 in Section 5.4) to minimize power consumption between conversions. Following the CNVRT pulse, however, the clock is enabled, and the SRx signals of the shift register are successively asserted, enabling their associated logic. CNVRT PREB a_ CLR CLR CR1 SRIO S9 PREB PREB SR8 _ CLOCK MTSTBL CMPRTRDONE CMPRTR OB PREB OB ' r 0B - r o toco r 0 Figure 5-35: SAR state machine circuit. The initial signals, enforcing the purging, auto-zeroing, sampling, and MSB bitcycling phases are synchronous, and are derived in a straight-forward manner from the associated shift register signal. The remaining bit-cycling signals are self-timed using the comparator outputs. The corresponding logic, which must be able to recover from comparator metastability, is described below. Additionally, the logic to enable resolution scaling is also described below. Self-Timing Logic The comparator evaluates while CLOCK is low. Self-timed bit-cycling allows the subsequent bit phase to begin immediately after a decision is made. Consequently, the shift register flip-flops, associated with bit-cycling, are triggered on the negative clock edge. The comparator provides pseudo-complementary outputs, which are both precharged high, in the level-restorers, before the latch resolves. A comparator decision 109 can then be detected by simply NAND'ing the outputs. This derives CMPRTRDONE, which is gated with the SRx signals to assert preset on the appropriate bit-cycling flip-flop. The rising BCYCLx signal is used to clock the previous bit-cycling flipflop, which, accordingly, stores at "1" or "0" depending on the comparator decision. Comparator Metastabiliy Detector However, in the scheme described above, if the comparator fails to make a decision, the entire bit-cycling process is stalled. To recover from such metastability conditions, the circuit in Figure 5-36a is used. Here, the CMPRTR-DONE signal is registered at the rising-edge of CLOCK. At this point, a comparator decision should have been made. If it has not, however, the MTSTBL signal gets asserted and enforces bit-cycling. MTSTBL_RST CNVRT CMPRTROUT TSTBL_RST PREB MPRTRDON CMPRTROUT R11 0 0B MTSTBL 10 YCL_11 -CYCL_10 TSTBL TSTBL * SR9 CYCL_9 TSTBL CLOCK (a) (b) Figure 5-36: Metastability recovery circuits (a)Metastability detector (b)Metastbility reset. Since metastability is detected at the rising clock edge, but bit-cycling logic is enabled by falling-edge SRx signals, MTSTBL must be reset before erroneously affecting the next bit. This is done by asynchronously asserting the flip-flop's clear signal using the logic shown in Figure 5-36b. Here, the BCYCLEV signal, which was just asserted by MTSTBL, propagates back to generate the reset. Resolution and Sampling Rate Scaling Scaling the ADC resolution from 12 bits to 8 bits requires bypassing the last four shift register flip-flops. Figure 5-37 shows the logic associated with the last flip-flop in the shift register, which is used to enable assertion of the SLEEP (once again, 110 reset of the lower flip-flop is not shown). The SLEEP signal indicates the end of the conversion cycle. It is used internally to power gates all of the active circuitry as well as to gate the ADC clock. Similar logic is used to bypass one of the shift register flip-flops associated with auto-zeroing, since only one preamplifier exists in the 8 bit comparator path. CNVRT 12BITMD CLR 00 S9R4 SRO CLOCK MTSTBL CMPRTR DONE PREB 0 SLEEP OB- Figure 5-37: Circuitry supporting resolution scaling. SLEEP signal is used to enforce power-gating. 5.4 Clock Manager To minimize the power consumed between ADC conversions, the internal clock signal is gated as shown in Figure 5-38. The inverters, driving CLOCK and CLOCK have been sized with cosideration to the internal clock load. CLOCK SLEEPCLOCK CLOCKEXT- Figure 5-38: Internal clock gating circuitry. 111 5.5 Charge Pump and Voltage Multiplier To increase the overdrive of NMOS analog switches, the charge pump circuit shown in Figure 5-39 is used. An unloaded output voltage of 2 VDD - VTN1 is generated. With loading, the output level depends on the relative size of Cpump and the load capacitance. Consequently, Cpump has been sized differently for the auto-zeroing switch drivers in the comparator and the input switch drivers in the DAC. VDO MNI MPI CBYPA- BAKBIAS BYPASS :OUT MN2 IN Figure 5-39: Charge pump circuit to generate voltages beyond VDD In the implementation shown, the junction implants of the PMOS device can be forward biased if the voltage of its bulk remains at VDD. To avoid this, a separate charge pump can be used to derive the bulk voltage, BAKBIAS [38. In this design, the Dickson voltage multiplier, shown in Figure 5-40, is used [39]. Since it uses only NMOS devices, its junctions are not at risk of being forward biased. The steady-state output voltage of this circuit is given by Equation 5.12. BAKBIAS = VDD + VDD Cboost - 3 VTN CbSt + Ccntri Thus the desired bulk bias is achieved by setting the ratio of Cboost to approximately 1. (5.12) Ccftri to Figure 5-41 shows a simulation of this circuit during startup. As shown, BAKBIAS is continually pumped higher each clock cycle, reaching a steady value beyond VDD (1v). 112 MNI MN3 MN2 CLOCK CLOCK= CLOCKG CbO.l,. cc.I .. Cto(Jt BAKBIAS cenf Figure 5-40: NMOS Dickson voltage multiplier [39]. Clock Signals 1 - - - - - - '- -' '--- CLK LR - a> 0)0.5 I -- 1 0 ) - - -U 2 -. - - - -J 4 -, - -U 6 8 6 8 0 BAKBIAS ,1.5 ) 1 -F 0.5 0 0 2 4 time (ji s) Figure 5-41: Voltage muliplier simulation. 