<<

AN ABSTRACT OF THE THESIS OF

Hirokazu Yoshizawa for the degree of Doctor of Philosophy in Electrical and Computer

Engineering presented on May 15, 1997

Title: High-Linearity Switched- Circuits in Digital CMOS Technologies

Abstract approved: Redacted for Privacy Gabor C. Temes

In this thesis, novel design techniques have been proposed for implementing high- linearity SC circuits in a standard digital CMOS process. They use nonlinear MOSFET instead of linear double-poly capacitors. To reduce their nonlinearities, a bias voltage is applied to keep MOSFET capacitors in their accumulation regions. For further reduction of distortion, two capacitors can be connected in series or in parallel so that a first-order cancellation of the nonlinearity can be achieved. Experimental results demonstrated that the among these techniques series compensation is the most effective for reducing the nonlinearity of MOSFET capacitors. A novel predictive SC has been proposed for its insensitivity to op-amp imperfections. Experimental results show that the S/THD of the predictive SC amplifier was 10 dB larger than that of the non-predictive one. It was also shown that a predictive circuit was effective for reducing the nonlinearity caused by the op-amp and/or the

MOSFET capacitors. It has been demonstrated that a two-stage op-amp with a large output swing can be fabricated in a standard digital CMOS process. The frequency compensation was accomplished using a source follower and a MOSFET capacitor. An SC amplifier using this two-stage op-amp and double-poly capacitors was fabricated, and it exhibited a large linear output range. A MOSFET-only digitally controlled gain/loss circuit was designed and fabricated in a 1.2 f_tm CMOS process. It demonstrated that the series compensation is effective not only for a large output swing in an amplifier, but also for a large input swing in an attenuator. A pipeline D/A converter utilizing MOSFET capacitors was designed as another application of charge processing technique. It consisted of three parts: a V-Q conversion stage, a charge transfer stage, and a Q-V converter.

A new configuration which enables the series compensation to have a large bias voltage has also been proposed. It was shown that it works well, and it will be helpful for low-voltage operation, too. © Copyright by Hirokazu Yoshizawa May 15, 1997 All Rights Reserved High-Linearity Switched-Capacitor Circuits in Digital CMOS Technologies

by

Hirokazu Yoshizawa

A THESIS

submitted to

Oregon State University

in partial fulfillment of

the requirements for the degree of

Doctor of Philosophy

Completed May 15, 1997

Commencement June 1998 Doctor of Philosophy thesis of Hirokazu Yoshizawa presented on May 15, 1997

Approved:

Redacted for Privacy

Major Pre es,sor, representing Electrical and Computer Engineering Redacted for Privacy

Chair of DepOment of Electrical and Computer Engineering

Redacted for Privacy

Dean of Graduat chool

I understand that my thesis will become part of the permanent collection of Oregon State University libraries. My signature below authorizes release of my thesis to any reader upon request.

Redacted for Privacy

irokazu Yoshizawa, Author ACKNOWLEDGMENT

I wish to express my sincere gratitude to Professor Gabor C. Temes, for his advice, discussions, patience, support, and encouragement . I learned lots of technical things from him, and also his great personality gave a big influence to my life. I am grateful to Professor Jack Kenney and Professor Richard Schreier for technical discussions and the review of the dissertation. I would like to thank to Professor Shih-Lien Lu for help with the fabrication process and for the review of the dissertation. I thank to Professor Kiaei for serving as a committe member. I also thank to Professor Vikram Saletore of the Computer

Science Department for serving as a graduate respresentative. I would like to thank to Orbit Semiconductor Inc. for the courtesy of fabrication of chips. I acknowledge the support of Dr. George Lewicky. I would like to express my special thanks to Paul Ferguson of Analog Devices Inc. for helpful discussions and suggestions and sharing his knowledge on practical experimental issues. My fellow graduate students in the analog IC group have been very helpful. Special thanks are due to Ayse Yesilyurt for sharing her folded-cascode op-amp; Yunteng Huang and Yuming Cao for helpful discussions; Charlie Chairatanatrai for layout support. Special thanks go to Seiko Instruments Inc. for supporting me under their sponsorship program. I would like to express my deep thanks to Mr. Yasunori Ebihara for giving me this great chance to study in the U.S. I would also like to thank to Mr. Kiyohiko Kato, Mr. Hiroichi Teruyama, Dr. Masaaki Kamiya, and Mr. Kenichi Kobayashi for their support. Finally, I would like to give deep thanks to my parents for their love and support. TABLE OF CONTENTS

Page

1. INTRODUCTION 1

1.1 Introduction 1

1.2 Thesis Outline 3

2. CAPACITORS IN CMOS TECHNOLOGY 5

2.1 Capacitors in Analog CMOS Technology 5

2.2 Capacitors in Digital CMOS Technology 8

2.3 MOSFET Capacitor 11

2.3.1 Gate Structures 11 2.3.2 C-V Characteristic 13 2.3.3 Parasitics 22 2.3.4 HSPICE Models 26 2.3.4.1 MOSFET Capacitor Model 27 2.3.4.2 PN Junction Model 30 2.3.5 Accuracy Considerations 30 2.3.6 Layout 34

3. EARLIER TECHNIQUES FOR LINEAR ANALOG SIGNAL PROCESSING IN DIGITAL CMOS TECHNOLOGY 37

3.1 Simple Biasing 37

3.2 Charge Processing 38

3.3 Switched-Current Circuits 40

3.4 Sample-and-Hold Circuits 41 iii

TABLE OF CONTENTS (Continued)

Page

4. MOSFET-ONLY SWITCHED-CAPACITOR USING NONLINEARITY COMPENSATION TECHNIQUES 43

4.1 Early Work and Applications 43

4.2 Simple Biasing Technique 43

4.3 Parallel Compensation 46

4.4 Series Compensation 49

4.5 Simulation Results 57

4.5.1 Simple Biasing 57 4.5.2 Parallel Compensation 62 4.5.3 Series Compensation 64

5. PREDICTIVE CDS TECHNIQUE 68

5.1 Background of the Predictive CDS Technique 68

5.2 Novel Predictive SC Amplifier 71

5.3 MOSFET-Only Predictive SC amplifier 77 5.3.1 Parallel Compensation 77 5.3.2 Series Compensation 77

5.4 Simulation Results 78

6. EXPERIMENTAL RESULTS FOR SWITCHED-CAPACITOR AMPLIFIERS 86

6.1 Test Circuit 86

6.2 Op-Amp 89

6.3 Clock Generator 91 iv

TABLE OF CONTENTS (Continued)

Page

6.4 Experimental Results 94 6.4.1 Non-Predictive SC Amplifier with Simple Biasing 94 6.4.2 Non-Predictive SC Amplifier with Series Compensation 102 6.4.3 Non-Predictive SC Amplifier Using Double-Poly Capacitors....102 6.4.4 Non-Predictive SC Amplifier with Parallel Compensation 102 6.4.5 Predictive SC Amplifier with Parallel Compensation 105 6.4.6 Output Swing (Inverting Mode) 106 6.4.7 Frequency Dependence (Inverting Mode) 107 6.4.8 Non-inverting Mode 108 6.4.9 Continuous Output Waveform without Clock Signal 111 6.4.10 1-1-4 Analysis Using an A/D Converter and NEXT Workstation 114 6.4.11 Noise 119

7. OPERATIONAL AMPLIFIERS IN DIGITAL CMOS TECHNOLOGY 120

7.1 Earlier Work 120

7.2 Folded-Cascode Op-Amp 120

7.3 Two-Stage Op-Amp 125

7.4 Simulation Results 131

7.5 Layout and Post-Layout Simulation 131

8. DIGITALLY CONTROLLED GAIN/LOSS CIRCUIT 134

8.1 Research Background 134

8.2 Circuit Design 136

8.3 Simulation Results 143 V

TABLE OF CONTENTS (Continued)

Page

8.4 Experimental Results 145

8.4.1 SC Amplifier Using Double-Poly Capacitors 145 8.4.2 MOSFET-Only Gain/Loss Circuit (Amplifier Mode) 147 8.4.3 MOSFET-Only Gain/Loss Circuit (Attenuator Mode) 151 8.4.4 Linearity Dependence on the Bias Voltage 152

9. PIPELINE DIGITAL-TO-ANALOG CONVERTER 163

9.1 Background of Pipeline D/A Converter 163

9.2 Circuit Design 165

9.3 Simulation Results 169

10. COMPENSATION IN DEEP SUBMICRON AND LOW-VOLTAGE PROCESS 173

10.1 Series Compensation in Low-Voltage Operation 173

10.2 New Switch Configuration for Series Compensation 176

10.3 Simulation Results 177

11. CONCLUSIONS AND SUGGESTED FUTURE WORK 179

11.1 Conclusions 179

11.2 Suggested Future Work: SC Integrators and Filters 181

BIBLIOGRAPHY 189

APPENDIX 192

SPICE Parameters for the Simulations Performed in this Thesis 193 vi

LIST OF FIGURES

Figure Page

2.1 Double-poly capacitor structure 6

2.2 Poly-diffusion MOS capacitor structure 7 2.3 Metal-metal capacitor 9 2.4 Metal-poly capacitor 9

2.5 MOSFET capacitors 12

2.6 C-V curve measurement of MOSFET capacitors 14

2.7 Measured C-V curves of MOSFET capacitors 15

2.8 Space-charge layers 18 2.9 Charges at the interface of Si02 and single-crystal silicon 20 2.10 C-V curve shift due to the positive fixed oxide charge 21

2.11 Parasitic capacitances of MOSFET capacitor 23 2.12 MOSFET unit capacitor 24 2.13 Profile of MOSFET capacitor 26 2.14 Simulated C-V curves of PMOSFET for different CAPOP values 29 2.15 Simulated C-V curves for CAPOP=2 31 2.16 Simulated C-V curve of the pn junction diode 32

2.17 Unit capacitors with common-centroid geometry 33

2.18 Layout of unit MOSFET capacitor 35

3.1 Lossy integrator using simple biasing technique 38 3.2 Charge processing 39 3.3 Current processing 40 3.4 Simple S/H circuit 42 3.5 Improved S/H circuit 42

4.1 Gain- and offset-compensated SC amplifier 44 4.2 SC amplifier using simple biasing arrangement 45 vii

LIST OF FIGURES (Continued)

Figure Page 4.3 Conceptual structure of the parallel compensation 46 4.4 Practical realization of the parallel compensation 47 4.5 SC amplifier using the parallel compensation 48 4.6 Series compensation SC branch 50

4.7 Capacitance linearization 51 4.8 SC amplifier using series compensation 54 4.9 Practical realization of a PMOS switch 55 4.10 Nonlinear parasitic capacitances of the seires SC branch 56

4.11 Another possible structure of the series SC branch 57 4.12 The output of the circuit of Figure 4.2 with Vin = 200 mV and Vb = 0 V 59 4.13 The output of the circuit of Figure 4.2 with Vin = 200 mV and Vb = 1.5 V 60

4.14 The output of the circuit of Figure 4.2 with Vin = 400 mV and Vb = 1.5 V 61 4.15 The output spectrum of the circuit of Figure 4.5 63 4.16 The output spectrum of the circuit of Figure 4.8 65

4.17 SC amplifier using the bottom-to-bottom-connected series compensation 66 4.18 The output spectrum of the circuit of Figure 4.17 with Vin = 200 mV 67

5.1 Predictive SC amplifier 70 5.2 Signal waveform of the predictive SC amplifier 70 5.3 Novel predictive SC amplifiers 72 5.4 Signal waveform of the predictive SC amplifiers for a DC input 74 5.5 Gain-uncompensated non-predictive SC amplifier 75 5.6 Gain error of SC amplifiers 76 5.7 Predictive SC amplifier using parallel compensation 79

5.8 Predictive SC amplifier using parallel compensation based on the circuit of Figure 5.1. 80 LIST OF FIGURES (Continued)

Figure Page

5.9 Predictive SC amplifier using parallel compensation based on the circuit of Figure 5.3a 81 5.10 Predictive SC amplifier using MOSFET capacitors with series compensation 82

5.11 The output spectrum of the circuit 83

5.12 The output spectrum of the circuit 85

6.1 Test circuit for SC amplifiers 87

6.2 Differential-to-single-ended converter 88

6.3 Folded-cascode op-amp 90 6.4 Common-mode feedback circuit using MOSFET capacitors 91 6.5 Clock generator circuit 92

6.6 Layout of the clock generator circuit 93 6.7 SC amplifiers (1) (3) 95

6.8 SC amplifiers (2) (4) 97

6.9 SC amplifiers (5) (6) 99

6.10 The measured output spectra of the SC amplifier using the simple biasing 101 6.11 The measured output spectrum of the SC amplifier using the series compensation with Vb=1.5 V 103

6.12 The output waveform of the SC amplifier using the series compensation with Vb=1.5 V 103

6.13 The measured output spectrum of the SC amplifier using double-poly capacitors 104 6.14 The measured output spectrum of the SC amplifier using the parallel compensation with Vbp=-Vbn=2.5 V 104

6.15 The measured output spectrum of the predictive SC amplifier using the parallel compensation with Vbp=-Vbn=2-5 V 105

6.16 The output waveform of the non-inverting SC amplifier using double-poly capacitors 110 6.17 The output waveform of the non-inverting SC amplifier using series compensation 111 ix

LIST OF FIGURES (Continued)

Figure Page

6.18 SC amplifier using the parallel compensation at (1)2=1 112

6.19 SC amplifier using the series compensation at 02=1 113

6.20 Data acquisition using an A/D converter 115

6.21 Output waveform and timing of sampling 117

7.1 A digital-process compatible op-amp 121 7.2 HSPICE simulation results of the SC amplifier 123 7.3 Folded-cascode op-amp using MOSFET capacitors with series compensation 124 7.4 Op-amp compensation (1) 126 7.5 Op-amp compensation (2) 126 7.6 Op-amp compensation (3) 128 7.7 Practical realization of Figure 7.6 128 7.8 Fully differential two-stage op-amp using MOSFET capacitors 129 7.9 Bode plot of the two-stage op-amp 130 7.10 HSPICE simulations of the SC amplifier by various kinds of op-amp 132

7.11 Layout of the two-stage op-amp 133

8.1 Digitally controld attenuator 135 8.2 Digitally controlled gain/loss circuit 135 8.3 One-bit gain/loss stage 137

8.4 MOSFET-only digitally controlled gain/loss circuit 138

8.5 Control circuit 139

8.6 Digitally controlled gain/loss circuit 141

8.7 Post-layout simulation results by HSPICE 144

8.8 Die photo of the digitally-controlled gain/loss circuit 145 8.9 The measured output spectrum of the SC amplifier using double-poly capacitors 146 x

LIST OF FIGURES (Continued)

Figure Page

8.10 S/THD ratios of the MOSFET-only amplifier 148

8.11 The measured output spectrum of the MOSFET-only SC amplifier with a gain of 10 149

8.12 The measured output spectrum of the MOSFET-only SC amplifier with a gain of unity 150 8.13 S/THD ratios of the MOSFET-only attenuator 152

8.14 Linearity dependence on bias voltage (Vout = 4 Vp-p) 153

8.15 The measured output spectrum of the MOSFET-only SC amplifier with Vb = 2.0 V 154

8.16 Linearity dependence on bias voltage (Vout = 2 Vp-p) 155

8.17 S/THD ratios for the MOSFET-only amplifier with Vb = 2.0 V 156 8.18 Linearity dependence on the bias voltage for the MOSFET gain circuit (Vout = 4 Vp-p) 157

8.19 Measured output spectra in amplifier mode: Vout = 4 Vp-p; Vb = 1.5 V 159 8.20 Measured output spectra in amplifier mode: Vout = 4 Vp-p; Vb = 2.0 V 160

8.21 Measured output spectra in attenuator mode: Vin = 4 Vp-p; Vb = 1.5 V 161 8.22 Measured output spectra in attenuator mode: Vin = 4 Vp-p; Vb = 2.0 V 162 9.1 10-bit pipeline DAC 164 9.2 The concept of charge processing 165 9.3 Charge-voltage converter 166 9.4 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.3 167

9.5 Pipeline D/A converter using an improved charge-voltage converter 168

9.6 Pipeline D/A converter using an alternative charge-voltage converter 170 9.7 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.5 171

9.8 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.6 171 xi

LIST OF FIGURES (Continued)

Figure Page 9.9 Integral nonlinearity (INL) of the 10-bit pipeline D/A converter using the Q-V converter of Figure 9.5 172 9.10 Integral nonlinearity (INL) of the 10-bit pipeline D/A converter using the Q-V converter of Figure 9.6 172

10.1 Switch configuration in series compensation 174

10.2 New switch configuration in series compensation 175

10.3 The output spectrum of the circuit of Figure 4.8 with VDD=-Vss=Vb=1.5 V 178

11.1 SC integrator using parallel compensation 182

11.2 SC integrator using a simple biasing arrangement 182

11.3 SC integrator using series compensation 183

11.4 SC integrator using series compensation 184

11.5 Realization of a high-Q biquad 185

11.6 Alternative realization of a high-Q biquad 187

11.7 An input branch combined with a voltage-to-charge converter using series compensation 188 xii

LIST OF TABLES

Table Page

1-1 Oxide thickness and equivalent capacitor size 2

2-1 Voltage coefficients of MOS capacitors 8

2-2 Capacitors in CMOS processes 10

2-3 Measured voltage coefficients of MOSFET capacitor structures 17

2-4 Voltage coefficients of MOSFET capacitor structures 22

2-5 Voltage coefficients of the HSPICE PMOSFET capacitor model 28

2-6 Voltage coefficients of the HSPICE NMOSFET capacitor model 28

5-1 Gain errors of SC amplifiers 75

5-2 Summary of the key features of the three predictive amplifiers 77

6-1 SC amplifiers on chip 94

6-2 Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz 106

6-3 Input signal frequency (fin): 100 Hz; clock frequency (fs): 1.6 kHz 108

6-4 Input signal frequency (fin): 100 Hz; clock frequency (fs): 1.6 kHz 109

6-5 Input signal frequency (fin): 10 kHz 112

6-6 Sinusoidal signal from Tektronix SG5010: input signal frequency (fin) was 1 kHz 116

6-7 Inverting mode sampled at the end of 02; Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz 118

6-8 Non-inverting mode sampled at the end of 4)2; Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz 118

7-1 Characteristics of the op-amps 130

7-2 Comparison of the SC amplifier performances with various op-amps 131

8-1 S/THD ratios of the MOSFET-only gain circuit 148

8-2 S/THD ratios of MOSFET-only loss circuit 152

8-3 S/THD ratios of the MOSFET-only gain circuit with Vb=1.5 V and 2.0V (Vout = 4 Vp-p) 153 LIST OF TABLES (Continued)

Table Page

8-4 S/THD ratios for the MOSFET-only gain circuit with Vb=1.5 V and 2.0 V (Vout = 2 Vp-p) 155

8-5 Linearity dependence of the MOSFET-only gain circuit on the bias voltage (S/THD ratio at Vout = 4 Vp-p) 157

8-6 S/THD ratios for SC amplifiers with double-poly capacitors and with MOSFET capacitors 158 xiv

To my parents High-Linearity Switched-Capacitor Circuits in Digital CMOS Technologies

Chapter 1 INTRODUCTION

1.1 Introduction

There is an increasing demand for combining analog and digital integrated circuits on a single chip. Because of their great capabilities for dense integration and low power consumption, complementary metal-oxide-semiconductor (CMOS) technologies are dominantly used in digital circuits and memory devices. To integrate analog circuits on the same chip with digital ones, it is preferable to realize the analog circuits also by the CMOS technologies. For realizing CMOS analog circuit functions such as amplifiers, filters, and data converters, switched-capacitor (SC) circuits have been the principal technique for

nearly two decades.

When analog and digital circuits are integrated on the same chip, an extra processing step creating a second polysilicon (poly) layer is usually added to fabricate linear double- poly capacitors for the analog circuit. This represents a significant added cost, even though

the chip area required for the analog circuitry is usually much smaller than that of the digital part. In addition, double-poly capacitors are not available in some inexpensive digital CMOS processes. In such a process, metal-1 to poly or metal-1 to metal-2 capacitors are often used to implement linear capacitors. However, these capacitors require a large chip area, because their Si-oxide layers are very thick. Also, their matching properties are inferior due to the rough surface of their bottom plates and the poor uniformity of the chemical-vapor-deposition (CVD) oxide [1], [2]. In addition, these capacitors have large 2 parasitic capacitance to the substrate, which affects the signal processing, and increases the dc power loss. The other possibility is to use MOSFET capacitors*. These have a large capacitance per unit area because the gate oxide is thinner than the oxide layer of any other capacitors in the CMOS process. As an example, for a typical 1.2 pm CMOS process the properties of several kinds of capacitors are shown in Table 1-1.

Table 1-1 Oxide thickness and equivalent capacitor size

Thickness Capacitance Equivalent size of oxide per unit area [Angstrom] [fF/ptm2]

MOSFET capacitor 225 1.53 1

Double-poly capacitor 700 0.49 3.1 Ml-poly capacitor 5200 0.066 23

M1 -M2 capacitor 7500 0.046 33

In the process considered in Table 1-1, a MOSFET capacitor requires a chip area which is less than 1/3 of that needed by a double-poly capacitor with the same capacitance and a small fraction of the size of equivalent metal-poly or metal-metal capacitors. In other processes the area saving is even larger. The MOSFET capacitor also offers good matching accuracy, due to the uniformity of its thermally-grown gate oxide and the smooth surface of its single-crystal silicon bottom plate [1]. Finally, MOSFET capacitors are available in any CMOS process because they are intrinsic elements of MOS technology. Nevertheless, they

normally cannot be directly used in SC circuits because of their strong voltage dependence.

* Throughout this thesis,"MOSFET capacitor" is used to refer to a poly-gate-to-channel capacitor without a heavily-doped diffusion layer under the gate oxide. By contrast, the term "MOS capacitor" is used to refer to any capacitor realized in metal-oxide-semiconductor technology such as a double-poly, metal-poly, poly- diffusion, and MOSFET capacitor. 3

In this thesis, circuit design techniques are described for linearizing the operation of CMOS analog ICs using MOSFET capacitors in digital CMOS processes. The usefulness of these techniques will be illustrated by applying them to some common SC building blocks.

1.2 Thesis Outline

Chapter 2 of this thesis explores capacitors in CMOS technology. It begins with a description of capacitors available in analog and digital CMOS processes. Then, MOSFET capacitors are introduced. The characteristics of the MOSFET capacitor will be described such as its C-V curve, parasitics, HSPICE models. A practical layouts will be also shown. Chapter 3 describes the background of this research work. Simple biasing and charge processing techniques using MOSFET capacitors will be introduced. Next, switched-current circuits will be described as a signal-processing technique which does not require high-linearity capacitors. Finally, sample-and-hold circuits are introduced as an example illustrating the feasibility of high-precision SC circuits using nonlinear capacitors. Chapter 4 shows design techniques for MOSFET-only switched-capacitor circuits. Three design techniques will be described for reducing the nonlinearity of the MOSFET capacitors. Chapter 5 introduces the predictive correlated-double-sampling (CDS) technique, and a new predictive SC amplifier will be described. Finally, the effects of CDS on the circuit nonidealities will be discussed. Chapter 6 provides experimental results obtained for the switched-capacitor amplifiers. A test circuit and a clock generator circuit will be shown. Finally, the measured

results will be discussed. Chapter 7 describes the circuitry of some operational amplifiers suitable for digital

CMOS processes. 4

Chapter 8 introduces a digitally controlled gain/loss circuit using MOSFET capacitors. First, an overview of some relevant earlier work will be given. Then the circuit design and experimental results will be shown. Chapter 9 discusses a MOSFET-only pipeline digital-to-analog converter. The design and simulation results will be given.

Chapter 10 describes a compensation technique for deep submicron and low-voltage process. New switch configurations suitable for low-voltage processes are proposed.

Finally, Chapter 11 gives the conclusions of this thesis and suggests future research work. 5

Chapter 2 CAPACITORS IN CMOS TECHNOLOGY

In this chapter, the various capacitor realizations available in both analog and digital CMOS processes are described. Then, MOSFET capacitors are introduced. The characteristics and models of MOSFET capacitor will be described, such as C-V curves, parasitics, and HSPICE models. Finally, the practical layout of the MOSFET capacitor will be also shown.

2.1 Capacitors in Analog CMOS Technology

The double-poly capacitor is most widely used for realizing high-linearity SC circuits because of its good linearity properties. Practical realization of the double-poly capacitor is illustrated in Figure 2.1a. Both poly layers are highly doped, and one surface

accumulates while the other depletes so that its voltage dependence is greatly reduced [1]. The total capacitance Ct is modeled in Figure 2.1b, and is given by the series combination

of the oxide and space-charge capacitances:

-1 = 1_ + 1-+ 1 (2-1) Ct Cox Cs1 Cs2

This capacitor has a relatively small parasitic capacitance (about 15 % of the main capacitor)

associated with its bottom plate because the field oxide between the lower poly layer and the

single-crystal silicon substrate is much thicker than the poly-poly oxide. In some CMOS processes, a capacitor with thermally grown oxide on heavily- doped single-crystal silicon is available as shown in Figure 2.2a. This capacitor has good matching properties and relatively small voltage coefficient for high doping density in the

bottom plate [1]. For metal-gate processes, the heavily-doped diffusion layer under the gate

oxide can be realized without an extra process step because the source/drain diffusions are 6

Metal

Poly Si

Si02 (field oxide)

p- substrate

O

Poly Si Cs2 Ct Si02 Cox

CS1 Poly Si

(b)(b)

Figure 2.1 Double-poly capacitor structure: (a) practical realization; (b) model with two space-charge capacitances [1]. 7

Metal

Si02 Poly Si

Si02 Si02 Si02 (thermally grown (Field oxide) (Field oxide) n+

p- substrate

Figure 2.2 Poly-diffusion MOS capacitor structure.

formed before fabricating the metal gate. By contrast, for the self-aligned Si-gate process, which is the current standard for CMOS process, additional process steps are needed for the formation of the heavily-doped diffusion layer under the poly gate. Again, the total capacitance Ct is given by equation 2-1. Notice that both the poly layer and the single-crystal silicon layer must be equally heavily doped to achieve high linearity. If the single-crystal silicon layer, for example, is more lightly doped than the poly layer, then this layer provides the dominant space charge capacitance (Cst in Figure 2.1b) [1]. In this case, the total capacitance decreases with decreasing bias voltage because the crystalline silicon layer is depleting, its space-charge layer is widening, and consequently

Cst decreases. 8

As the gate oxide becomes thinner in advanced CMOS processes, it becomes more difficult to achieve a low voltage coefficient because the higher-dose capacitor implant associated with the use of thinner oxides can lead to a reduced dielectric breakdown [3]. Table 2-1 shows an example of voltage coefficients of poly/heavily-doped single- crystal silicon capacitors.

