a ripple-based ultra-low power buck converter with ... · topics and the questions you ask pushed...

135
A Ripple-based Ultra-low Power Buck Converter with Constant On-Time Control BY Saikiran Reddy Ramidi, B.E A technical report submitted to the Graduate School in partial fulfillment of the requirements for the degree Master of Sciences, Engineering Specialization in: Electrical Engineering New Mexico State University Las Cruces, New Mexico July 2017

Upload: ngoduong

Post on 25-Jun-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

A Ripple-based Ultra-low Power Buck Converter with Constant On-Time

Control

BY

Saikiran Reddy Ramidi, B.E

A technical report submitted to the Graduate School in partial

fulfillment of the requirements

for the degree

Master of Sciences, Engineering

Specialization in: Electrical Engineering

New Mexico State University

Las Cruces, New Mexico

July 2017

“A Ripple-based Ultra-Low Power Buck Converter with Constant On-Time Con-

trol,” a technical report prepared by Saikiran Reddy Ramidi in partial fulfillment

of the requirements for the degree, Master of Sciences has been approved and

accepted by the following:

Dr. Louis ReyesDean of the Graduate School

Chair of the Examining Committee

Date

Committee in charge:

Dr. Paul M. Furth, Associate Professor, Chair.

Dr. Wei Tang, Assistant Professor.

Dr. Rolfe Sassenfeld, Associate Professor.

ii

DEDICATION

Dedicated to my father Bhoopal Reddy Ramidi, mother Lakshmi Ramidi,

sister Neha Ramidi, my advisor Dr. Paul Furth and my family members.

iii

ACKNOWLEDGMENTS

I thank my parents Bhoopal Reddy Ramidi and Lakshmi Ramidi for encour-

aging me to pursue Masters program. Your strong support and encouragement

kept me focused on my academics.

It is a great privilege to have Dr. Paul Furth as my advisor. He is not only

a great teacher but also a wonderful human being. His way of teaching not only

helped me gain great intuition of several concepts, but more importantly taught

me a new way of approaching a problem and new ways of learning.

Thank you my friends and roommates Saikrishna Nelluri, Suresh Badavath,

Karthik Panuganti, Mahender Manda and Harsha Gadde for making las cruces

feel like home. It was great joy being with you all.

Mahender Manda was a great company. The discussions we had on several

topics and the questions you ask pushed me to have more intuition on any topic.

I am also privileged to have Sri Harsh Pakala as my mentor and an advisor.

I thank him for having patience in answering even the silliest questions and for

guiding me both academically and personally.

I would like to specially thank my VLSI team friends Ravindra Jonnala-

gadda, Venu Siripurapu, Rohith Gaddam, Pradeep kumar Polavarapu, Kumar Pal

iv

Mandoth and Arun Kumar Bijjala. You made the work environment (classes and

lab) more fun.

I would like to thank Vamshidhar Reddy Rajannagari for his guidance and

support.

Being with Yeshwanth Puppala was great fun. I would like to thank him

for his guidance and support.

I would like to thank Harish Nammi for being my mentor. It is great

working with you and talking to you.

I would like to thank all other friends in las cruces who filled this place with

fun and joy, Saikiran Golconda, Akhilesh Kumar, Om Rameshwar Gatla, Ankith

Nadella, Hemanth Pendyala, Niranjan Eshappa, Chaitanya Kukutla, Sachin Sunka,

Narapa Bommu, Bala Kesavaraju Nadikatla, Bhanu Pinnapu, Yashwanth Madadi,

Nagendra Kalava Phanidhar Kukutla, Harvind Kumar, Gangadhar Pulipelli, Phani

Raj Dandamudi, Sujith Dandamudi, Avinash Savvy, Jyoteesh, Sreyas Kundurpi,

Charan Yellanki, Hemanth Bondilli, Manikanta Ponnam, Dinesh, Srinath Pin-

napu, Pranay Kumar Lingala, Bhumika Parikh, Tapaswy Muppaneni and Pavan

Chaturvedi.

v

ABSTRACT

A Ripple-based Ultra-low Power Buck Converter with Constant On-Time

Control

BY

Saikiran Reddy Ramidi, B.E

Master of Sciences, Engineering

Specialization in Electrical Engineering

New Mexico State University

Las Cruces, New Mexico, 2017

Dr. Paul M. Furth, Chair

MS Technical Report defense scheduled on 07/07/2017, 9 AM

Thomas & Brown Hall, Room 207.

Ultra-low power applications such as fitness tracking devices, which are

always carried by the users, so that the body activity is monitored round the clock,

requires the battery to run continually. As such, there is a demand for long battery

run-time (time to drain a fully charged battery before needing to recharge it) for

these devices. However, the battery size is limited which also limits the amount of

battery energy available. Therefore, a highly efficient DC-DC converter is needed

to improve the battery run-time. A highly efficient DC-DC buck converter for

vi

ultra-low power applications is proposed in this work. A ripple-based ultra-low

power buck converter with constant on-time control is implemented in IBM 180

nm CMOS technology. The implemented DC-DC buck converter can drive loads

from 20 µA to 200 µA. Power losses in the buck converter are understood before

implementing the design. A simulated peak efficiency of 87.1% is achieved with

the implemented design.

vii

TABLE OF CONTENTS

LIST OF TABLES xi

LIST OF FIGURES xii

1 INTRODUCTION 1

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Objectives and Unique Contributions . . . . . . . . . . . . . . . . 2

1.3 Report Organization . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 LITERATURE REVIEW 4

2.1 DC-DC Buck Converter . . . . . . . . . . . . . . . . . . . . . . . 4

2.1.1 Conventional DC-DC Buck Converter . . . . . . . . . . . . 4

2.1.2 Synchronous DC-DC Buck Converter . . . . . . . . . . . . 8

2.2 Modes of Operation . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.3 Control Techniques in a Buck Converter . . . . . . . . . . . . . . 13

2.3.1 Pulse Width Modulation Control (PWM) . . . . . . . . . 14

2.3.2 Pulse Frequency Modulation (PFM) . . . . . . . . . . . . 16

2.4 Buck Converter Operation in DCM . . . . . . . . . . . . . . . . . 18

2.5 Previous Low Power Buck Converters . . . . . . . . . . . . . . . . 21

2.6 PFM Controller Circuit Blocks . . . . . . . . . . . . . . . . . . . 28

2.6.1 Comparator . . . . . . . . . . . . . . . . . . . . . . . . . . 28

viii

2.6.2 Zero-Current Detector . . . . . . . . . . . . . . . . . . . . 29

2.6.3 S-R Latch . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.6.4 D Flip-Flop . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2.6.5 Current-Starved Delay Line . . . . . . . . . . . . . . . . . 33

2.6.6 Non-Overlapping Clock Generators . . . . . . . . . . . . . 35

2.6.7 Gate Drivers . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3 DESIGN AND BLOCK-LEVEL SIMULATION RESULTS 39

3.1 Design Specifications and Motivation . . . . . . . . . . . . . . . . 39

3.2 Buck Converter Design in Discontinuous Conduction Mode . . . . 39

3.2.1 Derivations Duty Cycle of PMOS and NMOS Switches . . 40

3.2.2 Derivation to Calculate Inductor Value . . . . . . . . . . . 43

3.2.3 Derivations to Calculate Capacitor Value . . . . . . . . . . 46

3.2.4 Optimizing the Power Switches . . . . . . . . . . . . . . . 50

3.2.5 Choosing the Inductor Value . . . . . . . . . . . . . . . . . 55

3.2.6 Ripple-Based Constant On-Time Control Logic . . . . . . 57

3.3 Design and Simulation Results of Control Circuitry . . . . . . . . 68

3.3.1 Comparator . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3.3.2 Zero-Current Detector (ZCD) . . . . . . . . . . . . . . . . 71

3.3.3 Non-Overlapping Clock Generator (NOCG) . . . . . . . . 72

3.3.4 Current-Starved Delay-Line (CSD) . . . . . . . . . . . . . 72

4 SYSTEM-LEVEL SIMULATIONS OF BUCK CONVERTER 78

4.1 Simulation Results at Minimum Load . . . . . . . . . . . . . . . . 78

4.2 Simulation Results Across the Load Range . . . . . . . . . . . . . 82

4.3 Load Transient Response . . . . . . . . . . . . . . . . . . . . . . . 84

ix

4.4 Line Transient Response . . . . . . . . . . . . . . . . . . . . . . . 85

5 HARDWARE IMPLEMENTATION AND MEASURED RESULTS100

5.1 Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

5.2 Measured Results at Minimum Load . . . . . . . . . . . . . . . . 104

5.3 Peak Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6 CONCLUSIONS 111

6.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.2 Comparison with the State-of-the-Art . . . . . . . . . . . . . . . . 112

6.3 Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

6.4 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

REFERENCES 117

x

LIST OF TABLES

2.1 Truth Table of SR Latch using NOR Gates. . . . . . . . . . . . . 32

3.1 Buck Converter Specifications. . . . . . . . . . . . . . . . . . . . . 40

3.2 Several Inductors Available in the Market . . . . . . . . . . . . . 46

3.3 Calculated Design Parameters . . . . . . . . . . . . . . . . . . . . 47

3.4 Calculated values of C . . . . . . . . . . . . . . . . . . . . . . . . 50

3.5 Optimized switch sizes and efficiency for L = 22 µH ; RDCR = 0.81Ω 55

3.6 Optimized switch sizes and efficiency for L = 100 µH ; RDCR = 3Ω 56

3.7 Final Component Values of this design . . . . . . . . . . . . . . . 56

3.8 MOSFET sizing in the comparator . . . . . . . . . . . . . . . . . 71

3.9 MOSFET sizing in the ZCD . . . . . . . . . . . . . . . . . . . . . 72

3.10 MOSFET sizing in the NOC . . . . . . . . . . . . . . . . . . . . . 74

3.11 MOSFET sizing in the ZCD . . . . . . . . . . . . . . . . . . . . . 76

4.1 Simulation results at 20 µA load current . . . . . . . . . . . . . . 79

4.2 Simulated Performance Across Load Current Range . . . . . . . . 83

5.1 Area occupied by major blocks . . . . . . . . . . . . . . . . . . . 102

5.2 Comparision of simulated and measured results for a 20 µA load . 108

5.3 Comparision of simulated and measured results at 96 µA load . . 110

6.1 Comparison with state-of-the-art topologies. . . . . . . . . . . . . 114

xi

LIST OF FIGURES

2.1 Conventional Inductor-based Buck Converter . . . . . . . . . . . . 5

2.2 Inductor Current . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.3 Switching node . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.4 Synchronous Buck Converter . . . . . . . . . . . . . . . . . . . . . 9

2.5 Synchronous Buck Converter in ON state . . . . . . . . . . . . . . 9

2.6 Synchronous Buck Converter in OFF state . . . . . . . . . . . . . 10

2.7 Dead-time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.8 Switching node in synchronous buck converter . . . . . . . . . . . 12

2.9 Inductor current in three operation modes with constant L. . . . . 13

2.10 Inductor current in three operation modes with constant averageload current. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.11 Conventional Pulse Width Modulation Control Technique . . . . 15

2.12 Conventional Pulse Frequency Modulation Control Technique basedon [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.13 Synchronous Buck Converter . . . . . . . . . . . . . . . . . . . . . 19

2.14 Gate Signals and Inductor Current in DCM . . . . . . . . . . . . 20

2.15 Switching node with ringing . . . . . . . . . . . . . . . . . . . . . 21

2.16 Implementation of PFM mode controller from [2] . . . . . . . . . 23

2.17 Implementation of retention mode controller from [2] . . . . . . . 25

2.18 Implementation of PFM controller from [3] . . . . . . . . . . . . . 26

xii

2.19 Implementation of of ST-ZCD [3] . . . . . . . . . . . . . . . . . . 27

2.20 A two-stage differential amplifier with an output inverter imple-mented as a comparator based on [4] . . . . . . . . . . . . . . . . 28