113 10 Chapter 6 Testing and Characterization The low-power ADC was fabricated in a 0.18prm CMOS process. We acknowledge Nation Semiconductor for providing the fabrication services. a 0.5mm pitch TQFP package. Figure 6-1. It was packaged in A micrograph of the entire test chip is shown in It includes two complete ADCs, as well as a separate layout of the offset calibrating latch. The first ADC, is the full design described in this document. The second ADC uses a standard, sense-amplifier based latch [40], instead of the experimental one. The additional latch layout allows controlled characterization of offset. Including pads, the test chip occupies an area of 2mm x 2mm. Figure 6-2 shows an annotated micrograph of the full ADC with the offset calibrating latch. It occupies a total area of 900pm x 700pim. 6.1 Test Setup Figure 6-3 shows the test setup and equipment used to characterize the ADC. The Audio Precision System One signal source was used to generate an ultra low-distortion sine wave [41]. It provides fully-balanced, differential outputs, which can be floated and set to the desired common-mode voltage. A Tektronix PS280 power supply is used for this purpose. A Tektronix TLA7NA3 Logic Analyzer module samples the digital output of the ADC, and a Tektronix TLA7PG2 generates the clock signal and conversion pulse. An additional PS280 power supply is used to power the ADC I/O 115 Figure 6-1: Micrograph of fabricated test chip. A k F 900um Figure 6-2: Micrograph of full ADC with offset calibrating latch. 116 drivers. Finally, a Keithley 6517A electrometer is used to power the ADC core as well as take accurate current measurements. Tektronix PS280 Keithley 6517A 0 Power Core ?ower Audio Precision System One Analog Input J _L r--]_ ADC (under test) Tektronix STLA7NA3 InputDigital __pu Common-Mode Convert Pulse CTektronix C k Tektronix PS280 TLA 7PG2 Tektronix TLA715 (mainframe) Figure 6-3: ADC Test setup. A custom printed-circuit board was designed and fabricated to facilitate testing the ADC. The four layer PCB uses two signal layers and two power/ground layers to make the board connections. On-board switches allow configuration of the ADC's resolution, and on-board potentiometers and biasing circuits allow adjustment of bias currents and voltages. A picture of the test board is shown in Figure 6-4. The complete ADC, with offset canceling latch, is functional at the target sampling rate. The biasing resistors require no adjustment beyond their expected, simulated values. In 12 bit mode, the circuit operates with a 2MHz clock, and, in 8 bit mode, it operates with a 4MHz clock, allowing a sampling rate of up to 200kS/s. The entire ADC operates from a 1V supply, as expected. Note, these parameters will be adjusted to characterize the maximum and optimum performance points of the ADC beyond expectations. 117 Figure 6-4: ADC Test PCB photograph. 6.2 Characterization This section describes the specific tests performed on the ADC. Additionally, results are presented along with some brief analysis. All of the tests used here are prescribed by the IEEE Standard 1241, Test Methods for Analog-to-Digital Converters [42]. 6.2.1 Static Linearity A variety of tests exist for extracting the static transfer characteristics of an ADC. The code density test involves deriving the digital output code histogram associated with a low-frequency sinusoidal input [43]. Since sine functions are well characterized mathematically, the ideal code density histogram is well known and can be compared with that observed. Thus, linearity errors can be quantified. The code density test was conducted using a full-swing, differential sinusoidal input with amplitude of 1V. The sampling rate of the ADC was 1OOkS/s, and the 118 frequency of the input signal was 111.381Hz. To test the 12 bit mode, approximately four million samples were taken (30 records each with a size of 131,072). Figure 6-5 shows the resulting code density histogram. Here, no bins are empty, suggesting a very low likely-hood for missing codes. Further, the offset of the ADC, determined using the method prescribed in [421, is approximately 830p-V. X 104 12 Code Density histogram 108 -2 IL 6 - 04 0420 0 1000 2000 3000 4000 Code Figure 6-5: Code density histogram of ADC in 12 bit mode. From the code density histogram, the INL and DNL can be determined. Figure 6-6 shows these with respect to the output code. The maximum INL is +0.68LSB/0.56LSB, while the maximum DNL is +0.58LSB/-0.66LSB. The INL plot has a sawtooth characteristic with a period of 64 (6 bits). This is a manifestation of imperfect sub-DAC interpolation, and is likely caused by a small error in the prediction of its top-plate parasitic capacitance (as described in Section 5.1.2). The peak value of the saw-tooth, however, is only 0.2LSB, confirming that compensation of the sub-DAC transmission gain was quite effective. Abrupt changes in the INL are also observed at the 1024 and 3072 code transitions, indicating mismatch in the MSB/2 capacitor (i.e. C4 in Figure 5-16b). Recall that the largest capacitors were arranged in an attempt to equalize their edge-coupling to active-capacitance ratio. This ratio for 119 C4, although much improved compared to the conventional layout, is most poorly matched. As a result, small corresponding INL changes are observed. Nonetheless, the static linearity of the ADC is quite good, well within t1LSB. DNL 1 0.5 M C/) _j L ALA~jW' -. .Ak -N 0 -0.5 -1 0 500 1000 1500 2000 2500 3000 3500 4000 2500 3000 3500 4000 Code INL 1 0.51 mj 0 -0.5 -1 0 500 1000 