Table 2-1 Voltage coefficients of MOS capacitors

Nd Oxide thickness MOS structure ay [Angstrom] [ppm/V] [cm-3]

McCreary [1] 7 1.1e20 > 1000 * poly /SiO2 /n+ 340 5.5e18 > 1000* poly /SiO2/n+ (Figure 2.2a)

Slater, 70 2e20 225 silicide/poly/Si02/n+ Pau los [3] 590 4e19 225 silicide/poly/Si02/n+

ay: the first-order voltage coefficient. Nd: impurity concentration in heavily-doped silicon.

2.2 Capacitors in Digital CMOS Technology

Several realization for capacitors are available in digital CMOS technologies. However, most of these require large chip area, and have large parasitics and poor

matching properties. The metal-metal capacitor (shown in Figure 2.3) has a low voltage dependence

because the metal plate surfaces do not accumulate or deplete. However, the bottom plate, which is aluminum in most cases, has a low melting point, and that mandates the use of low-temperature CVD oxide, which is known to be inferior compared to the thermally-

grown oxide in terms of thickness uniformity, dielectric strength, and leakage [3].

* This is an estimate, based on the state of the art at the time of ref. [1]. 9

Metal 2

Si02 (CVD deposited)

.4o. Metal 1

Si02 (field oxide)

p- substrate

Figure 2.3 Metal-metal capacitor.

Metal 1

Si02 (CVD deposited) Poly Si

Si02 (field oxide)

p- substrate

Figure 2.4 Metal-poly capacitor. 10

The metal-poly capacitor illustrated in Figure 2.4 has a relatively large voltage dependence compared to the metal-metal one because the surface of the poly accumulates or depletes while the surface of the metal does not, and therefore the space-charge capacitance cannot be compensated [3]. In addition, CVD oxide is used as an insulator between the metal and the poly layers, and this results in poor matching property because of its inferior uniformity. For the capacitors described above, the parasitic capacitance associated with their bottom plates is about 50 to 100 % of the main capacitance, because their parasitic oxide

thickness is of the same order as the main capacitor oxide thickness.

These two capacitors can be combined into a sandwich structure by connecting the poly and the upper metal layers to increase the capacitance per unit area. Even in this case,

the chip area required is much larger than for double-poly capacitors.

Table 2-2 summarizes the properties of the capacitors described above.

Table 2-2 Capacitors in CMOS processes

Capacitor Advantages Disadvantages

Double-poly Excellent linearity. Requires additional process steps to a standard digital Good matching property. process.

Poly-diffusion Moderate linearity. Requires additional process steps to a standard digital Good matching property. process.

Large capacitance per unit area. Small breakdown voltage.

M1 -M2 Available in a standard digital Poor matching property . process. Requires large area. Good linearity. Large parasitics. 11

Table 2-2 continued

Capacitor Advantages Disadvantages

MI -poly Available in a standard digital Poor matching property. process. Requires large area. Moderate linearity. Large parasitics.

2.3 MOSFET Capacitor

2.3.1 Gate Structure

In an n-well CMOS process, four types of gate-structure capacitors (MOSFET capacitors) are available [4] as shown in Figure 2.5. (Note that similar capacitors can also be realized in a p-well CMOS process. However, since n-well processes are more often used for fine-line-width processes, only n-well CMOS processes are discussed in this thesis.) Figure 2.5a shows a floating capacitor which operates in its accumulation region.

The capacitor of Figure 2.5b is a grounded capacitor which also works in its accumulation region. The bottom plate (p+ channel) has the same potential as the p-substrate, which is permanently connected to Vss. The capacitors of Figures 2.5c-d operate in their strong inversion regions. In Figure 2.5c, the bottom plate of the MOSFET capacitor can be shielded from the substrate noise by connecting the n-well to a clean fixed potential such as VDD. However, this shielding will boost the threshold voltage of this MOSFET capacitor

due to the body effect so that higher bias voltage will be required to keep this capacitor in its strong inversion region. In Figure 2.5d, an increase of threshold voltage due to the body

effect is unavoidable unless the bottom plate is tied to Vss. 12

Top plate OH EC) Bottom (Poly) plate (Channel)

\los > 0

VSS

n- well

p- substrate

(a) (b)

+

n- well

p- substrate

(c) (d) Figure 2.5 MOSFET capacitors. 13

To implement high-linearity SC circuits, these capacitors have to be sufficiently biased to operate in their accumulation or strong inversion region. The total small-signal gate capacitance is given by

-1 = cox-, + Cs-, (2-2) 24)t Cg C [1 g OX I ivGs vi + 20t where t is the thermal voltage, VGS is the gate-to-source voltage, V is the flatband voltage

(VFB) for accumulation or the threshold voltage (Vm) for strong inversion, Cox is the gate- oxide capacitance, and Cs is the space-charge capacitance of the single-crystal silicon [4]. Therefore, VGS has to be large enough to achieve Cg = Cox for good linearity. (The space- charge capacitance of the poly layer is neglected in equation 2-2 because of the high poly doping density normally used.) Notice that all four capacitors can be realized in any conventional n-well CMOS process without additional process steps such as the second poly layer. Thus they are compatible with digital CMOS processes. For the capacitors of Figures 2.5a,c and d, the bottom plate (n-well) has a parasitic pn junction diode, which has a nonlinear C-V characteristic. This parasitic capacitance and its effects will be discussed in Sections 2.3.3 and 4.3.

2.3.2. C-V Characteristic

The capacitance-voltage (C-V) characteristics of MOSFET capacitors are among the most important issues in this research because the linearity of MOSFET capacitors can limit the linearity of SC circuits. The C-V characteristics of two MOSFET capacitors with the structures of Figure 2.5a-b were measured using the HP4061 C-V meter as illustrated in Figure 2.6, and the measured C-V curves are shown in Figure 2.7. The DC bias voltage was swept between -5 14

HP 4061 C-V mesurement system I

10 mV Top f=1 MHz

Bottom

Figure 2.6 C-V curve measurement of MOSFET capacitors

and 5 V while a sinusoidal voltage signal of 10 mV peak amplitude with a frequency of 1

MHz was applied. And the capacitance C was defined as C=dq/dv. The electron-accumulated structure of Figure 2.5a can be operated in its accumulation region by applying a positive bias voltage or in its strong inversion region by using a negative bias voltage. As shown in Figure 2.7a, its capacitance is close to gate- oxide capacitance Cox even without a bias voltage (V=0) because near the interface between the single-crystal silicon and SiO2 positive fixed oxide charges exist [5] and they attract electrons to the surface of the single-crystal silicon. 15

Accumulation Strong inversion

.5

0 -5 -q -3 -a -1 L1 1 a 3 5 BIFIS (V1

(a)

Accumulation

Strong inversion .sr

0 -5 3 4 BIAS (V)

(b) Figure 2.7 Measured C-V curves of MOSFET capacitors: (a) electron-accumulated MOS structure (Figure 2.5a); (b) hole-accumulated MOS structure (Figure 2.5b). The insert shows the frequency effect. 16

The first-order voltage coefficient was 850 ppm/V for the operating dc bias range between 0.5 V and 5 V, which corresponds to the range with sufficiently accumulated channel at the surface of the single-crystal silicon. When a negative DC bias voltage is applied to this capacitor, the space-charge layer which is in series with the gate-oxide capacitance depletes, and finally the surface of the n-well inverts as the negative bias voltage increases. In this structure, the only source for positive charge is minority carriers in the n-well (holes in this case) and the recombination-generation rates of the minority carriers cannot keep up with the high-frequency small-signal variations [5]. Hence in strong inversion, the total capacitance decreases for high-frequency signal and has a strong frequency dependence. (If the AC signal frequency is very low, e.g. 100 Hz, the total capacitance rises to Cox because the minor carriers can follow the signal and the depletion- layer capacitance increases. See the insert in Figure 2.7) The measured C-V curve for the hole-accumulated structure of Figure 2.5b is shown in Figure 2.7b. With a negative bias voltage, this capacitor is kept in its accumulation region and the capacitance approaches to Cox. The first-order voltage coefficient was 12,300 ppm/V for the operating dc bias range between -2 V and -5 V, in which holes are accumulated at the surface of the single-crystal silicon. With a positive dc bias voltage, the surface of the p-substrate inverts and the electron channel is formed (strong inversion). A small dc bias voltage is required to keep this capacitor in the strong inversion region, and its linearity is much better there than in its accumulation region. However, again, the capacitance in the strong inversion is frequency-dependent and it decreases for high-frequency signal because its minor carriers in the p-substrate (in this case, electrons) cannot follow the high-frequency signal, and the depletion layer capacitance, which is in series with Cox, limits the total capacitance. Therefore, both structures of Figures 2.5a-b have to operate in their accumulation

region when they are used in SC circuits; otherwise they will show a frequency dependence

along with the dc bias voltage dependence. 17

As shown in Table 2-3, the voltage coefficient in the accumulation region of Figure 2.7a was much smaller than that of Figure 2.7b. The reason is as follows. Figure 2.8 illustrates two poly/Si02/Si structures and their space-charge layers. In Figure 2.8a, the surfaces of the poly and n-well are depleting and space-charge layers are facing the SiO2 surface. In a typical CMOS process, the poly layer is n+ doped. Therefore the impurity polarity of the poly is the same as of the n-well. When a positive gate voltage VG is applied, the space-charge layer in the poly increases while it decreases in the n-well. When a negative gate voltage is applied, the space-charge layer in the poly decreases while it increases in the n-well. Thus, the variation of their space-charge layer thickness is compensated. (Since the poly is much heavily doped more than the n-well, they do not compensate completely. However, even the partial compensation results in a reduced voltage coefficient.) In Figure 2.8b, the surfaces of the poly and the p-substrate are also depleting. Note that the poly and the p-substrate have opposite impurity polarities. When a positive gate voltage VG is applied, both space-charge layers in the poly and in the n-well increase. When a negative gate voltage is applied, both space-charge layers in the poly and in the n-well decrease. Therefore, the variation of their space-charge layer thickness is much larger than of the structure in Figure 2.8a, and that results in the large voltage coefficient.

Table 2-3 Measured voltage coefficients of MOSFET capacitor structures

Operating dc bias MOSFET structure av voltage [V] [PPrnAl poly/Si02/n-well 850 0.5 < V < 5.0 (Figure 2.5a)

poly/Si02/p-substrate 12,300 -5.0 < V < -2.0 (Figure 2.5b) 18

VG = 0 V VG > 0 V VG < 0 V

tsl

SiO2 Space-charge layers

ts2 t

tsl ts2 = cons tan t n- well =t) Low voltage coefficient

VG = 0 V VG > 0 V VG < 0 V

tsl"

SiO2 Space-charge layers

ts2

Ip- substrate

Figure 2.8 Space-charge layers: (a) poly-Si/n-well; (b) poly-Si/p-substrate. 19

It should be also pointed out that the electron-accumulated MOS structure requires a smaller minimum bias voltage (around 0.5 V) than the hole-accumulated MOS structure (around -2 V) to operate in the accumulation region. This can be explained as follows.

Figure 2.9 illustrates the charges at the interface of SiO2 and single-crystal silicon: Qit is the interface-trapped charge and Qf is the fixed oxide charge. The interface-trapped charge

(sometimes called fast state or surface state charge) is due to the interruption of the periodic lattice structure at the surface of the silicon [5] and is known to cause 1/f noise. (The effect of the 1/f noise will be discussed in Section 6.4.11) The fixed oxide charge Qf originates from excess silicon or the loss of an electron out of excess oxygen centers near the Si-SiO2 interface, and it has the following properties [5]: it is fixed and generally positive; it cannot be charged or discharged over a wide variation of surface potential. When the fixed oxide charge is present at the interface, a MOS C-V curve shifts along the voltage axis as shown in Figure 2.10. In Figure 2.10a, the positive fixed oxide charges help the electron- accumulated capacitor operate in accumulation region with smaller bias voltage because they

attract negative charges to the surface of the crystal silicon. By contrast, due to the positive fixed oxide charges, the hole-accumulated capacitor must have a larger negative dc bias

voltage to operate in its accumulation region (Figure 2.10b).

Next consider the two inverted-channel MOSFET structures of Figures 2.5c-d. Since we didn't have test structures of Figure 2.5c-d, no measured results are available. However, in refs.[4][9][24], measured C-V curves of MOSFET capacitors with

the structure of Figure 2.5c-d are shown. For example, in reference [9], the measured first- order voltage coefficient av in strong inversion was given as 42.7 kppmN and 27.9

kppmN for the structures of Figure 2.5c and d, respectively (Table 2-4). Their C-V curves

do not have a frequency dependence in strong inversion, and their total capacitances in the

strong inversion region approach the gate-oxide capacitance Cox even for a high-frequency signal, because they have source/drain p+ or n+ diffusions, and channel carriers are

provided rapidly enough to follow the high-frequency signal. 2020

1_ Poly-Si

SiO2

li ii a a Si 0,,

Si

X: interface-trapped charge (Qit) ei : fixed oxide charge (Q0

Figure 2.9 Charges at the interface of SiO2 and single-crystal silicon [5]. 21

C

C-V curve with oxide charge effect

0

(a)

C

Ideal C-V curve

C-V curve with oxide charge effect

0

(b)

Figure 2.10 C-V curve shift due to positive fixed oxide charge [5]: (a) electron-accumulated MOS structure; (b) hole-accumulated MOS structure. 22

Table 2-4 Voltage coefficients of MOSFET capacitor structures [9]

Operating dc bias MOSFET structure [pp]0:V voltage [V] PMOSFET 42,700 -3.0 < V < -2.0 (Figure 2.5c)

NMOSFET 27,900 2.0 < V < 3.0 (Figure 2.5d)

It is easily realized that these structures require bias voltages at least equal to their threshold voltages to operate in their strong inversion regions. When the body effect is incorporated, even larger voltage will be needed to achieve good linearity. Among the four structures described above, the MOSFET capacitor with

accumulated electrons at the surface of the single-crystal silicon (Figure 2.5a) was found to be the most useful because it shows a smaller voltage coefficient than other structures and

requires a small dc bias voltage to operate in a good linearity region. Therefore throughout

this thesis, it will be used to design SC circuits in digital CMOS processes.

2.3.3 Parasitics

In this section, the parasitic capacitances of the MOSFET capacitor are discussed. As already mentioned, the electron-accumulated MOSFET capacitor will be used in this

thesis. This capacitor has a nonlinear parasitic capacitance at its bottom plate and a relatively linear stray capacitance at its top plate (Figure 2.11). We shall estimate the parasitic capacitance at the bottom plate first. Then the effect of the top parasitic capacitance will be

discussed. Figure 2.12 illustrates a 1 pF unit MOSFET capacitor laid out with the MOSIS 1.2 gm design rule. The width of the depletion layer between the p-substrate and the n-well is

given by 23

=112Es; NA + ND (vbi v) (2-3) d q NAND where Esi is the permittivity of silicon, q is the electron charge, NA is the acceptor impurity concentration, ND is the donor impurity concentration, Vbi is the built-in potential, and V is the reverse voltage for the pn junction [5]. We assume the following physical parameters for a standard CMOS process:

ND = 2 x 1016[cm-3], NA = 5 x 1014[cm-3], Vbi = 0.64[V], q =1.6 x10-19[C],

Esi =11.7.8.854 x10-14[F / cm]

Top (poly-Si)

Cpara_top

Bottom (accumulated channel)

Cpara_bottom

Figure 2.11 Parasitic capacitances of MOSFET capacitor 24

Top

p-

Figure 2.12 MOSFET unit capacitor

For V=0 V, the depletion-layer width is given by

2 -11.7. 8.854 x10-14(2 x1016+ 5 x1014) d =1 (0.64 + 0) 1.6 x10-19 .2 x1016 .5 x1014 (2-4)

= 1.31 x 10-4cm

The pn junction capacitance per unit area is given by 25

Esi 11.7 x 8.854 x10-14 -9 2] C= = = 7.9 x 10 [F/ cm d 1.31 x 10-4 (2-5)

= 7.9 x 10-5[F / m2]

Then the parasitic capacitance Cpar is calculated as

Cpar = Cbottom + Csidewall

= 7.9 x10-5 .37.2 x10-6 . 33.6 x10-6 (2-6) +7.9 x10-5 .3 x 10-6 . 2 (37.2 + 33.6) x10-6

= 0.13pF

This is a worst case because the reverse voltage V was assumed to be equal to 0 V. As the reverse voltage increases, the depletion layer widens, and consequently Cpar decreases. For example, when V is increased to 2.5 V, the depletion-layer width increases to 2.88 p.m, and Cpar=0.06 pF. In this case, the parasitic capacitance is only 6 % of the

main capacitance. Thus, the parasitic capacitance at the bottom plate of this unit capacitor is estimated

as only 15 % or less of the main capacitance. This value is of the same order as for the

double-poly capacitor and is much smaller than for the metal-metal or metal-poly capacitor. However, this parasitic capacitance is nonlinear, and hence the bottom plate should not be a floating node to avoid nonlinear distortion. The effect of the nonlinear parasitic capacitance

on SC circuits will be discussed in Section 4.3. Next consider the parasitic capacitance associated with the top plate of the MOSFET

capacitor. As illustrated in Figure 2.13, the top plate (poly layer) overlaps the field oxide. Its thickness is not constant due to the birds-beak shape of the field oxide, hence it is not straightforward to estimate the parasitic capacitance quantitatively. However, since the

thickness of the field oxide is 4000 to 8000 angstrom, and the overlap width is about 10 % of the MOSFET width, the parasitic capacitance is expected to be less than 10 % of the 26 main capacitance. To minimize the matching error due to these parasitics, a parallel combination of several smaller unit capacitors will be used in the layout [6].

= 0.6 pm

44X "II' 2X -11

'OF-2X 11111' Poly Si SiO2

SiO2 (thermally grown

Field oxide Field oxide

n- well

p- substrate

Figure 2.13 Profile of MOSFET capacitor

2.3.4 HSPICE Models

In this section, the models for the MOSFET capacitor and the pn junction diode

which have been used in HSPICE simulations will be discussed. 27

2.3.4.1 MOSFET Capacitor Model

The Level-2 HSPICE models optimized for the Orbit 1.2 gm CMOS process have been used throughout this thesis. HSPICE provides several MOS gate capacitance models, and they are selected by the CAPOP command [7] as follows:

CAPOP=0 Original Meyer model in SPICE

CAPOP=1 Modified Meyer model

CAPOP=2 Parameterized modified Meyer model (default) CAPOP=3 Parameterized modified Meyer model with Simpson integration

CAPOP=4 Charge conservation model Figure 2.14 shows simulated PMOSFET C-V curves for different CAPOP values: (a) 0; (b) 1; (c) 2; (d) 3; (e) 4. For CAPOP = 0 and 4, the PMOSFET capacitance was constant in both accumulation (V > 0.5 V) and strong inversion regions (V < -1.5 V), that is, they did not show any nonlinear behavior in those regions (Figure 2.14a, e). By contrast, the MOSFET capacitances with CAPOP = 1, 2, 3 (Figure 2.14b, c, d) show weakly nonlinear behavior in the accumulation region (V > 0.5 V), relatively strongly

nonlinear behavior in the strong inversion (V < -1 V), and very strongly nonlinear behavior

in the depletion region (-1 < V < 0.5 V). No big differences were observed among the three CAPOPs. To include the weak nonlinearity of the MOSFET capacitor in the accumulation

region, CAPOP=2 was chosen for HSPICE simulations in this work. Next, the PMOSFET and NMOSFET C-V curves are compared. Figure 2-15 shows the C-V curves of the PMOSFET and NMOSFET for CAPOP=2. Their first-order voltage coefficients of these are shown in Table 2-5 and 2-6, respectively. Both MOSFETs have a much better linearity in their accumulation region than in the strong

inversion region. The first-order voltage coefficient of the PMOSFET is 750 ppm/V for the

dc bias range of 0.5-2.5 V, and that of the NMOSFET is 610 ppm/V between -3 and -1 V. In the experimental results described in Section 2.3.2, the first-order voltage coefficient of the MOSFET capacitor with the structure of Figure 2.5a was 850 ppm/V for the operating 28 dc bias range between 0.5 and 5 V. Therefore the simulated PMOSFET C-V curve is closer to that of the electron-accumulated MOSFET capacitor in accumulation region, and the PMOSFET model was adopted to simulate the MOSFET capacitor in HSPICE, even though the actual structure used (Figure 2.5a) was an n-channel one.

Table 2-5 Voltage coefficients of the HSPICE PMOSFET capacitor model

Operating mode av [PPInN] Operating dc bias voltage [V]

Accumulation 750 0.5 < V < 2.5

Strong inversion 21,200 -3.0 < V < -1.0

Table 2-6 Voltage coefficients of the HSPICE NMOSFET capacitor model

Operating mode av [ppmN] Operating dc bias voltage [V]

Strong inversion 23,500 1.0 < V < 3.0

Accumulation 610 -2.7 < V < -0.7 29

I.OP 7

800.0F

600.01_ (a)

400.0F_

L I

1.0P 7 --A

900.0F_ (b) 600.01_

400.0F I 3 L I

1.0P

800.0F (c) GOO.OF

400.0F I 1 1 I 1 4.0 -4,0 -2 01 0_ 2.0 -5.0 VOLTS (LIN) 5 0

1 OP

900 . OF (d)

1 OP

9 00 Or

600 OF 7 (e)

4 0 0 . 0 1

1 -4.0' -2 01 0_ 0 4.0 -5.0 VOLTS (LIN) 5 0

Figure 2.14 Simulated C-V curves of PMOSFET for different CAPOP values: (a) 0; (b) 1; (c) 2; (d) 3; (e) 4 30

Note that this model has discontinuities at the boundary of the depletion and accumulation (or strong inversion) regions. However, if MOSFET capacitors operate only in their accumulation regions in HSPICE simulations, this model can achieve a good agreement with actual devices. (Remember that the accumulation region of the electron- accumulated MOSFET capacitor is the most important one among all MOSFET capacitor configurations.) In addition, it should be pointed out that the slope d(C/Cox)/dV in the accumulation region of the simulated PMOSFET C-V curve is negative as shown in Figure 2.15c, while it is positive in a real device. However, the absolute value of the voltage coefficient is more important than the polarity of the slope when the nonlinearity of the MOSFET capacitor is taken into account for circuit simulations, and hence using the wrong polarity will not give a serious problem on simulation results.

2.3.4.2 PN Junction Diode Model

Since a MOSFET capacitor has a parasitic pn junction diode at its bottom plate, a pn junction diode model should be connected at the bottom plate of each MOSFET capacitor in

HSPICE simulations. Based on the physical parameters described in Section 2.3.3, the C-V curve for a pn junction diode was obtained. The bottom plate area is 37.2 pm x 33.6 inn, the peripheral length of the sidewall pn junction is 2 x (37.2 pm + 33.6 pm), and the n-well depth is estimated as 3 pm. Figure 2.16 shows a simulation result of C-V curve for the pn junction diode. This simulation result agrees with the analytical result in Section 2.3.3.

2.3.5 Accuracy Considerations

Two factors determine the accuracy of MOS capacitors: systematic and random errors. These errors have been extensively investigated in refs.[1][2][6][8]. 31

1 0 =

950.0M 7

900.0M _7

B50.0M _

B00.0M 7

750.0M 7

700.0M _7

650.0M 7

600.0M =

550.0M 7

500.0M _7

450.0M

400.0M 7 II 11 350.0M (7 I 2.0 . 0 -20 0 . 5 0 -4. 0 VOLTS [LIN) 5 0 (a)

1.0422 f

950.0M _

900.0M _

850.0M 7 eonnm

750.0M

700.0M

650.0M _

600.0M _

550.0M _

5 0 0 0M _

450.0M _

400.0M , 1 4.0 -4.01 -2.01 O. -5. 0 VOLTS (LIN) (b) Figure 2.15 Simulated C-V curves for CAPOP=2: (a) PMOSFET; (b) NMOSFET; (c) magnified PMOSFET C-V curve (0.5 < V < 5 [V]). 32

996.20M

986 .10/.1 986.0M 985.90M 7 985.80M ­

985.70M ! 985.60M: 985.50M ­ E 985.40M .7 985.30M: 985.20M­ : 985.10M: 985.0M 984.90M : 984.80M: 7 984.70M LI. 984.60M : 984.50M­ = 984.40M:

CI , I i 1 1 I i I 1 1.0 2.0 3.0 4.0 500.0M VOLTS (LIN) 5 0

(c)

Figure 2.15 continued.

200.01 190.0F 180.0F 170.0F 160.0F 150.0F 140.0F 130.01 120.0F _= 110.0F 100.0F 90.0F 80.0F 70.0F :

60.01 -7= 50.0F 40.0F 30.0F = 20.01 10.01 2.0 3.0 4.0 O. 5.0 Reverse voltage [V] Figure 2.16 Simulated C-V curve of the pn junction diode 33

Systematic errors affect capacitors with identical geometries equally, hence they can be reduced by unit-capacitor layout and common-centroid geometry [6] as shown in Figure

2 -17. By contrast, random errors are uncorrelated from capacitor to capacitor and they cannot be improved by layout techniques. They are divided into two categories: random edge effect and random oxide thickness effect.