2.21 Synchronous Buck Converter . . . . . . . . . . . . . . . . . . . . . 30

2.22 Three-stage comparator using common-gate differential amplifier [5] 31

2.23 S-R Latch with NOR Gates . . . . . . . . . . . . . . . . . . . . . 32

2.24 A positive-edge triggered D Flip-Flop with asynchronous-low reset 33

2.25 Current-starved delay line based on [4] . . . . . . . . . . . . . . . 34

2.26 Non-overlapping clock generator based on [4] . . . . . . . . . . . . 35

2.27 Creating Dead-Time using Non-Overlapping Clock Generator fromFig. 2.26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.28 Gate Drivers with scale factor S . . . . . . . . . . . . . . . . . . . 38

3.1 Synchronous Buck Converter in phase 1 . . . . . . . . . . . . . . . 41

3.2 Synchronous Buck Converter in phase 2 . . . . . . . . . . . . . . . 42

3.3 Synchronous Buck Converter in phase 3) . . . . . . . . . . . . . . 42

3.4 Inductor current in DCM . . . . . . . . . . . . . . . . . . . . . . . 45

3.5 Voltage and current in a capacitor . . . . . . . . . . . . . . . . . . 48

3.6 Voltage and current in a capacitor that has ESR . . . . . . . . . 49

3.7 Parasitic capacitances at the gate terminal of switches . . . . . . . 51

3.8 A Constant on-time control loop implementation . . . . . . . . . . 57

3.9 Start-up operation without start-up circuitry . . . . . . . . . . . . 61

3.10 Start-up operation with start-up circuitry . . . . . . . . . . . . . 63

3.11 Start-up when the clock misses . . . . . . . . . . . . . . . . . . . 64

3.12 D flip-flop clock misses as RST is low when the clock edge occurs 64

3.13 Start-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

xiii

3.14 A three-stage CMOS differential amplifier implemented as a com-parator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3.15 A three-stage comparator using common-gate differential amplifier 67

3.16 A current-starved delay line . . . . . . . . . . . . . . . . . . . . . 68

3.17 A three-stage CMOS differential amplifier implemented as a com-parator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

3.18 Comparator Operation . . . . . . . . . . . . . . . . . . . . . . . . 70

3.19 A three-stage comparator using common-gate differential amplifier 71

3.20 Inputs and output signals of ZCD . . . . . . . . . . . . . . . . . . 73

3.21 An SR latch with NAND gates implemented as a non-overlappingclock generator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

3.22 Dead-time generated by non-overlapping clock generator . . . . . 75

3.23 A current-starved delay line . . . . . . . . . . . . . . . . . . . . . 76

3.24 Delay from input to output . . . . . . . . . . . . . . . . . . . . . 77

4.1 Test bench for system level buck converter . . . . . . . . . . . . . 79

4.2 Output voltage and gate signals at 20 µA load current. . . . . . . 80

4.3 Inductor current and switching node at 20 µA load current. . . . 81

4.4 Output voltage and gate signals in steady-state at 20 µA load current. 82

4.5 Inductor current and switching node in steady-state at 20 µA loadcurrent. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

4.6 Detecting zero volts at node VSW at 20 µA load current. . . . . . 88

4.7 Ringing at the switching node in phase 3 at 20 µA load current. . 89

4.8 Output voltage and gate signals in steady-state at a load currentof 200 µA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

4.9 Switching node and inductor current in steady-state at a load cur-rent of 200 µA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

4.10 Load regulation testbench . . . . . . . . . . . . . . . . . . . . . . 92

xiv

4.11 Load transient response . . . . . . . . . . . . . . . . . . . . . . . 93

4.12 20 µA to 200 µA load transient . . . . . . . . . . . . . . . . . . . 94

4.13 200 µA to 20 µA load transient . . . . . . . . . . . . . . . . . . . 95

4.14 Testbench for line transient simulation . . . . . . . . . . . . . . . 96

4.15 Line transient simulation response . . . . . . . . . . . . . . . . . . 97

4.16 Line transient simulation response (3.6 V - 3.0 V) . . . . . . . . . 98

4.17 Line transient simulation response (3.0 V - 3.6 V) . . . . . . . . . 99

5.1 Complete layout of the DCM buck converter with pad frame . . . 101

5.2 Comparator layout . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.3 Zero Current Detector layout . . . . . . . . . . . . . . . . . . . . 103

5.4 Current-starved delay line . . . . . . . . . . . . . . . . . . . . . . 103

5.5 Current-starved delay line . . . . . . . . . . . . . . . . . . . . . . 104

5.6 D flip-flop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.7 Non-overlapping clock generator . . . . . . . . . . . . . . . . . . . 105

5.8 Output voltage ripple . . . . . . . . . . . . . . . . . . . . . . . . . 105

5.9 Output signals of non-overlapping clock generators . . . . . . . . 106

5.10 Output voltage and input signals to the gate drivers . . . . . . . . 107

5.11 Output of ZCD and the NMOS gate driver input . . . . . . . . . 107

xv

Chapter 1

INTRODUCTION

1.1 Motivation

Wearable electronic devices such as smart watches and fitness trackers have

gained vast significance over the past few years and transformed the way we live.

Smart watches have brought the power of smartphones onto a person’s wrist [6],

whereas fitness tracking wearables are used to monitor the body activity of the

person. The small size and light weight of these devices make these devices more

attractive among users.

Wearable electronics, which are portable in nature, are generally powered

by a rechargeable battery. Users do not want to recharge their batteries often,

which will demand these devices to have longer battery run time (time to drain

a fully charged battery before needing to recharge it). Fitness tracking devices,

which are always carried by the users so that the body activity is monitored

round the clock, requires the battery to run continually. However, the small sizes

of these devices also limits the battery size, which thereby limits the battery

energy. In order to achieve long battery run time with the battery running all the

time and with limited battery energy, an efficient energy management system is

required. Increasing the battery life is not the only solution. There are alternative

ways to supplement the battery, such as energy harvesting or wireless charging [7].

Regardless, designing an efficient energy management system can reduce the power

1

delivered from the battery, which thereby increases the time between successive

recharges.

The rechargeable batteries that are used in these devices have a fixed volt-

age range. However, the functional blocks integrated in these devices work under

different voltage domains and also have unique current requirements. Therefore,

voltage regulators, which convert the input battery voltage to different DC volt-

ages, are required for each voltage domain. Apart from the DC-DC conversion, the

efficient management of battery power is also determined by the DC-DC voltage

regulators. Therefore, all the voltage regulators together form a power manage-

ment system.

Voltage regulators are primarily classified as linear regulators and switching

regulators. A Low Drop-Out regulator (LDO) is a linear regulator, which is highly

efficient if the output voltage is close to the input voltage. However, an LDO will

lose its efficiency if the input to output voltage ratio increases. Switched-capacitor

and inductor-based converters fall under switching regulators, which are highly

efficient for any voltage conversion ratio. Inductor-based converters are preferred

over switched-capacitor converters, because of their better efficiency over wide

load current and input voltage range. The main disadvantage of the inductor-

based switching regulator is that the passive components, such as the inductor

and capacitor, consume more board area.

1.2 Objectives and Unique Contributions

The objective of this work is to design a highly efficient DC-DC buck

converter for ultra-low power applications in the IBM 180-nm CMOS process.

Power losses in the buck converter when it drives ultra light load currents are

studied before implementing the design.

The unique contributions of this project are:

2

1. Implemented the first DCM-mode synchronous buck converter at NMSU.

2. Implemented a highly efficient DC-DC buck converter for ultra-low power ap-

plications.

1.3 Report Organization

This technical report is organized into five chapters as follows.

Chapter 2 provides the literature review, where the basic operation of

the buck converter and the conventional control mechanisms used with the buck

converter are discussed. Published work on ultra-low power applications and sub-

circuit blocks used in the proposed design are also discussed here.

Chapter 3 includes the design of the buck converter, which includes the

derivations for calculating the inductor and capacitor values. This chapter also

describes the proposed control mechanism and the simulated results of the sub-

circuit blocks.

Chapter 4 presents the system-level simulation results at various load cur-

rents.

Chapter 5 shows the layout and presents the hardware measurement re-

sults. The layout techniques used and the area consumed by different blocks are

discussed.

Chapter 6 summarizes the proposed work and the results are compared

with other published work. Finally, the outstanding issues in this design are

analyzed.

3

Chapter 2

LITERATURE REVIEW

Ultra-low power applications, such as wearable medical devices, are widely used,

and their use is increasing exponentially. These devices should function contin-

uously, meaning, all the time, to achieve their purpose. The continuous func-

tionality demands more power from the battery and reduces battery run time.

Therefore, power from the battery should be used efficiently.

The aim of this chapter is to review inductor-based DC - DC buck con-

verters from the literature for ultra-light loads.

2.1 DC-DC Buck Converter

A DC-DC buck converter is a switching regulator that steps down a DC

input voltage to a lower DC output voltage. This section describes the steady-state

operation of a DC - DC buck converter, and the difference between conventional

and synchronous topologies.

2.1.1 Conventional DC-DC Buck Converter

This section gives an understanding of the operation of conventional inductor-

based buck converter in steady-state.

The open-loop topology of a conventional DC-DC buck converter is shown

in Fig. 2.1, in which the switch, diode, inductor and capacitor together form the

power stage and the load is powered from the power stage. The switch, which

4

is implemented as a MOSFET, is controlled by an external signal VP (t), whereas

the diode inherently acts as a switch without any external control.

iOUT

VIN

iL

VD

iIN

iC

LOAD

L

C

VSW

VLVOUT

VP(t)iD

Switch

+

Figure 2.1: Conventional Inductor-based Buck Converter

When the MOSFET switch is turned ON, current from the battery flows

through the inductor, to the output capacitor and load. The current does not

rise instantaneously even if the switch is turned ON instantaneously, because

of the inherent inductor property to resist sudden changes in current. Instead,

current starts increasing linearly and will continue to increase as long as the

MOSFET switch is turned ON. An increase in current causes the energy stored

in the inductor to increase. In this phase, the inductor is said to be charging.

The inductor current behavior during the ON time of the MOSFET switch is

graphically represented in Fig. 2.2

The currents and voltages in each circuit element are indicated in Fig. 2.1.

During the ON time of the switch, the inductor current iL is equal to the current

supplied from the battery iIN and, assuming no drop across the switch, the voltage

at node VSW is VIN . A positive voltage at VSW makes the diode reverse biased

and non-conducting, when the switch is turned ON.

5

iL(t)

< iL >

time

TSW

VP(t)

time

Switch

ON

Switch

OFF

VIN

0

0

Figure 2.2: Inductor Current

The ratio of the ON time of the switch (TON) to the switching period (TSW )

is defined as the duty cycle D

D ≡ TON

TSW(2.1)

Now, when the switch is turned OFF, current supplied to the inductor from

the battery stops and the inductor resists any instantaneous changes in current.

So, as soon as the switch is turned OFF, the diode is forced to turn ON by the

inductor to create a new current path. During this phase, the inductor acts as

a supply to the load and releases its energy. The inductor is discharging in this

phase and the current in the inductor linearly decreases. The inductor current

6

behavior during the OFF time of the MOSFET switch is graphically represented

in Fig. 2.2.

During the OFF time, the voltage at node VSW is pulled to –VD to make

the diode forward-biased. The diode starts conducting at this time and the diode

current iD is equal to the inductor current iL.

In Continuous Conduction Mode (CCM) and Boundary Conduction Mode

(BCM) (described in Section 2.2), the switching cycle ends just before turning the

MOSFET switch ON again. When in steady-state, the inductor current reaches

the same value at the end of every switching cycle.

From all of the above discussion, it can be observed that the voltage at

node VSW switches between VIN and – VD as shown in Fig. 2.3. The average value

of VSW is denoted as < VSW >.