1500 2000 Code Figure 6-6: DNL and INL of ADC in 12 bit mode. Figure 6-7 shows the histogram corresponding to the ADC in 8 bit mode, and Figure 6-8 shows the DNL and INL. Here, the sampling rate is 200kS/s, and the input frequency is, once again, set to 111.381Hz. The maximum INL is +0.19LSB/- 0.16LSB, while the maximum DNL is +0.16LSB/-0.14LSB 6.2.2 Dynamic Noise and Linearity The dynamic performance of the ADC is also characterized by means of tone testing. The signal-to-noise-plus-distortion ratio (SNDR) is derived by varying the input frequency from DC to one-half the sampling rate, so that the Nyquist performance 120 6 x 10 4 Code Density histogram 5 c4 C) =3 2 _-.2h6,16010) 1 0 0 50 100 Codi 150 200 250 Figure 6-7: Code density histogram of ADC in 8 bit mode. DNL 1 0.5 M C/) -j 0 -0.5 -1 ) 50 100 Code 150 0.5 Cl) _j 250 INL 1 M 200 l il''IIl) 1111lll11111111il W -i 0 -0.5 -1 0 50 100 150 200 250 Code Figure 6-8: DNL and INL of ADC in 8 bit mode. of the ADC may be evaluated. The resulting output, at each frequency, is used to derive a least-squares fit to an ideal sinusoid. The RMS error between the output and the fit quantifies the noise and distortion [42]. The resulting SNDR can be used, 121 as in Equation 6.1, to express the effective number of bits (ENOB) achieved by the ADC. ENOB = SNDR(dB) - 1.76 6.02 (6.1) Figure 6-9 shows the ENOB of this ADC with respect to the input frequency. In 12 bit mode, the ADC samples at 1OOkS/s and achieves an ENOB of 10.55 bits (65.3dB SNDR) operating at its Nyquist rate. The SNDR is fairly consistent over the input frequency range. In the 8 bit mode, the ADC samples at 200kS/s and achieves an ENOB of 7.96 bits (49.8dB SNDR) at its Nyquist rate. ENOB VS Input Frequency 12 EU U * -U- 12B Mode 8B Mode ~n 10 84 En0 z 6 w 4 2 0 0 - - - - 20 40 60 80 100 Input Frequency (kHz) Figure 6-9: ENOB versus input frequency for the ADC in 12 bit mode and 8 bit mode. In 8 bit mode, the accuracy of the ADC is nearly ideal. The loss of precision in 12 bit mode can be analyzed with the aide of an FFT. Figure 6-10 shows an FFT of the ADC output for an input tone of frequency 47.3kHz, which is near half the sampling rate. The 3 rd 5 th, and 7 th harmonics are all visible. From the plot, the spurious-free dynamic range (SFDR) is 71dB. The combined power of the harmonics in Figure 6-10 is approximately -70.4dB 122 FFT 0-20 -40- - -603rd Harmonic (-71dB) -80- 5th Harmonic (-80dB) 7th Harmonic (-89dB) -100 -120 0 10 20 40 30 50 Frequency (kHz) Figure 6-10: FFT of ADC output with 47.3kHz input tone. with respect to the fundamental (i.e. -10logio(10 1 +10 1 -0 + 10 )). This suggests that a modest portion of the observed degradation in SNDR is due to harmonic distortion. Since no even-order harmonics are prominent, the source of the distortion must effect the positive and negative signal paths similarly. Thus, mismatch in the DAC capacitors, which is uncorrelated in the positive and negative arrays, is not a likely source. Further, good matching of the DAC capacitors was confirmed by the INL and DNL tests, and it is not dependant on input frequency. The most likely source is the non-linearity in the resistance of the CMOS input switch. In simulation, the resulting power of the distortion introduced by the sample-and-hold circuit is -69dB with respect to the fundamental. This was measured by a applying a 50kHz input signal to the distributed circuit consisting of six input switches and six binary weighted capacitors. An ideal sinusoid was fit to the output voltage, and the meansquared power of the error was determined. 123 6.3 Power Consumption At the highest 12 bit performance point, corresponding to 1OOkS/s, the ADC core (not including I/O) consumes 25puW from the 1V supply. At the highest 8 bit performance point, corresponding to 200kS/s, the ADC core consumes 39pW from the 1V supply. The power consumption in 8 bit mode is higher than expected and is currently under investigation. In both 12 bit and 8 bit modes, the linear power scaling with respect to sampling rate is observed as expected. Figure 6-11 shows the power consumption versus sampling rate, where it can be seen that the power consumption approaches zero for very low sampling rates. Power VS Sampling Rate 40 -E- 12B Mode - . 8B Mode 35. 30- Af 25 ifN 20- o1510F 5 50 100 150 200 Sampling Rate (kS/s) Figure 6-11: ADC power consumption with respect to sampling rate. The power consumption in 12 bit mode matches quite well with simulations. Table 6.1 shows the measured and simulated power consumption of the constituent blocks in this design. 124 Block(s) Simulated Power Measured Power SAR, Switch Matrices 8.7pW 8.4pW Charge Pumps 0.39pW 0.38p1W Pre-Amplifiers 4.9p1W 5.8pW Latch 3.9pW 5.2pW DAC 5.9pW 5.3pW Total 23.8pW 25pW Table 6.1: Simulated and measured power consumption of ADC blocks. 