The nominally straight edges of the top layer (poly or metal) of a MOS capacitor are wavy in a practical fabrication process due to the granular nature of the poly (or metal) and the jagged edges of the developed photoresist [8]. This is called random edge effect. The expected relative error AC/C is proportional to C-3/4 [6]. Random oxide effect is caused by fluctuations of the oxide thickness and permittivity. The expected relative error AC/C is proportional to C-112 [6].

C

C C c

Figure 2.17 Unit capacitors with common-centroid geometry 34

When a capacitance is small, the edge effect dominates; by contrast, oxide effects limits the accuracy for a large capacitance. Although the crossover value depends on technologies, it is estimated around L = 20 to 50 pm [6].

2,3.6 Layout

In this Section, the practical layout of a MOSFET capacitor will be presented. Some remarks associated with CAD tools will be also made. Figure 2.18a shows a possible layout of a unit MOSFET capacitor. To implement the structure of Figure 2.5a, n+ diffusions are formed inside the n-well. A p+ guard ring surrounds the capacitor to reduce substrate noise. The following remarks should be taken into account for the Mentor layout tool.

They can also be applied to other CAD tools. 1. When a design-rule check (DRC) is run on the MOSFET capacitor with the gate

structure of Figure 2.5a, a design error message "N-type gate must not be in n-well" will appear, because n+ regions are formed inside an n-well for the MOSFET capacitor. However, one can ignore this DRC error if it relates only to the MOSFET capacitor, and it

should not cause a problem in fabrication. 2. When the layout is completed, it is advisable to extract devices and parasitics to

run a post-layout simulation. Before extraction of the MOSFET capacitors, the n+ implant- select layer, which determines an area where n+ doping is given, must be replaced by the p+ implant-select layer (Figure 2.18b) so that the CAD tool can recognize them as PMOSFETs. Otherwise, the CAD tool will not be able to extract the MOSFET capacitors

because the n-type gate is formed in the n-well. This operation should be done only for the post-layout simulation. In the final layout file, all MOSFET capacitors must have the N+ IMPLANT-SELECT LAYERS

instead of the p+ implant-select layers as it was originally laid out. 35

Poly

Metal-1

Imuumuu minium' I mum Minium n minim imumn inim Metal-2

P+ implant-select --I1 N+ implant-select Active

r"--1 x7- -well IN ,1011111111111ummummilinimminimminominumile, t Contact

(a) N+ implant-select P+ implant-select

71,

un m n i inn

rrAI.InuuuuhIuIII IIIIIUIIIII inn ouiiiioiminrniiuoiiuiiii t

(b) Figure 2.18 Layout of unit MOSFET capacitor: (a) original; (b) with dummy p+ implant-select layer. 36

3. When a post-layout simulation is performed by HSPICE, a PMOSFET model should be used for the MOSFET capacitor as it is extracted. 4. In the HSPICE netlist file generated, parasitic capacitors can be extracted as linear components even if some capacitances, such as pn junction capacitance between n- well and p-substrate, are nonlinear. If the circuit is very sensitive to parasitic pn junction

capacitances and one wants to see their effects in post-layout simulations, it is necessary to

put them manually in the extracted netlist file. 37

Chapter 3 EARLIER TECHNIQUES FOR LINEAR ANALOG SIGNAL PROCESSING IN DIGITAL CMOS TECHNOLOGY

In this chapter, earlier research results on the realization of SC circuits using MOSFET capacitors and on other analog signal processing techniques without linear capacitors are described. For illustration, sample-and-hold circuits are described which can

be realized using nonlinear capacitors.

3.1 Simple Biasing

Figure 3.1 shows an example of how MOSFET capacitors can be operated in their accumulation or strong inversion regions to achieve high linearity in a lossy SC integrator

circuit [4]. Since the non-inverting node of the operational amplifier (op-amp) is connected

to a bias voltage Vb, the dc voltage of the virtual ground also becomes Vb assuming that the

op-amp is ideal. Thus, if Vb is large enough, all capacitors in this circuit can be kept in their accumulation or strong inversion regions according to the polarity of Vb and the MOS gate structure. Notice that the voltage difference between Vb and the input (or output) voltage must be large enough to keep all MOSFET capacitors in the high-linearity regions over the

full signal range. Bazarjani et al. designed a fourth-order SC bandpass delta-sigma modulator using such simple biasing technique [10]. The bandpass modulator consists of cascaded resonators. Each resonator was implemented with two SC delay cells in a negative feedback loop. PMOSFET capacitors were used in the SC delay cells and were kept in their strong inversion region using a simple biasing technique. Their simulation results showed a

S/(THD+noise) ratio 28.5 dB (vs. a simulated 36 dB) over a band of 5 MHz. 38

C3

Top plate I (poly gate)

TBottom plate (channel)

Vin

Vout

Vb

Figure 3.1 Lossy integrator using simple biasing technique [4].

3.2 Charge Processing

Bermudez et al. proposed a concept of charge processing [11], which is similar to current processing in switched-current techniques [12]. In Figure 3.2, suppose that a charge iq 1 is transferred to capacitor C1 from an input branch. Then the increment of output AVO is given by

AV0=Aqi/Ci (3-1) assuming the op-amp is ideal. The charge across the load capacitor C2 is given by

Aq2--.-C2AVo (3-2)

Notice that the ratio of charges is given by 39

Aqi Ci AVo C10 fi[Vo] AVo Aq2 C2 AV° C20 f2[Vo] AV° (3-3)

= if f1H f 2[1. %--20 where the Cio are constant scale factors determined by capacitor geometries, and the fij.] are nonlinear functions of the voltages across the capacitors. Thus, even if these capacitors are voltage-dependent, this circuit provides a linear charge mirror with a gain defined by capacitor geometries if C1 and C2 have matched nonlinearities.

Cl

Co

Q2

Figure 3.2 Charge processing.

To achieve this, two conditions have to be satisfied [11]: (a) after the settling time corresponding to each clock phase, the voltage across any capacitor must depend only on a single node voltage; (b) all capacitors connected to the same op-amp output must exhibit identical nonlinearities. In SC circuits, these conditions are usually satisfied, and hence charge processing can be applied to general SC circuits. Charge processing is useful for internal charge transfers in SC circuits. However, just as a switched-current circuit requires a linear V-I transfer circuit at its input and a linear I-V transfer circuit at its output, a switched-capacitor circuit using charge processing requires a linear V-Q circuit at its input and a linear Q-V circuit at its output. Therefore, the 40 input and output capacitors have to be linear*, and in digital CMOS processes, the choice

for such linear capacitors would be metal-metal, metal-poly, or MOSFET capacitors. It was reported that a S/THD ratio of 50 dB for an input frequency of 1 kHz was experimentally

obtained at the output of a third-order lowpass SC filter with MOSFET capacitors [11].

3.3 Switched-Current Circuits

Switched-current (SI) circuits were proposed as an inexpensive alternative to SC circuits [12]. Figure 3.3 shows a simple current mirror, in which a current iA is linearly scaled to a current iB in spite of the nonlinear I-V device characteristics. Analog signal processing steps such as summation, integration, inversion, and scaling can be achieved in the current mode, and highly linear capacitors are not required. Hence, SI circuits are compatible with a standard digital CMOS process, and that is the main advantage of this

technique.

Figure 3.3 Current processing.

* Or have matched nonlinearities. In general, this is not feasible since Vont is not equal to Vin. 41

A comparison of SI and SC circuits can be found in reference [13]. SC circuits have advantages such as high accuracy, low power consumption, and high speed under the same power consumption. By contrast, SI circuits have higher speed (because SI circuits are not constrained by the op-amp gain-bandwidth products [14]) but at the expense of more power, and small chip area. However, SI circuits generally require V-I and I-V conversion circuits because most of the signals are available or needed as voltages. In addition, the signal-dependent charge injection caused by limits the accuracy of SI circuits.

3.4 Sample-and-Hold Circuits

The sample-and-hold (S/H) circuit is one of the most important analog building blocks. In this section, some S/H circuits are introduced to show the feasibility of SC circuit realization without using highly linear capacitors or special nonlinearity

compensation technique. Figure 3.4 shows a simple offset-free S/H circuit [15]. In this circuit, the same capacitor is used for sampling the input voltage and holding the stored charge for the output. Therefore, that capacitor can be realized even by a nonlinear capacitor such as a

MOSFET capacitor because capacitance matching is not needed. An improved offset-free S/H circuit for fast settling [16] is illustrated in Figure 3.5. In this circuit, capacitor Cs is used only for holding the output voltage during ch =1, and

the nonlinearity of Cs is also irrelevant to the output linearity. Hence, this S/H circuit can be

also realized using nonlinear capacitors. 42

Top plate I(poly gate)

Bottom plate (channel)

yin oi o

Figure 3.4 Simple S/H circuit.

yin 4)1 0

Figure 3.5 Improved S/H circuit. 43

Chapter 4 MOSFET-ONLY SWITCHED-CAPACITOR AMPLIFIERS USING NONLINEARITY COMPENSATION TECHNIQUES

In this chapter, novel design techniques will be presented to reduce the effects of MOSFET capacitor nonlinearity. MOSFET-only SC amplifiers will be described using these techniques. The validity of the method will be verified by HSPICE simulations.

4.1 Early Work and Applications

The SC amplifier is a fundamental analog building block, which is used not only as an amplifier, but also widely used in many applications such as digital-to-analog (D/A) converters, multiplying circuits, and programmable gain/loss circuits [17]. Various kinds of SC amplifiers were already proposed in refs.[18]-[21]. The SC amplifier shown in Figure 4.1 has several special features [21]: a) offset voltage cancellation; b) 1/f noise reduction; c) insensitivity to finite op-amp gain and slew rate.

This SC amplifier is a suitable test vehicle for estimating the capacitor nonlinearity effects because of its simplicity. Therefore, we used MOSFET capacitors in the SC amplifier of Figure 4.1 for an estimation of their nonlinearities.

4.2 Simple Biasing Technique

Figure 4.2 shows an SC amplifier using MOSFET capacitors and the simple biasing technique proposed in ref.[4]. MOSFET capacitors in the signal paths of this circuit operate in their accumulation regions by having a positive dc bias voltage at the inputs of the op­ amp (virtual grounds). To establish a positive dc bias voltage at the virtual grounds, the analog ground which was connected to the virtual ground via an SC branch in Figure 4.1

was replaced by a positive dc bias voltage Vb. Fully differentialconfiguration was used to 44

I( Top (2nd poly) Cdg 1`

Bottom plate (1st poly)

Cs

\Tout

=

Figure 4.1 Gain- and offset-compensated SC amplifier [20]

A 45

I ( Top (poly) Cdg I± Bottom (accumulated channel) Cs

Figure 4.2 SC amplifier using simple biasing arrangement (only half of the fully differential circuit is shown).

reduce the effect of the supply and substrate noise, as well as clock feedthrough and charge injection [22]. (In Figure 4.2 and other figures of SC amplifiers in this chapter, only half of the fully differential circuit is shown for simplicity.)

The deglitching capacitor Cdg was connected between the input and output terminals to reduce the generation of output spikes during the non-overlapping period of clock phases

[23]. 46

4.3 Parallel Compensation

To achieve better linearity, a novel design technique, called parallel compensation, was developed [25]. Figure 4.3 illustrates a conceptual structure of the parallel compensation. Two MOSFET capacitors Ci are connected with floating bias voltages so that these capacitors operate in their accumulation regions. Whench is high, the capacitors are discharged. In the next clock phase 02, these capacitors are connected to the input. Then the voltage across one of the two capacitors increases, while the voltage across the other decreases; hence one of these capacitances increases while the other decreases. When the charges are transferred to the feedback path via the virtual ground, their nonlinearities are canceled to a first-order approximation.

Cl To the feedback path ) + r I 02 A V in / cr­ \­ / / Virtual C1 ground I +

( /1 I / I/

Floating dc bias voltage

Figure 4.3 Conceptual structure of the parallel compensation.

Figure 4.4 shows a practical realization of the circuit of Figure 4.3. Two extra capacitors C11 are added to the main capacitors C1. When 01 is high, node A is connected to 47 the positive bias voltage while node B is connected to the negative bias voltage so that all four capacitors operate in their accumulation regions. When (1)1 goes low, nodes A and B will be floating. Since the switch connected to the virtual ground is turned off, no charge flows into or out of the virtual ground node, and the charges stored on the holding capacitors CH will be conserved. Thus, the voltages at nodes A and B will be constant and those two nodes will become ac virtual grounds. The main capacitors CI will be kept in their accumulation regions throughout the operation.

Vbp (positive bias voltage)

ITo the feedback path Cl 01 f A ) 1

02 Vin Virtual o \ ground Cl

Vbn (negative bias voltage) Figure 4.4 Practical realization of the parallel compensation

Figure 4.5 illustrates the resulting SC amplifier using the parallel compensation. In practical SC circuits, the input and feedback branches can share the holding capacitors CH. The value of CH can be much smaller than Ci, because CH's, Ci's and C2's are connected 48 in parallel, and hence the kT/C noise of the CH branches can be neglected. Therefore, although the capacitor area increases compared to the simple biasing circuit, the increase is insignificant.

Cdg

is Vbp

/ C2

CH

Cl Vont+ 0

Vbn 0 \Tout-

Figure 4.5 SC amplifier using parallel compensation 49

4.4 Series Compensation

Figure 4.6a shows another novel design technique which was developed to reduce the nonlinearity of capacitors, called series compensation [26]. Two MOSFET capacitors are connected at their top plates, and when 01 goes high, the top plates are connected to the dc bias voltage Vb. The bottom plates of these capacitors are connected to the ground (one to the analog ground and the other to the virtual ground) during the phase 01, and the two capacitors are biased to operate in their accumulation region. In the next clock phase 02, the switch connected between the bias voltage Vb and the top plates is open, and node A will

be floating. Since the input is connected to the bottom plate of C1, the potential of the floating node A will be changed. Then, one of the two capacitances increases while the other decreases. Since the two capacitors are connected in series, the nonlinearity of their total capacitance will be canceled out to a first-order approximation. A possible implementation is shown in Figure 4.6b. Since the two bottom plates (n-wells) have

different potentials, the n-wells have to be separated. Let us analyze how linearly charges will be transferred to the feedback path from this series SC branch. Using the notations used in [11], the nonlinear capacitances Ci and

C2 in Figure 4.7 can be expressed as

Cl = C10 fl[V1] (4-1) C2 =C20 f2[V2] where Cip is a scale factor independent of the applied voltage, and f[.] is a nonlinear function of the voltage across the capacitor. Assuming that C10 = C20 = C and that f1[] :±: f2[], i.e. the nonlinearity of capacitor Ci matches that of C2, the charge delivered

by this series branch during the phase 4)2 is given by

Aq C f[VA(n)] f[VA(n)- Vin(11)1 Vin(n) f[VA(n)]+ f[VA(n)- Yin(n)] (4-2) +C f[Vb] (f[VA(n)]- f[VA(n)- Yin(n)]) Vb f[VA(n)] + f[VA(n) Vin(n)] 50

C1 C2 $2 B A C Vin 0 Virtual 0 \ +I ( I )1 + ground -V 1 V2

01 (11 r+ 1 _ Vb _

(a)

A 0

Vss C Vb B

n-well n-well

p-substrate

(b)

Figure 4.6 Series compensation SC branch: (a) circuit diagram; (b) practical implementation.

where VA is the voltage at node A in Figure 4.7. If the MOSFET capacitors operate in their accumulation region during the phase 02, VA can be approximated as 51

(a)

f [Vi]

yin yin 2 2

--S Vb vi (b)

Figure 4.7 Capacitance linearization: (a) Charge transfer from the series SC branch; (b) nonlinear function f[V] 52

VA(11) Vb + 21 Vin(n) (4-3)

Then,

1 f[VA(n)] f[Vb Vin (n)]

(4-4) f[Vb] +d1 6V(n) dV vb

f[VA(n) Vin(n)] = f[Vb Vin(n)]

(4-5) df f[Vb] 8V(n) dV vb where df is the first-order derivative of the nonlinear function f[] at Vb , dV Vb

1 SV(n) = Vin(n) is the deviation of VA(n) from Vb, and 8V(n) « Vb is assumed. From 2 equations (4-4) and (4-5),

f[VA (n)] tIVA (n) Vin (n)] f[VA(n)] + f[VA(n) Vin(n)]

2 df (f[Vb])2 8V(n) dV vb (4-6) 2 f[Vb]

f[Vb] 2

Here the second-order small term was neglected. Also,

f[VA(n)] f[VA(n) Vin(n)] f[VA(n)]+f[VA(n)Vin(n)]

df (4-7) 8V(n) dV Vb f[Vb] 53

By equations (4-2), (4-6) and (4-7),

C df Aq = c Vib(n) f[Vb1+--i Vin(n) 2 Vb dV

(4-8) f[Vbj +df Vb]Vin(n) 2 dV Vb

In equation (4-8), the expression in the square bracket is constant because Vb is a fixed potential. Therefore, to a first-order approximation, Aq is a linear function of Nib,.

We applied the MOS SC branch of Figure 4-6 in the fully differential SC voltage

amplifier. Figure 4-8 shows a circuit diagram of the SC amplifier, in which all capacitors in the signal path were replaced by the series-compensation SC branch. When 01 is high, all

capacitors are biased and kept in their accumulation region. In the subsequent phase 02=1,

the capacitors still stay in the accumulation region. Then, the output voltage of the SC

amplifier can be estimated by

Ci Vout(n) E --Vib(n) C2 (4-9) 2 {C1 _ uv bcif SVi(n) 8V2(n)} f[Vb] dV Vb

1 1 where Mil (n) Vin(n) and 8V2(n) Vout (n) . Since 8V2(n) = 8Vi(n), the 2 C2

nonlinearity at the output will be further reduced. Notice that voltage at the gate node of the feedback MOS SC branch increases as the output swing increases. This node voltage should not go beyond the positive supply

voltage because then the pn junction between the drain and the n-well of the PMOS switch

may turn on, and that results in a leakage of charges (Figure 4.9). This limits the output swing in this configuration. (A switch configuration to overcome this constraint will be

described in Chapter 10.) 54

Cdg Top plate I­ (poly gate) I eel I y2 \ Cs --r­ I \401 IC C2 T 2 02 Bottom plate

(accumulated channel) I

S2 oi I b Vb 7:­

vout+ S

\Tout-

Figure 4.8 SC amplifier using series compensation. 55

C2 C2 (1)2 Vout Virtual p ) ground

.01

Vb O Vss VDD ? Li 1 p+ II n+ I n-well

p-substrate

Figure 4.9 Practical realization of a PMOS switch

Since no dc bias voltage is given to the capacitor Cs in Figure 4.8, it operates in a strongly nonlinear region. However, it does not affect the linearity of the SC amplifier because its role is just to provide a feedback to the op-amp during the phase (11.1 to keep the op-amp from saturation. (When 4i is high, the top plate of Cs is connected to the virtual ground, and the charges from C1 and C2 flow into Cs. In the next clock phase 02, which is the valid phase, the top plate of Cs is switched to the ground, and Cs does not contribute to the charge redistribution with Ci and C2.) Note that a nonlinear parasitic capacitance (pn junction diode) is connected to the bottom plate of each MOSFET capacitor as shown in Figure 4.10. Let us consider the effect of the parasitic capacitance. In Figure 4.10, node 1 is switched to the ground or to the input, both of which are low impedance nodes, and hence the charge on Cdi does not contribute to the charge redistribution of the series-compensation SC branch. Since node 3 is the virtual ground, the charge on Cd2 is constant. Thus, these nonlinear parasitic 56

2 C2 3 yin0 Virtual - I ground

V2 I (+01

Cd2

4 I -Vb Vss Vss

Figure 4.10 Nonlinear parasitic capacitances of the series SC branch.

capacitances will not generate distortion in the SC amplifier. The same explanation can be applied to the SC branch in the feedback path. Another possible structure of the series-compensation SC branch is illustrated in Figure 4.11. The bottom plates of the two MOSFET capacitors are tied together and switched to a negative dc bias voltage during the phase 01 to keep these capacitors in their accumulation regions. The basic operation is the same as the other series SC branch described earlier. However in this configuration, the nonlinear parasitic capacitance of the MOSFET bottom plate will cause a serious distortion problem. When (1)2 is high, the input is connected to the top plate of C1, and the potential of node 2 changes in consequence of the charge redistribution between C1, C2 and Cd. Since Cd is nonlinear, the resulting charge redistribution causes distortion, which is generally unacceptable. 57

CI C2 Vin 102 1 2 3 0 N- Virtual + C ground V1 ' -)12 1

01 1 1 f 01

I

1 Cd # Vb

i T-_­ i VSS

Figure 4.11 Another possible structure of the series SC branch.

4.5 Simulation Results

To verify the operation of these SC amplifiers and the effectiveness of the proposed design techniques, the SC amplifiers were simulated using HSPICE. A voltage gain of 10 was assumed.

4.5.1 Simple Biasing

First, the parameters and conditions used in HSPICE simulations are described. They remain the same in Sections 4.5.2 and 4.5.3, unless otherwise specified. The PMOSFET model of CAPOP=2 described in Section 2.3.4 was used for all capacitors. The size of the unit capacitor was 18 pm x 18 p.m, which corresponds to 0.5 pF. Ci consisted of the parallel connection of 10 unit capacitors, while C2 and Cs were single-unit capacitors. The deglitching capacitor Cdg was connected between the input and output terminals to reduce the generation of output spikes during the non-overlapping period of clock phases [23]. The MOSFET capacitor size for Cdg was 8.4 pm x 8.4 pm (about 0.1 58 pF). In each MOSFET capacitor, the well was connected to the source and the drain. A CMOS transmission gate was used for the output node, PMOS switches were used for the biased SC branches, and NMOS switches for the rest of the circuit. The switch size was 3.6 pm x 1.2 pm. We used a SPICE macro model for the op-amp with the following conditions: DC gain 70 dB; unity-gain bandwidth 30 MHz; offset voltage 5 mV. The frequencies of the input sinusoidal and clock signals were 10.4 kHz and 312.5 kHz, respectively. The bottom plate of an MOSFET capacitor has a parasitic capacitance between n-well and p-substrate. To incorporate an effect of the nonlinear capacitance into the simulation, a pn junction diode was connected to the bottom plate of each MOSFET capacitor. The power supply voltages were +/- 2.5 V. We assumed a sinusoidal signal with an amplitude 200 mVp-p for each differential input, and the resulting swing was expected to be 2 Vp-p for each output node. To investigate the nonlinear effects due to the MOSFET capacitors, an FFT analysis of the output was performed in MATLAB. For the FFT analysis, the differential outputs were sampled at the end of each clock phase 02. The amplitude in the FFT analysis was normalized in such a way that a sinusoidal waveform with an amplitude of 4 Vp-p corresponds to 0 dB. Figure 4.12 shows the time-domain and frequency-domain outputs of the SC

amplifier of Figure 4.2 with Vb = 0 V. Since no dc bias voltage is provided, the MOSFET capacitors operate mostly in their depletion regions, and hence the gain was smaller than expected (Figure 4.12a) and the output was badly distorted (Figure 4.12b). When a dc bias voltage of 1.5 V is applied in the circuit, the linearity is greatly improved as shown in Figure 4.13 because each MOSFET capacitor in the signal paths now operates in its accumulation region. In Figure 4.13b, the signal-to-total-harmonic­

distortion ratio (S/THD) was 86 dB. When the input amplitude is doubled, the harmonic distortion increases as shown in Figure 4.14. The S/THD is down to only 26 dB. This is because the output amplitude

increases (now the maximum output amplitude is 2 V), while the dc bias voltage across the 59

1.20 TEST. TR 0 YCOUT - I 011( IN2 1 0

800

G00.0/1

400. ON

200 OM

-200 OM

-000 0

-GOO

-BOO N

1 3 I I I -1.20 " 54 0 .01.1 500.05 580.0U 000.0U 620.00 525 OU TIME (LIN) 025.0U

(a)

20

-20­

-80

100­

-120

- 1400 20 40 60 80 100 120 140 160 Frequency [kHz]

(b)

Figure 4.12 The output of the circuit of Figure 4.2 with Vin = 200 mVp_p and Vb = 0 V: (a) time domain. The dashed line represents Vin multiplied by a gain of 10; (b) frequency domain. 60

1.20 WBMO.TRO V(OUT IOwV(VIN21 I 0

000.0M

600.0M

400.0M

V 0 L 200.00

0.

-200.0M

-400. OM

-600.0H

-900.0M

-I.0

I t I -1.2° " 5140.00 560.0U 580.0U 600.0U 620.0U 525.0U TIME (LIN] 625.0U

(a)

20

0

-20

-40

-60

-80

100 Ar1tOlm-f) 120

140 0 20 40 60 80 100 120 140 160 Frequency [kHz]

(b)

Figure 4.13 The output of the circuit of Figure 4.2 with Vin = 200 mVp_p and Vb = 1.5 V: (a) time domain. The dashed line represents Vin multiplied by a gain of 10; (b) frequency domain. 61

2 -50 . . weno,TRo VIOUT 2.250 : = 10IVIVIN21 2.0: 13 1.750: 1.50 :

1 250 :

750.0M V 0 500.02 : 250.0M :

0.:­ N 250 OM -500 OM -750.02; -1.0:

-1 .250 :

-1 .50 :

-1 750 -2.0 : -2.250 -2.50 ' .1. .1 540.00 560 OU 520.0U GOO CU G20.0U 525 OU TIME (LIN) 625.90

(a)

20

0

-20

-40 an

-60

< -80

100

120

- 140 0 20 40 60 80 100 120 140 160 Frequency (kHz]

(b) Figure 4.14 The output of the circuit of Figure 4.2 with Vin = 400 mVp.p and Vb = 1.5 V: (a) time domain. The dashed line represents Vin multiplied by a gain of 10; (b) frequency domain. 62

MOSFET capacitor C2 decreases (the minimum dc bias voltage is -0.5 V), and hence the capacitor operates in the depletion region instead of the accumulation region. Therefore, the simple biasing technique is not very effective for a large output swing*.