VSW(t)

< VSW >

time0

VIN

-VD

TSW

Switch

ON

Switch

OFF

Figure 2.3: Switching node

7

The high frequencies in VSW are removed by the second-order low-pass L-C

filter and the average DC value appears at VOUT [8], as

VOUT = D · VIN − (1−D) · VD (2.2)

From the above equation it can be inferred that the output voltage can be set

using the duty cycle.

The major disadvantage of this topology is the diode drop voltage of VD,

(ranging from 0.6 V–1.0 V for a silicon diode) which is significant in battery

powered devices and will limit the efficiency of the converter. Schottky diodes,

which have a lower voltage drop of 0.15 V–0.45 V, can help in improving the

converter efficiency, but this voltage drop is still considered significant. Voltage

drops in the range of 25 mV–100 mV allow the converter to achieve the highest

efficiency. This is possible by replacing the diode with a second MOSFET. The

topology is discussed in the next section.

2.1.2 Synchronous DC-DC Buck Converter

The topology of a synchronous DC-DC converter is shown in Fig. 2.4,

where the diode in the conventional buck converter is replaced with an NMOS

switch.

The operation is similar to the conventional buck converter except an ad-

ditional control signal is required to turn ON and OFF the NMOS switch at the

right time. The switches operate in two different phases. Fig. 2.5 shows the syn-

chronous buck converter in phase 1, where the PMOS switch is turned ON and

NMOS switch is turned OFF. During phase 1, the inductor current increases lin-

8

iOUT

VIN

iL

iIN

LOAD

iC

M1

M2

L

C

VP(t)

VN(t)

VL VOUTVSW

Figure 2.4: Synchronous Buck Converter

early and stores energy in the inductor. The time period for which the converter

is in phase 1 is called the ON time, denoted as TON .

iOUT

VIN

iL

iIN

LOAD

iC

L

C

VL VOUT

+

RSW1

VSW1 VSW

Figure 2.5: Synchronous Buck Converter in ON state

Phase 2 operation is shown in Fig. 2.6, where the NMOS switch is turned

ON and PMOS switch is turned OFF. During phase 2, the inductor current de-

9

creases linearly and releases energy stored during phase 1. The time period for

which the converter is in phase 2 is called the OFF-time.

iOUT

VIN

iL

LOAD

iC

L

C

VL VOUT

VSW2

RSW2

iSW2

VSW

Figure 2.6: Synchronous Buck Converter in OFF state

While transitioning between phase 1 and phase 2, the PMOS and NMOS

switches should not be turned ON simultaneously. If both switches are turned

ON at the same time, they create a high current path from the input battery to

ground. This will cause a huge power loss. To avoid this, a small time window

called dead-time (see Fig. 2.7) is created between the two phases, in which both

transistors are turned OFF. However, during the dead-time, the inductor current

forces the voltage at node VSW to –0.7 V by turning ON the body diode of the

NMOS switch. This condition causes power loss, but is far lower than the power

loss caused by a high current path from the battery to ground. The switching

node waveform for a synchronous buck converter is shown in Fig. 2.8.

10

VN(t)

time

TSW

VP(t)

time

VIN

0

0

VIN

Dead-time

Switch1

ON

Switch2

ON

Figure 2.7: Dead-time

2.2 Modes of Operation

A buck converter is operated in three modes that are classified based on

the behavior of inductor current.

• Continuous Conduction Mode (CCM):

In CCM, the inductor current charges and discharges and never reaches zero.

As shown in the Fig. 2.9, at the end of the switching cycle, the inductor

stops discharging and starts charging immediately. Converters are generally

operated in CCM for moderate to high load currents (see Fig. 2.10). Lower

inductor current ripple and lower output voltage ripple are achieved through

11

VSW(t)

time0

VIN

-VDS

TSW

-VD

Switch1

ON

Switch2

ON

Dead-time

Figure 2.8: Switching node in synchronous buck converter

CCM. However, converters in CCM may end up requiring large inductor

values.

• Boundary Conduction Mode (BCM): In BCM, the inductor current

will completely discharge to zero. Although the inductor current reaches

zero, it will immediately start charging and never stays at zero. Inductor

current operating in BCM is represented in Fig. 2.9. Converters are operated

in BCM for moderate loads. The inductor current ripple and output voltage

ripple are higher than in CCM. When operated in BCM, converters can be

implemented with lower inductor values.

• Discontinuous Conduction Mode (DCM): In DCM, the inductor cur-

rent is fully discharged to zero and stays at zero for a certain period of

time before a new switching cycle starts. The inductor current in DCM is

represented in Fig. 2.9. For low load currents, converters are implemented

12

in DCM. The inductor current ripple and output voltage ripple are higher

than the other two modes. When operated in DCM, lowest inductor values

are achieved.

iL(t)

time

< iL,CCM >

TSW

0

< iL,BCM >

< iL,DCM >

(a)

(b)

(c)

Figure 2.9: Inductor current in three operation modes with constant L.

2.3 Control Techniques in a Buck Converter

DC-DC converters power various functional blocks that require different

currents and voltages to operate. The input to the converters is also not a constant

and changes with time. These requirements demand the converter to operate for

wide load current and input voltage ranges.

Whenever there are load current or input voltage changes, the output de-

viates from the desired voltage. To maintain the output voltage at a desired value

13

iL(t)

time

TSW

0

< iL >

(a)

(b)

(c)

Figure 2.10: Inductor current in three operation modes with constant average

load current.

in all conditions, a control mechanism is needed for the converter to detect any

changes at the output, and regulate it. Generally, feedback control is used to

achieve regulation and there are several techniques to implement it.

The feedback control techniques are primarily classified into Pulse Width

Modulation control and Pulse Frequency Modulation control.

2.3.1 Pulse Width Modulation Control (PWM)

PWM control operates with a fixed frequency for the entire load current

and input voltage range. An external clock is required to achieve fixed frequency.

As the name of the control technique suggests, during a load transient, the pulse

width, i.e., the duty cycle of the clock signal that drives the power switches, is

14

modulated to regulate the output. The duty cycle remains the same after reaching

steady-state.

The implementation of voltage-mode PWM control is shown in Fig. 2.11,

where the output voltage of the converter is fed back and compared with a refer-

ence voltage to detect any deviations in the output. The error detected is amplified

and this amplified signal causes the duty cycle to change in such as way as to bring

the output voltage back to the desired value.

L

C

LO

AD

VREF

Error

AmplifierComparator

Non-Overlap

Clock Gen

Compensator

Ramp

Genearator

Gate-Driver

Gate-Driver

Ramp

Waveform

Clock Signal

D

VOUT

VC

VFB

R1

R2

Power

Switches

Figure 2.11: Conventional Pulse Width Modulation Control Technique

As shown in the Fig. 2.11, VOUT is fed back to the input of an error amplifier

as VFB after sampling it through a resistor divider. The other input of the error

amplifier is an external reference voltage VREF . The difference between VFB and

15

VREF is integrated across the error amplifier as VC . The fixed frequency of the

clock signal CLK is used to generate a ramp signal. The output of the error

amplifier, VC , and the ramp signal are compared to generate a pulse signal at the

comparator output. The duty cycle of this pulse signal is determined by the value

of VC . In particular, D = VC/VIN

In steady-state, VFB and VREF are equal which produces a constant VC

and therefore maintains a constant duty-cycle. For high-to-low load changes,

overshoot occurs at VOUT and this deviation in VOUT will cause VC to decrease,

thereby reducing the duty-cycle. The change in duty-cycle will bring down the

voltage at VOUT and will eventually make VFB equal to VREF when the converter

reaches steady-state.

The comparator output which contains the duty-cycle information is fed

into non-overlapping clock generators, so that two synchronous signals with dead-

time are generated to drive the MOSFET switches. Because the switches are huge

in size, the output stage of the non-overlapping clock generators could not drive

them. To increase the drive strength of the signals, gate-drivers are used.

A compensation network is needed to maintain stability of the loop. How-

ever the speed of the loop (i.e., speed at which the duty-cycle changes) is limited

by the unity-gain frequency determined by the compensation network. Therefore,

this control technique tends to have a slow transient response.

2.3.2 Pulse Frequency Modulation (PFM)

PFM is another control technique where the steady-state frequency scales

proportionally with the load current. This control technique is most often imple-

mented when the converters operate in Discontinuous Conduction Mode (DCM).

16

The conventional PFM control technique is shown in Fig. 2.12. The sam-

pled output voltage VFB is fed back to the comparator and compared with VREF .

Whenever VFB goes below VREF , comparator output VC goes high and sets the SR

latch to turn ON the PMOS switch. The turn-OFF decision of the PMOS switch is

determined by the current comparator, whose inputs are the sensed inductor cur-

rent and a reference current limit. When the inductor current reaches a maximum

value, the output of the current comparator VCL goes high and resets the SR latch

to turn OFF the PMOS switch. Subsequently, the NMOS switch is turned ON

after a short dead-time determined by the non-overlapping clock generators. The

turn-OFF decision of the NMOS switch is determined by the Zero-Current Detec-

tor (ZCD), which compares the switching node VX with GND. When switching

node VX crosses GND (i.e., when the inductor current reaches zero), the output

of the ZCD goes high and the NMOS switch is turned OFF. Both PMOS and

NMOS switches remain OFF until VFB again goes below VREF .

When both switches are turned OFF, the converter enters DCM where the

converter is completely OFF and there is no energy supplied from the inductor.

During this period, the inductor current stays at zero and the output capacitor

discharges to supply current to the load.

When the load current changes from high to low, the increase in load resis-

tance will cause the RC time constant at the output to increase, which translates

into a reduced discharge rate of the output capacitor. Slow discharge of the out-

put capacitor will thereby delay the turn-ON of the PMOS switch, causing the

switching frequency to decrease.

The main advantage of PFM control is that the converter can have a wide

load current range because of its property of frequency scaling proportionally with

the load. The main disadvantage of converters implemented with PFM is that they

17

L

C

LO

AD

m

Comparator

Non-Overlap

Clock Gen

Gate-Driver

Gate-Driver

VOUT

VFB

R1

R2

Power

Switches

VREF

m

RQ

SQ

VX

Current

Comparator

Sensed Inductor

Current

Current Limit

ZCD and

Control Logic

PMOS

NMOS

VC

VCL

S01

Q

Figure 2.12: Conventional Pulse Frequency Modulation Control Technique based

on [1]

cannot be used to power circuits that are sensitive to varying switching frequencies.

Another disadvantage of PFM control, which comes with its implementation in

DCM, is higher output voltage ripple.

2.4 Buck Converter Operation in DCM

CCM and BCM (described in Section 2.2) modes have two phases of oper-

ation in one switching period where phase 1 is immediately started at the end of

phase 2. However, in DCM (also described in Section 2.2), at the end of phase 2,

a third phase of operation comes into the picture where the inductor current stays

18

at zero. The open-loop topology of the converter remains the same and can be

referred to in Fig. 2.13.

iOUT

VIN

iL

iIN

LOAD

iC

M1

M2

L

C

VP(t)

VN(t)

VL VOUTVSW

Figure 2.13: Synchronous Buck Converter

In DCM, phase 1 and phase 2 operation are similar to that described in

Section 2.1.2. At the end of phase 2, the inductor discharges completely when

the inductor current reaches zero. When the inductor current reaches zero, the

NMOS switch is turned OFF (the PMOS switch is already turned OFF) and

the converter enters phase 3. During phase 3, both switches remain OFF and

the inductor current stays at zero. The inductor current in DCM is shown in

Fig. 2.14. At the end of phase 3, the converter enters into phase 1 again, starting

a new switching cycle.