6.4 Comparision Study The power consumption of an ADC may be normalized by its effective dynamic range and speed, to derive a figure-of-merit (FOM) for comparing a broad range of implementations. The FOM commonly used is shown in Equation 6.2. FOM = P 2 fP2 ENOB Here, P is the power consumption of the ADC, and fin (6.2) is in input frequency at which the ENOB is measured. For the ADC presented, the FOM is equal to 167fJ/conversion step in the 12 bit mode. Figure 6-12 plots the FOM of previously demonstrated implementations against their resolution. As shown, this design is the most efficient known to the author, and, so far as Nyquist rate converters, of at least 12 bits, are concerned, the FOM is over five times better than previous implementations. In addition to good efficiency, this ADC achieves low absolute power consumption. Figure 6-13 plots the power consumption of previously demonstrated SAR ADCs against the input bandwidth they can digitize, while Figure 6-14 plots their power consumption against resolution. Architecture comparison in Section 1.2 suggested that SAR ADCs are able to achieve the lowest absolute power. At the 12 bit level however, their power consumption increased faster than other architectures (namely oversampling ADCs). As shown, the increase in power consumption of this 12 bit design, compared with low-resolution (8 bit), micro-power solutions, is modest. 125 FOM I x * Resolution 0 x AL x~ X xx Ax xA C,, C- 12 Flash SAR x Pipelined A AE * o ) x 0810- x A 0 L- A A 0 10-13L 6 8 O-* This Work 12 14 10 16 18 Resolution (Bits) Figure 6-12: Figure-of-merit of this ADC compared with previous implementations (data courtesy B. Ginsburg, MIT). 126 Power of SAR ADCs VS Input Frequency 0 0 0 coo0 10-2 00 0 0 10-4 0- 0-0 This Work 0 10~6 10 104 102 108 Input Frequency (Hz) Figure 6-13: Power consumption of SAR ADCs with respect to input frequency (data courtesy B. Ginsburg, MIT). Power of SAR ADCs VS Resolution 0 00 10-2 8 @ a) 10~0- Work 0-This 10-6 [ 6 8 10 12 14 16 Resolution (Bits) Figure 6-14: Power consumption of SAR ADCs with respect to resolution (data courtesy B. Ginsburg, MIT). 127 6.5 Summary Table 6.2 summarizes the performance and features of the ADC. 8 Bit Mode 12 Bit Mode Process 0.18pm CMOS, National Semiconductor Area 900pm X 700pm Voltage Supply 1V Clock Frequency 4MHz 2MHz Resolution 8 Bits 12 Bits Maximum Sampling Rate 200kS/s 1 00kS/s Minimum Sampling Rate OS/s OS/s Power Consumption 1 9pW @ 100kS/s 25pW @ 100kS/s SNDR (at Nyquist) 49.7dB (fin=100Hz) 65.3dB (fin=50Hz) ENOB (at Nyquist) 7.96 Bits (fin=100Hz) 10.55 Bits (fin=50Hz) SFDR (at Nyquist) 63.2dB (fin=100Hz) 71dB (fin=50Hz) Table 6.2: ADC performance summary. 128 Chapter 7 Discussions and Future Work A Nyquist rate ADC, operating from a 1V supply, was presented. The SAR architecture, which allows for a mostly passive implementation, was leveraged to achieve micro-power operation. Additionally, the selected architecture enabled efficient power management, allowing the power consumption to scale linearly as the sampling rate varied between 0 and 1OOkS/s. For further power savings, 8 bit or 12 bit operating modes could be selected dynamically. Improvements in the efficiency of the converter were achieved by employing a variety of techniques. The low supply voltage significantly reduced the overall power consumption of the ADC. Although 1V, 12 bit designs have previously been demonstrated [44][45], achieving a high SNDR at these voltage levels has shown be exceedingly difficult. Specifically, the low voltage imposes severe restrictions on analog implementation. Most notably, in this design, the performance of analog switches was degraded, amplifier gain enhancement methods (such as cascoding) were restricted, and the dynamic range of noise-limited stages was reduced. The effect of these drawbacks was largely offset by a comparator implementation that relied on efficient regenerative amplification. To make this implementation viable, an offset calibrating circuit was developed. Further, the power consumption of the charge redistribution DAC was minimized by employing a fully capacitive structure that draws no static current. The elements were sized aggressively, but within matching limitations. Layout and transmission gain compensation techniques were required to minimize the 129 ensuing sensitivity to parasitics. The following sections analyze the benefits of the techniques applied and look ahead to improvements that can be made to enhance the ADC subsystem further. Many of the possibilities for future work focus on the ADC performance and features, however, others also consider optimizations at higher system level. 7.1 Effects of Applied optimizations Generally speaking, it's difficult to quantify the precise effect of the individual optimizations applied. Nonetheless, a rough analysis of the power improvements achieved as a result of the specific techniques can be considered. First, designing the ADC to operate from a 1V supply, instead of 1.8V, which is the nominal limit for this technology, has a predictable advantage. Specifically, the power consumption of all blocks reduces by at least a linear factor of 1.8. In the case of the digital state machine and the charge redistribution DAC, quadratic reduction is expected. With consideration to Table 6.1, the power consumption with a 1.8V supply would be approximately 65pW. Although an increased supply voltage can ease thermal noise limitations, only the first comparator preamplifier would benefit. As a result, this estimate of power savings from reduced supply voltage is reasonable. Second, without self-timed bit-cycling, the power consumption of the comparator preamplifiers increases due to stringent settling requirements. As mentioned in Section 5.2.1, 350ns has been allocated for preamplifier settling. The remainder of the clock period is consumed by logic delay, DAC settling, and latch resolution. Without self-timed bit-cycling, the preamplifiers would be required to settle in approximately 150ns. Accordingly, the power consumption of the cascade would increase by a factor of 350ns to 14pW. Third, the low offset regenerative latch reduces preamplifier power consumption by easing