4.5.2 Parallel Compensation

When the SC amplifier using the parallel compensation was simulated, the capacitor size was chosen to have the same kT/C noise as in the simple biasing. (Although HSPICE does not simulate noise, the capacitor size was chosen in this way for fabrication.) The size of the unit capacitor was 13.2 pm x 13.2 lim, which corresponds to 0.25 pF. CI consisted

of the parallel connection of 10 unit capacitors, while C2 was a single unit capacitor. Since

two CI capacitors are connected in parallel, the total input capacitance is 5 pF. In the same

manner, the total feedback capacitance is calculated as 0.5 pF. Other capacitors were chosen as follows: CH = 1 pF; Cs = 0.5 pF; Cdg = 0.1 pF. A CMOS transmission gate was used for the output node, a PMOSFET was used for any switch connected to Vbp, and NMOS

switches for the rest of the circuit. Vbp and Vbn were +/- 2.5 V, respectively. Figure 4.15 shows the output spectrum of the SC amplifier illustrated in Figure 4.5. For a sinusoidal signal with an amplitude 200 mVp-p, the S/THD was 86 dB (Figure

4.15a). When the input amplitude is doubled, capacitors in this circuit still operate in their accumulation regions, and the S/THD of 74 dB was obtained (Figure 4.15b). Hence, the parallel compensation is effective even for a large signal swing. A 12-dB decrease of S/THD implies that the second-order voltage coefficient of the MOSFET capacitors

increased 4 times when the signal was doubled.

* If the input capacitance C1 is smaller than the feedback capacitance C2, the voltage swing across C1 will be larger than across C2. In such a case, the maximum input voltage swing limits the linearity of the circuit. 63

20

-20

-40

-60

-80 ­

100

-120

, 1400 , 20 40 60 80 100 120 140 160 Frequency [kHz]

(a)

20

- 20

-40­

60 ­

- 80

100

- 120

- 1400 20 40 60 80 100 120 140 160 Frequency [kHz]

(b)

Figure 4.15 The output spectrum of the circuit of Figure 4.5: (a) Vin = 200 mVp_p; (b) V in = 400 mVp_p. 64

4.5.3 Series Compensation

When the SC amplifier using the series compensation shown in Figure 4.8 was simulated, the capacitor size was chosen as follows. The size of the unit capacitor was 26.4 pm x 26.4 pm, which corresponds to 1 pF. C1 consisted of the parallel connection of 10 unit capacitors, while C2 was a single unit capacitor. Since two Ci capacitors are connected in series, the total input capacitance is 5 pF. In the same manner, the total feedback capacitance is estimated as 0.5 pF. Hence, the total capacitance is the same as in the SC amplifier using the simple biasing or parallel compensation. Cs was a single unit capacitor and the capacitor size for Cdg was 8.4 pm x 8.4 pm. A CMOS transmission gate was used for the output node, a PMOSFET was used for a switch connected to Vb, which was equal to 1.5 V, and NMOS switches for the rest of the circuit.

Figure 4.16 shows the output spectrum of the SC amplifier illustrated in Figure 4.8. For a sinusoidal signal with an amplitude 200 mVp-p, the S/THD was 89 dB (Figure

4.16a). When the input amplitude is doubled, capacitors in this circuit still operate in their accumulation regions, and the S/THD of 82 dB was obtained (Figure 4.16b). To have an equivalent kT/C noise, the unit capacitance should have been four times larger here than for the simple biasing or the parallel compensation; hence the capacitor area increases considerably. However, the best linearity among these design techniques was achieved by the series compensation because the voltage swing across each capacitor is divided by half, and that results in a smaller operating region and hence less capacitor nonlinearity. For a comparison, the SC amplifier using the bottom-to-bottom connected series SC branch, shown in Figure 4.17, was also simulated using HSPICE. Capacitor sizes were the same as for the SC amplifier of Figure 4.8. NMOSFETs were used for all switches except for the one connected to the output, which was realized by a CMOS switch. 65

20 Series compensation Each ouput swings 2.Vp-p Vos = 5 mV 20

-40 ­

-60 ­

80 ­

-120­ ov,vpol'iA

140 20 40 60 BO 100 120 140 160 Frequency [kHz]

(a)

20

- 20

-80

- 100 IN\ Ti(iooksivIllow

- 120

1400 20 40 80 80 100 120 140 160 Frequency[kHz]

(b)

Figure 4.16 The output spectrum of the circuit of Figure 4.8: (a) Vin = 200 mVp_p; (b) Vin = 400 mVp_p. 66

Top plate I e I (poly gate) Cdg Cs I .--L­ I .01 T C2 C2 Bottom plate (accumulated channel) T+ V b CMOS switch

411,. 01 4)2 Vin Vout+ 0_ \ L' i ( 1 ) 1 ci C1 Vout­ NMOS switch for the rest

Figure 4.17 SC amplifier using the bottom-to-bottom-connected series compensation.

Figure 4.18 shows the output spectrum of the SC amplifier of Figure 4.17. For a sinusoidal signal with an amplitude 200 mVp-p, the S/THD was 72 dB and large harmonic components can be seen because the nonlinear parasitic capacitance connected to the MOSFET capacitors generated distortion as discussed in Section 4.4. 67

20

-20

-40 ­ V m

-60

-80

-100

-120

1400 20 40 60 80 100 120 140 160 Frequency [kHz]

Figure 4.18 The output spectrum of the circuit of Figure 4.17 with Vin = 200 mVp.p. 68

Chapter 5 PREDICTIVE CDS TECHNIQUE

In this chapter, predictive correlated-double-sampling (CDS) techniques will be presented. First, a background of predictive CDS will be described. Then, novel predictive SC amplifiers will be introduced. Next, MOSFET-only predictive SC amplifier stages based on the novel amplifiers will be shown. Finally, the usefulness of the predictive CDS technique will be verified by HSPICE simulations.

5.1 Background of the Predictive CDS Technique

In the SC amplifier shown in Figure 4.1, the DC gain is obtained from the formula

[21]

C2 1[1(z)lz=i 1+ (1+ ) 2 C2 (5-1)

[1_11+a),,2] C2 C? where 11 << 1 is the reciprocal of the DC gain of the op-amp. The second term in the bracket represents the DC gain error. Since the DC gain error is proportional to a square of the reciprocal of the op-amp DC gain, the requirement for the op-amp DC gain is greatly alleviated. This is called the gain-squaring effect [21]. The gain squaring is very effective when Vout(n-1/2) at the phase 01=1 and Vout(n) at the phase 02=1 are close to each other, or when the input frequency is much lower than the sampling frequency. The gain error increases with the input signal frequency, because the difference between Vout(n-1/2) and

Vout(n) widens. 69

For wideband applications, predictive SC amplifiers, such as the one illustrated in

Figure 5.1, are more suitable because they have a small gain error over a wide frequency range [27]. In Figure 5.1 we assume that C1 /C2=C3/C4. It was also assumed that the input signal is already sampled and held, and changes only when 01 goes high as shown in

Figure 5.2. During the phase 01, the predictive path containing capacitors C3 and C4 is active while the main path including capacitors Ci and C2 is open, and the SC amplifier output anticipates the output voltage of the next half clock cycle k. When 02 goes high, the predictive path is open while the main signal path is active, and the valid output is obtained.

Since the capacitors Ci and C2 are permanently tied to the virtual ground node, the main signal path can cancel the offset voltage and reduce the 1/f noise. Also, because the output voltage was already predicted during the previous half clock cycle (:01=1), the output voltage varies only slightly during the transition from 01 to 02 (Figure 5.2), and the effects of finite op-amp DC gain are greatly reduced. In addition, the op-amp does not require a high slew rate during the valid (02=1) phase.

The DC gain error of the circuit shown in Figure 5.1 is given [27] by

Ci C2 IH(z)lz=i = 1+ + 1242 C2 (5-2)

2 - Cl [1- (C1+ R2] C2 C2

Since in the second term bothii and (1+C1 /C2) are squared, and since Ci/C2 is larger than 1 in amplifiers, that makes the DC gain error larger than that of the non-predictive SC amplifier of Figure 4.1. However, the gain error is nearly independent of signal frequency, because the outputs at 41 and (02 are nearly equal to each other even for high signal frequency. Hence, when the input frequency becomes higher, the predictive SC amplifier can achieve a better gain accuracy than the non-predictive one. 70

Cdg Main path

(02 Vout

01 01 C3 C4

\-1---) I 1 C1 C3

_C (1)2 C2 C2 C4 Predictive path

Figure 5.1 Predictive SC amplifier [27]

(1)1 (02 (I)1 42 4) 1 02 (01 (02 tr-inn 1-1 El t/T

yin t/T t/T Vout

1

Figure 5.2 Signal waveform of the predictive SC amplifier

In ref.[28], it was shown that predictive SC amplifiers are also effective for reducing the nonlinear distortion caused by the op-amp. The output voltage of the predictive

SC amplifier in Figure 5.1 can be expressed as 71

1 vout(n)= vin (n)+ ve(n) (5-3) 2c where ve(n) is the output error voltage, given by

ve(n) = (1+ [v(n) (5-4) C2 2111

In equation (5-4), v(n) and v(n-1/2) are the voltages at the virtual ground node during 02=1 and 01=1, respectively. In the predictive CDS circuit, v(n)---v(n-1/2) because vout(n-1/2) is nearly equal to vout(n) and v(n)=-1,tvout(n). Hence, the error voltage ve(n) in equation (5-3) can be neglected in the predictive circuit. In a predictive circuit using nonlinear capacitors, the linearity of the circuit is determined by the coefficient of the first term of the RHS in equation (5-3), Ci/C2. By applying an appropriate compensation for these capacitors, the nonlinearity of C1 /C2 will be also suppressed, and consequently an output with low distortion can be obtained.

5.2 Novel Predictive SC Amplifier

Figure 5.3a shows a novel predictive SC amplifier [29] which has a simpler configuration than other predictive SC amplifiers in refs.[27] and [30]. This circuit is obtained simply by adding capacitor C3 between the input node and the node at the top plate of capacitor Cs in the non-predictive SC amplifier of Figure 4.1. Hence, it uses fewer switches than the circuit in Figure 5.1 and has a simpler configuration. Assuming that the input signal is sampled and held as shown in Figure 5.2, the DC gain error of this circuit is also given by the equation (5-2). Figure 5.3b shows another possible structure [29]. A switch is inserted between the input and the capacitor C3 in the circuit of Figure 5.3a. If capacitor sizes are chosen as

C1=C3 (and C2=C4), the capacitance seen from the input node is constant during (I)i and 02.

That is, the load capacitance seen by the previous stage is constant. That makes it easier to 72

Cdg 1 Main path 1

(1)1 i Cl (I)2 Vout

Predictive path / (a)

Main path Vin \ 02

Predictive/ path (b)

Figure 5.3 Novel predictive SC amplifiers: (a) simplified predictive SC amplifier; (b) improved predictive SC amplifier.

design the op-amp of the previous stage. By contrast, the load capacitance of the op-amp in this circuit (and of the circuit of Figure 5.3a) is not constant: C4 at 01; C2+C4 at 02. 73

However, C2 and C4 are smaller than Ci and C3 in amplifiers, and the variation of the load capacitance is usually not a major problem for op-amp design. A very important feature of this circuit is that the DC gain error is given by equation (5-1), hence it is the same as that of Figure 4.1 and smaller than those of Figures 5.1 and 5.3a. In addition, it also has a small gain error over a wide frequency range due to the predictive property. The reason why the circuit of Figure 5.3b works better than that of Figure 5.1 or Figure 5.3a is explained as follows. Assuming that the input signal is constant (DC), the output of the predictive amplifiers is shown in Figure 5.4. For the circuit of Figure 5.1, Vout (4)1) is determined by the charge Aq coming from Cl, C2 and C3 to C4 during 4)1=1, and by the voltage of the virtual ground node, i.e.,

V., (40 = + Vx (4)1) 4 where Vx is the voltage of the virtual ground node. Since the charges coming from Cl and

C2 cancel each other, Aq is equal to the charge coming from C3. Then Aq is given by Aq = C3(V,(4)i) Vin)

vou, (00 = 1(vx(4)1) via) + vx (4)1 4 C = C3 + 1+ (C) C4 in C4 x

Assuming that V,out (02) C4

AVout = Vold (01 ) \Tout ) (5-5) (1 + C: )Vx (11)1)

For the circuit of Figure 5.3a, AVout is also given by equation (5-5). By contrast, for the circuit of Figure 5.3b, AVout is given by

AV out = V.(01) (5-6) 74 because the charges coming from Cl and C2 cancel each other, and no charge is coming from C3 to C4. (Note that the input signal and Vx(01) are constant.) Hence, AVout is smaller in Figure 5.3b than in Figure 5.1 or Figure 5.3a. Since the voltage at the virtual ground node is less affected by the smaller AVout and it is stored on the capacitors Cl and

C2, the DC gain error becomes smaller in Figure 5.3b. Hence, the circuit of Figure 5.3b gives a better performance than that of Figure 5.1 or 5.3a for the DC input signal. And the predictive CDS helps all these predictive SC amplifiers to keep their gain errors small in the wide range of the input frequency.

01 02 (1)1 02 01 02 (1)1 02 IFIFIE1171-11-11-1El t/T

yin VT 1 t/T Vout # AVout 1

Figure 5.4 Signal waveform of the predictive SC amplifiers for a DC input

To verify the behavior of the gain error of these novel predictive circuits, they were simulated using SWITCAP. For comparison, the non-predictive and no-gain-compensated

SC amplifier of Figure 5.5 and the non-predictive gain-compensated SC amplifier of Figure 4.1 were also simulated. The conditions used in the simulations were as follows: DC gain of the op-amp = 40 dB; offset voltage = 5 mV; gain of the SC amp = 10; clock frequency =

312.5 kHz. To the three predictive SC amplifiers, the sampled-and-held input illustrated in Figure 5.2 was applied. The gain errors of these SC amplifiers are shown in Table 5-1 and

Figure 5.6. 75

Top (2nd poly) T Bottom plate (1st poly)

V in

Figure 5.5 Gain-uncompensated non-predictive SC amplifier

Table 5-1 Gain errors of SC amplifiers

SC amplifier Input frequency Input frequency 10.4 kHz 61.6 kHz

Gain-uncompensated non-predictive 0.91 dB 0.91 dB amplifier (Figure 5.5)

Gain-compensated non-predictive 0.04 dB 0.76 dB amplifier (Figure 4.1)

Gain-compensated predictive amplifier 0.09 dB 0.14 dB (Figure 5.1)

Gain-compensated predictive amplifier 0.10 dB 0.15 dB (Figure 5.3a)

Gain-compensated predictive amplifier 0.02 dB 0.15 dB (Figure 5.3b) 76

Gain error of SC amplifiers

0.8

-o

6 0.6

0.4

0.2

10 20 30 40 50 60 70 Input signal frequency [kHz] (fs = 312.5 [kHz])

Figure 5.6 Gain error of SC amplifiers: solid line for Figure 5.1; dashdot line for Figure 5.3a; dashed line for Figure 5.3b; + + Figure 5.5; x x Figure 4.1.

As can be seen from Figure 5.6 and Table 5-1, the gain-uncompensated non- predictive amplifier of Figure 5.5 has a constant but large gain error over the operating frequency range. The gain-compensated non-predictive amplifier of Figure 4.1 has a very small gain error at low frequency, however the error increases rapidly with the signal frequency. The gain errors of the gain-compensated predictive amplifiers of Figures 5.1 and 5.3a are nearly identical: the error is slightly larger than that of the non-predictive amplifier of Figure 4.1 at low frequency, and it remains small even for a high-frequency input signal. The circuit of Figure 5.3b has a very small error over a wide range, because of the gain squaring given by equation (5-1) and the predictive feature.

The key features of the three predictive amplifiers are summarized in Table 5-2. 77

Table 5-2 Summary of the key features of the three predictive amplifiers

Number SC amplifier DC gain error of Cin(4:11)=Cin(02) Cout(01)=Cout(02) switches

Figure 5.1 112(l +C i/C2)2 10 possible possible

Figure 5.3a -112(1+C1 /C2)2 6 not possible not possible

Figure 5.3b 112(1+C1 /C2) 7 possible not possible

5.3 MOSFET-Only Predictive SC amplifier

To realize the novel predictive SC amplifiers in a digital CMOS process, MOSFET capacitors using the parallel and series compensations may be used, as discussed next.

5.3.1 Parallel Compensation

Figure 5.7 shows a predictive SC amplifier based on the circuit of Figure 5.3b using the parallel compensation technique. In this circuit, the main path uses parallel

compensation while the predictive path is realized by only half of the parallel compensation circuit (Figure 5.7b), because the predicted output (during 01=1) does not have to be so accurate as the valid output during 02=1. The other predictive circuits (Figures 5.1 and

5.3a) can be also implemented using MOSFET capacitors, as shown in Figure 5.8 and 5.9,

respectively. For comparison, they will be simulated as well.

5.3.2 Series Compensation

Figure 5.10 illustrates predictive SC amplifiers using the series compensation

technique. In both circuits shown in Figure 5.10, the predictive path is realized by simple 78 biasing, because high accuracy is not required for the output in the predictive phase and also because the simple biasing requires less capacitor area.

5.4 Simulation Results

To verify the operation of the predictive SC amplifiers using MOSFET capacitors, the SC amplifiers were simulated using HSPICE. The parameters and conditions used in the HSPICE simulations are the same as in Section 4.5, unless otherwise specified. Figure 5.11a shows the output spectrum of the SC amplifier with the parallel compensation of Figure 5.7. A S/THD ratio of 101 dB was obtained for a 4 Vp-p differential output swing. The linearity is further improved compared to that of the non- predictive SC amplifier of Figure 4.5 (the S/THD ratio is 86 dB).

Figures 5.11b and c show the output spectra of the circuits of Figures 5.8 and 5.9.

The S/THD ratios were 96 dB and 98 dB, respectively. Among the three circuits, the circuit of Figure 5.7 had the best S/THD ratio. However, it is fair to say that these results are nearly equivalent because the few dB differences at this high linearity level, obtained in HSPICE, are not significant. Figure 5.12 shows the simulation results for the predictive amplifiers with series compensation shown in Figure 5.10. The S/THD ratios in Figure 5.12a and b were 97 dB and 99 dB, respectively. Again, these results are nearly the same.

In all of these predictive circuits, a very high linearity was obtained in spite of using nonlinear MOSFET capacitors. 79

Main path m I Vbp Cdg j /4)1 C2 1;(1)1 Vi- Y2 / \ ) IVbp 02 Cl

C3 C4

4)2 CH CH Predictive path

CH

Cl C2 4)2 Vont+

Cl C3 C2 C4 Vbn /°vout (a)

Vbp

r C3 `-`1 Vont Vin I I H-CI

01 1)2 Virtual ground (b) Figure 5.7 Predictive SC amplifier using parallel compensation: (a) half of the whole circuit; (b) predictive path 80

I K Cdg Main path Vbp

1 j 1 TO I C2 1-/-01 Vin- (02 // \ i ) /' Vbp K 02 Ci /

4)2 71 C3 C4 01 / ) 1 I E-1-\----1

021 CH 11)2 / CH Predictive path -=

CH

Ci 4:12 Vout+ \..___0 Ci - C3 C2 C4 Vbn 1 (t--°r2 Vout-

Figure 5.8 Predictive SC amplifier using parallel compensation based on the circuit of Figure 5.1 81

- Main path Vbp Cdg 01j C2 1;01 Vin

Vbp I/ 02

L 02 C3 C4 ) i K 02 CH CH Predictive path 01

CH

Cl C2 4)2 Vout+ 1 Ci C3 C2 C4 Vbn /--°Vout-

Figure 5.9 Predictive SC amplifier using parallel compensation based on the circuit of Figure 5.3a 82

Vbn Top plate Predictive path (poly gate) I 01 C3 02 C4 J_ \---I ( I' I )1 01 Bottom plate 2 C2 4)2 (accumulated channel)

(01 101 Vbp

Vbp .. To 4)1 yin­ 4)2 C1 I C 1 Vout+ 4 ) 4 1E__.____ 11)----0 0 Cl C3 1-7out­ Main path = (a) C2 C4

Vbn Top plate Predictive path (poly gate) I C3 0'2 C4 I 1( )1 -r Bottom plate (accumulated channel)

Vbp

,i, 01 Cu ,(1)1 y2 1 Vin 1 %-- 1 Vont+ 4 ) 4 l( 1P---0

ID 0

C3 I Vont­ Main/ path C1 C2 C4 (b) Figure 5.10 Predictive SC amplifiers using MOSFET capacitors with series compensation based on: (a) Figure 5.3b; (b) Figure 5.3a. 83

20

20

-40

-60

E -80

100

-120

140o 20 40 60 80 100 120 140 160 Frequency [kHz]

(a)

20

0

-20

-60

E -80

- 100

- 120

- 1400 20 40 60 80 100 120 140 160 Frequency [kHz]

(b)

Figure 5.11 The output spectrum of the circuit: (a) Figure 5.7; (b) Figure 5.8; (c) Figure 5.9. 84

20

20

40 co -o

a) 73 60 E 80

100

120

1400 120 140 160 20 40 60 80 100 Frequency [kHz]

(c) Figure 5.11 continued 85

20

20

-40

-60­

80

- 100

- 120 W

- 140 0 20 40 60 80 100 120 140 160 Frequency [kHz]

(a)

20

20

40

60 ­

80

-100

- 1201 Vr$

- 1400 20 40 60 80 100 120 140 160 Frequency [kHz]

(b)

Figure 5.12 The output spectrum of the circuit: (a) Figure 5.10a; (b) Figure 5.10b. 86

Chapter 6

EXPERIMENTAL RESULTS FOR SWITCHED-CAPACITOR AMPLIFIERS

The SC amplifiers described in Chapters 4 and 5 were fabricated in the Orbit 1.2 CMOS process. In this chapter, their experimental results will be shown. The test

circuit, op-amp, and clock generator circuit will also be described.

6.1 Test Circuit

Figure 6.1 shows the test circuit used for SC amplifiers. Since the SC amplifiers are fully differential, a single-ended-to-differential converter was needed at the front end of the

test circuit. It consists of two inverting op-amps, and it also works as a low-pass filter. The

loss is 0.2 dB at 20 kHz. Linear ceramic capacitors (NPO capacitors*) were used in this test

circuit to avoid signal distortion. A 0.01 µF NPO capacitor was used from each input pin of the chip to the analog ground plane, and a 0.0047 .1.F NPO capacitor was placed between the input pins to

minimize the RF transmission and reception capability of the inputs [31]. These capacitors

will not degrade the input signal. Several bias voltages needed by the op-amp were externally applied to the SC

amplifier chip.

A buffer was connected to the output of the SC amplifier followed by a differential­ to-single-ended converter. The buffer was added to drive the low input resistance of the converter because the on-chip op-amp in the SC amplifier cannot drive a resistive load. Figure 6.2 shows the differential-to-single-ended converter without its 100 pF NPO

capacitors. For a

* Originally an NPO capacitor is for low temperature-coefficient use. Generally it has a low voltage- coefficient, too. Single-ended to SC Buffer Differential to single-ended Spectrum differential converter amplifier converter analysis

tr) 180 p (NPO) CN1 ,--; 6 c?