In section 2.1.2, the need for dead-time when transitioning between phase 1

and phase 2 is discussed, where dead-time is required at both transitions (phase 1

to phase 2 and phase 2 to phase 1). However in DCM, dead-time is only needed

when transitioning from phase 1 to phase 2. When transitioning from phase 2

to phase 3 and phase 3 to phase 1, switching action takes place only for one

19

switch. Therefore, the possibility for both switches turning ON at the same time

is avoided. The dead-time in DCM can be observed in Fig. 2.14.

VN(t)

time

TSW

VP(t)

time

VIN

0

0

VIN

Dead-time

iL(t)

time0

Both Switches

OFF

Phase 1

Phase 2 Phase 3 Phase 1 Phase 2 Phase 3

Both Switches

OFF

Figure 2.14: Gate Signals and Inductor Current in DCM

At the starting of phase 3, when the inductor current is zero, switching

node VSW is at zero volts and VOUT is a finite positive voltage. The voltage

difference between VOUT and VSW will cause charge on the output capacitor to

flow through the inductor to charge the parasitic capacitance CPAR,SW at node

20

VSW . The inductor and parasitic capacitance form an LC tank causing the charge

to flow back and forth. The charge flow appears as ringing at the switching node

VSW with ringing frequency expressed as,

fringing =1

2 · π ·√L · CPAR,SW

(2.3)

The ringing can be observed in Fig. 2.15. This ringing is damped by the parasitic

series resistance of the inductor. VSW eventually settles to a voltage equal to

VOUT .

Ringing at node VSW during phase 3 is unavoidable, but can be damped

using one of several techniques.

VSW(t)

time

0

VIN

-VDS

-VD

Phase 1 Phase 2

Dead-time

Ringing

<VOUT>

Phase 3

Both Switches OFF

Phase 1 Phase 2 Phase 3

Figure 2.15: Switching node with ringing

2.5 Previous Low Power Buck Converters

Switching DC-DC converters used in ultra-low power applications have to

be efficient across the entire load range for enhanced battery life [9]. The efficiency

21

of a DC-DC converter according to [10] is given as

η ≡ POUT

PIN

(2.4)

where

PIN = POUT + PLOSS (2.5)

and

PLOSS = Pconduction + Pswitching + Pcontroller (2.6)

Power losses in a buck converter are primarily classified into conduction

losses, switching losses and quiescent (i.e., controller) losses [10]. When all of these

losses are minimized, the converter is designed to achieve the highest efficiency at a

specific load current. However, over a wide load range, efficiency of the converter

will vary. As the load current goes low, conduction losses will scale down but

controller and switching losses will remain constant [11], assuming PWM control.

Therefore, at low load currents, controller and switching losses will dominate and

degrade the efficiency of the converter.

Scaling switching frequency with load current will minimize switching losses,

which improves efficiency at light loads. This can be achieved through the PFM

control technique which is implemented in [2, 3]. If the converter is operated in

CCM at light loads, reverse inductor current will degrade the efficiency. Therefore,

converters are operated in DCM at light loads.

The work in [2], which is designed for IOT/wearable applications, uses a

DCM enabled PFM control technique at light loads (500 µA to 10 mA). As shown

in Fig. 2.16, the PFM control consists of two comparators, one-shot triggers, an

SR latch and a Zero Current Detector (ZCD). One of the comparators makes the

22

Figure 2.16: Implementation of PFM mode controller from [2]

decision to turn-ON the PMOS switch when VFB is lower than VBGR. The turn-

OFF decision of the PMOS is made by a second comparator when the inductor

current reaches its peak value and subsequently, after a small dead-time, the

NMOS switch is turned ON. The turn-OFF decision of the NMOS switch is made

by the ZCD when the inductor current reaches zero. Both switches remain turned

OFF until VFB goes below VBGR to turn ON the PMOS device again.

23

As the load current goes low, the increase in load resistance will reduce

the discharge rate of the output capacitor which will delay the turn-ON event of

the PMOS transistor. The delay in the turn-ON event of the PMOS transistor

translates into an increased switching period and reduced switching frequency.

The efficiency of the converter is further increased by reducing the current

consumption in the controller. To reduce the power consumption in the Zero

Current Detector, an Adaptive Zero Current Detector is implemented which turns

ON only when the switching node VX is negative and turns OFF immediately after

the NMOS is turned OFF. Power consumption is reduced because the turn-ON

duration of the ZCD is reduced. However, the other two comparators are ON all

the time and will consume a significant amount of power.

At ultra-light loads, the converter switches to a new control technique

called retention mode control. The circuit blocks used in the retention mode

controller (shown in Fig. 2.17) are operated in the sub-threshold region to reduce

power consumption. The switching frequency goes as low as 32 Hz to achieve high

efficiency even at 10 µA load.

Even though the efficiency remains high at ultra-light loads, because of

very low switching frequency, the output voltage ripple is very high. Therefore,

this design is not suitable for powering ultra-low power applications that work

with tens of µA of load current. The transient response is also slow because of

very low switching frequency.

Another work in [3] uses a PFM control technique with constant on-time

(COT). As shown in Fig. 2.18, the comparator makes the decision to turn-ON the

PMOS switch and the constant on-time block ensures that the switch is turned

ON only for a fixed amount of time. After the PMOS is turned OFF, the NMOS

switch is turned ON. The turn-OFF of the NMOS switch is controlled by a self-

24

Figure 2.17: Implementation of retention mode controller from [2]

tracking zero current detector (ST-ZCD) by monitoring the switching node VX

and output voltage VOUT . The implementation of ST-ZCD is shown in Fig. 2.19.

The ST-ZCD is enabled only after the NMOS switch is turned OFF and

monitors if the NMOS switch is turned OFF at the right time. An inductor

current sampler is used to monitor the inductor current direction. The voltage

polarity across MFW determines the direction of the inductor current. If the

NMOS switch is turned OFF early, VOUT is higher than VX and the output of

ST-ZCD goes high, which will increase the pulse width of the NMOS gate signal.

25

Figure 2.18: Implementation of PFM controller from [3]

If the NMOS is turned OFF too early, VX is higher than VOUT and the output

of the ST-ZCD goes low to decrease the pulse width of the NMOS gate. The

output of the ST-ZCD toggles between high and low and adaptively determines

the appropriate pulse width.

Implementation of PFM control will cause the switching frequency to scale

down with load, thereby increasing efficiency at low loads. ST-ZCD and the

OFF-time controller are implemented digitally to reduce power consumption in

the controller. To achieve high efficiency even at ultra light loads, the control

technique switches to asynchronous mode. In asynchronous mode, the NMOS

switch is completely turned OFF and during the OFF-time, the inductor current

is conducted by the body diode of the NMOS switch. The forward voltage drop of

the diode will cause more conduction loss, but disabling the ST-ZCD and OFF-

time controller will significantly reduce the controller loss and improve overall

efficiency.

26

Figure 2.19: Implementation of of ST-ZCD [3]

The ST-ZCD ensures the precise turn-OFF of the NMOS switch. However,

after a transient event, the converter needs additional cycles to determine the

precise turn-OFF time of the NMOS switch, thereby increasing the settling time

of the converter. Another disadvantage that comes with the ST-ZCD is the need

for the inductor current sampler, which introduces additional power loss in the

control-loop.

27

2.6 PFM Controller Circuit Blocks

This section gives a detailed description of all of the circuit blocks used in

a PFM controller.

2.6.1 Comparator

A comparator compares the magnitude of two inputs and determines the

larger one at the output. Comparators can be implemented using an operational

amplifier (op-amp) with no negative feedback. Because of the high open-loop gain

of an op-amp, a small difference between the inputs can amplify to a large value at

the output. However, the output value is limited by the positive and negative rails

applied to the circuit. Therefore, the output is saturated either at the positive or

negative rail.

VIN

VSS

VP VM

M1 M2

M3 M4

M5 M6

M7

M8

M9

M10

VOUT

IBIAS

Figure 2.20: A two-stage differential amplifier with an output inverter imple-

mented as a comparator based on [4]

28

The topology of a comparator is shown in Fig. 2.20. It is a two-stage

CMOS op-amp with a third stage as an inverter. The first stage is a differential

amplifier and the second stage is a common-source amplifier, which is added to

increase the overall open-loop gain. An inverter at the output is added for fast

rise and fall times. The positive and negative inputs are named based on the effect

they have on the output, i.e., if the positive input (VP ) is more than the negative

input (VM), then the output goes to the positive rail and if it is less than VM , the

output goes to the negative rail.

2.6.2 Zero-Current Detector

A Zero-Current Detector (ZCD), as the name suggests, detects a zero volt-

age at a node and toggles its output whenever the node crosses zero volts.

In this design, a ZCD is used to prevent negative current in the induc-

tor, and ensures that the converter operates in Discontinuous Conduction Mode

(DCM).

Referring to Fig. 2.21, in DCM mode, when the inductor current reaches

zero, the voltage at node VSW reaches zero, whereas the output voltage VOUT is

at a finite positive voltage. At this point, if the NMOS switch remains ON, the

voltage difference between VSW and VOUT will discharge the output capacitor to

ground through the inductor and NMOS switch. The charge flow, i.e., the current,

in this direction introduces an additional power loss which needs to be avoided. In

order to prevent this negative current, inductor current information is monitored

to turn-OFF the NMOS switch at a precise time. This can be achieved either by

sensing the inductor current directly or monitoring the voltage at the switching

node (VSW ) that contains inductor current information.

29

iOUT

VIN

iL

iIN

LOAD

iC

M1

M2

L

C

VP(t)

VN(t)

VL VOUTVSW

Figure 2.21: Synchronous Buck Converter

The topology of a ZCD is shown in Fig. 2.22. It has three cascaded stages.

The first stage is a differential amplifier with inputs at the source terminals of

M2 and M3. The second stage is a common-source amplifier, which is added to

increase the overall gain. The third stage is an inverter for faster rise and fall

times at the output. The voltage VSW is fed to input V+, while V− is at ground.

The output of the ZCD is triggered whenever VSW crosses ground.

As shown in the Fig. 2.22, voltages V+ and V− are at source terminals of

M2 and M3. Because of varying VSW , the gate-source voltages of transistors M2

and M3 are generally different and will cause different currents to flow through

M2 and M3. The current through M5 is set by the current through M2, which

is then mirrored to M4. Because of the unequal currents flowing in M3 and M4,

node A is pulled high or low and this causes the output to trigger.

Initially, in phase 2, when the NMOS switch is ON, VSW is at a small

negative voltage. At the end of phase 2, the inductor discharges fully and when

30

VSW goes slightly above zero, the output of the ZCD triggers and turns OFF the

NMOS switch immediately, thereby preventing negative inductor current.

M1

M4IBIAS

VIN

M2 M3

M5

V+

M6

M7

M8

M9

VSS

VOUTA

V-

Figure 2.22: Three-stage comparator using common-gate differential amplifier [5]

2.6.3 S-R Latch

A Set-Reset or S-R Latch is a sequential logic circuit that is used as a basic

data storage element. It has two stable output states that change depending on

input and previous output states.

The topology of S-R latch shown in Fig. 2.23 has two NOR gates where

the outputs of two NOR gates are fed back as input to the other NOR gate. S

and R are external input signals and outputs are Q and Q.

There are four possible combinations of inputs S and R, which lets the

circuit operate in four different conditions. When S is high and R is low, output

Q goes high and Q goes low, and the circuit is said to be set. If R is high and S

is low, output Q goes low and Q goes high and resets the circuit. Now, if inputs

S and R remain low, both outputs hold the value from the previous state. This

31

VSS

S

R

Q Q

M1

M2

M3 M4

M5

M6

M7 M8

VIN

Figure 2.23: S-R Latch with NOR Gates

condition is called latch, where the output latches the previous state. The fourth

input state is generally avoided.

Table 2.1: Truth Table of SR Latch using NOR Gates.