the linear gain requirement in the comparator. Without calibration, a latch offset of approximately 50mV can be expected. Since the simulated offset of the calibrated latch used here is approximately 1.5mV (as shown in Section 5.2.2), an 130 additional gain of roughly 30 is required. This can be achieved efficiently using a second, three stage cascade, identical to the existing preamplifiers. The additional contribution to the preamplifier chain's time-constant is dependant on fixed parasitic output capacitance, not noise limitations. Accordingly, the resulting total time constant is approximately 85ns, instead of 55ns in the original case. Since the number of stages has doubled, the preamplifier power consumption increases by a factor of 28 in order to restore the previous settling characteristic. In this case, the preamplifiers consume approximately 18paW. The cumulative benefit of all of these techniques can be estimated. In the case where the offset-canceling latch is replaced with linear gain stages, the increase in preamplifier power is countered somewhat by reduced latch offset requirements. For instance, simulations show that, at the desired speed, a non-calibrating, sense amplifier latch consumes a third the power of the offset-calibrating latch. Consequently, preamplifier power would increase to approximately 18puW, but latch power would decrease to 1.7puW. Without self-timed bit-cycling, the preamplifier power consumption would increase by an additional factor of 30,, to 42W. Finally, if the power supply were increased to 1.8V, the entire ADC would consume 124pW. Other techniques, such as sub-DAC transmission gain adjustment, which enables an aggressively sized capacitive DAC implementation, also provide power savings in this design. Nonetheless, based on the figure-of-merit of previous work, 124pW is inline with the power consumption expected for an ADC having the precision and speed of this design. 7.2 Future Work This section examines further optimizations that might improve the power efficiency, dynamic linearity, and scalability of the ADC. Additionally, the possibility of adding post-processing capability to ADC to improve system efficiency is considered. 131 7.2.1 Digital Optimization Although the digital standard cells were easily fast enough for this design, they did not result in the most power efficient implementation for the SAR state machine. Specifically, non-minimum sized devices were often used unnecessarily. Additionally, during the conversion process transitions in the digital blocks occur fairly infrequently, and in a highly deterministic manner. Consequently, the logic can be optimized to ensure minimum activity. 7.2.2 Programmable On-Chip Post-Filtering As mentioned in Section 1.1, the ADC may over-sample the input data to ease antialiasing filter requirements. This requires that the digital output be appropriately filtered and decimated before further processing. If the ADC is not integrated with the subsequent DSP, it is beneficial to implement the decimation filter on the ADC chip. This will minimize the bandwidth on the heavily loaded off-chip drivers. Of course, to maintain scalability in the effective sampling rate, the digital filter might also be programmable 7.2.3 Resolution Scalable DAC Section 4.2.2 discussed the challenges associated with achieving efficient scalability in the DAC. In this design, no special techniques were employed to overcome the parasitic loading imposed by MSB capacitors as the resolution is reduced. A fu- ture design might investigate de-coupling the inactive MSB capacitors using reliable, charge-pump boosted switches. 7.2.4 Active Input Switch Section 6.2.2 considered the distortion contributed by the sample-and-hold input switch. Based on simulation, it was determined that the linearity of the ADC would have benefited from some active biasing to reduce the input dependant variation in 132 its resistance. A variety of constant-VGS switch biasing techniques exist and may be employed in future versions of this design. 133 Appendix A ADC Fundamentals This chapter examines fundamentals of the analog-to-digital conversion process. Since modern ADC architectures and implementations make use of a plethora of new technologies and circuit techniques, it is worth examining the impact these have on the signal processing characteristics of the converter. This chapter serves to place the goals of the design efforts pursued, during the development of this ADC, into a system context. Section A.1 develops an ideal model of the ADC as a signal processing unit. Section A.2 analyzes the effect of practical non-idealities on that model. Finally, Section A.3 examines the physical constraints associated with ADC implementation. A.1 Linear Signal Processing ADCs perform a highly non-linear function. Converting a continuous-time (CT), analog signal into a discrete-time (DT), digital representation involves quantization and sampling. Nonetheless, it is valuable to model an ADC as a linear block, at least partially, so that well-understood, linear signal processing theory may be applied to data-conversion systems. The following sections develop that model. A characteristic voltage associated with ADCs, know as the LSB voltage, will frequently be referred to. This specifies the smallest voltage differences resolvable. Specifically, it is defined as !, where Vrej is the supported full-scale input voltage, and N is the number of 135 bits output by the ADC. A.1.1 Ideal ADC Model Figure A-1 shows an ideal