I F-­ II II 11 r.1Z.1"1 100 p Vin 10 k 10 k 0.0111 I (NPO) 30 (NPO) 111_ 100 p 30 (NPO) Spectrum analyzer 50 560 LM833 560 1-1 1004711. SC LM833 A/D 180 p (NPO) (NPO) = amp converter NEXT LM833 Workstation chip Vout AD1671 LM833 10 k 10 k 50 100p

0.0112 (NPO) vLM833 30

0.01,,

CLK

Figure 6.1 Test circuit for SC amplifiers 88

100 pF

I

R2

R1 Vout

R3 LM833

CI: parasitic op-amp input capacitance RI: 560 Q

Figure 6.2 Differential-to-single-ended converter. high-frequency signal, the voltage at node A rolls off due to R2 and parasitic capacitance CI. To prevent this, a 100 pF NPO capacitor was connected in parallel with R2, and also in parallel with the other 560 CI to have an equivalent impedance for a high-frequency signal. In addition, a 30 CI resistor is connected in series with each 100 pF NPO capacitor to prevent oscillation in the op-amp in the converter due to the large capacitive load. After the differential-to-single-ended conversion, the output signal is evaluated by the spectrum analyzer. Also, it was fed into a 12 bit A/D converter. The digital output was fed into a NEXT workstation, then FFT analysis was performed by MATLAB. 89

6.2 Op-Amp

Figure 6.3 shows the fully differential folded-cascode op-amp [32] which was used in the SC amplifiers. Since this is a single-stage op-amp, internal frequency compensation is not necessary. However, it requires a common-mode feedback (CMFB) circuit to keep the output common-mode voltage at the middle of the supply voltages. The CMFB circuit was realized by a switched-capacitor circuit [33]. In most of these SC amplifiers, the CMFB of the op-amp was realized using double-poly capacitors to prevent op-amp nonlinearity. In one SC amplifier, MOSFET capacitors were used in the CMFB of the op­ amp for comparison. In this case, the top and bottom plate of capacitors were interchanged, as shown in Figure 6.4, so that all MOSFET capacitors had a positive dc bias voltage and operated in their accumulation region. 90

Vdd = 2.5 V

600/1.2 1200/1.2 600/1.2 1.2 V

600/1.2 1200/1.2 600 1.2 0.5 V

0 11 t­

in+ in­ 600/1.2 600/1.2

600/1.2 600/1.i

600/1.2 60071.21

Vss = -2.5 V

)I ( 1 pF 1 pF 01) \

)1 IC 0.25 pF 0.25 pF

4)

Vrefl = 0 V Vref2 = -1.3 V Vrefl = 0 V

Figure 6.3 Folded-cascode op-amp [32] 91

out- out+

IC )1 0.25 pF 0.25 pF

02

Vref1 = 0 V Vref2 = -1.3 V Vrefl = 0 V

Figure 6.4 Common-mode feedback circuit using MOSFET capacitors

6.3 Clock Generator

Figure 6.5 shows the clock generator circuit. For the last two inverters, analog VDD and Vss were used. For the rest, digital VDD and Vss were applied. The layout of the clock generator circuit is shown in Figure 6.6. 4)2

PAD 3.6/1.2 3.6/1.2 3.6/1.2 10.8/1.2 10.8/1.2 32.4/1.2 32.4/1.2

CLK

3.6/1.2

3.6/1.2 3.6/1.2 3.6/1.2 10.8/1.2 10.8/1.2 32.4/1.2 32.4/1.2

DVDD, DVSS AVDD, AVSS Figure 6.5 Clock generator circuit IMMINSIIIMMINIIMINE p"..., Elf 1=11 ...::;:....illEl liliJ ni3 NEAP 1.1:1I nit:1111NQ r .. ituquejinegm : Uli! IJIIIIIENRI III rAtilitAllikif Vi[ri 1217 rilfant'LliffIlL]lif:j --atlYLNErilhlr aiiil­ ,-,r,/ //,/,,,,,, //./ // -`..X;''' // /lr -1M-TM-1E! 13,. it1101:1615tIt1191 4. 1 ii .1 :31. y,lifFRITIFI=IRIE"); IP 11111111111111140111111 lifiliiiiiiliTin NI.,,,,..! ea i irRin la `11.711EiiittlirillP11121 1 :,,,..,:-,,,,,..nirriii, , WILL al6a-L am MINN =IA / Mil $ MS Liii gallgi lit i iii /7/ ice', Iii liMii .1 111 4)2

Figure 6.6 Layout of the clock generator circuit 94

6.4 Experimental Results

Six kinds of SC amplifiers, described in Table 6-1, were fabricated in the Orbit 1.2 CMOS process. Their die photos and pad locations are shown in Figures 6.7-6.9. A spectrum analyzer was used to obtain the experimental results described in Section 6.4.1­ 6.4.9. Section 6.4.10 gives the results obtained by using an A/D converter and NEXT workstation.

Table 6-1 SC amplifiers on chip

Capacitor compensation Capacitor CMFB of Figure SC amplifier code name method array op-amp

(1) 6.8 DP Without compensation Double-poly Double-poly

(2) 6.7 SCOMP_1 Series compensation MOSFET Double-poly

(3) 6.8 SCOMP_2 Series compensation MOSFET MOSFET

(4) 6.7 SB Simple biasing MOSFET Double-poly

(5) 6.9 PCOMP_1 Parallel compensation MOSFET Double-poly

(6) 6.9 PCOMP_2 Parallel compensation MOSFET Double-poly (Predictive SC amp)

6.4.1 Non-Predictive SC Amplifier with Simple Biasing

Figure 6.10 shows the output spectra measured using the spectrum analyzer of the simpler circuit of Figure 4.2 with Vb = 0 V and 1.5 V. The input signal (Vin in Figure 6.1) was a 1 kHz, 200 mVp-p sinewave. Hence, after the differential-to-single-ended conversion, the output signal was a 4 Vp-p sinewave. This was a large amplitude for the +/- 2.5 V bias voltages used. The clock frequency was 16 kHz. Without the bias voltage, the output waveform was totally distorted (Figure 6.10a). With a bias voltage of 1.5 V

(Figure 6.10b), the largest harmonic component was 58 dB below the input signal. Thus, it was shown that the simple biasing could reduce the nonlinearity caused by the MOSFET 95 capacitors. In Figure 6.10b, we can see both the second and third harmonic components at the same level. Even though this SC amplifier is a fully differential circuit, the second harmonic component was not cancelled. It is likely that it was due to the effect of the offset voltage of the SC amplifier.

a ma caloadmag-a.mt a tTal

II I JX/ :ff 11.1111.01110. AC (> 7Il NII 11W11:1470MMEICNIIIIii

I , I S 1.174,4.1 1 r F

(a)

Figure 6.7 SC amplifiers (1)(3): a) die photo; b) floor plan. 96

SC amplifier using double- poly capacitors (DP)

VDD Vref2 Vrefl Vn2 Vinp VSS nip Vn2 Vrefl Vref2 VDD

E1 14 9 8 U 6

16 4

Folded­ Vpl cascode Vp1 18 Folded- op-amp cascode op-amp Vp2 19 fc_op_2 Vp2 fc_op_l Vinn 20 40 Vinn MOSFET capacitor 39 Vb Eu Double-poly array with capacitor the series Vcont array compensation 38

Voutp2 Voutpl

Clock Voutnl Voutn2 24 generator

VSS 26 Ea 28 29 30 El VSS Clk Gnd

MOSFET-only SC amplifier using the series compensation (SCOMP_2)

(b)

Figure 6.7 continued. 97

(a) Figure 6.8 SC amplifiers (2)(4): a) die photo; b) floor plan. 98

SC amplifier using the simple biasing (SB)

VDD Vref2 Vrefl Vn2 Vinp VSS inp Vn2 Vrefl Vref2 VDD

cc 14 El le el 9 8 si 6 B

16 4

iii Folded­ la cascode Vp1 18 Folded- op-amp in Vpi cascode op-amp fc_op_ 1 Vp2 19 Ei Vp2 fc_op_ 1 Vinn 20 40 Vinn MOSFET MOSFET capacitor 39 Vb al capacitor array with array with the series the simple compensation 38 biasing Voutp2 Es Voutpl Clock Voutnl Vouln2 24 generator

VSS sa 26 Ea 28 29 30 El VSS Clk Gnd

SC amplifier using the series compensation (SCOMP_1)

(b)

Figure 6.8 continued. 99

(a)

Figure 6.9 SC amplifiers (5)(6): a) die photo; b) floor plan. 100

Non-predictive SC amplifier using the parallel compensation (PCOMP_1)

VDD Vref2 Vrefl Vn2 Vinp VSS inp Vn2 Vrefl Vref2 VDD

111 14 Ea ma al 9 8 El 6 li 16 Fl ni Folded­ u cascode Vp1 Folded­ Vp1 18 op-amp cascode op-amp Vp2 19 fc_op_1 II Vp2 fc_op_1 Vinn 20 MOSFET 40 Vinn capacitor array Vbn MOSFET with 39 Vbp El capacitor the parallel array with compensation Ea the parallel (with 38 compensation predictive path) Voutp2 Voutp1 111 El

Voutn2 Clock Voutnl 24 generator

VSS E1 26 Ea 28 29 30 au VSS C lk Gnd

Predictive SC amplifier using the parallel compensation (PCOMP_2)

(b)

Figure 6.9 continued. 101

REF 10.0 dBm MARKER 5 240.0 Hz 10 dB/DIV RANGE 10.0 dBm -81.9 am 0

-50dB

-100dB START .0 Hz STOP 20 000.0 Hz ROW 30 Hz VBW 100 Hz ST 44.6 SEC =16 kHz (a)

REF 15.0 dBm MARKER 4 880.0 Hz 10 dB/DIV RANGE 15.0 am -78.7 dBm

-50dB

-100dB START .0 Hz STOP 20 000.0 Hz ROW 30 Hz VBW 100 Hz ST 44.6 SEC

(b) fs = 16 kHz

Figure 6.10 The measured output spectra of the SC amplifier using simple biasing: a) Vb = 0 V; b) Vb = 1.5 V; 102

6,4,2 Non-Predictive SC Amplifier with Series Compensation

Figure 6.11 shows the output spectrum measured with the spectrum analyzer of the circuit of Figure 4.8 using the series compensation scheme, with Vb = 1.5 V. Now the largest harmonic component was 72 dB below the fundamental of the input signal; the signal-to-THD ratio was 70 dB. Thus, the linearity of the circuit was further improved by the series compensation scheme. In Figure 6.11, we see higher-order harmonic components. It is likely that they were caused by the interactive effects of the op-amp and

MOSFET capacitor nonlinearities.

Figure 6.12 shows the output waveform observed on an oscilloscope.

6.4.3 Non-Predictive SC Amplifier Using Double-Poly Capacitors

Figure 6.13 shows the output spectrum measured using the spectrum analyzer of the circuit of Figure 4.1 using double-poly capacitors, with Vb = 1.5 V. Now the largest harmonic component was 70 dB below the fundamental of the input signal; the signal-to- THD ratio was 69 dB. Therefore, a large part of this small remaining nonlinear distortion

can be attributed to the op-amp, which had a narrow linear output voltage range.

6.4.4 Non-Predictive SC Amplifier with Parallel Compensation

Figure 6.14 shows the output spectrum measured using the spectrum analyzer of the circuit of Figure 4.5 using the parallel compensation scheme, with Vbp = -Vbn = 2.5 V. The largest harmonic component was 57 dB below the signal level, which is nearly the same as the result obtained by the simple biasing arrangement. However, HSPICE simulations show that a S/THD of 86 dB can be obtained using an ideal op-amp. The post- layout simulation showed that the S/THD is limited to 67 dB because of the small linear output range of the folded-cascode op-amp used in the chip. 103

;5.0 d8m MARKER 1 780.0 Hz ICI dB/DIV RANGE 15.0 am -75.5 dEm 0 r

-50dB

START .0 Hz STOP 20 000.0 Hz REL.' 30 Hz VOW 100 Hz ST $4.6 SEC

fS = 16 kHz

Figure 6.11 The measured output spectrum of the SC amplifier using the series compensation with Vb = 1.5 V.

rrj fl 4

4-4g, V

44^ 1 ms

Figure 6.12 The output waveform of the SC amplifier using the series compensation with Vb = 1.5 V. 104

REF 00.0 :Sr, 000.,) 10 dE'0::.1 S.0 -7a.E cDr, 0

-50dB

Wt,40,4,011.44/44-0410,14EAAAtJ,4 '4`,144 -100dB 7,.-flF =0 000.0 R:7

;7.0, 7t L4.5 F.,

fs = 16 kHz Figure 6.13 The measured output spectrum of the SC amplifier using double-poly capacitors.

REF ;5.0 dErn HARKER b 460.0 Hz iO E/Dis) RAi4GE iS.0 dErn -72.4 dEA 0

57 dB

-50dB

1 1 a

r 1 11 1

1 1 I

1 I1 i 1 1 t i 1 II n LIn 4jilifittftklaiii.kikilikilifi iiiikh 1111101161 ir II, , -AI, 'fly r-opere 0 41i killi tti)lomi , , r I rrlirlrrit 14 r Yi 1 r

-100dB START .0 Hz STOP 20 000.0 Hz REW 30 Hz RF5W 100 Hz ST 44.6 SEC fs = 16 kHz Figure 6.14 The measured output spectrum of the SC amplifier using the parallel compensation with Vbp = -Vbn = 2.5 V. 105

6.4.5 Predictive SC Amplifier with Parallel Compensation

Figure 6.15 shows the output spectrum measured with the spectrum analyzer of the predictive SC amplifier using the parallel compensation scheme shown in Figure 5.7, with Vbp = -Vbn = 2.5 V. Its largest harmonic component was 66 dB below the signal level.

Thus, the predictive SC amplifier had a 9 dB linearity improvement compared to the non- predictive SC amplifier of Figure 4.5. The post-layout simulation by HSPICE also showed that a 9 dB linearity improvement can be obtained by the predictive SC amplifier compared to the non-predictive one.

PEF­ 1 5.0 dBm HiIRKER 3 000.0 Hz EiCI.; FiriNCE i5.0 dEm -50.8 dOm 0

-50dB

-100dB STFIRT .0 Hz STOP 20 000.0 Hz new DO Hz VEW TOO Hz ST 44.6 ]Cl. fs = 16 kHz

Figure 6.15 The measured output spectrum of the predictive SC amplifier using the parallel compensation with Vbp = -Vbn = 2.5 V. 106

6.4,6 Output Swing (Inverting Mode)

Table 6-2 shows the largest harmonic components and S/THD ratios obtained using a spectrum analyzer HP 3585B for a differential output swing of 2 Vp-p and 4 Vp-p. An input signal frequency was 1 kHz and a clock frequency was 16 kHz. All SC amplifiers were in their inverting mode.

Table 6-2 Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz

Largest harmonic component STIED ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -79 dB -70 dB 76 dB 69 dB

SCOMP_1 -67 dB -71 dB 64 dB 67 dB

SCOMP_2 -70 dB -64 dB 66 dB 62 dB

SB -71 dB -58 dB 68 dB 57 dB

PCOMP_1 -56 dB -57 dB 55 dB 56 dB

PCOMP_2 -73 dB -66 dB 68 dB 65 dB

For most of these results, an S/THD ratio decreases as an output increases. For example, an S/THD of SC amplifier DP is limited to 69 dB for 4 Vp-p output swing. It was caused by the small linear output range of the op-amp. In SCOMP_1, the common-mode feedback (CMFB) circuit of the op-amp consists of double-poly capacitors and the rest of the SC amplifier were MOSFET capacitors. On the other hand, in SCOMP_2, all capacitors in the SC amplifier were MOSFET capacitors including the CMFB circuit of the op-amp. For a large output swing, SCOMP_1 shows a better result than SCOMP_2. This is because in SCOMP_2 no specific compensation but 107

simple biasing was given in MOSFET capacitors of CMFB circuit, and these capacitors

were out of the linear (accumulation) region when the output swing became large. For an output swing of 2 Vp-p, SC amplifier SB showed comparable results with SCOMP_1 and SCOMP_2 because the nonlinearity of MOSFET capacitors were considerably smaller in this operating range. However, the linearity of SB is much lower

than that of SCOMPs for a large signal swing. Parallel compensation was not effective in this testing. For 4 Vp-p, PCOMP_1 is only comparable to SB. For 2 Vp-p, PCOMP_1 was much worse than other amplifiers. The linear output range of the op-amp should be fine, therefore the capacitance was limiting

the performance. However, once a predictive path is added, the linearity is much improved.

At 2 Vp-p voltage swing, the largest harmonic component of PCOMP_2 was 14 to 17 dB

smaller than that of PCOMP_1.

6.4.7 Frequency Dependence (Inverting Mode)

To investigate a frequency dependence of these circuits, measurements were also performed with an input signal frequency of 100 Hz and 10 kHz. The test results are shown in Table 6-3. The results for an input signal frequency of 100 Hz were nearly the

same as those for 1 kHz. However, linearity worsened at 10 kHz even for the SC amplifier using double-poly capacitors. Since MOSFET capacitors are operating in their accumulation region, these capacitors should be just as frequency independent as the double-poly capacitors. Therefore, it is presumed that the degradation of linearity at 10 kHz were caused by the op-amp slew rate and/or the on-resistance of the switches. 108

Table 6-3 (a) Input signal frequency (fin): 100 Hz; clock frequency (fs): 1.6 kHz

Largest harmonic component STIED ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -79 dB -71 dB 75 dB 71 dB

SCOMPi -65 dB -73 dB 62 dB 67 dB

SCOMP_2 -71 dB -66 dB 67 dB 64 dB

SB -72 dB -58 dB 69 dB 57 dB

PCOMP_1 -56 dB -58 dB 55 dB 57 dB

PCOMP_2 -73 dB -63 dB 69 dB 62 dB

Table 6-3 (b) Input signal frequency (fin): 10 kHz; clock frequency (fs): 160 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -66 dB -57 dB 65 dB 57 dB

SCOMP_1 -67 dB -53 dB 65 dB 51 dB

SCOMP_2 -61 dB -52 dB 60 dB 50 dB

SB -66 dB -55 dB 63 dB 54 dB PC OMP_1 -55 dB -51 dB 54 dB 50 dB

PCOMP_2 -69 dB -59 dB 66 dB 58 dB

6.4.8 Non-inverting Mode

Table 6-4 shows the experimental results for SC amplifiers in the non-inverting mode. Again, the largest harmonic components and S/THD ratios are shown for various input signal frequencies and output swings. 109

Table 6-4 (a) Input signal frequency (fin): 100 Hz; clock frequency (fs): 1.6 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -76 dB -66 dB 74 dB 66 dB

SCOMP_2 -69 dB -66 dB 65 dB 64 dB

Table 6-4 (b) Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz

Largest harmonic component STIED ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -72 dB -64 dB 71 dB 63 dB

SCOMP_2 -68 dB -67 dB 65 dB 63 dB

Table 6-4 (c) Input signal frequency (fin): 10 kHz; clock frequency (fs): 160 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -57 dB -51 dB 57 dB 50 dB

L;COMP2 -57 dB -54 dB 56 dB 50 dB

For SC amp using double-poly capacitors, the output waveform has double steps as shown in Figure 6.16 because during 01=1, charge coming from the input capacitor and charge coming from the feedback capacitor do not cancel each other. In addition, the output voltage keeps changing as the input voltage changes. 110

11, 1 ail

'114i

Figure 6.16 The output waveform of the non-inverting SC amplifier using double-poly capacitors.

By contrast, for the MOSFET-only SC amplifier SCOMP_2, the output waveform is kept constant even during the reset phase 01 as shown in Figure 6.17. This is because charge coming from the input capacitor and charge coming from the feedback capacitor have opposite polarities and the same amount of charges. Hence they cancel each other, and no charge is flowing into the supporting capacitor Cs in Figure 4.8. Therefore, CDS is less effective for SC amplifier DP in the non-inverting mode compared with the inverting mode. And hence, experimental results in the non-inverting mode were not so good as those in the inverting mode. By contrast, the MOSFET-only SC amplifier SCOMP_2 in the non-inverting mode has a comparable linearity with that in the inverting mode because CDS is working effectively. 111

Figure 6.17 The output waveform of the non-inverting SC amplifier using series compensation.

6.4.9 Continuous Output Waveform without Clock Signal

SC amplifier circuits were evaluated without having clock signals. Suppose that all circuits have inverting-mode configuration. After working in a regular condition for a while, the clock signals were changed in such a way that 02 was kept high while 01 remained low. Then these amplifiers worked as continuous-time inverting amplifiers, as shown in Figures 6.18 and 6.19. Since there is no DC path to their virtual grounds, the op­ amp saturates within 20 to 30 seconds after the clock was stopped, and then these circuits could not work properly anymore. However, that period was enough to store the output waveforms for FFT analysis. Table 6-5 shows the linearity of SC amplifiers without clock signals. The input signal frequency was 10 kHz. 112 4 Regular operation 4-0 Clock stopped

4)2

I 1 1 (' Cdg C2 Vin- ci (vbp)

1 T CH

(Vbn) Ci Vout+ 0 Vout-

Figure 6.18 SC amplifier using the parallel compensation at 02=1.

Table 6-5 Input signal frequency (fin): 10 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing D ifferential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -83 dB -73 dB 79 dB 72 dB

SCOMP_2 -84 dB -72 dB 81 dB 71 dB

SB -69 dB -57 dB 67 dB 56 dB

PCOMP_1 -80 dB -72 dB 77 dB 70 dB

PCOMP_2 -84 dB -76 dB 81 dB 73 dB 113

Figure 6.19 SC amplifier using the series compensation at 02=1.

In these experiments, the S/THD values were nearly the same for both SC amplifiers, those with double-poly capacitors and those with MOSFET capacitors using series compensation: 79 dB and 81 dB for vout = 2 Vp-p, and 72 dB and 71 dB for vout = 4 Vp-p, respectively. (MOSFET capacitors seems to have obtained a better result than double-poly, however, the differences are considered to be in the error range.) The limiting factors in these results are the signal source and the op-amp. The linearity was much better for a 2 Vp-p output swing than for 4 Vp-p. The S/THD ratio was 72 dB for DP (double-poly capacitors), and it is presumed that it was limited by the linearity of the op-amp. Notice that for the SCOMP_2 (series compensation) and PCOMP_1 (parallel compensation), the S/THD ratios were comparable to that of the DP. In these two circuits, a pair of MOSFET capacitors compensate each other's capacitance variations due to voltage change. 114

The S/THD ratio of SB is only 56 dB for a 4 Vp-p output swing, because the

capacitance variation due to the voltage swing is not compensated in the SB.

Notice again that the predictive amplifier PCOMP_2 is better than the non-predictive one PCOMP_1. This can be explained as follows. Once (02 is kept high level, the resulting circuit is the same for both PCOMP_1 and PCOMP_2. However, during the regular clocked operation, the potentials of their virtual grounds are slightly different from each

other because they are related to their output voltage. (PCOMP_2 has a better linearity than

PCOMP_1.) Therefore, in each amplifier, the error voltage at the virtual ground is stored on the capacitors connected to the virtual ground. There is no charge flowing in or out of the virtual ground node after 4)2 becomes constantly high. Therefore, based on the error

voltage at the virtual ground, the linearity of SC amplifier is determined. It is even better than for the DP because predictive CDS can reduce the nonlinearity caused by the op-amp

128].

6.4.10 1.1-4 Analysis Using an A/D Converter and NEXT Workstation

Experimental results described in previous sections were measured using a spectrum analyzer. Hence, all undesirable effects such as glitches and slewing were incorporated in the spectra.

To eliminate these effects from the measured spectra, another spectrum analysis was completed. The data acquisition system is shown in Figure 6.20. The output waveform of SC amplifier was sampled by an A/D converter (AD1671) at the end of the valid phase 02, and it was converted into 12-bit (parallel) digital data. Then the parallel digital data were converted to serial data and taken into a NEXT workstation for data acquisition. Finally, the

1+ 1 analysis was accomplished by using MATLAB.

First, a sinusoidal signal from the signal source Tektronix SG5010 was fed into the A/D converter and it was tested how much linearity can be obtained using the A/D converter. Its result is shown in Table 6-6. 115

5V

To SC amp test board 16 kHz Analog data from SC amp test board -2.5 V 0 v

QAQB QCQD Ain ENCODE T A/D converter FG5010 CLK LOAD Counter AD 1671 fs=256 kHz C 74163 CARRY MSB LSB OUT ABCD bl b8 b9 b12 5V D1 D8 D1 D4 OWL Latch 256 kHz CLK CLK 74374 Q8 Q1 Q4

To NEXT Computer

Dl D8 D1 D4 D5 D8

,-, Shift register Shift register DS QH DS 74166 74166 0 LOAD CLK 0 LOAD CLK ]

Figure 6.20 Data acquisition using an A/D converter 116

Then, FET analyses of the SC amplifiers DP, SCOMP_1, SCOMP_2, and SB were performed. The output waveforms of the SC amplifiers are shown in Figure 6.21. In the inverting mode, the output still changes during the phase 02, and hence clock jitter can cause distortion when it is sampled during 02. By contrast, the phase 01 is not valid but stable, hence the jitter effects can be avoided. Measurements in the inverting mode were completed using both phases: Table 6-7a shows the test results sampled at the end of I:02; Table 6-7b shows results sampled at the end of 01. In the non-inverting mode, 4)2 is the valid and stable phase as shown in Figure 6.21. Table 6-8 shows experimental results for outputs sampled at the end of 02.

No big difference was found in these results compared with those obtained by the spectrum analyzer. Also, the results taken during 01 were nearly the same as those taken during 4)2.

From these experimental results, it is concluded that linearity of these SC amplifiers were limited not by the nonlinearity of MOSFET capacitors, but by other factors such as nonlinearity of the op-amp.

Table 6-6 Sinusoidal signal from Tektronix SG5010: the input signal frequency (fin) was 1 kHz.

Largest harmonic component S/THD ratio Differential output signal swing Differential output signal swing Signal source 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

SG5010 -75 dB -85 dB 70 dB 80 dB 117

Inverting mode

02: Valid phase 01: Stable phase

02 4)1 02 (01

Non-inverting mode

Figure 6.21 Output waveforms and timing of sampling time 118

Table 6-7 (a) Inverting mode sampled at the end of 02; Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing D ifferential ou tP ut signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -75 dB -72 dB 70 dB 71 dB SCOMP_1 -58 dB -64 dB 55 dB 61 dB

SCOMP_2 -64 dB -65 dB 62 dB 61 dB

SB -65 dB -59 dB 60 dB 58 dB

Table 6-7 (b) Inverting mode sampled at the end of 01; Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -66 dB -70 dB 59 dB 67 dB

SCOMP_1 -68 dB -58 dB 61 dB 57 dB

SCOMP_2 -64 dB -67 dB 60 dB 66 dB

SB -65 dB -56 dB 60 dB 55 dB

Table 6-8 Non-inverting mode sampled at the end of (1)2; Input signal frequency (fin): 1 kHz; clock frequency (fs): 16 kHz

Largest harmonic component S/THD ratio SC amplifier Differential output signal swing Differential output signal swing code name 2 Vp-p 4 Vp-p 2 Vp-p 4 Vp-p

DP -66 dB -66 dB 62 dB 66 dB

SCOMP_2 -66 dB -65 dB 62 dB 62 dB 119

6.4.11 Noise

1/f noise is a major factor which limits the S/N ratio in SC circuits. As already described in Section 2.3.2, 1/f noise is caused by interface-trapped charges (surface state charges) due to the interruption of the periodic lattice structure at the surface of the silicon

[5]. In a MOSFET capacitor, the periodic lattice structure is interrupted between the bottom silicon bulk layer (single crystalline) and gate-oxide layer (amorphous), hence interface- trapped charges can cause 1/f noise. In a double-poly capacitor, interface-trapped charges are also generated because phases changes between the bottom poly layer (poly-crystalline) and insulated oxide layer (amorphous). Therefore, the 1/f noise level is about the same for an SC amplifier using double-poly capacitors as for the one using MOSFET capacitors.