S R Q Q

0 0 latch latch

0 1 0 1

1 0 1 0

1 1 0 0

2.6.4 D Flip-Flop

A D flip-flop is a data storage element that tracks its input data only when

the clock triggers, i.e., only on a rising or falling clock edge. The topology of the

D flip-flop shown in Fig. 2.24 is a positive-edge triggered flip-flop, which means

that the output tracks the input data only when the clock goes from low to high.

32

D

clk

clk

clkclk

RST

RST

Q

Q

TG1

TG2

TG3

TG4

clk clk

clk

clk

Figure 2.24: A positive-edge triggered D Flip-Flop with asynchronous-low reset

Initially, when the clock is low, TG3 is OFF, which will block the data

at input D from passing to output Q. When clk goes from low to high, i.e.,

when a positive edge occurs, TG3 is turned ON and will pass the information at

D to Q. After the positive edge, when clk remains high, and if D changes, the

output doesn’t track the information, because TG1 is turned off. At the negative

edge, i.e., when the clock goes from high to low, because TG3 will turn OFF, the

output does not track the input. RST is active low. If it goes low, the output Q

immediately goes low and the flip-flop is reset.

2.6.5 Current-Starved Delay Line

As the name suggests, a current-starved delay line has a delay that is

controlled by currents flowing in it. The delay from the input to output of a

digital inverter circuit is determined by the NFET or PFET ON resistance and

the capacitance at its output node. If current sources are placed in series with each

transistor, the amount of current through the MOS transistors becomes limited. In

this case, the value of the current sources establishes the charging and discharging

33

rate of the output capacitor, which translates into delay. The topology of the

current-starved delay line is shown in Fig. 2.25.

VIN

VSS

OUTIBIAS

M1 M2 M3 M4

M11 M12 M13 M14

INM5

M8

M6

M9

M7

M10

M15

M16

Figure 2.25: Current-starved delay line based on [4]

As shown in Fig. 2.25, four inverters are placed in series. Above and below

these inverters are the transistors that limit the current in the first three inverter

stages. A bias current is generated in bias transistors M1–M11 and mirrored to

transistors M2–M4 and M12–M14. The mirrored current will limit the current in

each inverter branch, thereby determining the charging and discharging rate of

the output capacitor. The more the current, the lower the amount of time to

charge or discharge the output capacitor and, hence, the lower the delay. Delay

can also be increased by increasing the number of current-starved inverter stages.

Transistors M15–M16 are used for fast rise and fall times at the output and

also to restore the logic levels.

34

2.6.6 Non-Overlapping Clock Generators

The need for dead-time between the clocks that drive the power switches

is discussed in Section 2.1.2. Dead-time is introduced between the clocks so that

both switches are not turned ON at the same time, to avoid shoot-through current

from the input supply to ground. A Non-Overlapping clock generator is used to

generate two synchronized signals with dead-time in between them.

The topology shown in Fig. 2.26 is an SR latch with NAND gates.

IN

VIN

VSS

Phi1

Phi2

Phi1

Phi2

P1

P2

NAND1

NAND2

Figure 2.26: Non-overlapping clock generator based on [4]

Inverters and resistors placed at the output of the NAND gates create

dead-time between Phi1 and Phi2. The resistors between the inverter stages will

increase the RC time constant, thereby increasing the delay. Delay can also be

increased by increasing the length of the transistors in the inverters.

To explain how the dead-time is created, initially the input is considered

high and the outputs P1 and P2 as low and high, respectively. As shown in

Fig. 2.27, at t1, when the input goes low, the output of NAND1 changes and this

change appears at P1 after some delay, while the output P2 remains unchanged.

After P1 goes high, then the output of NAND2 goes to zero and after some delay,

this is reflected at the output P2. The delay is created by the resistors and inverter

35

stages at the output of the NAND gates. This delay creates the dead-time between

the two signals Phi1 and Phi2. When the input again goes back high at t2, P2

changes first and only after P2 changes its state, after a certain delay P1 changes

its state.

In order to achieve equal dead-time at both t1 and t2, all inverters at the

output of NAND gates should have the same size, and also all the resistors should

be of the same value.

2.6.7 Gate Drivers

Power switches are sized huge because of the huge currents that flow

through them. Therefore, a huge capacitance is present at the gate. The out-

put stage of the non-overlapping clock generator is a small size inverter, which

lacks the necessary drive strength to drive such a huge capacitance. If the output

the non-overlapping clock generator drove the power switch, large rise and fall

times at the gate would result in huge losses during the switching event.

In order to drive such a huge gate capacitance, gate drivers are used. As

shown in Fig. 2.28, gate drivers are a series of inverters in which each successive

inverter stage has higher input capacitance compared to the previous stage. So,

in Fig. 2.28, the capacitance is scaled higher by a certain factor in each successive

stage. Therefore, the last inverter stage will have the highest capacitance that

can drive the huge gate capacitance of the MOS switch. In summary, the required

drive strength is provided by the gate driver. The topology of a gate-driver is

shown in Fig. 2.28.

36

p2

time

p1

time

VIN

0

0

VIN

IN

time

VIN

0

Dead-time

t2t1

Figure 2.27: Creating Dead-Time using Non-Overlapping Clock Generator fromFig. 2.26

37

IN

M1

M5

M2

M6

M3

M7

M4

M8

OUT

VIN

VSS

1 S S2 S

3

Figure 2.28: Gate Drivers with scale factor S

38

Chapter 3

DESIGN AND BLOCK-LEVEL SIMULATION RESULTS

The aim of this work is to design a highly efficient DC-DC buck converter for

ultra-low power applications.

3.1 Design Specifications and Motivation

Wearable devices are most often run by batteries because of their porta-

bility. The most widely used battery type is Li-ion because of its high energy

density. As such, the input voltage range of this design is 2.8 V – 4.2 V, which

is the voltage range of a Li-ion battery. The functional blocks in ultra-low power

applications, which consume ultra-low currents, will act as a load for the DC-DC

converter. Therefore, an ultra-low current range of 20 µA – 200 µA is chosen

for this design. The nominal operating voltage of the processor core in wearable

devices is 1.8 V [12], which is the output voltage of the converter. The output

voltage ripple should be < 2% of the output voltage to match with state-of-the-

art design constraints. To achieve high efficiency across the entire load range, the

switching frequency (fSW ) is kept variable. The targeted efficiency of this design

is 90%. Specifications of the proposed buck converter are tabulated in Table 3.1.

3.2 Buck Converter Design in Discontinuous Conduction Mode

This design is implemented using a synchronous buck converter (see Section

2.1.2) in Discontinuous Conduction Mode (DCM) operation.

With the specifications given, other parameters such as duty cycle, inductor

value and capacitor value should be calculated to complete the open-loop DC-

39

Table 3.1: Buck Converter Specifications.

Technology IBM 180 nm CMOS process

Input Voltage 2.8 V – 4.2 V

Output Voltage 1.8 V

Output Current 20 µA – 200 µA

Switching Frequency Variable

Output Voltage Ripple < 2% of VOUT

Targeted efficiency 90%

DC converter design. The derivations and calculations for these parameters are

detailed in following sections.

3.2.1 Derivations Duty Cycle of PMOS and NMOS Switches

The steady-state DCM operation of a buck converter is described in section

2.4. To calculate the duty cycle of the switches, the voltage across the inductor

is calculated in each phase and averaged across the entire switching period. The

voltage across inductor can be determined by applying KVL in each phase of

operation.

In phase 1, the PMOS switch is ON and NMOS switch is OFF, as shown

in Fig. 3.1. Applying KVL around the loop will result in

VL,P ON = VIN − VSW,1 − VOUT (3.1)

where VSW,1 is the voltage drop across the PMOS switch (VSD,P ).

40

If TON,P is the ON-time of the PMOS switch, the ratio of TON,P to the

switching period (TSW ) is the duty cycle (D1) of the PMOS switch. Thus,

D1 ≡TON,P

TSW(3.2)

iOUT

VIN

iL

iIN

LOAD

iC

L

C

VL VOUT

+

RSW1

VSW1 VSW

Figure 3.1: Synchronous Buck Converter in phase 1

Phase 2 operation, when the NMOS switch is ON and the PMOS is OFF,

is depicted in Fig. 3.2. Applying KVL across the loop will result in

VL,N ON = −VOUT − VSW,2 (3.3)

where VSW,2 is the voltage across the NMOS switch (VDS,N).

If TON,N is the ON-time of the NMOS switch, the ratio of TON,N to switch-

ing period (TSW ) is the duty cycle (D2) of the NMOS switch. Thus,

D2 ≡TON,N

TSW(3.4)

Fig. 3.3 shows phase 3 operation when both switches are OFF. Applying

KVL across the loop will result to

VL,OFF = 0 (3.5)

41

iOUT

VIN

iL

LOAD

iC

L

C

VL VOUT

VSW2

RSW2

iSW2

VSW

Figure 3.2: Synchronous Buck Converter in phase 2

The duration of phase 3 operation is TSW − TON,P − TON,N and the fraction of

time the converter is in phase 3 is 1−D1 −D2.

iOUT

VIN

iL

LOAD

iC

L

C

VL VOUTVSW

Figure 3.3: Synchronous Buck Converter in phase 3)

According to the inductor volt-second balance, the average voltage across

the inductor is zero in steady-state. The average voltage across the inductor in

42

DCM for one switching period is expressed as

< VL >=VL,P ON ·D1 · TSW + VL,N ON ·D2 · TSW + VL,OFF · (1−D1 −D2) · TSW

TSW(3.6)

Assuming VSW,1 ≈ VSW,2 and simultaneously solving (3.6), (3.1) and (3.3) for

VOUT , we get

VOUT =VIN ·D1

D1 +D2

− VSW,1 (3.7)

where 0 < (D1 +D2) < 1 in DCM.

In (3.7), because D1/(D1 +D2) is always less than one, it can be deduced

that VOUT is always lower than VIN , which also confirms that the converter acts

as a buck converter. Given a particular value for D1 + D2 between 0 and 1, D1

can be calculated directly from (3.7).

After further solving (3.7) for D2/D1, D2 can be calculated using

D2

D1

=VIN

VOUT + VSW,1

− 1 (3.8)

3.2.2 Derivation to Calculate Inductor Value

The voltage-current (V-I) relation of an inductor with constant voltage is

expressed as

VL = L · ∆iL∆t

(3.9)

where VL is the voltage across the inductor, L is the inductance, and ∆iL is the

change in the inductor current in ∆t seconds.

In phase 1, when the PMOS switch is ON, the inductor current starts

at zero and linearly increases to reach its peak value in D1 · TSW seconds. The

difference between the peak value and its starting value is defined as the inductor

current ripple and denoted as ∆iL. In phase 1, the voltage across the inductor,

VL,P ON , is given by (3.1). Therefore, for any given switching period TSW , the V-I

43

relation of the inductor in phase 1 can be expressed as

VIN − VSW,1 − VOUT = L · ∆iLD1 · TSW

(3.10)

The above equation can be further solved for inductance L

L =(VIN − VSW,1 − VOUT ) ·D1 · TSW

∆iL(3.11)

The inductor current ripple ∆iL in terms of IOUT can be derived from

Fig. 3.4. The average value of the inductor current, as observed from Fig. 3.4, is

expressed as

< iL >=1

2· (D1 +D2) ·∆iL (3.12)

Since the output current is equal to the average inductor current (< iL >=

IOUT ), the above equation can be rewritten as

IOUT =1

2· (D1 +D2) ·∆iL (3.13)

From the above equation, ∆iL can be expressed as

∆iL =2 · IOUT

D1 +D2

(3.14)

Rewriting (3.11) by replacing ∆iL from (3.14), we get

L =(VIN − VSW,1 − VOUT ) ·D1 · TSW · (D1 +D2)

2 · IOUT

(3.15)

Therefore, for a fixed switching period (TSW ), the inductance value can be

calculated from the above equation, since the parameters VIN , VOUT and IOUT

are known values from Table 3.1. D1 and D2 values can be calculated from (3.8)

and (3.7) by assuming a particular value for D1 + D2 between 0 and 1. VSW,1 is

assumed to be approximately equal to 50 mV.