model of an ADC. Here, V, represents the quantized signal, and Vet is the final digital output. The operation of quantization is modeled by the addition of an error signal, V, to the input signal, V,. In this sense, all ideal ADCs introduce error. Specifically, this error is called quantization error or quantization noise. Vn t. VIN Vout + X VXAL Impulse train to discrete sequence N n t t Figure A-1: Ideal model of an ADC. Quantization noise is explored further in the next subsection. However, notice, due to abrupt changes at the quantization boundaries, V has a very wide bandwidth. In particular, V, is the sum of Vi, and VqrL. While Vi may be bandlimited, V, is not. As a result, when the Vq, portion of V, is subsequently sampled, aliased components will fold into the baseband, defining a noise floor. A.1.2 Quantization Noise In general, the quantization noise signal, Vqri, is dependant on the input signal. Further, this dependence is highly non-linear. However, if the input signal varies rapidly, with a peak value larger than an LSB voltage, some approximations may be made regarding the quantization noise appearing at the ADC output. In particular, V , 136 can be treated as a random variable with an equal probability density between the values - LSB equal to , and LSB. In this case, the RMS value of the quantization error, Vqr, is [46]. As mentioned in Chapter 6, sinusoids are typically used to characterize ADCs. In such cases, the maximum signal-to-noise ratio (SNR) of the ADC is given by Equation A.1, where the amplitude of the sinusoid is SNR ref v'2 2V2oo(i~ 201ogl0 ( 2 N ~)(A. 1) From this relation, the SNR (in dB) is approximately equal to 6.02N + 1.76, where N is the resolution of the ADC. A.2 Non-Ideal ADC Model Practical ADCs introduce error in addition to quantization noise. The main categories of errors observed in an ADC are offset, gain-error, linearity error, and random noise. Offset and gain error are typically easy to correct, but they may effect the SNR slightly by degrading the full-scale input range. Linearity errors depend on the input, and as a result are not easy to correct. In the case of sinusoidal inputs, linearity errors manifest themselves as harmonic spurs in the frequency domain. The power of the spurs corresponds to the distortion introduced by the ADC, and it may added to the power of the other noise sources to quantify the signal-to-noise-plus-distortion ratio (SNDR), which expresses the total error. Random noise, typically originating in the constituent physical devices of analog circuits, does not manifest itself as spurious noise. Instead, it, along with quantization noise, contributes to the noise floor of the converter. The following subsections consider the effect of practical error sources on the ADC model. 137 A.2.1 Effective Resolution By applying a simple redefinition, the ideal ADC model developed earlier can account for practical errors. In particular, all errors present in the converter, in addition to inherent quantization, can be referenced back to the quantization noise signal. If the actual SNDR is known, the following relation, derived from Equation A.1, can be used to determine the effective number of bits (ENOB) of the ADC. ENOB = SNDR(dB) - 1.76 6.02 (A.2) (A2 As a result, non-idealities in a real ADC, degrade the effective resolution of the conversion beyond the finite resolution determined by the number of output bits. A.2.2 Random Noise in SAR ADCs Determining the SNDR of an ADC involves deriving the mean-squared error between its input signal (in digital representation) and its output signal. Thus, SNDR doesn't explicitly say anything about how the measured error appears, in terms of deviation of the actual ADC output code form the ideal digital output code. The effect of random noise on the output code of a SAR ADC is analyzed in [23]. Here, it is shown that, although random noise at the comparator input may be presumed to be Gaussian, the distribution of resulting output codes is not Gaussian. The probability of any code depends on the joint probability of each bit in that code. For example if the five LSBs of the ideal output code are 01000 (8), an output code of 10000 (16) may be more likely than 01111 (15), since it relies on an erroneous decision on only the 5 th bit rather than all of the 4 th to 1" bits. Although, it is true that an error on the 5 th bit is less likely than one on the lower bits, the overall probability may be higher. As a result, the likelihood of a particular code does not necessarily decrease monotonically from that of the ideal code. 138 A.3 Performance Normalization Three critical parameters associated with ADCs are sampling rate, resolution, and power consumption. This section examines the relationship between these, which Section A.3.1 examines the power cost of is rooted in inherent circuit tradeoffs. increasing the sampling rate of an ADC. Section A.3.2 examines the power cost of increasing resolution. Finally, Section A.3.3 uses the previous relations to rationalize a widely used figure-of-merit (FOM) for evaluating ADCs. A.3.1 Sampling Rate Generally speaking, a linear power-speed relationship exists in both digital and analog circuits. In the case of digital circuits, this can be seen in Equation A.3, where a is the switching activity factor, CL is the load capacitance, VDD is the supply voltage, fclk is the clock frequency, and d is the duty cycle (i.e. ratio of time the circuit is actively processing) [24]. Often, d is combined with a and may