In experimental results, no difference in noise level was observed between double- poly capacitors and MOSFET capacitors, and it was shown that the 1/f noise of MOSFET capacitor are at the same level as for double-poly capacitors. 120

Chapter 7

OPERATIONAL AMPLIFIERS IN DIGITAL CMOS TECHNOLOGY

In this chapter, operational amplifier (op-amp) configurations for realization in a digital CMOS process are discussed. First, the earlier work will be summarized. Then, a folded-cascode op-amp used in our chips will be described. Finally, a novel two-stage op­

amp using MOSFET capacitors will be proposed.

7.1 Earlier Work

In a digital CMOS process, the op-amp also has to be implemented without the second poly layer. A digital-compatible op-amp was proposed in ref. [34]. As shown in Figure 7.1, the authors used a set of PMOSFET and NMOSFET capacitors in the frequency compensation branches of a class-AB op-amp so that their nonlinearities could be compensated. However, the threshold voltage of an NMOSFET capacitor in this circuit is increased because of the body effect, and that results in a C-V curve shift. Each time the output voltage changes, the threshold voltage of the NMOSFET capacitor changes as well as its C-V curve. Tho-efore, the nonlinearity of the PMOSFET and NMOSFET capacitors in this configuration cannot be fully cancelled. The experimental S/THD was only 55 to 65 dB for a 10 kHz signal.

7.2 Folded-Cascode Op-Amp

In our fabricated chips, a folded-cascode op-amp was used. As shown in Chapter

6, the MOSFET capacitors in the CMFB circuitry generate distortion when the output signal has a large swing and they operate out of their accumulation region. 121

Metal _2 1,777771

Metal 1

p-substrate

Figure 7.1 A digital- process compatible op-amp [34]: (a) whole op-amp schematic; (b) practical implementation of Miller compensation by MOSFET capacitors. 122

Figure 7.2 shows a post-layout simulation of the output spectrum of the MOSFET- only SC amplifier using the folded-cascode op-amp of Figure 6.4. In this op-amp, the

CMFB was realized by MOSFET capacitors. For a single-ended output, large-order even

harmonic components are observed (Figure 7.2a). However, since it is a fully differential circuit, these harmonic components are canceled at the differential output (Figure 7.2b).

By applying series compensation to the CMFB, the linearity at a single output can

be improved. Figures 7.2c and d show a post-layout simulation of the SC amplifier using the folded-cascode op-amp of Figure 7.3. For single-ended output, the linearity is improved compared to that of Figure 7.2a as shown in Figure 7.2c. However, for the differential output (Figure 7.2d), the S/THD ratio was the same as in Figure 7.2b under ideal symmetry conditions.

Figures 7.2e and f show the post-layout simulation of the SC amplifier using the folded-cascode op-amp of Figure 6.3. Here, the CMFB was realized by double-poly capacitors. As shown in Figures 7.2e and f, the linearity was as same as that shown in Figures 7.2c and d even if linear capacitors are used. By contrast, when the op-amp is

replaced by an ideal op-amp, the linearity is much improved. Therefore, it is likely that the linearity of the SC amplifier is limited by the relatively small linear output range of the folded-cascode op-amp due to the cascoded output . 123

Single-ended output Fully differential output

o

-60

80

00

-120

-140 40 60 80 100 120 140 160 60 80 100 120 140 180 Frequency 11H41 Frequency NW) (a) (b)

(c) (d)

-40

-60

-80

- 100

- 120

-140 80 100 120 140 160 20 40 60 80 100 120 140 160 60 Frequency (kHz' Frequency (khal

(e) (0 Figure 7.2 HSPICE simulation results of the SC amplifier: (a)(b) op-amp of Figure 6.4; (c)(d) op-amp of Figure 7.3; (e)(f) op-amp of Figure 6.3; (a)(c)(e) single output; (b)(d)(f) differential output. 124

Vdcl = 5 V

600/1.2 I 1200/1.2 600/1.2 1.2 V

600/1.2 I 1200/1.2 600/1.2 0.5 V T

out- out+ -.0-i in­ 600/1.2 _600/1.2 -0.65 V

in:_16 00/1.2 600/1.21

--I 1 600/1.2 600/1.2 1 Vss = -2.5 V 2pF 2pF 2pF 2pF

)1 IC

/ 4)1\ 4)1 0.5 pF 0.5 pF 0.5 pF 0.5 pF 1( ) IC ) 1

(1)) (1)2

Vb = 2.0 V Vref = -1.3 V V = 2.0 V

Figure 7.3 Folded-cascode op-amp using MOSFET capacitors with series compensation 125

7.3 Two-Stage Op-Amp

It was found that the folded-cascode op-amp used in our past chips had a relatively narrow linear output range. To widen the linear output range of the op-amp, a two-stage op-amp with a class-A output stage was designed. The two-stage op-amp requires frequency compensation to avoid oscillation, and since we are attempting to fabricate our chips in a digital CMOS process, the capacitors used in the frequency compensation also had to be MOSFET capacitors, rather than double-poly ones.

Figure 7.4 shows a two-stage op-amp with the Miller compensation capacitor Cc. In

this op-amp, the dominant pole, the non-dominant pole and the zero are given by [35]

1 P1 gnu' R1R11 Cc

grim P2 = (7-1) CII

gmII Cc

respectively, where grail is the transconductance of the second stage, and RI (R11) is the output resistance of the first (second) stage. Notice that the zero is in the right-half-plane, and hence it degrades the phase margin of the op-amp. (In a bipolar op-amp, the transconductance of the is large enough to place the zero at a very high frequency in the right-half-plane so that the zero does not affect the phase margin. However, in a CMOS op-amp, the transconductance of the transistor is much smaller, hence the phase margin may be degraded. Therefore, the right-half-plane zero has to be shifted to a very high frequency or to the left half plane.) In addition, if the capacitor is realized by a

MOSFET capacitor, it will operate out of the accumulation region when the output swing becomes large. 126

Two frequency compensation methods are widely used to move the right-half-plane zero to the left half plane or remove it altogether. One is to use both a resistor and a capacitor in the feedback branch, as illustrated in Figure 7.5.

VDD

Vout

Vss Figure 7.4 Op-amp compensation (1)

VDD

Vout

vss Figure 7.5 Op-amp compensation (2) 127

The dominant and non-dominant poles are given by (7-1), and the zero and an additional pole are given by [35]

z= 1 1 Cc Rz) gam (7-2)

1 P3 = RzCi

where CI is the capacitance to ground seen from the output of the first stage. By controlling

Rz, the zero z and pole P2 can be made to cancel each other. The other frequency compensation method is to use a buffer and a capacitor as shown in Figure 7.6. Again, the dominant and non-dominant poles are given by (7-1), and

the zero and another pole is given by [35]

z= 1 Ro Cc (7-3) 1 P3 = Ro CI

where Ro is the output resistance of the buffer amplifier. This method is convenient for a digital CMOS process because the unity-gain buffer (source follower) provides a dc bias voltage for the MOSFET capacitor, and thug keeps it in its accumulation region. The practical realization of this compensation is illustrated in Figure 7.7. A fully differential configuration of this op-amp is shown in Figure 7.8 [36]. Because the output impedance is relatively small in the two-stage op-amp, resistors can be used for the CMFB. Although this op-amp was fabricated in such a way that the CMFB resistors were connected to the output nodes Voutp and Voutn as shown in Figure 7.8, these resistors could have been connected to the nodes Voutp' and Voutn' because the impedances of these nodes are further smaller than those of Voutp and Voutn. 128

VDD

Vout

Vss Figure 7.6 Op-amp compensation (3).

VDD

Vss

Figure 7.7 Practical realization of Figure 7.6. Voutp Vcmin Voutn

20 k 20 k

5011 5011 5011 60011 5411 VDD 200[1. 5411 600[1. 20011 336/1.8 84/1.8 F­ 336/1.8

120/4.8 504/1.8 2/1.8 504/1.8 120/4.) Voutn' voutn 0 Voutp 108/12 Vcmref k_. 12q2] 120/1.2 Vin/071 jinn 126/1.2 I` 120/1.2 294/3 294/3 126/1.2 IL IL .51 378/1.2 15/12 15/12 15/12 378/1.2

21/3 21/3 21/3 21/3

Vss

Figure 7.8 Fully differential two-stage op-amp using MOSFET capacitors 130

Table 7-1 summarizes the characteristics of the op-amp simulated using HSPICE. The Bode plot is shown in Figure 7.9. In this Figure, two curves are shown for the positive and negative output signals.

60.017 PLS__RDAMP7. YOBCVOUTP

V 50.0 VE113(VOUTN 0 '40.0 2 0 30.0

20 .0

1 10.0

0. 7

E-­ 1 1 1 111111 1 1111111 1 1111111 1 1 111111 1 1 11111 PLS_ADAHP7. 150.0 ­ VP(1/OUTP . V VP(YOUTN , 100.0 7

50.0 \. 2

0.

-50.0.7

-100.0

1 1111111 1 111,11 1 1111111 11.1 11,11 149.139 1.0K 10.0K 100.0K 1.0X io.ox 100.0 HERTZ CLOG) 100 .ox

Figure 7.9 Bode plot of the two-stage op-amp

Table 7-1 Characteristics of the op-amps

Op-amp of Figure 7.4 Op-amp of Figure 7.8 using linear capacitors using MOSFET capacitors

DC gain [dB] 67 67

Gain-bandwidth [MHz] 20 28

Phase margin [deg] 73 99 131

7.4 Simulation Results

To compare the op-amps described above, the SC amplifier of Figure 4.8 incorporating these op-amps was simulated using HSPICE. The simulation results are shown in Table 7-2 and Figure 7.10.

Table 7-2 Comparison of the SC amplifier performances with various op-amps

Peak harmonic Op-amp component below Signal/THD [dB] the signal [dB]

(a) Macro-model op-amp (ideal op-amp) -94 88

(b) Folded-cascode op-amp; -71 68 CMFB using linear capacitors

(c) Two-stage op-amp; compensation -89 85 using linear capacitors (Figure 7.4)

(d) Two-stage op-amp; compensation -89 85 using MOSFET capacitors and source followers (Figure 7.8)

Clearly the use of the two-stage op-amp with MOSFET capacitors improves the linearity of the series-compensated SC amplifier by more than 15 dB compared to the folded-cascode op-amp which was employed in the previous chips.

7.5 Layout and Post-Layout Simulation

An SC amplifier with series compensation was laid out using the two-stage op-amp, and the chip is in the process of fabrication. The layout plot of the two-stage op-amp is shown in Figure 7.11.

Post-layout simulations by HSPICE show that an S/THD of 91 dB can be obtained for the predictive SC amplifier using series compensation.

Experimental results will be described in Section 8.4.1. 132

-1C0

-120

-1400 20 GO 10 100 122 140 140 Froquercy11012)

(a) (b)

20

0

-20

IT -4°

I-40

-100

-120

-140 20 40 b 40 100 120 140 0 20 40 SO 110 Ho 120 140 Frequency110121 2,442/4.2 WWI

(c) (d)

Figure 7.10 HSPICE simulations of the SC amplifier by various kinds of op-amp: (a) macro model; (b) folded-cascode op-amp; (c) two-stage op-amp with linear capacitors; (d) two-stage op-amp using MOSFET capacitors 133

Figure 7.11 Layout of the two-stage op-amp 134

Chapter 8 DIGITALLY CONTROLLED GAIN/LOSS CIRCUIT

In this chapter, the design of a MOSFE,T-only digitally controlled gain/loss circuit will be described. First, the research background will be summarized. Then, the design

details of the MOSFET-only digitally controlled gain/loss circuit will be presented. Finally, simulation and experimental results will be provided.

8.1 Research Background

Digitally controlled gain/loss circuits are useful in telecommunication and other applications. For example, when the amplitude of a signal is too large in a telecommunication channel, the amplitude has to be reduced by an attenuator. By contrast, the input signal to an A/D converter may need to be amplified to increase the dynamic range. The gain/loss circuit functions as a multiplying D/A converter (MDAC) by generating the product of the analog input signal and the digital signal [17]. If the input signal is a constant reference voltage, the circuit operates as a D/A converter.

Baldwin and McCreary used precision capacitors for their digitally controlled analog attenuator [37]. In Figure 8.1, 2k-1C denotes the feedback capacitor, and all other capacitors can be switched either to the input or to the output node depending on the digital word. The loss of this circuit is given by

Vout _ N (8-1) Vin 2k N where N=1, 2, ..., 2k-1. Therefore, e.g. for k=4, its loss can be 1/15, 2/14, 3/13, ..., 1.

Babanezhad and Gregorian utilized resistors for their programmable gain/loss circuit [38] as shown in Figure 8.2. Gain or loss can be selected by a digital word, and the magnitude of gain (or loss) is determined by other digital words controlling a multiplexer. 135

2k-1C

vout Cl Cl 12k 2C oo3 Vim

Figure 8.1 Digitally controlled attenuator [37]

GAIN Vin

GAIN

Figure 8.2 Digitally controlled gain/loss circuit [38] 136

8.2 Circuit Design

We have designed a MOSFET-only 4-bit digitally controlled gain/loss circuit [36]. This is useful not only for the applications described above, but also for verifying that

series compensation is effective for large signal swings at both input and output nodes. For simple explanation, Figure 8.3a shows a one-bit gain/loss circuit. When the GAIN option is selected, this circuit works as an amplifier with a gain of 8 as shown in Figure 8.3b. If LOSS is chosen, the circuit of Figure 8.3a operates as an attenuator with a loss of 1/8 as illustrated in Figure 8.3c. To utilize MOSFET capacitors in the circuit, all capacitors in the signal paths are series compensated. Figure 8.4 shows the details of a MOSFET-only digitally controlled gain/loss circuit, and Figure 8.5 illustrates its control circuits. In the circuit of Figure 8.4, the binary signal d3 is the MSB, and dO is the LSB of the control input word. Digital signal G determines whether the circuit works as an amplifier or as an attenuator: when G is 1, the circuit is an amplifier whose gain varies from 1 to 15 with a step of 1, while for G =O, it is an attenuator with a loss of 1/15, 1/14, 1/13, ..., 1. Since both input and output swings can be large, CMOS transmission gates are used in all switches connected to the input and output nodes.

This gain/loss circuit was laid out using the two-stage op-amp described in Section 7.3. For comparison, an SC amplifier using double-poly capacitors and the two-stage op­ amp was also laid out. Figure 8.6 shows the layout and floor plan of this circuit. 137

$2 cs

GAIN- 02 GAIN. 4)2

vin o 1 8)YC(---.114( LOSS' 4)2 LOSS' 4)2 (1)1. (01 (1)1 0- - Vb Vb \-40 Vin- vout+ vout+

Vout­

1 (a)

(I)2 V> cs

(01

4)2 8C 8C C C 1)2 vin 01),. (01 - Vb Vb vout+

V o ut

yin­

(c)

Figure 8.3 (a) One-bit gain/loss stage; (b) amplifier; (c) attenuator. 138

top plate (poly gate)

1._ T Cs bottom plate (channel)

yin- A T OAO OH (I) 1-i VOUt+ 4131 0 M'''''.. Vb

7473.0 OA 1 2C 2C 1 T111 Om 01-41171-1 OB 1 --filT Vb (1)B I 40A21 4 -I2­

T (1)B2 4>A31 ETT (1)A3 diT­ I. 4:1B3

Figure 8.4 MOSFET-only digitally controlled gain/loss circuit 139

G. di di G. di G. di G. di I I VDD

Vs s G- di

6- di 6. di ,, T. di G. di v DD I _±.

6. di 4Bi IG

Vss

G G G VDD I I_ I T VSS T G G

OD OD

Figure 8.5 Control circuit 140 do di d2 d3 G Gd3 Gd3

Gd2 Gil d2

Gdo

6°d3 G d3

G d2 60 d2

G di Gdi

d Zdo

Figure 8.5 continued 141

W4:

(a)

Figure 8.6 Digitallly controlled gain/loss circuit: (a) layout; (b) floor plan. 142

SC amplifier using double-poly capacitors

VDD P2 P1 VDD

E1 14 so 9 8 El 6 EI

16 4

is Two-stage Bi op-amp 18 Two-stage ts_op_1 op-amp 19 ts_op_1 u Digitally- 20 40 Vion controlled Vinn MOSFET Ei Double-poly capacitor 39 Vb capacitor array with array the series m compensation 38 G Voutp2 el El Vontpl Clock Voutn2 24 generator 36 Voutn1 Vss El 26 Ea 28 29 30 au El Ei Vss Clk Gnd dO dl d2 d3

MOSFET-only digitally controlled gain/loss circuit using the series compensation

(b) Figure 8.6 continued. 143

8.3 Simulation Results

Figure 8.7 shows the S/THD ratios for various gains and losses obtained by post- layout simulations using HSPICE. In this figure, the amplitude of the fully differential input signal was 0.4 Vp-p for the amplifier mode, and it was 4 Vp-p for the attenuator mode. In Figure 8.7a, the S/THD ratio decreases as the gain increases. When the gain increases, so does the output signal level. However, the increase of the voltage across the capacitor in the feedback path induces a distortion because the nonlinear operating range of the MOSFET capacitor widens. As the gain increases, the increase of distortion is faster than that of signal, hence the S/THD ratio decreases. On the other hand, in Figure 8.7b, the S/THD ratio increases slightly as the loss increases. In this attenuator configuration, the amplitude of the input signal is much larger than the output swing, and hence the nonlinearity caused by the input capacitors is the dominant factor for generating distortion rather than the nonlinearity caused by the feedback capacitors. The increase of the output signal is larger than that of the distortion; therefore, as the amplitude of the output signal increases, the S/THD ratio also increases. 144

Amplifier mode (Vin = 0.4 Vp-p)

95

07 90 li 85 v) 80

75 2 5 10 15 Gain

(a)

Attenuator mode (Vin = 4 Vp-p)

95

90 a7 775 a) 85 ' 80

75 1/15 1/10 1/5 1/2 Loss

(b) Figure 8.7 Post-layout simulation results by HSPICE (a) S/THD vs. gain; (b) S/THD vs. loss. 145

8.4 Experimental Results

The SC amplifier using double-poly capacitors and the MOSFET-only gain/loss circuit was fabricated in the Orbit 1.2 p.m CMOS process. Figure 8.8 shows the die photo

of this chip. In the following experimental results, the signal source HP8903A and the spectrum analyzer HP4195A were used. The test circuit and other experimental conditions

were similar to those described in Chapter 6.

Figure 8.8 Die photo of the digitally-controlled gain/loss circuit.

8.4.1 SC Amplifier Using Double-Poly Capacitors

Figure 8.9a shows an output spectrum of the SC amplifier using double-poly capacitors for a fully differential output of 2 Vp-p. The input signal was a 1 kHz sinewave 146

SPECTRUM A:REF B:REF A MKR 3 999.999 Hz -30.00 -10.00 LIMAG -81.379S dB 0 dB dBm 4MAG

-50 dB

- 100dB DIV DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBI.4: 30 Hz ST:1.98 min RONGE:R=-20,T=-10dBm RBW=_3Q HZ (a)

SPECTRUM A:REF B:REF A MKR 2 000.000 Hz -30.00 -10.00 AMOG -77.2099 dB 0 dB C dBrn IC I JMAG

-50 dB

- 100 dB o 4, DIV DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBW: 30 Hz ST:1.58 min RANGE:R=-20,T=-10dBm RBW=_30 HZ (b)

Figure 8.9 The measured output spectrum of the SC amplifier using double-poly capacitors: a) Vout = 2 Vp-p; b) Vout = 4 Vp-p. 147 and the clock frequency was 16 kHz. The largest harmonic component was -81 dB below the fundamental signal, almost hidden in the noise level, and a S/THD ratio of 76 dB was obtained. For a 4 Vp-p output swing, the third harmonic component was -77 dB below the signal, and the S/THD was 76 dB (Figure 8.9b). For a 6 Vp-p output swing, the S/THD ratio decreased to 68 dB. It is likely that the distortion was caused by the nonlinearity of the op-amp. However, compared to the folded-cascode op-amp used in previous chips, this two-stage op-amp has a better linearity. For example, in Sec. 6.4.6, the SC amplifier with double-poly capacitor had a S/THD ratio of 69 dB for the output of 4 Vp-p. Hence, in this case, the two-stage op-amp contributed 7 dB improvement in linearity over the folded­ cascode op-amp.

8.4.2 MOSFET-Only Gain/Loss Circuit (Amplifier Mode)

First, the MOSFET-only digitally controlled gain/loss circuit was tested in the amplifier mode. A bias voltage of 1.5 V was applied to keep all MOSFET capacitors in their accumulation regions. The amplitude of the input signal was set in such a way that the amplitude of the output signal would be 2, 4, and 6 Vp-p. The test results are summarized in Table 8-1 and Figure 8.10. Figure 8.11a shows the output spectrum of the gain circuit with a gain of 10 for a fully differential output of 2 Vp-p. The largest harmonic component was -78 dB below the signal level, and a S/THD of

74 dB was obtained. For a 4 Vp-p output swing, the third harmonic component was -73 dB below the signal, and the S/THD was 71 dB (Figure 8.11b). These few dB discrepancies compared to the results of the SC amplifier using double-poly capacitors can be attributed to the nonlinearity of the MOSFET capacitors. For a 6 Vp-p output swing, the S/THD ratio decreases to 63 dB. It is likely that this distortion was caused by both the limited output range of the op-amp and the nonlinearity of the MOSFET capacitors. 148

Table 8-1 S/THD ratios of the MOSFET-only gain circuit

Vout Gain = 1 Gain = 2 Gain = 5 Gain = 10 Gain = 15 2 Vp-p 76 dB 75 dB 75 dB 74 dB 73 dB 4 Vp-p 80 dB 74 dB 72 dB 71 dB 70 dB 6 Vp-p 75 dB 65 dB 63 dB 63 dB 63 dB

Amplifier mode Vb = 1.5 V

85

80 IL

FEIn 1:2, 75 t 40Vout = 2 Vp-p -Vout = 4 Vp-p 70 -A-- VOUt = 6 Vp-p ;75 65 A ­ 60

1 2 5 10 15 Gain

Figure 8.10 S/THD ratios of the MOSFET-only amplifier

The linearity of the gain circuit improves when the circuit operates with a unity gain. A S/THD ratio of 76 dB and 80 dB were obtained for 2 Vp-p and 4 Vp-p output swings, respectively, as shown in Figure 8.12. A S/THD of 75 dB was obtained even for a 6 Vp-p output swing while that of the SC amp using double-poly capacitors was only 68 dB. In addition, for a gain of unity, its S/THD ratios are better for 4 Vp-p than for 2 Vp-p, while the S/THD ratios are better for 2 Vp-p than for 4 Vp-p in other gain configurations with a gain of 2 to 15. The reason is considered as follows. In general (such as for a gain of 2 to 15), the distortion increases as the output swing increases, because the operating range of MOSFET capacitors widens. In addition, because the minimum voltage across the

MOSFET capacitors decreases, the nonlinearity caused by the capacitors becomes larger. 149

SPECTRUM A:REF B:REF A MKR 3 999.999 Hz -30.00 -10.00 AMAG -78.4689 dB OdB [ dBm ][ AMAG

50 dB

-100dB DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBW: 30 Hz ST:1.68 min RANGE:R=-20,T=-10dBm RBW=_30 HZ (a)

SPECTRUM A:REF B:REF A MKR 2 000.000 Hz -30.00 -10.00 AMAG -72.9513 dB 0 dB [ dBm ][ ] AMAG

50 dB

-100dB DD, DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBA: 30 Hz ST:1.58 min RANGE:R=-20,T=-10dBm RBW=_30 HZ (b)

Figure 8.11 The measured output spectrum of the MOSFET-only SC amplifier with a gain of 10: a) Vout = 2 Vp-p; b) Vout = 4 Vp-p. 150

SPECTRUM A:REF B:REF J MKR 2 000.000 Hz -30.00 -10.00 JMAG -80.2387 dB 0 dB E dBm JE ] LIMAG

50 dB

100dB DIV DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBA: 30 Hz ST:1.58 min RANGE:R=-20,T=-10dBm PBW=_30 HZ (a)

LFEOTRUN A:PEF B:REF J MKR 1 000.000 Hz -30.00 -10.00 JMAG -85.5278 dB ] LIMAG OdB E dBm 11

50 dB

: I

100dB .1,-411111 P DIV DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBw: 3Q Hz ST:1.58 min RANGE:R=-20,T=-10dBm RBA= - 30 HZ (b)

Figure 8.12 The measured output spectrum of the MOSFET-only SC amplifier with a gain of unity: a) Vout = 2 Vp-p; b) Vout = 4 Vp-p. 151

By contrast, a gain of unity is a special case because the voltage across the input capacitor and the output capacitor becomes equal. If the nonlinearities of the two capacitors match well, their capacitances also become equal even if they are nonlinear, and then their nonlinearities cancel each other. In this special case, linearity degradation by increased voltage swing seems to be insignificant. On the contrary, an increased signal level helps to improve the S/THD ratio in the unity-gain SC amplifier. Thus, the linearity of the unity- gain configuration is extremely good, even better than that of the double-poly SC amp with a gain of 10.

8.4.3 MOSFET-Only Gain/Loss Circuit (Attenuator Mode)

Next, the MOSFET-only digitally controlled gain/loss circuit was tested in the attenuator mode. The amplitude of the input signal was set at 2, 4 and 6 Vp-p. The experimental results are summarized in Table 8.2 and Figure 8.13. Their linearity was lower than those obtained in the amplifier mode. Since an attenuator reduces the signal level, the S/THD ratio reduces as the output swings smaller, or as the loss gets smaller. In an attenuator with a loss of 1/10 for a fully differential input of 2 Vp-p, the largest harmonic component was -58 dB below the signal, and it was nearly the same as the noise level. The S/THD ratio was 55 dB. For a 4 Vp-p input, both the signal and the third harmonic increased, and its S/THD was nearly the same level as for 2 Vp-p. When the circuit operated with a loss of 1/2, a S/THD ratio increased to 67 dB and 72 dB for a 2 Vp­ p and 4 Vp-p input swing, respectively. In the unity-loss configuration, the linearity of the circuit further improved as in the case of the equivalent unity-gain configuration. When the input amplitude is constant for various loss values, the nonlinearity caused by the input capacitors dominated the distortion level because the input swing was much larger than the output swing, and nonlinearity caused by the op-amp and output capacitors was negligible compared to the nonlinearity caused by input capacitors as long as 152 an output swing was not very large. Thus, a larger output signal contributed to a better linearity. Hence, the linearity of the loss circuit improved as the loss approached 1.