Similarly in phase 2, when the PMOS switch is OFF and the NMOS switch

is ON, the voltage across the inductor is given in (3.3). The inductor current

44

iL(t)

time0

< iL >

D1 D2

D1+D2

TSW

D1+D2

D1 D2

TSW

Figure 3.4: Inductor current in DCM

linearly decreases from its peak value to reach zero in D2 · TSW seconds. ∆iL

is the same as in (3.14), since the inductor current varies by the same amount.

Therefore, the inductance can also be calculated from

L =(−VSW,1 − VOUT ) ·D2 · TSW · (D1 +D2)

2 · IOUT

(3.16)

Other than a sign change, (3.15) and (3.16) give identical results, so either of these

equations can be used to calculate the inductance.

Small size portable devices, such as wearables, severely limit the size of the

circuit components that can be integrated in these devices. Therefore, off-chip

components, such as inductors and capacitors, that are used in DC-DC converters

have stringent size and weight constraints during the design. The sizes of several

practical inductors available in the market are shown in Table 3.2.

As observed from Table 3.2, inductors of value 10 µH, 22 µH and 100 µH

have approximately the same size, whereas the inductor of value 220 µH is signif-

icantly larger than the others. Therefore, with the size as the limiting parameter,

the maximum inductor value that can be used in a practical design is limited to

100 µH.

45

Table 3.2: Several Inductors Available in the Market

Inductor (µH) Size, LxWxH (mm) RDCR (Ω)

10 2.0x1.9x0.6 1.4

22 3.0x3.0x1.1 0.81

100 3.0x3.0x1.3 3

220 7.3x7.3x4.1 0.7

The other parameter of the inductor that needs to be considered is the

parasitic series resistance of an inductor called the DCR (DC Resistance). Because

DCR is in series with the inductor, the inductor current (iL) that flows through

DCR will cause a power loss equal to

PLoss,DCR = i2L,RMS ·RDCR (3.17)

where

i2L,RMS =∆i2L · (D1 +D2)

3(3.18)

Assuming different values of L, values for D1 +D2 and ∆iL are calculated

from (3.15) and tabulated in Table 3.3. We note that higher inductor value results

in lower inductor current ripple, which, in general, results in lower output voltage

ripple. The other advantage is higher D1 + D2, which results in more relaxed

timing constraints.

3.2.3 Derivations to Calculate Capacitor Value

In a DC-DC converter, the output voltage is the voltage across the capac-

itor which consists of a DC value, VOUT and the ripple, ∆VOUT . In steady-state,

the voltage across the capacitor varies near to VOUT . The varying voltage across

the capacitor is defined as ∆VOUT

46

Table 3.3: Calculated Design Parameters

L D1 + D2 ∆iL

10 µH 0.01 1.88 mA

22 µH 0.0157 1.27 mA

100 µH 0.033 0.6 mA

The voltage-current relation of a capacitor is expressed as

iC(t) = C · dVC(t)

dt(3.19)

where iC(t) is the current through a capacitor, C is the capacitance and dVC(t) is

the change in the voltage across capacitor in dt amount of time.

The voltage ripple across a capacitor can be calculated from

∆VC(t) =1

C·∫iC(t)dt (3.20)

In a buck converter, most of the inductor current ripple goes into the capacitor

and the average inductor current goes to the load. The steady-state capacitor

current in a buck converter is represented in Fig. 3.5 which exactly looks like the

inductor current, except that the average capacitor current is zero.

The integral of capacitor current in the above equation is represented as

the triangle of current greater than zero. After integrating the current greater

than zero,

∆VC =1

C· 1

2· ((α · (D1 +D2) · TSW ) · (∆iL − IOUT )) (3.21)

where ∆iL−IOUT is the height of the triangular region and α·(D1+D2)·TSW is the

base of the triangular region which is the scaled time duration of (D1 +D2) ·TSW .

47

iC(t)

time0

D1+D2

TSW

D1+D2

TSW

<iL>

VOUT(t)

0

<VOUT>

time

Figure 3.5: Voltage and current in a capacitor

Parameter α is the scaling constant and can be expressed as

α =∆iL − IOUT

∆iL(3.22)

After solving (3.21) and (3.22), and since ∆VC = ∆VOUT , the output volt-

age ripple ∆VOUT is

∆VOUT =((α · (D1 +D2) · TSW ) · (∆iL − IOUT )2)

2 · C ·∆iL(3.23)

Capacitors are not ideal and every capacitor comes with a parasitic series

resistance called the Equivalent Series Resistance (ESR). The voltage across the

48

capacitor when there is ESR is shown in Fig. 3.6. So, ∆VOUT is modified and now

expressed as

∆VOUT = ∆VC + ∆iL ·RESR (3.24)

After replacing ∆VC from (3.21)

∆VOUT =((α · (D1 +D2) · TSW ) · (∆iL − IOUT )2)

2 · C ·∆iL+ ∆iL ·RESR (3.25)

Capacitance C can be calculated from

C =(α · (D1 +D2) · TSW ) · (∆iL − IOUT )2

2 ·∆iL · (∆VOUT −∆iL ·RESR)(3.26)

iL(t)

time0

D1+D2

TSW

D1+D2

TSW

<iL>

VOUT(t)

0

<VOUT>

time

Figure 3.6: Voltage and current in a capacitor that has ESR

49

Assuming a particular fSW , C can be calculated from (3.26), if D1 + D2

and ∆iL are known. D1 +D2 and ∆iL vary with each inductor value and can be

inferred from Table 3.3. Assuming fSW as 1 MHz, the corresponding values of C

for each inductor value are tabulated in Table 3.4.

Table 3.4: Calculated values of C

L C

10 µH 1.64 nF

22 µH 1.62 nF

100 µH 1.56 nF

3.2.4 Optimizing the Power Switches

Every MOSFET switch has two parameters associated with it, ON-resistance

(RON) and gate capacitance (CG). These two parameters introduce power losses

when the MOSFET is used as a switch in the buck converter. Conduction loss is

caused by RON , and is proportional to RON . Switching loss due to CG is also pro-

portional to CG. These two losses together are the major contributors of the total

power loss. The switches with their parasitic capacitances are shown in Fig. 3.7.

Parameters RON and CSG + CDG vary with switch size. RON is inversely

proportional to the width of the MOSFET, whereas CSG + CDG is proportional

to the width. The length of the MOSFET is always kept minimum to achieve

maximum speed. Therefore, when the MOSFET width increases, conduction loss

decreases and switching loss is increased. Minimum loss, i.e., maximum efficiency,

is achieved when both conduction and switching losses are equal. The MOSFET

width for which minimum losses are achieved is the optimum size of the switch.

50

VIN

LOAD

M1

M2

L

C

VP(t)

VN(t)

VOUTVSW

CSG, M1 CDG, M1

CGD, M1

CGS, M1

RDCR

RESR

RON, M1

RON, M2

Figure 3.7: Parasitic capacitances at the gate terminal of switches

Since a synchronous buck converter has two switches, PMOS and NMOS,

minimum losses are achieved when the combined conduction losses of both switches

equal the combined switching losses, that is,

Pconduction,PMOS + Pconduction,NMOS = Pswitching,PMOS + Pswitching,NMOS (3.27)

Conduction loss due to the PMOS switch is expressed as

Pconduction,PMOS = i2L,RMS ·RdsON,PMOS ·D1 (3.28)

where iL,RMS is the inductor RMS current flowing through the PMOS switch,

RdsON,PMOS is the ON resistance of the PMOS switch and D1 is the duty cycle of

the PMOS switch.

The RMS current of the inductor, i2L,RMS, in DCM is

i2L,RMS =∆i2L · (D1 +D2)

3(3.29)

After replacing (3.29) in (3.30), conduction loss is

Pcond,PMOS =∆i2L ·RdsON,PMOS ·D1 · (D1 +D2)

3(3.30)

51

Similarly, conduction loss in the NMOS switch is

Pcond,NMOS = i2L,RMS ·RdsON,NMOS ·D2 =∆i2L ·RdsON,NMOS ·D2 · (D1 +D2)

3

(3.31)

Switching losses of a MOSFET switch can be expressed as

PSwitching,Loss = CG · fSW · ∆V 2 (3.32)

where CG is the gate capacitance that consists of the gate-source capacitance

(CGS) and gate-drain capacitance (CGD), fSW is the switching frequency and ∆V

is the change in voltage across the capacitor.

At the turn-ON event of the PMOS switch, the switching loss due to the

gate-source capacitance of the PMOS switch is

PCSGP (ON),Loss =1

2· CGS,P · fSW · V 2

IN (3.33)

and the switching loss due to the gate-drain capacitance of the PMOS is

PCDGP (ON),Loss =1

2· CDG,P · fSW · (2 · VIN − VOUT )2 (3.34)

In DCM, before the turn-ON of the PMOS switch, when both switches are turned

OFF, the switching node VSW , which is the drain terminal of the PMOS, settles

to VOUT and the gate terminal is at VIN . So, the voltage across CDG is VOUT -VIN .

After the turn-ON, VSW is approximately VIN and the gate terminal is at zero

Volts, which makes the voltage across CDG −VIN . Therefore, the change in the

voltage across the capacitor is 2 · VIN − VOUT .

At the turn-OFF event of the PMOS switch, the switching loss due to the

gate-source capacitance is

PCSGP (OFF ),Loss =1

2· CGS,P · fSW · V 2

IN (3.35)

52

The source terminal of the PMOS switch is always a constant voltage VIN .

During the turn-ON or turn-OFF event of the switch, only the gate voltage toggles

between 0 and VIN . Therefore, the voltage change across CGS during both turn-

ON and turn-OFF is VIN

The switching loss due to the gate-drain capacitance is

PCDGP (OFF ),Loss =1

2· CDG,P · fSW · (2 · VIN − VD)2 (3.36)

When the PMOS switch is ON, the gate terminal is at zero volts and the voltage

at the drain terminal, which is the switching node VSW , is VIN . So, the voltage

across CDG is VIN . After the turn-OFF, during the dead-time, the gate terminal

is at zero volts and node VSW is at -VD (body diode drop voltage) which makes

the voltage across CDG as −VD − VIN . Therefore, the voltage change across CDG

is 2 · VIN − VD.

The total switching losses due to the PMOS switch are

Pswitching,PMOS = PCSGP (ON),Loss + PCDGP (ON),Loss (3.37)

+ PCSGP (OFF ),Loss + PCDGP (OFF ),Loss

which gives

Pswitching,PMOS = CGS,P · fSW · V 2IN +

1

2· CDG,P · fSW · (2 · VIN − VOUT )2

+1

2· CDG,P · fSW · (2 · VIN − VD)2 (3.38)

For the NMOS switch, at the turn-ON event of the NMOS switch, the

switching loss due to CGS of the NMOS is

PCGSN(ON),Loss =1

2· CGS,N · fSW · V 2

IN (3.39)

53

The switching loss due to CGD,N of the NMOS is

PCGDN(ON),Loss =1

2· CGD,N · fSW · (VIN − VD)2 (3.40)

Before turning ON the NMOS switch, during the dead-time, the voltage across

CGD,N is VD because the drain terminal of the NMOS switch is -VD and the gate

terminal is zero volts. After the NMOS switch is turned ON, the drain terminal is

at zero volts and the gate is at VIN which makes the voltage across CGD,N equal

to VIN . Therefore, the voltage change across CGD,N is VIN -VD.

At the turn-OFF event of the NMOS switch, the switching loss due to

CGS,N is

PCGSN(OFF ),Loss =1

2· CGS,N · fSW · V 2

IN (3.41)

Because the source terminal of the NMOS switch is always at ground, at both

turn-ON and turn-OFF events, only the gate terminal toggles between 0 and VIN .