be considered a specific parameter associated with switching activity. Pdig = aCLVJDfclkd (A.3) Additional forms of power consumption, namely direct-path power and leakage power, are also present. However, these typically represent a smaller portion of the total power consumption. In the case of analog circuits, we can consider, as an example, a simple single stage amplifier. Here, the required gain, A, and bandwidth, f-MB, are given by the following expressions, where GM is the amplifier's transconductance, and R0 ,t is the amplifier's output resistance. A = GMRout 1 f-3B = I 27rRout CL 139 (A.4) (A.5) For a simple differential pair implementation, the amplifier's Gm coincides with the transconductance of the input devices, gm, current, bis, and can be expressed in terms of bias , electron charge, q, subthreshold slope, n, Boltzmann's constant, k, and absolute temperature, T (note, this relationship is true in the weak-inversion regime [18] where efficiency is highest). gM = Ibiasq nkT (A.6) Panaq nkTVDD _ Combining these results, the power consumption can be expressed as in Equation A.7, where its proportional dependence on signal frequency is shown. Pana = 2WfnkTVDDCLAf- 3 dB (A.7) q Although the results above were derived with consideration to specific examples, they, in fact, appear quite generally. The key point to infer, is that increasing the sampling rate of an ADC, which requires the constituent analog and digital circuits to process faster, inflicts a proportional increase in power. A.3.2 Resolution Resolution quantifies the loss in precision, due to noise, whether that noise be from quantization error or other non-idealities. Dynamic range is defined as the ratio of maximum signal power to the minimum signal power when the SNR has degraded to unity. Note that signal power, at unity SNR, is equal to the accumulated power of all noise sources. The ability of an ADC to resolve a higher number of bits is ultimately limited by intrinsic device noise in the analog circuits. To characterize the power-resolution relationship in analog circuits, the effect of thermal noise, on limiting dynamic range, can, once again, be analyzed for a simple one stage amplifier. Equation A.8 relates dynamic range, DR, to the power of a full-swing sine wave, and the noise power, v2. V 2 AMP 2' noV2 DR = - 2V20 140 (A.8) The value of vt2 is given by -y , where - is the equivalent number of noise sources in the amplifier and CL is the load capacitance. Then, by applying Equation A.4, Equation A.6, and Equation A.5, we get the following expression. DR = VmpPanaq 4lrn-y(kT) 2 AVDDf- 3 dB (A.9) This results shows that power consumption in analog circuits is directly proportional to dynamic range. Extrapolating to ADCs, where dynamic range is exponentially related to resolution (namely 2 N, where N is the number of bits), power consumption can be expected to vary exponentially with the number of bits to be resolved. A.3.3 Figure of Merit Based on the power-sampling rate and power-resolution relationships developed above, the power consumption of an ADC can be normalized. The figure-of-merit (FOM), shown in Equation A.10, was suggested to allow comparison of different ADC architectures and implementations. FOM= P(. 2fIN2ENOB Here, P is the power consumption of the converter, fIN 0 (A.10) is the input signal frequency, and ENOB is the effective resolution at that frequency. Note, fIN has been multiplied by two to relate the input signal frequency to the Nyquist rate. A comprehensive survey of ADCs, using a similar metric, confirms that the linear speed, and exponential resolution, normalizations are appropriate for comparing a broad range of designs [47]. 141 Bibliography [1] I. F. Akyildiz, Y. S. W. Su, and E. Cayirci, "Wireless sensor networks: A survey," Computer Networks, vol. 38, no. 4, pp. 393-422, March 2002. [2] R. Min, M. Bhardwaj, S.-H. Cho, E. Shih, A. Sinha, A. Wang, and A. Chndrakasan, "Low-power wireless sensor networks," white-paper on sensor network architectures. [3] T. Pering, T. Burd, and R. Brodersen, "The simulation and evalutation of dynamic voltage scaling algorithms," in Proc. Int. Symp. Low Power Electronics and Design, Aug. 1998, pp. 76-81. [4] Condenser Microphone Cartridges - Types 4133 to 4181, Bruel and Kjaer, Naerum, Denmark, product datasheet. [5] B. Olson, April 2005, e-Mail Correspondence. [6] K. Gulati and H.-S. Lee, "A low-power reconfigurable analog-to-digital converter," IEEE Journal of Solid-State Circuits, vol. 36, no. 12, pp. 1900-1911, Dec. 2001. [7] I. Ahmed and D. Johns, "A 50MS/s (35mw) to lkS/s (15pw) power scaleable pipeline 10b ADC with minimal bias current variation," in IEEE Int. Solid-State Circuits Conf. Dig. Tech. Papers,San Francisco, CA, Feb. 2005, pp. 280-281. [8] V. Peluso, P. Vancorenland, A. Marques, M. Steyaert, and W. Sansen, "A 900mV low-power AE A/D converter with 77dB dynamic range," IEEE Journal of Solid-State Circuits, vol. 33, no. 12, pp. 1887-1897, Dec. 1998. 