Table 8-2 S/THD ratios for the MOSFET-only loss circuit

Vin Loss = 1/15 Loss = 1/10 Loss = 1/5 Loss = 1/2 Loss = 1 2 Vp-p 52 dB 55 dB 62 dB 67 dB 75 dB 4 Vp-p 55 dB 55 dB 67 dB 72 dB 81 dB 6 Vp-p 58 dB 57 dB 64 dB 63 dB 74 dB

Attenuator mode Vb = 1.5 V

90

80 as 72_, 70 Vin -= 2 Vp-p -0 -Vin = 4 Vp-p e- ­ 60 ---A-- Vin = 6 Vp-p 50 40

1/15 1/10 1/5 1/2 1 Loss

Figure 8.13 S/THD ratios of the MOSFET -only attenuator

8.4.4 Linearity Dependence on the Bias Voltage

So far, all experimental results of the MOSFET-only gain/loss circuit were achieved with a bias voltage of 1.5 V to enable a large output swing. If the bias voltage increases to

2.0 V, the amplitude of the fully differential output and/or input will be limited to 4 Vp-p: otherwise a pn junction in PMOS switches will turn on and charges stored on MOSFET capacitors will be lost. (A solution to increase a bias voltage while keeping a large signal 153 swing will be described in Chapter 10.) However, by applying a larger bias voltage, MOSFET capacitors will be kept in their accumulation region more deeply so that some linearity improvement will be expected.

Table 8-3 and Figure 8.14 show the S/THD ratios of the MOSFET-only gain circuit with Vb = 2.0 V and Vout = 4 Vp-p as well as with Vb = 1.5 V for comparison. In the entire gain range, we can see linearity improvements with increasing Vb. Figure 8.15a shows the output spectrum for a gain of 10. The largest harmonic component was -81.7 dB

Table 8-3 S/THD ratios of the MOSFET-only gain circuit with Vb=1.5 V and 2.0V (Vout = 4 Vp-p)

Vb Wil Gain = 1 Gain = 2 Gain = 5 Gain = 10 Gain = 15 1.5 80 dB 74 dB 72 dB 71 dB 70 dB 2.0 82 dB 78 dB 77 dB 76 dB 74 dB

Amplifier mode (Vout = 4 Vp-p)

85 II, 80 -It._ _ - - it °4v 75 ---*--Vb = 1.5 V M --E--Vb = 2.0 V 70

65

60

1 2 5 10 15 Gain

Figure 8.14 Linearity dependence on bias voltage (Vout = 4 Vp-p) 154

SPECTRUM A:REF B:REF J MKR 1 000.000 Hz -30.00 -10.00 AMAG -81.6583 dB OdB [ dBm JMOG

- 50 dB

* , I AMlifikr)A1* tr4IMVOSY114/101 ,::,1 :

. . -100dB : . DIV DIV START 0.n81 Hz to.00 in.00 STOP 10 000.000 Hz RPW: 30 Hz ST:1.58 min RANGE:R=-20,T=-10dBm RDW=_30 HZ (a)

SPECTRUM A:REF B:REF J MKR 1 000.000 Hz -30.00 -10.00 JMOG -86.90E6 dD OdB [ dBm ][ JMAG

- 50 dB

-100dB *Om 001240w DIV DIV START 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz RBW: 30 Hz ST:i.c..P, min RONGE:R=-20,T=-10dBm REW=_30 HZ (b)

Figure 8.15 The measured output spectrum of the MOSFET-only SC amplifier with Vb = 2.0 V: a) gain = 10; b) gain = 1. 155 below the signal level, and its S/THD was 76 dB, which is the same as the S/THD for the SC amplifier with double-poly capacitors. Figure 8.15b shows the output spectrum for a gain of 1. The harmonic components are almost hidden in the noise, and the S/THD was 82 dB. Table 8-4 and Figure 8.16 show a comparison of S/THD ratios of the MOSFET-only gain circuit with Vb = 1.5 V and 2.0 V at Vout = 2 Vp-p. The improvement of linearity for

Vout = 2 Vp-p was not so outstanding as for Vout = 4 Vp-p, however, a slight increase of

S/THD was observed.

Table 8-4 S/THD ratios for the MOSFET-only gain circuit with Vb=1.5 V and 2.0 V (Vout = 2 Vp-p)

Vb [V] Gain = 1 Gain = 2 Gain = 5 Gain = 10 Gain = 15 1.5 76 dB 75 dB 75 dB 74 dB 73 dB 2.0 77 dB 77 dB 75 dB 76 dB 71 dB

Amplifier mode (Vout = 2 Vp-p)

*Vb = 1.5 V -m -Vb = 2.0 V

Figure 8.16 Linearity dependence on bias voltage (Vout = 2 Vp-p) 156

Figure 8.17 shows a comparison of S/THD ratios for the MOSFET-only gain circuit with Vout = 2 Vp-p and 4 Vp-p at Vb = 2.0 V. In the case of Vb = 2.0V, the S/THD ratios were equal or better for 4 Vp-p than for 2 Vp-p even with gains of 2 to 15. The reason is as follows. All MOSFET capacitors are biased deeply enough to operate in their linear region even for a 4 Vp-p swing. Thus, a larger signal level contributes to a better linearity.

85

80

75 = 2 Vp-p -. -Vout = 4 Vp-p 70

65

60

1 2 5 10 15 Gain

Figure 8.17 S/THD ratios for the MOSFET-only amplifier with Vb = 2.0 V

Figure 8.18 and Table 8-5 show the linearity dependence on bias voltage for the MOSFET gain circuit. For a gain of 10, the output was totally distorted with Vb = 0.5 V. As the bias voltage Vb increases to 1.0 V, a S/THD of 59 dB was obtained. However, this was still not enough to keep the MOSFET capacitors in their deep accumulation regions. By contrast, for a gain of 1, a S/THD was 78 dB even for Vb = 1.0 V. As already described, the nonlinearities of the input and output capacitors cancel each other in the unity-gain configuration, thus this high linearity can be obtained even with a small bias voltage. 157

Table 8-5 Linearity dependence of the MOSFET-only gain circuit on the bias voltage (S/THD ratio at Vout = 4 Vp-p)

Gain Vb= 0.5 V Vb =1.0 V Vb= 1.5 V Vb = 2.0 V

1 51 dB 78 dB 80 dB 82 dB

10 18 dB 59 dB 71 dB 77 dB

Dependence on bias voltage

100 80 --­ Fr5' MO. wININ 77) 60 il..... ---*-- Gain ,-- 1 40 -s -Gain = 10 ... /' 20

0 0.5 1.0 1.5 2.0 Bias voltage Vb [V]

Figure 8.18 Linearity dependence on the bias voltage for the MOSFET gain circuit (Vout = 4 Vp-p)

Table 8-6 and Figures 8.19-8.22 summarize the experimental results with different gains and Vb. When applying a bias voltage of 2.0 V, the linearity of the MOSFET-only gain-controlled amplifier with gain of 10 was equivalent to that of the SC amplifier with double-poly capacitors. For a gain of 1, further improvement in linearity was obtained. 158

Table 8-6 S/THD ratios for SC amplifiers with double-poly capacitors and with MOSFET capacitors

Circuit Gain Vb Vout = 2.0 V Vout = 4.0 V scar! 2,1Nyitp_91211 10 o V 76 dB 76 dB

MOSFET-only gain-controlled amp 10 2.0 V 76 dB 76 dB

MOSFET-only gain-controlled amp 10 1.5 V 74 dB 71 dB

MOSFET-only gain-controlled amp 1 2.0 V 77 dB 82 dB

MOSFET-only gain-controlled am p 1 1.5 V 76 dB 80 dB 159

FFATI1r1 :FE,. rRuri A:EEF REF A NV.F 1 000.000 H: 3:PEF F.REF J rftJ. 000.000 Hz ',0 00 :1,00 AA111tA -215.5276 dE -16.22 10.00 INAH dE I ,a, 1 J03'-, .1Em II AMH0

85 dB 76 dB

W.0410,6y, (V( '0111 voosi,01,Lkv'eY'f''''0W116s1* 0.001 Hz DIV DIV START 0.001 Hz I 00 10 1_10 II U33 10 .0 10.00 10.00 STOP 10 000.000 Hz FPH: 30 H= F,1 E0 12311u11F--D0,1 ---10dP. RBW: 30 Hz ST:1 S8 min RHNGE:F1-20,T--10dBm FEU- IFO H­ RBW-_30 HZ Gain=1 Gain=2

SPEFTRAM 3RECT200 A,REE E1PEE J FIEF 2 000.000 01REF F:PEE A IIFF 2 000.000 H: -30.00 10.001 An:1,S -72 :1 1 Se -30.00 -10.00 AWAG -73.9113 dE 11 11-13u. dEm 11 J113G

172 dB 73 dB

4,1e. r Ak'ihr41104"

'1-1F 0.001 OJ 1 THE T 0.001 10 000.000 H: 10 00 0 10 U0 11 00 IU 000.00U H: mIr. F0H0E1E--.-?U.F--11J3F0, EF1A: 5n 11: -1:1 =.10 ,In P3NF,E1P--ZO.T--10.11.3, F.EW: 30 Hz 1 9 PEW-30 P 1 1 . 1 = 3 2 30: Gain=5 Gain=10

9PECTROW H:EACF E A ERF 2 202 -000 3: 50.00 ID -O6 JI-IU 1630 dF

I dEm .1 JFIFIG

72 dB

040ykitoolfoflfrwv110,114i\miJi

DIV 1HFT 0.001 H.: 1 0 . 0 0 1 0 . 0 0 .1011 10 030 000 H: .,..aa. T = 1 thj i ,,, i.po: =a) 14: I: I ..1.1 m, I, oorq,,F:: FPF1=JO Ji-:. Gain=15

Figure 8.19 Measured output spectra in amplifier mode: Vout = 4 Vp-p; Vb = 1.5 V. 160

/- I Ir'l ,PLLTPUN

A REF L . ji to 000 Hz -,:011 B'. REF j ru,r., t 000.000 ri, 111.00 -10 .00 I11R6 01.13010 11 11

IE lt 11-1Hc, I -IE fl. 11 illHH

86 dB 83 dB

s,,Niclm9p100008t010* 't,11'1,\HV\4Vitiv,c),111*Ny0,11(.2141,0ovet)o,ifp,

DIV DIV '.1.-TORT 0.001 Hz DIV DIV 0.001 Hz 10.00 10.00 0.1-,-.1F 10 000.000 HE 10.00 1E1.00 TIFF 10 000.000 Hz RBW, 30 Hz '-:-.T:I.,8 ,In PHNGE,P=-20.T-10dEm RE,14-30 HZ Gain=1 Gain=2

:FL. 'FUN 0:PER P:PEF HrR 00.000 H, R:PR, 01L1 -30.00 -lo.no Mr,. 01 7,0 00 -10 00 ,11 ] I dDrn 1( 1 ,000 I d iNHE

81 dB 82 dB

l'iq\*lYk114\"rtIrl'''°:"Wq

DIV DIV START 0.001 Hz Di: THisT 0.0U1 H: 10.00 10.00 ".TOE 10 000.000 Hz IC 0:1 10.CU tti 1 10u 111 PH.]: -50 Hz '0-1:1.00 min P,:.INGE:1-'=-20,7=-10dPm FEU: 50 0: '77,1 ;,7, R REk- 30 HZ PPl.1-30 HZ Gain=5 Gain=10

f-''ECTRUM 943 10800 8:800 A NKR nrirq H7 -30.00 -10.00 MIHG ,^0.0-109 dB

1 clEm IC ] JNi00

78 dB

:,:Yiilli(*01014,711011frW0d

0.0E11 He 10.00 10 -00 10 000.000 0, RBI-I­ 3E1 Hz I .S9 mtn RONGE F.­ z'O. T 1 0dPH RE4.1-_30 HZ Gain=15

Figure 8.20 Measured output spectra in amplifier mode: Vout = 4 Vp-p; Vb = 2.0 V. 161

,DE,. 14.00 SPECTRUM i 6[PE! I 'NI 01-1­ i no, 000 H: A:REF B,REF 3 NKR Z 000.000 H­ :30 00 10 00 _]m,:.6 ,1,1 01001 dE 30.00 -10.00 3NAG 7.1..1912 c1F.

[ ,IL,: I f dEm If I JMAG

85 dB 74 dB

1 1 t

I j, 1, 0. 11 I

Ei [0 U U01 ZI0PT U.001 Hz 10 00' 19 U UOU 91,1, He OU 10.00 [zi0P 10 000.000 Hz lUzIEH F[E1.0 0-10 Hz ':1,1.1.H PHNC,L,1.--LO.T--110dP. PE:1-1-110 9: PD,1--_01 HI Loss=1 Loss=1/2

-.0E1 11-00 0PECTRUN iF:FLEF I H-F- 010..1 000 1i­ r[0PEF E0PF.F 71 01:F: 2 000.000 H, .1[ -3000 -10.00 3110G -D7.10-11E ilk

( dE, jE JHHG

71 dB 57 dB

NY,W4'0(4

1.101F1 0001 Hz DIV [1.T0RT 0.001 10.00 10 00 11111- 10 000.000 Hz 10.00 10.00 STOP 10 000.000 Hz FKW, 'ICI 9, _T:1.S6 F0[4111E:P=-20.1i--10c1Fr. PEW, 70 Hz ST:1.50 RHNGE,R--20,T--10dEm PD1.1.7,0 HT: EII_ 3171 117 Loss=1/5 Loss=1/10

LEE: [EVIL HIP[EF P,PEF MP 2 000.000 H: '10 0[1t 10 OH SE.COS7 IL ]

57 dB

. U.OUI He

1 00 10.01 I F 0 000.0011 H: 110.0 7,0 10 IT'j 0,0 1-,440[ 1-11,0- H: Loss=1/15

Figure 8.21 Measured output spectra in attenuator mode: Vin = 4 Vp-p; Vb = 1.5 V. 162

'PECTFON -.S.F. FP, N: REF E: REF d MFR 2 000 -000 H.­ H.EH E REF A MEP 000.000 H. -30.00 10.00 JMO.-., OH "'36 dE, 311 1111 1.1 111W, FE I dPm 11 1 FEWF., ( Fr, ANHI.

84 dB 78 dB

kttpti.N, 10A,'", IN ,1 yr!' i f 0 l'EYO'fl",,rf4VokOd'vymok,i),Viy . .

+F II 0.701 H.: 10 U1; 10.00 470E 000 3011 H7 10 000.G00 HF 11' 11, PPIF. 30 H-F -T =0.T=-10dPm PE. -?0.F=-10dB, EBFF- 30 HZ. Loss=1 Loss=1/2

iEFFIFON :.PE._ UPUN iuFEF E.FD­ H,PEF P:FF-F j 0..F 2 000 0347 H,, 00 00 IIH 30.00 -10 0. 31,,, E-17 F E FF1H.,

( d Efi. I [ i 30r-6

66 dB 156 dB

0.001 Hz DIV PL; ':..TEIET i0.00 10.00 10.00 'STOF 10 000.000 Hz EF.1, 30 Hz -1.1 't REL.'. 30 0: '7.7.1.0 mln P,-)NC,E,R--20,T--10dEm REI-1-3n HZ 4 E1.1-30 H­ Loss=1/5 Loss=1/10

,4REF P,REF MkR 1 000.000 Hz -30.00 -10.00 drIHG -57.1210 dP

I dBm 11 J JMHG

57 dB

iieivqyorr101)ftyrirlOm*41

DIV DIV STEIRT 0.001 Hz 10.00 10.00 STOP 10 000.000 Hz PELF, 30 Hi- T -1.58 min R1NF4 ,P--20.-I--10dPm P.P47--30 IT: Loss=1/15

Figure 8.22 Measured output spectra in attenuator mode: Vin= 4 Vp-p; Vb = 2.0 V. 163

Chapter 9 PIPELINE DIGITAL-TO-ANALOG CONVERTER

An analog signal processing system which consists of an A/D converter, digital signal (DSP), and a D/A converter has become a popular structure because the development of a fine-line-width CMOS process has made digital circuits smaller, faster and cheaper. Consequently, A/D and D/A converters have become very important building blocks because the accuracy of the system is often determined by the performance of these data converters. In this chapter, the design of a MOSFET-only pipeline D/A converter will be described as an example of charge processing technique. First, the background of the pipeline D/A converter will be shown. Then, circuit design using MOSFET capacitors will be presented. Finally, HSPICE simulation results will be provided.

9.1 Background of the Pipeline D/A Converter

The pipeline D/A converter (DAC) shown in Figure 9.1 has several favorable features: a) high accuracy because of small capacitor mismatch error; b) low power consumption compared to resistive array or current source DAC; c) small area because of using equal-valued capacitors instead of binary-weighted ones; d) high speed improved by a pipeline structure [39]. In addition, it can be realized using MOSFET capacitors because the charge-processing technique can be easily applied. Figure 9.2 shows the concept of charge processing. First, an input voltage is linearly converted to charge (V-Q conversion), then the charge is transferred to the later stage, and finally the charge is linearly converted to a voltage (Q-V conversion). In this pipeline DAC, when a MOS capacitor in the capacitor array is charged to a reference voltage, it is biased in its accumulation region, and hence the voltage-to-charge 3 gs 4-11 b1 (LSB) II­ I. - I

b2 1

b3

b4

b10 (MSB) 43 gs .bik-3 bik-2 0.41 bik-1 Vr = 5 V bik --' A.G = 2.5 V bikOi bikoi b2k02 b24,2 b3k03 b3k03 b4k -141 b4k-101 biok-341 biok-341

Vr Vr Vr A.G. A. 1 41 4)2 03 \ 4)3 4)1 1 C C Analog Buffer Ground TC T (2.5 V) GND (0 V) Figure 9.1 10-bit pipeline DAC [39] 165 conversion is performed nearly linearly. Also, the charge acquired is nominally the same for all capacitors. When the capacitor is switched to ground, no nonlinear effects will be encountered. Finally, when a capacitor is connected to its neighbor, the two capacitors share their charges and their total charge is conserved. Thus by means of charge processing, the DAC converts digital inputs to an analog output linearly even with nonlinear MOSFET capacitors. However, in order to have an output voltage signal, a charge-to-voltage (Q-V) converter has to follow the pipelined capacitors. This Q-V conversion can be achieved by using one of the proposed SC branches.

Voltage-to-charge AQ Charge-to-voltage converter (V-Q) converter (Q-V) 01°' Charge conserved

Figure 9.2 The concept of charge processing

Theoretically, when a digital input is high, its corresponding capacitor of the pipeline should be charged between Vr and VSS (instead of the analog ground) to have a large dynamic range. However, in HSPICE simulation, when a voltage across a MOSFET capacitor reduced to 0 V, the charge stored on the capacitor was not zero, and that resulted in an error. It is believed that this was caused by the integration algorithm of HSPICE. Therefore, to avoid the nonlinear depletion region in the simulation, all capacitors in the pipeline are switched either the reference voltage Vr (VDD) or the analog ground (instead of

Vss) as shown in Figure 9.1.

9.2 Circuit Design

Figure 9.3 shows a simple Q-V conversion circuit using MOSFET capacitors. When 02 goes high (reset phase), two series-compensated MOSFET capacitors are biased 166 in their accumulation region. In the next clock phase 03, these MOSFET capacitors are connected between the output and virtual ground of the op-amp. At the same time, charge is transferred from the pipelined capacitor array to these capacitors. Since the series- compensated branch converts charge to voltage linearly, a desired output signal is obtained

(valid phase).

02

2.5C 2.5C

.----) I 1 IF

b10k-3(01 b10k-34:01 02

Vr Vb I­

03 \__,0Vout + i Analog ground (A.G.)

Figure 9.3 Charge-voltage converter

Figure 9.4 shows the output waveform of the pipeline DAC using this Q-V converter. For simplicity, only 4-bit configuration is used here. As can be seen from Figure 9.4, each time when 02 goes high, the output is reset to the virtual ground level (2.5 V).

Consequently, a high slew-rate op-amp will be required. 167

2.50 CIC_MO52 TRO VI OUT 2.40 2.30 2 20,

2 1

1.90,

1

1 70 1.00= I.50= 1.4 1.3

I 2 1.10, 1.0= 900.09 900.09

7 0 0 011, 000.0M =­

I 500 OM`' glob 4 Oil 5 OU 0.0U 7 01.1' 9 OU 2 70U TIME OLIN) Figure 9.4 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.3

To reduce the output change rate, and to alleviate the slew-rate requirements for the op-amp, another Q-V converter (with a 4-bit pipelined capacitor array) can be used. It is illustrated in Figure 9.5. When 4)1 and 4)2 are high, the two series-compensated capacitors are reset (and biased to operate in their accumulation region). A capacitor Cc is biased to 2.5 V during 4)1 and 4)2. Capacitor Cs provides a feedback path to the op-amp to prevent it from saturating during this period. When 4)3 goes high, the output becomes available. Now the capacitor Cc is charged to the output level. In the next clock 4)1, Cc is connected to the virtual ground, and the charge flow from the top plate of Cc into the virtual ground node is negative because the voltage across Cc during 4)3 was less than 2.5 V. On the other hand, the charge flow from the series-connected capacitor C1 is positive. The change in the voltage across Cc between 4)2 and 4)3 is twice as large as that for C1, but C1 is twice as large as Cc. Thus, the charge flows from Cc and from C1 when 4)3 goes high cancel each other. Cs

4)1+02 01+02 03

1.25 C N+ diffusion inside n-well (accumulation) 2.5 C 2.5 C

Ea I C1 F.01+402 b1k411 bikOi b242 b242b3k03 b3k03 b4k-101 b4k-101 J1-01+02 Switch sizes Vr = 5 V Vr Vr Vr _L 1_ Vr VDD 6.0 [trn/ 1.2 p.m 121 A.G.v22 A.G. vout

A.G. (2.5 V) A.G (2.5 V) GND (0 V) Figure 9.5 Pipeline D/A converter using an improved charge-voltage converter 169

An alternative Q-V converter is shown in Figure 9.6. In this circuit, an auxiliary capacitor Cv is added to create an improved virtual ground, and the last capacitor of the pipelined array is connected to that node instead of the real ground. The role of Cs is the same as in Figure 9.5 and Cdg works as a deglitching capacitor to prevent spikes in the output signal. Figures 9.7 and 9.8 show the simulated output waveforms for the circuits of Figure 9.5 and 9.6, respectively. In Figure 9.8, the output changes when 03 goes high, and it keeps that level in the consecutive clock phases 01 and 02. Thus, the slew-rate requirement for the op-amp is greatly reduced. We see glitches, however, during the non-overlapping period because no deglitching capacitor is used. In Figure 9.8, the glitches are reduced by the deglitching capacitor Cdg. Also, in this circuit, the output during the valid phase 03 is held during the consecutive clock phases

01 and 02.

9.3 Simulation Results

The circuits shown in Figure 9.5 and 9.6 were simulated using HSPICE. In these simulations, the capacitor array was extended to 10 bits. The size of a MOSFET capacitor C was 12 x 18 tm x 18 Jim, which corresponds to 6 pF. The size of switches was 6.0 iim x 1.2 Jim for b3th PMOS and NMOS devices. An op-amp macro model was used in the simulations. Figure 9.9 and 9.10 show an integral non-linearity (INL) of the 10-bit pipeline D/A converter using the Q-V converters of Figure 9.5 and 9.6, respectively. The INL was +/­ 1.30 mV for Figure 9.5 and +/- 1.55 mV for Figure 9.6. Both of them were slightly larger than 1 LSB for a 10 bit accuracy (+/- 1 mV). Figure 9.6 is a glitch-free circuit. However, the last capacitor of the pipelined array in Figure 9.6 may have a mismatch with the other capacitors in the array because its bottom plate is switched from the real ground to the virtual ground while the bottom plates of other capacitors are fixed to the real ground. 2.5 C 2.5 C

I

I-01+02 N+ diffusion VDD inside n-well A.G (accumulation) 01+02 _L­ a.] bikOi bikOi b242 b242b3k03 b3k03 b4k-101 b4k-1,01 Switch sizes Vr =5 Vr -L Vr -L -1­ Vr Vr 6.0 pm / 1.2 gm

vout C C1c A.G C. (2.5 V) 01+0 A.G

Cdg Figure 9.6 Pipeline D/A converter using an alternative charge-voltage converter 171

2 50 : 4B_MOS5 TRO 2.40 = 2COUT 2.30 =

2 20 2_10 = 2_0 90=

1 eo

0 1 70 = 1.60= 1_50=

1 90­ 1.30= 1.20: 1.10= 0=­ 900 011= 000 on = 700.0M : 500 omhilI

I 500 OM 3.0U 4.0 5.06 6 OU 7.06 2 70U TIME (LIN) 9.00

Figure 9.7 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.5

4.50 49_M054 TRO 4.40= _f4VCOUT 4.30= 4.20= 4.10 4.0: 3_90: 3.80= 3.70: 3 GO: 3.50:

N 3.40: 3.30: 3.20 : 3_10-­ 3.0 2.90= 2.90E 2.70 == 2.50= 2.50 1111 11 3 4 OU 5.0U 6.00 7 OU 2.40U TIME (LIN) 9.0U

Figure 9.8 The output waveform of 4-bit pipeline D/A converter using the charge-voltage converter of Figure 9.6 172

Integral Non-Linearity of 10 Bit Pipeline DAC

100 200 300 400 500 600 700 800 900 1000

Digital input

Figure 9.9 Integral nonlinearity (INL) of the 10-bit pipeline D/A converter using the Q-V converter of Figure 9.5

INL of 10 Bit Pipeline DAC

1.5

0.5

0.5

1.5

200 400 600 800 1000 1200

Digital input

Figure 9.10 Integral nonlinearity (INL) of the 10-bit pipeline D/A converter using the Q-V converter of Figure 9.6 173

Chapter 10

NONLINEARITY COMPENSATION IN DEEP SUBMICRON AND LOW-VOLTAGE PROCESSES

A demand for higher integration density forces CMOS processes to use deep submicron line widths. In a deep submicron process, the gate oxide becomes thinner and electrical field increases. Consequently, the carrier distribution in the MOSFET channel will be affected, and that results in an increase of the MOSFET capacitor nonlinearity. Hence it is preferred that the bias voltage across MOSFET capacitors be increased to have a better linearity. However, in such a process, the supply voltage has to be lowered for the reliability of the MOS devices. In addition, the need for battery-operated equipment also demands low-voltage operation. In this chapter, a capacitor nonlinearity compensation scheme suitable for deep submicron and low-voltage processes will be discussed. First, series compensation in low- voltage operation will be described. Then, a new switch configuration which provides larger bias voltage to series-compensated MOSFET capacitors will be presented, along with

HSPICE simulation results.