Therefore, the voltage change across CGS,N during both events is VIN

The switching loss due to CGD,N is

PCGDN(OFF ),Loss =1

2· CGD,N · fSW · (VIN + VOUT )2 (3.42)

Before turning OFF the NMOS switch, the voltage across CGD,N is VIN because

the drain terminal of the NMOS switch is zero volts and the gate terminal is at

VIN . After the NMOS switch is turned ON, the drain terminal settles to VOUT and

the gate is at zero volts, which makes the voltage across CGD,N equal to −VOUT .

Therefore, the voltage change across CGD,N is VIN + VOUT .

The total switching losses in the NMOS switch are

Pswitching,NMOS = PCGSN(ON),Loss + PCGDN(ON),Loss (3.43)

+ PCGSN(OFF ),Loss + PCGDN(OFF ),Loss

54

which results in

Pswitching,NMOS = CGS,N · fSW · V 2IN +

1

2· CGD,N · fSW · (VIN − VD)2

+1

2· CGD,N · fSW · (VIN + VOUT )2 (3.44)

3.2.5 Choosing the Inductor Value

As described in Section. 3.2.2, the maximum size of the inductor is limited

to 100 µH and the best inductor has to be chosen among 10 µH, 22 µH and 100

µH.

For fixed design values of L and C, the converter will achieve maximum

efficiency when (3.27) is satisfied i.e., when the combined conduction losses of both

switches equals the combined switching losses. The losses introduced by RDCR of

the inductor will also reduce the efficiency.

The open-loop converter is optimized when (3.27) is satisfied. Assuming

fSW is 1 MHz, the corresponding capacitor value for a 22 µH inductor is 1.62 nF.

The optimized switch sizes are tabulated in Table 3.5 and the open-loop efficiency

of the converter is 88.8%

However, when fSW changes, switching losses will change and the converter

needs to be optimized again to satisfy (3.27) . The optimized switch sizes and

efficiency at different frequencies are tabulated in Table 3.5.

Table 3.5: Optimized switch sizes and efficiency for L = 22 µH ; RDCR = 0.81Ω

fSW C PMOS size (W/L) NMOS size (W/L) Efficiency

1 MHz 1.62 nF 40 µm/0.6 µm 20 µm/0.7 µm 86.4%

500 kHz 3.27 nF 100 µm/0.6 µm 50 µm/0.7 µm 88.76%

250 kHz 6.59 nF 150 µm/0.6 µm 75 µm/0.7 µm 90.8%

55

The optimized switch sizes and efficiency at different switching frequencies

for a 100 µH inductor are tabulated in Table. 3.6.

Table 3.6: Optimized switch sizes and efficiency for L = 100 µH ; RDCR = 3Ω

fSW C PMOS size (W/L) m NMOS size (W/L) m Efficiency

1 MHz 1.56 nF 40 µm/0.6 µm 20 µm/0.7 µm 91.12%

500 kHz 3.17 nF 40 µm/0.6 µm 20 µm/0.7 µm 92.1%

250 kHz 6.47 nF 100 µm/0.6 µm 50 µm/0.7 µm 93.1%

It can be observed that the converter implemented with a 100 µH inductor

has achieved higher efficiency in each case when compared to a 22 µH inductor.

Even though the RDCR of the 100 µH inductor is significantly larger, the power

loss caused by the RDCR is negligible at ultra-light loads.

The targeted efficiency for the converter is 90%. For a 100 µH inductor,

the open-loop converter achieved an efficiency of 92.1% at a switching frequency

of 500 kHz. If we assume 2% losses in the controller, then the converter achieves

the targeted efficiency of 90%. Even though the maximum efficiency achieved

is 93.1% at 250 kHz, operating at higher fSW would make the converter react

faster to transient events. Therefore, the converter is chosen to operate at fSW of

500 kHz.

The final component values used in the design are given in Table 3.7.

Table 3.7: Final Component Values of this design

L 100 µH

C 3.17 nF

56

3.2.6 Ripple-Based Constant On-Time Control Logic

The selected control-loop implementation is shown in Fig. 3.8. The com-

parator makes the turn-ON decision of the PMOS switch and the constant ON-

time generator will determine the ON-time of the PMOS switch. The circuit

blocks ZCD, SR latch and NAND logic together will determine the OFF-time of

the PMOS switch. The start-up circuitry shown is needed initially during the

start-up before the converter reaches the steady-state.

VREF

ANDNAND

EXT_RESET

PMOS_Gate_

Drivers

NMOS_Gate_

Drivers

Current-

Controlled

Delay Line

CLK

L

C

VIN

M1

M2

RDCR

RESR

VOUTVSW

LOAD

P_G

N_G

phi1

VIN

RQ

SQ

NA

ND

VSWD

CLK

CLK

RST

Q

Q

D

CLK

CLK

RST

Q

Q

Non-Overlap

Clock Gen

VIN

Current-

Controlled

Delay Line

ZCD_OUT

ZCD_OUT

ZCD_OUT

N_G

Start-Up Circuitry

mComparator

mZCD

PHI2

Constant On-time generator

SR

_O

UT

PHI1

PHI1

Figure 3.8: A Constant on-time control loop implementation

• Steady-State Operation

As shown in the Fig. 3.8, VOUT is fed back to the comparator and compared

with VREF . In steady-state, when VOUT goes below VREF , the output of

57

the comparator goes high. The signal ZCD OUT is always high and goes

low only at the end of phase 2, whereas the signal EXT RESET is always

high during the steady-state. The comparator output along with the start-

up circuitry will generate the clock for the D flip-flop. As the comparator

output is high, the clock of the positive-edge triggered flip-flop also goes

high, which generates a constant on-time signal at the output of the flip-

flop Q. Q is further fed as an input to the non-overlapping clock generator to

generate two output signals, phi1 and phi2 with a short dead-time in between

them. Signals phi1 and phi2 are complementary to each other. phi1, which

contains the constant on-time information, is fed as an input to the PMOS

gate driver. The inverted phi1 signal at the output of the PMOS gate driver

turns ON the PMOS switch for a fixed time, i.e., for a constant on-time.

In the steady-state, whenever the comparator output goes high, a positive

clock edge occurs at D flip-flop.

At the positive clock-edge of the D flip-flop, Q goes high and Q goes low.

Q is delayed through a current-starved delay line and the delayed signal is

fed to the asynchronous low reset of the flip-flop. So, after a predetermined

delay set by the current-starved delay line, the flip-flop is reset, which will

make Q go low. Therefore, Q stays high for a fixed time and then goes low,

which makes it a constant on-time signal. Whenever a positive clock-edge

occurs, a constant pulse width signal is generated at Q.

The ZCD, SR latch and NAND gate ensures that the NMOS switch remains

turned OFF while the PMOS switch is turned ON. The signals SR OUT and

phi2, which are inputs to the NAND gate, determine the turn-ON and turn-

OFF events of the NMOS switch. phi1, which is the input to PMOS gate

driver, is also fed to the SET input of the SR latch, whereas, ZCD OUT

58

is fed to the RESET input. When the PMOS switch is turned ON, phi1 is

high and ZCD OUT is low, which sets the SR latch and makes SR OUT

high. When phi1 is high, phi2 is low. With SR OUT high and phi2 low,

the output of the NAND gate is high. The signal at the NAND gate output

is inverted at the output of the NMOS gate drivers, which turns OFF the

NMOS switch.

When phi1 goes low after a constant on-time, the PMOS switch is turned

OFF. SR OUT , which was already high, latches in the previous output

state and remains high. After phi1 goes low, phi2 goes high after a short

dead-time. As both phi2 and SR OUT are high, the output of the NAND

gate goes low, which gets inverted at the output of the NMOS gate driver

to turn ON the NMOS switch. Therefore, the NMOS is turned-ON after a

short dead-time after the PMOS turn-OFF.

After the turn-ON of the NMOS switch, the inductor current starts decreas-

ing and switching node VSW is at a negative voltage. VSW is fed as an input

to the ZCD, while the other input is ground. The ZCD is enabled only after

the NMOS is turned ON and is disabled when the NMOS turns OFF. When

the ZCD is turned ON, VSW is at a small negative voltage, which gradu-

ally increases towards a positive voltage. As soon as the inductor current

reaches zero, VSW reaches zero volts. When VSW crosses zero, the output of

the ZCD goes high and resets the SR latch. This makes the output of the

NAND gate high and turns OFF the NMOS switch.

After the NMOS switch is turned OFF, both switches remain OFF and

during this period, the output capacitor discharges to supply power to the

load. During this period, the slope of VOUT , i.e., the discharge rate of the

59

output capacitor, is determined by the RC time constant of the output

capacitor (C) and the load resistance (R). The PMOS switch is again turned

ON when VOUT goes below VREF . If the discharge rate is slow, the time for

VOUT to go below VREF increases and the PMOS turn-ON is delayed. This

translates into an increased switching period (TSW ) and decreased switching

frequency (fSW ).

The steady-state DC-DC converter operation in one complete switching cy-

cle is explained in the above discussion. When VOUT again goes below VREF ,

comparator output goes high and a new switching cycle starts. The fre-

quency at which this cycle repeats is the switching frequency (fSW ).

• Start-Up Operation

If only for the steady-state operation, the comparator output can be directly

fed as a clock to the D flip-flop. However, directly feeding the comparator

output to the D flip-flop will never allow the converter to start-up. This

start-up issue and the solution to this issue are discussed below.

It is important to observe that it is the positive clock-edge to the D flip-

flop that generates the constant on-time for the PMOS switch and starts a

new switching cycle every time. In the steady-state, the comparator output

goes high when VOUT goes below VREF and the comparator output goes low

when VOUT goes above VREF . If the comparator output is directly fed to

the D flip-flop clock input, the positive clock-edge to the D flip-flop in every

switching cycle can be provided by the comparator when VOUT goes below

VREF .

However, at start-up, when VOUT is far below VREF , the comparator output

goes high once initially and will never switch to low until VOUT goes above

60

VREF . The start-up operation without the start-up circuitry can be observed

in Fig. 3.10. Initially, when the comparator output goes high, the D flip-flop

generates a constant on-time and turns ON the PMOS switch. The turn-ON

of the PMOS switch will charge the output capacitor, which increases VOUT

only by a small amount. Subsequently, the NMOS turns ON and when the

inductor current reaches zero, the output of the ZCD goes high and turns

OFF the NMOS. After the NMOS switch turns OFF, the comparator output

remains high because VOUT is still far below VREF . Because the comparator

output never toggles, a positive clock-edge never occurs, if the comparator

output is directly fed to the D flip-flop clock. Therefore, the PMOS switch

never turns ON and VOUT never gets charged. To overcome the clock issue for

the D flip-flop at the start-up, additional logic is needed that generates the

clock irrespective of the comparator. This additional logic is implemented

through the start-up circuitry shown in Fig. 3.8.

Figure 3.9: Start-up operation without start-up circuitry

61

The start-up circuit makes use of the inverted output of the ZCD (ZCD OUT )

to generate the clock for the D flip-flop. As shown in the figure, ZCD OUT

and the comparator output are the two inputs of the AND gate. Since

the comparator output is always high, the output of the AND gate toggles

whenever ZCD OUT toggles. The AND gate output can act as a clock

for the D flip-flop. However, the output of the AND gate is enabled only

when the EXTRESET signal is high. The EXTRESET is the externally

controlled reset signal which is initially kept low for a short period of time

until the flip-flop reaches a known state. ZCD OUT goes from high to low

when the inductor current reaches zero. Whenever ZCD OUT toggles, the

output of the NAND gate toggles, which acts as a clock to the D flip-flop.