143 [9] L. Yao, M. Steyaert, and W. Sansen, "A 1V 88dB 20kHz EA modulator in 90nm CMOS," in IEEE Int. Solid-State Circuits Conf. Dig. Tech. Papers, San Francisco, CA, Feb. 2004, pp. 80-81. [10] J. Sauerbrey, D. Schmitt-Landsiedel, and R. Thewes, "A 0.5-V 1-pW successive approximation ADC," IEEE Journal of Solid-State Circuits, vol. 38, no. 7, pp. 1261-1265, July 2003. [11] M. Scott, B. Boser, and K. Pister, "An ultralow-energy ADC for smart dust," IEEE Journal of Solid-State Circuits, vol. 38, no. 7, pp. 1123-1129, July 2003. [12] B. J. Sheu and C. Hu, "Switch-induced error voltage on a switch capacitor," IEEE Journal of Solid-State Circuits,vol. SC-19, no. 4, pp. 519-525, Aug. 1984. [13] S. Aghtar, J. W. Haslett, and F. N. Trofimenkoff, "Subthreshold analysis of an MOS analog switch," IEEE Transactions on Electron Devices, vol. 44, no. 1, pp. 89-96, Jan. 1997. [14] M. J. M. Pelgrom, A. C. J. Duinmaijer, and A. P. G. Welbers, "Matching properties of MOS transistors," IEEE Journal of Solid-State Circuits, vol. 24, no. 5, pp. 1433-1439, Oct. 1989. [15] Digital Offset CalibrationLoop, Exar Corporation, March 2000, application Note. [16] ADuC812 ADC Software Calibration,Analog Devices, Jan. 2000, technical Note. [17] R. Sarpeshkar, "Low power analog VLSI," MIT, Cambridge, MA, Sept. 2004, course notes. [18] Y. Tsividis, Operation and Modeling of the MOS transistor,2nd ed. McGraw Hill, 1987. [19] T. L. Tewksbury, H.-S. Lee, and G. A. Miller, "The effects of oxide traps on the large-signal transient response of analog MOS circuits," IEEE Journal of Solid-State Circuits, vol. 24, no. 2, pp. 542-544, April 1989. 144 [20] J. L. McCreary and P. R. Gray, "All-MOS charge redistribution analog-to-digital conversion techniues - part I," IEEE Journal of Solid-State Circuits, vol. SC-10, no. 6, pp. 371-379, Dec. 1975. [21] R. K. Hester, K.-S. Tan, M. D. Wit, J. W. Fattaruso, S. Kiriaki, and J. R. Hellums, "Fully differential ADC with rail-to-rail common-mode range and nonlinear capacitor compensation," IEEE Journal of Solid-State Circuits, vol. 25, no. 1, pp. 173-183, Feb. 1990. [22] B. Razavi, Data Conversion System Design. IEEE Press, 1995. [23] P. W. Yee, "Noise considerations in high-accuracy A/D converters," IEEE Jour- nal of Solid-State Circuits, vol. SC-21, no. 6, pp. 1011-1015, Dec. 1986. [24] J. Rabaey, A. P. Chandrakasan, and B. Nikolic, Digital Integrated Circuits, 2nd ed. Prentice Hall, 2003. [25] G. Promitzer, "12-bit low-power fully differential noncalibrating successive approximation ADC with lms/s," IEEE Journal of Solid-State Circuits, vol. 36, no. 7, pp. 1138-1143, July 2001. [26] J.-T. Wu and B. A. Wooley, "A 100MHz pipelined CMOS comparator," IEEE Journal of Solid-State Circuits, vol. 23, no. 6, pp. 1379-1385, Dec. 1988. [27] H.-S. Lee, "Design of analog MOS LSI," MIT, Cambridge, MA, Sept. 2003, course notes. [28] B. Razavi and B. A. Wooley, "Design techniques for high-speed, high-resolution comparators," IEEE Journal of Solid-State Circuits, vol. 27, no. 12, pp. 19161926, Dec. 1992. [29] C. C. Enz and G. C. Temes, "Ciruit techniques for reducing the effects of op-amp imperfections: Autozeroing, correlated double sampling, and chopper stabiliza- tion," Proc. of the IEEE, vol. 84, no. 11, pp. 1584-1614, Nov. 1996. 145 [30] C. C. Enz, E. A. Vittoz, and F. Krummenacher, "A CMOS chopper amplifier," IEEE Journal of Solid-State Circuits, vol. SC-22, no. 3, pp. 335-342, June 1987. [31] Y. P. Tsividis, P. R. Gray, D. A. Hodges, and J. Chacko, "A segmented p-255 law PCM voice encoder utilizing NMOS technology," IEEE Journal of SolidState Circuits,vol. SC-11, no. 6, pp. 740-747, Dec. 1976. [32] B. Fotouhi and D. A. Hodges, "High-resolution A/D conversion in MOS/LSI," IEEE Journal of Solid-State Circuits, vol. SC-14, no. 6, pp. 920-926, Dec. 1979. [33] K. B. Ohri and M. J. Callahan, "Integrated PCM codec," IEEE Journal of Solid-State Circuits, vol. SC-14, no. 1, pp. 38-46, Feb. 1979. [34] H.-S. Lee, D. A. Hodges, and P. R. Gray, "A self-calibrating 15 bit CMOS A/D converter," IEEE Journal of Solid-State Circuits, vol. SC-19, no. 6, pp. 813-819, Dec. 1984. [35] A. Hastings, The Art of Analog Layout. Prentice Hall, 2001. [36] E. A. Vittoz, Design of Analog-Digital VLSI Circuits for Telecommunications. Prentice Hall, 1994, ch. Micropower Techniques. [37] P. J. Lim and B. A. Wooley, "An 8-bit 200-MHz BiCMOS comparator," IEEE Journal of Solid-State Circuits,vol. 25, no. 1, pp. 192-199, Feb. 1990. [38] T. B. Cho and P. R. Gray, "A 10-bit, 20-MS/s, 35-mW pipeline A/D converter," in Proc. IEEE Custom Integrated Circuits Conference, San Diego, CA, May 1994, pp. 499-502. [39] J. F. Dickson, "On-chip high voltage generation in NMOS integrated circuits using an improved voltage multiplier technique," IEEE Journal of Solid-State Circuits,vol. SC-11, no. 3, pp. 374-378, June 1976. [40] M. Matsui, H. Hara, Y. Uetani, L.-S. Kim, T. Nagamatsu, Y. Watanabe, A. Chiba, K. Matsuda, and T. Sakurai, "A 200 MHz 13mm 2 2-D DCT macro146 cell using sense-amplifier pipeline flip-flop scheme," IEEE Journal of Solid-State Circuits, vol. 29, no. 12, pp. 1482-1490, Dec. 1994. [41] Audio PrecisionSystem One Operator's Manual, Audio Precision. [42] "IEEE standard for terminology and test methods for analog-to-digital convert- ers," IEEE Std 1241-2000. [43] J. Doernberg, H.-S. Lee, and D. A. Hodges, "Full-speed testing of A/D convert- ers," IEEE Journal of Solid-State Circuits, vol. SC-19, no. 6, pp. 820-827, Dec. 1984. [44] J. Li, G.-C. Ahn, D.-Y. Chang, and U.-K. Moon, "A 0.9-V 12-mW 5-MSPS algorithmic ADC with 77-dB SFDR," IEEE Journal of Solid-State Circuits, vol. 40, no. 4, pp. 960-969, April 2005. [45] D.-Y. Chang, G.-C. Ahn, and U.-K. Moon, "A 0.9V 9mW 1MSPS digitally calibrated ADC with 75dB SFDR," in Proc. IEEE Symp. VLSI Circuits, June 2003, pp. 67-70. [46] D. Johns and K. Martin, Analog Integrated Circuit Design. John Wiley and Sons, Inc., 1997. [47] R. H. Walden, "Analog-to-digital converter survey and analysis," IEEE Journal on Selected Areas in Communications, vol. 17, no. 4, pp. 539-550, April 1999. 147