10.1 Series Compensation in Low-Voltage Operation

In circuits using the series compensation technique described in Section 4.4, the voltage of the top capacitor plates (node A in Figure 10.1) should not exceed VDD, because any voltage larger than VDD at node A can forward bias the pn junction between a p+ diffusion and the n-well of the PMOS switch Ml, and that results in a leakage of the charge stored in the capacitors and harmonic distortion at the output. However, when a very large voltage swing is required, or for low-voltage operation, it is desirable to allow the top-plate 174 node potential to exceed VDD, so that a large bias voltage can be applied to the MOSFET capacitors to guarantee that all MOSFET capacitors operate in their accumulation regions.

A 0 Virtual ground vout

(or vin) ( 01

Ml Vb I­

(a)

B A

)I Virtual ground

vss VDD

Vb

ip+ I P+ I I P+ 1111+ Ml n-well

p-substrate

(b) Figure 10.1 (a) Switch configuration in series compensation; (b) Practical realization of switch MI. 175

4)2 A

vout Virtual (or vin) ground 01 \ 01 M1 D

02 M2 11\43

Vb=VDD

(a)

4)2 0

7+ in+ II p-1 I p-substrate O$

M3

Vb=VDD C vss

I n+ II P+ I P+ IP +I I P+ II n+ I n well n-well

p-substrate

(b) Figure 10.2 (a) New switch configuration for series compensation; (b) Practical realization of series switches M1 and M2. 176

For example, assume that Vb = VDD = -VSS = 1.5 V in Figure 10.1. During 4)1=1, the switch M1 turns on, and the voltage at node A becomes equal to Vb = 1.5 V. When vout increases to say 1.0 V during 4)2 = 1, the voltage potential of node A exceeds VDD.

Then the pn junction between the p+ diffusion of M1 at node A and the n-well of M1 turns on and leakage current starts to flow if the n-well of M1 is tied to VDD. If the n-well of M1 is connected to the p+ diffusion of M1 at node A, then the leakage problem can be avoided for a large positive output. However, if vout decreases to -1.0 V during 4)2 = 1, the voltage of node A becomes 1.0 V. Then the pn junction between the p+ diffusion of M1 connected to Vb (= 1.5 V) and the n-well of M1 (= 1.0 V) will get forward biased, and leakage current begins to flow.

10.2 New Switch Configuration for Series Compensation

To overcome this constraint, we have developed a series switch configuration, illustrated in Figure 10.2 [36]. An additional PMOS switch M2 is connected in series with the PMOS switch M1 to prevent the flow of leakage current, and an NMOS switch M3 is connected to node D to guarantee that the pn junctions at node D are reverse-biased during 4)2=1. The n-wells of M1 and M2 are tied to their sources. Assume that Vb = VDD = -VSS

= 1.5 V. During 4)1=1, Ml and M2 turn on, and the voltage at node A becomes equal to Vb = 1.5 V. During 4)2=1, these PMOS switches are off, and the NMOS switch M3 turns on. If vout is equal to 1.0 V when 02=1, the potential of node A goes up to 2.0 V as does the voltage of the n-well of Ml. Then, the pn junction between the p+ diffusion of M1 at node D and the n-well of M1 is reverse-biased, because node D is connected to the ground. (The other pn junction between the p+ diffusion of M2 at node D and its n-well is also reverse- biased.) The pn junction between node A and the n-well of M1 is short-circuited, thus it is not forward biased. If vout is equal to -1.0 V during 4)2 = 1, the potential of node A becomes 1.0 V. Again, the pn junction between p+ diffusion of Ml at node D and the n- well of M1 are reverse biased. Therefore, using this switch configuration, a large voltage 177 swing at output (and input) node can be achieved without the leakage problem even for low supply voltages. Notice that the additional PMOS and NMOS switches (M2 and M3) can be shared by all M1 PMOS switches in Figure 10.1. Thus, only two additional devices are required to realize this switch structure for the whole SC circuit.

10.3 Simulation Results

To verify the effectiveness of the new switch configuration, two SC amplifiers were simulated using HSPICE: a) the original series-compensated SC amplifier shown in

Figure 4.8; b) the improved circuit of Figure 4.8 in which series compensated SC branches were replaced by the new branch shown in Figure 10.2. In the new switch structure (b), a parasitic pn junction diode capacitance between the n-well and the p-substrate is connected to the node A as show in Figure 10.2. To include the effect of the parasitic capacitance in simulations, a pn junction diode model was added to the n-well of PMOS switches in the

HSPICE netlist*. The conditions assumed in this simulation were similar to those used in Section 4.5 except that Vb = VDD = -VSS = 1.5 V was used. In addition, to exclude the nonlinear distortion caused by the on-resistance of switches, the clock signal for CMOS switches connected to inputs and outputs swan +/- 2.5 V. Figure 10.3a shows an output spectrum of the original SC amplifier for a 3 V supply voltage. Large harmonic components are observed in this Figure and a S/THD ratio was only 33 dB because charges stored on capacitors were lost due to leakage current through the pn junctions. Figure 10.3 b shows an output spectrum of the improved SC amplifier for a 3 V supply voltage. An S/THD of 84 dB was now obtained. The effect of the parasitic capacitance connected to the n-well of PMOS switches were insignificant because the switch size was much smaller than the main capacitor area, and also because

*A parasitic pn junction diode capacitance is connected to the node A even in the original switch configuration as shown in Figure 10.1. However, the parasitic capacitance is originally included in the SPICE MOSFET model as the drain-bulk capacitance. 178 the depletion width of the pn junction diode was much thicker than the gate oxide thickness. Hence, it is clear that this new switch configuration is effective for low voltage operation.

20 40 60 80 100 120 140 160 Frequency [kHz/

(a)

0

-20

-40

-60

-80

100

-120 (t6

140 zo 40 60 80 100 12 0 140 160 Frequency (kHz]

(b)

Figure 10.3 The output spectrum of the circuit of Figure 4.8 with VDD = -VSS = Vb = 1.5 V: (a) using the original switch configuration; (b) using the novel series switch scheme. 179

Chapter 11

CONCLUSIONS AND SUGGESTED FUTURE WORK

11.1 Conclusions

In this thesis, novel design techniques have been proposed for implementing high- linearity SC circuits in a standard digital CMOS process. They use nonlinear MOSFET capacitors instead of linear double-poly capacitors. To reduce their nonlinearities, a bias voltage is applied to keep the MOSFET capacitors in their accumulation regions. For further reduction of distortion, two capacitors are connected in series or in parallel so that a first- order cancellation of the nonlinearity can be achieved. To realize high-linearity SC circuits, the series compensation has an advantage over the other two methods (parallel compensation and simple biasing), because the operating ranges of nonlinear capacitors are reduced by a factor of two. Several chips were fabricated in a 1.2 tm CMOS process using the series compensation, parallel compensation, and simple biasing circuits. It was demonstrated that among these techniques, the series compensation is the most effective for reducing the nonlinearity of MOSFET capacitors. If the requirement for linearity is not too tight, the parallel compensation or the simple biasing may be useful to save chip area. To have the equivalent kT/C noise, the unit capacitance can be four times smaller for these techniques than for the series compensation.

Hence the capacitor area decreases considerably. A novel predictive SC amplifier has been proposed for reduced sensitivity to op­ amp imperfections. An SC amplifier using this predictive structure and the parallel compensation was fabricated in the 1.2 µm CMOS process. Experimental results show that the S/THD of the predictive SC amplifier was 10 dB larger than that of the non-predictive one. It was shown that the predictive circuit was effective to reduce nonlinearity caused by the op-amp and/or MOSFET capacitors. 180

It was also demonstrated that a two stage op-amp with a large output swing can be fabricated in a standard digital CMOS process. The frequency compensation was accomplished using a source follower and a MOSFET capacitor. The source follower provides a bias voltage to the MOSFET capacitor and keeps it in its accumulation region.

An SC amplifier using this two-stage op-amp and double-poly capacitors was fabricated in a 1.2 tm CMOS process, and its S/THD was 76 dB for a differential output swing of 4

Vp-p. Using this two-stage op-amp and the series compensation, a MOSFET-only digitally controlled gain/loss circuit was designed and fabricated in a 1.2 tm CMOS process. It was demonstrated that the series compensation is effective not only for a large output swing in an amplifier, but also for a large input swing in an attenuator. For an amplifier gain of 10, the SC circuit had a S/THD of 76 dB for a 4 Vp-p output swing, which was comparable to that of the SC amplifier using double-poly capacitors. For an attenuator with a loss of 1/2, the S/THD was 72 dB for a 4 Vp-p input swing. For a unity- gain configuration, a S/THD of over 80 dB was obtained for a 4 Vp-p input and output swing. A pipeline D/A converter utilizing MOSFET capacitors was described as an application of the charge processing technique. It consists of three parts, V-Q conversion, charge transfer, and Q-V conversion. The V-Q conversion can be linearly achieved by using two reference voltages to which MOSFET capacitors are charged or discharged. The charge transfer with pipeline operation does not require linear capacitors, because each capacitor just shares their charges with their neighbors. Finally, the V-Q conversion can be linearly performed by using series compensation. In HSPICE simulation, 9-bit accuracy has been obtained using only MOSFETs. Having a large bias voltage helps MOSFET capacitors operate in their deep accumulation region with high linearity. However, in the series compensation, this could have lead to a charge leakage because the parasitic pn junction diode of the PMOS switch 181 connected to the connecting node of two series-connected MOSFET capacitors turns on.

Therefore, the bias voltage was limited to some extent. A new switch configuration to solve this leakage problem and to enable the series compensation have a large bias voltage has been proposed. HSPICE simulation shows that it works well and it will be helpful for a

low-voltage operation, too.

11.2 Suggested Future Work: SC Integrators and Filters

The SC integrator is an important building block for realizing SC filters. Parallel compensation and simple biasing, which were described in Chapter 4, can be applied to an

SC integrator as shown in Figures 11.1 and 11.2. It is also possible to build an SC integrator by using series compensation. For example, Figure 11.3 shows a modified SC integrator whose transfer function is given by

C1 2 z-1 C2 1- z-1 However, it is difficult to apply series compensation to an SC integrator whose transfer

function is given by

1 (11-2) 1 z-1 One way to solve this problem is, as shown in Figure 1L4, to add a charge-to­ voltage (Q-V) converter to the output of an SC integrator. Here, the nonlinearity errors caused by nonlinear capacitors C2 and C3 cancel each other because the voltage across these capacitors become equal during 01=1. The concept of charge processing, described in

Chapter 3, can be extended to more complicated higher-order filters. Figure 11.5 shows a realization of high-Q biquad. The circuit surrounded by the broken line is a conventional high-Q biquad [17] followed by a charge-to-voltage (Q-V) converter. Once the input voltage is converted to charge, then the charges are transferred to the last Q-V converter

where they are reconverted into voltages. 182

vin­ o°

CH

CH Top (poly) ( I Bottom (accumulated channel)

Figure 11.1 SC integrator using parallel compensation

C2

Vin0-\ (1)2

Figure 11.2 SC integrator using a simple biasing arrangement 183

Cdg Top plate ,-­ (poly gate) I e I I 02 Cs I

I \ 1--- oi I I C2 C2 T $2 Bottom plate (accumulated channel) )I t (---­ ( 0,

I

I ,Vb / V in (1)2 / Vnut

Cl C1

Figure 11.3 SC integrator using series compensation SC integrator Charge-to-voltage conversion

Cs Top plate (poly gate) 1 7 Bottom plate (accumulated channel) C2

vout 02 1

Cdg

I(

Figure 11.4 SC integrator using series compensation Voltage-to-charge conversion Charge processing

C8 01 I( 1 \---1 C7 / 02 C2 C6 C1 C1 02 Vin 02 02 C3 -I(--­ `-r-41--r-''-r­oi oi fo I + r bp T

C9

Charge-to-voltage conversion Po' 02

Top plate (poly gate) 1 7 Bottom plate (accumulated channel)

Figure 11.5 Realization of a high-Q biquad 186

In this configuration, some nonlinear MOSFET capacitors operate in their accumulation region, but some in their depletion region. In the depletion region, capacitance mismatch can affect the performance much more than in the accumulation region, because the voltage coefficient of a MOSFET capacitor IdC/dVI is much larger there. To reduce the error caused by capacitance mismatch due to a voltage mismatch, a bias voltage may be applied to the feedback capacitors. To make sure that the capacitors operate in their linear region, a negative bias voltage Vbn may be applied to the positive input terminal of each op-amp as shown in Figure 11.6, so that all capacitors within the

broken line are simply biased*. Note that the feedback capacitors are turned over so that the polarity of the bias voltage across the MOSFET capacitors associated with their bottom plates becomes positive. In this configuration, the bottom plates of MOSFET capacitors are connected to the virtual grounds of op-amps. Hence, a fully differential circuit is recommended to suppress common-mode noise coupling from the substrate. A drawback

of the circuits shown in Figures 11.4-6 is that they require additional chip area for the V-Q

converter and the Q-V converter. In some applications such as in delta-sigma modulators, the V-Q converter may not be necessary due to the presence of a comparator. Huang realized a delta-sigma modulator

with a S/THD ratio of 94 dB by combining series compensation and charge processing [40]

[41]. When all input branches of a filter contain only switched capacitors (and no unswitched capacitors), i.e., there is no C9 in Figure 11.5, they can be replaced by a series compensated SC branch as illustrated in Figure 11.7. Hence, the voltage-to-charge conversion can be combined with the front end of the filter, and the chip area needed for the

voltage-to-charge conversion can be saved. Many other applications of the compensation techniques introduced in this work

may exist. It is hoped that they will be explored by future researchers.

* If a positive bias voltage Vbp is applied instead of Vbn, it is harder to keep the series-connected input capacitors in their accumulation regions. Voltage-to-charge conversion Charge processing C8 )1 C7

C6 Cl C1 4)2 Vin C3

rt01 Vbp fEl:(44)r1Vbn V n 1Vbn Vbn I +T _1°1Vbn Vbn C9 +T. T Charge-to-voltage conversion

Top plate (poly gate) Cli ClITi RF I +(1)1 101 T Bottom plate C10 ..h/bp" (accumulated channel) Vout Figure 11.6 Alternative realization of a high-Q biquad _T_V n d g I( (poi C8 I(

C7 ( CI 4 C6

(I)2 Vin C3' C3' 4)2 I( 0 Cs (1)2 °\ I ) I I I ( \--­ \ 1 )I \

01 01 1)1 I ( 02 01 0 _ I Vbp ' -1­

Figure 11.7 An input branch combined with a voltage-to-charge converter using series compensation 189

BIBLIOGRAPHY

[1] J. L. McCreary, "Matching Properties, and Voltage and Temperature Dependence of MOS Capacitors," IEEE J. Solid-State Circuits, vol. SC-16, pp. 608-616, 1981.

[2] D. J. Allstot and W. C. Black, "Technological Design Considerations for Monolithic MOS Switched-Capacitor Filtering Systems," Proc. IEEE, vol. 71, pp. 967-986, Aug. 1983.

[3] D. B. Slater and J. J. Paulos, "Low-Voltage Coefficient Capacitors for VLSI Processes," IEEE J. Solid-State Circuits, vol. 24, pp. 165-173, 1989.

[4] A. T. Behr, M. C. Schneider, S. N. Filho, and C. G. Montoro, "Harmonic Distortion Caused by Capacitors Implemented with MOSFET Gates," IEEE J. Solid- State Circuits, vol. 27, pp. 1470-1475, 1992.

[5] S. M. Sze, Physics of Semiconductor Devices, John Wiley & Sons, Inc., 1981.

[6] J. B Shyu, G. C. Temes, and K. Yao,"Random Errors in MOS Capacitors", IEEE J. Solid-State Circuits, vol. SC-17, pp. 1070-1076, 1982.

[7] HSPICE User's Manual, vol. 2, Meta-Software.

[8] J. B Shyu, G. C. Temes, and F. Krummenacher,"Random Error Effects in Matched MOS Capacitors and Current Sources", IEEE J. Solid-State Circuits, vol. SC-19, pp. 948-955, 1984.

[9] S. S. Bazarjani, Ph. D. dissertation, Carleton University, Ottawa, Canada, 1996.

[10] S. S. Bazarjani, and W. M. Snelgrove,"A 4th Order SC Bandpasss EA Modulator Designed on a Digital CMOS Process," Proc. 38th Midwest Symposium on Circuits and Systems, pp. 1345-1348, Aug. 1995. [11] J. C. M. Bermudez, M. C. Schneider, and C. G. Montoro, "Linearity of Switched Capacitor Filters Employing Nonlinear Capacitors," Proc. IEEE Int. Symp. on Circuits and Sys., pp. 1211-1214, 1992. [12] J. B. Hughes, N. C. Bird, and I. C. Macbeth,"Switched-Currents A New Technique for Analog Sampled-Data Signal Processing," Proc. IEEE Int. Symp. on Circuits and Sys., pp. 1584-1587, 1989. [13] G. C. Temes, P. Deval, and V. Valencic,"SC Circuits: The State of the Art Compared to SI Techniques," Proc. IEEE Int. Symp. on Circuits and Sys., pp. 1231-1234, 1993.

[14] J. B Hughes, K. W. Moulding,"Swtiched-Current Signal Processing for Video Frequencies and Beyond," IEEE J. Solid-State Circuits, vol. 28, pp. 314-322, 1993. [15] K. Matsui, T. Matsuura, S. Fukasawa, Y. Izawa, Y. Toba, N. Miyake, and K. Nagasawa,"CMOS Video Filters Using Switched Capacitor 14-MHz Circuits," IEEE J. Solid-State Circuits, vol. SC-20, pp. 1096-1102, 1985. 190

[16] F. J. Wang and G. C. Temes,"A Fast Offset-Free Sample-and-Hold Circuit," IEEE J. Solid-State Circuits, vol. 23, pp. 1270-1272, 1988. [17] R. Gregorian and G. C. Temes, Analog MOS Integrated Circuits for Signal Processing, John Wiley and Sons, Inc, 1986. [18] T. Foxall, R. Whitbread, L. Sellars, A. Aitken, and J. Morris,"A switched-capacitor bandsplit filter using double polysilicon oxide isolated CMOS," 1980 ISSCC Digest Tech. Papers, pp. 90-91, 1981. [19] R. Gregorian, K. Martin, and G. C. Temes,"Switched-Capacitor Circuit Design," Proc. IEEE, vol. 71, pp. 941-966, 1983. [20] K. Haug, G. C. Temes, and K. Martin,"Improved Offset Compensation Schemes for Switched Capacitor Circuits," Proc. IEEE Int. Symp. on Circuits and Sys., pp. 1054-1057, 1984. [21] K. Martin, L. Ozcolak, Y. S. Lee, and G. C. Temes,"A Differential Switched- Capacitor Amplifier," IEEE J. Solid-State Circuits, vol. SC-22, pp. 104-106, 1987. [22] K. L. Lee and R. G. Meyer,"Low-Distortion Switched-Capacitor Filter Design Techniques," IEEE J. Solid-State Circuits, vol. SC-20, pp. 1103-1113, 1985. [23] H. Matsumoto and K. Watanabe,"Spike-free switched-capacitor circuits," Electron. Lett., vol. 8 pp. 428-429, 1987. [24] A. S. Grove, B. E. Deal, E. H. Snow, and C. T. Sah,"Investigation of Thermally Oxidized Silicon Surfaces Using Metal-Oxide-Semiconductor Structures," Solid State Electron., vol. 8, pp.145, 1965. [25] H. Yoshizawa, G. C. Temes, P. Ferguson Jr., and F. Krummenacher,"Novel Design Techniques for High-Linearity MOSFET-Only Switched-Capacitor Circuits," Proc. of the IEEE VLSI Circuit Symposium, pp. 152-153, 1996. [26] H. Yoshizawa, G. C. Temes,"High-Linearity Switched-Capacitor Circuits in Digital CMOS Technology," Proc. IEEE Int. Symp. on Circuits and Sys., pp.1029-1032, 1995. [27] L. Larson, K. W. Martin, and G. C. Temes,"GaAs Switched-Capacitor Circuits for High-Speed Signal Processing," IEEE J. Solid-State Circuits, vol. SC-22, pp. 971­ 981, 1987. [28] Y. Huang, G. C. Temes, and P. F. Ferguson,"Reduced Nonlinear Distortion in Circuits with Correlated Double Sampling," Proc. IEEE Int. Symp. on Circuits and Sys., pp.159-162, 1996. [29] H. Yoshizawa, Y. Huang, and G. C. Temes,"Improved SC amplifiers with low sensitivity to op-amp imperfections," Electron. Lett., vol. 33, pp. 348-349, 1997. [30] K. Nagaraj, J. Vlach, T. R. Viswanathan, and K. Singhal," SC Circuits with Reduced Sensitivity to Amplifier Gain," IEEE Trans. Circuits. and Sys., vol. CAS­ 34, pp. 571-574, May 1987.

[31] ADMOD79 data sheet, Analog Devices Inc. 191

[32] A. Yesilyurt, Ph. D. dissertation, Oregon State University, Corvallis, Oregon, U.S.A., 1997. [33] G. Nicollini, P. Confalonieri, and D. Senderowicz, "A Fully Differential Sample- and-Hold Circuit for High-Speed Applications," IEEE J. Solid-State Circuits, vol. SC-24, pp. 1461-1465, 1989. [34] W. S. Wu, W. J. Helms, J. A. Kuhn, and B. E. Byrkett, "Digital-Compatible High- Performance Operational Amplifier with Rail-to-Rail Input and Output Ranges," IEEE J. Solid-State Circuits, vol. 29, pp. 63-66, 1994. [35] P. E. Allen and D. R. Holberg, CMOS Analog Circuit Design, Saunders College Publishing, 1987. [36] H. Yoshizawa, Y. Huang, and G. C. Temes, "MOSFET-Only Switched-Capacitor Circuits in Digital CMOS Technologies," Proc. IEEE Int. Symp. on Circuits and Sys., Hong Kong, June 1997, to appear. [37] G. L. Baldwin and J. L. McCreary, " An MOS Digitally Controlled Analog Attenuator for Voiceband Signals", IEEE Journal of Solid-State Circuits, vol. SC-14, no. 1, pp. 74-79, Feb. 1979.

[38] J. N. Babanezhad and R. Gregorian, "A Programmable Gain/Loss Circuit", IEEE Journal of Solid-State Circuits, vol. SC-22, no. 6, pp. 1082-1090, Dec. 1987. [39] F. J. Wang, G. C. Temes, and S. Law, "A Quasi-Passive CMOS Pipeline D/A Converter," IEEE Journal of Solid-State Circuits, vol. 24, pp. 1752-1755, 1989. [40] Y. Huang, G. C. Temes, and H. Yoshizawa, "A 16-bit 3-V All MOSFET Delta- Sigma Modulator," Proc. CICC, pp. 293-296, 1997. [41] Y. Huang, Ph. D. dissertation, Oregon State University, Corvallis, Oregon, U.S.A., 1997. 192

APPENDIX 193

APPENDIX

SPICE Parameters for the Simulations Performed in this Thesis

.MODEL n NMOS +(Level = 2 UO = 521.3 VTO = 0.9175 NFS = 5.9984E11 + TPG = 1 TOX= 225E-10 NSUB = 5.7E16 UCRIT= 5.0E4 + UEXP = 8.5652E-2 VMAX = 4.953E4 RSH = 349.7 + XJ = 3.0E-7 LD = 2.029E-7 DELTA= 5.281 + PB = 0.8 JS = 1.0E-5 NEFF = 3.09735 WD = 2.2278E-7 + CJ = 5.56037E-4 MJ = 4.349E-1 CJSW = 3.73283E-10 + MJSW = 3.4355E-1 CGS() = 3.1E-10 CGDO = 3.1E-10 + CGBO = 3.4E-10 FC = 0.5 XQC = 1.0 )

.MODEL p PMOS +(Level = 2 UO = 167.6 VTO =-0.9153 NFS = 6.5023E11 + TPG = -1 TOX= 225E-10 NSUB = 3.67E16 UCRIT= 1.0E4 + UEXP = 9.9314E-2 VMAX = 2.8585E4 RSH = 418.2 + XJ = 3.0E-7 LD = 7.093E-8 DELTA= 2.146 + PB = 0.8 JS = 1.0E-5 NEFF = 0.93116 WD = 4.082E-7 + CJ = 4.4864E-4 MJ = 4.039E-1 CJSW = 4.57E-10 + MJSW = 3.344E-1 CGSO = 1.08E-10 CGDO = 1.08E-10 + CGBO = 6.25E-10 FC = 0.5 XQC = 1.0 )