Initially, since VOUT is less than VREF , the comparator goes high and the

PMOS turns ON. Subsequently, the NMOS switch turns ON and when the

inductor current reaches zero, ZCD OUT goes high, whereas ZCD OUT

goes from high to low. Whenever ZCD OUT goes from high to low, a pos-

itive clock-edge occurs, which turns ON the PMOS switch again starting a

new cycle. The successive turn-ON of the PMOS switch after each cycle will

charge VOUT to eventually reach the steady-state. The start-up operation

after including start-up circuit is shown in Fig. At start-up, the converter

operates in Boundary Conduction Mode (BCM)

As discussed above, when the output of the ZCD goes high, the NMOS

switch is turned OFF. The gate signal of the NMOS switch is in turn used

as an enable signal for the ZCD. As soon as the NMOS is turned OFF, it

immediately disables the ZCD , which makes the ZCD output low. So, the

output signal of the ZCD has a very short pulse width. It has to be noticed

that the clock of the D flip-flop, which is generated based on ZCD OUT

62

Figure 3.10: Start-up operation with start-up circuitry

will also have a very short pulse width if the immediate output of the ZCD

is directly used. From simulation, it was observed that this short pulse

width causes the D flip-flop clock to miss after a few cycles during start-up.

Therefore, the pulse width of the ZCD output is increased using the D flip-

flop and current-starved delay line at the output of the ZCD. The start-up

operation before and after increasing the pulse width is shown in Fig. 3.11

and Fig. 3.13, respectively. The event when the clock misses is shown in

Fig. 3.12.

• Power Reduction Techniques in the Control Loop

The power consumption for any circuit block is expressed as

PQ = VIN · IQ (3.45)

where VIN is the supply voltage and IQ is the current consumed by that

particular circuit block. From the above equation, it can be observed that

63

Figure 3.11: Start-up when the clock misses

Figure 3.12: D flip-flop clock misses as RST is low when the clock edge occurs

the power consumption in the circuit is directly proportional to the current

consumed by that circuit.

Power consumption in the controller, which is not significant at high loads,

will dominate at light and ultra-light loads. Therefore, it is essential to

64

Figure 3.13: Start-up

minimize the power consumption in the controller when operating at light

loads.

The most power hungry elements in any mixed-signal design are analog

circuit blocks. Analog circuits consume current all the time, whereas digital

circuits consume current only at a switching event. In this design, the

comparator and the ZCD are the only analog circuit blocks, while all other

blocks are digital.

The comparator used in this design is a two-stage differential amplifier with

the third stage as an inverter, as shown in Fig. 3.14. The comparator is

required to operate all the time. Power consumption in the comparator

can be reduced to a certain extent by reducing the bias current. However,

the speed and the delay specifications of the comparator will limit the bias

current from going lower. The output inverter stage in the comparator is

not bias limited and the current consumption depends on the size of the

transistors. The inverter consumes current only when the output changes

65

its state. However, the average current consumption in the inverter stage

was significant as observed from the simulations. To reduce the power con-

sumption, the length of both PMOS and NMOS transistors in the inverter

stage of the comparator are increased.

VIN

VSS

VP VM

M1 M2

M3 M4

M5 M6

M7

M8

M9

M10

VOUT

IBIAS

Figure 3.14: A three-stage CMOS differential amplifier implemented as a com-parator

The other analog circuit in the control-loop is the Zero Current Detector

(ZCD) (see Fig. 3.15b), which is also a comparator. The limitation on

the bias current as discussed above is also applicable to the ZCD. In this

design, the ZCD is enabled by the NMOS gate signal so that it is turned

ON only when the NMOS is turned ON and is turned OFF in the remaining

time. Because the ZCD is turned ON only for a short duration, the power

consumption is greatly reduced.

The schematic of the current-starved delay line (CSD) is shown in Fig. 3.16.

There are two current-starved delay line blocks in this design, in which the

delay generated directly depends on the current flow through the individual

66

M1

M4IBIAS

VIN

M2 M3

M5

V+

M6

M7

M8

M9

VSS

VOUTA

V-

Figure 3.15: A three-stage comparator using common-gate differential amplifier

branches. To reduce the power consumption, the current consumption in

the first two branches is kept low. The third branch requires more current

to achieve fast rise and fall times and also to restore the logic levels. The

two CSD blocks share the same bias current in order to further reduce the

power consumption.

To reduce the power consumption in the digital blocks, the PMOS and

NMOS transistor ratio which are generally implemented with 2:1 ratio is vi-

olated. Instead the PMOS to NMOS ratio is maintained 1:1 in all the digital

blocks. Maintaining 1:1 ratio will decrease the capacitance, which thereby

reduces the power losses. The power loss in a digital block is expressed as

PSW = C · fSW · (VIN)2 (3.46)

67

VIN

VSS

OUTIBIAS

M1 M2 M3 M4

M11 M12 M13 M14

INM5

M8

M6

M9

M7

M10

M15

M16

2xIBIAS 4xIBIASIBIAS 8xIBIAS

IBIAS 2xIBIAS 4xIBIAS 8xIBIAS

Figure 3.16: A current-starved delay line

3.3 Design and Simulation Results of Control Circuitry

The simulations of the design are done in the IBM 7HV 180 nm CMOS

technology using Cadence Design Tools. All the MOSFET devices used in the

design are 5-V devices.

3.3.1 Comparator

The comparator used in this design is a two-stage PMOS differential am-

plifier cascaded with an inverter as the third stage (see Fig. 3.17). The MOSFET

sizes are designed to operate in the subthreshold region at a bias current of 50 nA

and are tabulated in Table 3.8.

The length of MOSFETs M3, M4, M7 and M8 are maintained minimum so

that the comparator operates with maximum speed. MOSFETs M9 and M10 form

an output inverter stage. Generally, the length of the MOSFETs in the inverters

is maintained minimum to achieve fast rise and fall-times. However, when M9 and

68

VIN

VSS

VP VM

M1 M2

M3 M4

M5 M6

M7

M8

M9

M10

VOUT

IBIAS

Figure 3.17: A three-stage CMOS differential amplifier implemented as a com-parator

M10 are implemented with minimum length, it was observed from the simulations

that the inverter is consuming a significant amount of shoot-through current.

The large current consumption will cause significant power loss and degrades the

efficiency of the whole converter. In order to reduce the current consumption, the

width of the MOSFETs is reduced and the length is increased. This makes the

rise and fall times larger, which is not good. However, because this work is a low

power design, power consumption is traded-off with the rise and fall times.

The comparator operation can be seen in Fig. 3.18, where the inputs are

VM and VP . Input VP is a constant reference voltage of 1.8 V and the actual VOUT

from the converter is fed to VM , which moves between 1.79 V and 1.83 V. It can

be observed from the plots that, when VM goes below VP , the comparator changes

its state.

69

1.78

1.79

1.8

1.81

1.82

1.83

1.84

(V)

VPVM

0

1

2

3

4

(V)

Time 0.5(µs/Div)

VOUT

Figure 3.18: Comparator Operation

70

Table 3.8: MOSFET sizing in the comparator

MOSFET Size (W/L) m

M0 and M1 1 µ/1.4 µ

M2, M5 and M6 2 µ/1.4 µ

M3 and M4 1 µ/0.7 µ

M7 and M8 2 µ/0.7 µ

M9 and M10 0.5 µ/8µ

3.3.2 Zero-Current Detector (ZCD)

The ZCD used in this design is a two stage common-gate differential ampli-

fier with the third stage cascaded as an inverter. The MOSFET sizes are tabulated

in Table 3.9.

The schematic of the ZCD is shown in Fig. 3.19. Transistors M2 and

M3, which form an input stage, are sized different to create an intentional offset

between the two inputs.

M1

M4IBIAS

VIN

M2 M3

M5

V+

M6

M7

M8

M9

VSS

VOUTA

V-

Figure 3.19: A three-stage comparator using common-gate differential amplifier

71

The working of the ZCD is shown in the Fig. 3.20. The negative input VM

is connected to the ground, whereas the actual switching node VSW is fed to the

positive input VP . From the figure, it can be observed that whenever VSW crosses

ground, ZCDOUT changes its state.

Table 3.9: MOSFET sizing in the ZCD

MOSFET Size (W/L) m

M1 1.4 µ/1.4 µ

M2 22.4 µ/1.4 µ

M3, M4 and M5 5.6 µ/1.4 µ

M6 2.8 µ/0.7 µ

M7 1.4 µ/0.7µ

M8 and M9 0.5 µ/0.7µ

3.3.3 Non-Overlapping Clock Generator (NOCG)

The NOCG is implemented using an SR-latch with NAND gates. The

topology is shown in Fig. 3.21.

All the resistors used in NOCG have the same value, which is equal to

1.2 kΩ. The sizes of the inverters are tabulated in Table 3.10. Signals phi1 and

phi2 and the dead-time generated between them is shown in Fig. 3.22. The dead-

time on both sides is the same and is equal to 4 ns. Signals phi1 and phi2 are

complementary to each other.

3.3.4 Current-Starved Delay-Line (CSD)

The schematic of the CSD is shown in Fig. 3.23 and the MOSFET sizes

are tabulated in Table. 3.11. The bias current is 65 nA.

72

−1

−0.5

0

0.5

1

(V)

VSWVSS

0

1

2

3

4

(V)

Time 0.05(µs)/Div

ZCD_OUT

Figure 3.20: Inputs and output signals of ZCD

There are three inverter stages in series, which are current limited by the

top and bottom transistors. The operation of the CSD is described in Sec. 2.6.5.

73

IN

VIN

VSS

Phi1

Phi2

P1

P2

NAND1

NAND2

invx1 invx2

invx3 invx4

invx5

invx6

Figure 3.21: An SR latch with NAND gates implemented as a non-overlappingclock generator

Table 3.10: MOSFET sizing in the NOC

Block PMOS size (W/L) m NMOS size (W/L) m

invx1, invx2, invx3 and invx4 0.5 µm/2 µm 0.5 µm/2 µm

invx, invx5 and invx6 0.5 µm/0.7 µm 0.5 µm/0.7 µm

NAND1 and NAND2 0.5 µm/2 µm 0.5 µm/2 µm

The current in the third-stage is more than the previous stages. The cur-

rent in the first two stages is maintained low in order to reduce the power con-

sumption. The current in the third stage is made higher for fast rise and fall

times.

The length of transistors M15 and M18 of an inverter stage is increased,

instead of maintaining minimum sizes in order to decrease the power consumption.

The delay achieved with the bias current of 65 nA is 94 ns, which can be

observed in Fig. 3.24.

74

0

0.5

1

1.5

2

2.5

3

3.5

4

(V)

Time 10(ns)/Div

PHI1PHI2

Figure 3.22: Dead-time generated by non-overlapping clock generator

75

VIN

VSS

OUTIBIAS

M1 M2 M3 M4

M11 M12 M13 M14

INM5

M8

M6

M9

M7

M10

M15

M16

2xIBIAS 4xIBIASIBIAS 8xIBIAS

IBIAS 2xIBIAS 4xIBIAS 8xIBIAS

Figure 3.23: A current-starved delay line

Table 3.11: MOSFET sizing in the ZCD

MOSFET Size (W/L) m

M1 and M11 1.4 µ/5.6 µ

M2 1.4 µ/2.8 µ

M3 1.4 µ/1.4 µ

M4 2.8 µ/1.4 µ

M5, M6, M7, M8, M9 and M10 0.5 µ/0.7µ

M15 and M16 0.5 µ/3µ

76

0.8 0.9 1 1.1 1.2 1.3 1.4 1.50

0.5

1

1.5

2

2.5

3

3.5

4

(V)

Time 0.1(µs/Div)

Delay

INOUT

Figure 3.24: Delay from input to output

77

Chapter 4

SYSTEM-LEVEL SIMULATIONS OF BUCK CONVERTER

This chapter discusses the simulation results of the converter and control-loop.

Simulations across a wide load current range are discussed here.

4.1 Simulation Results at Minimum Load

The converter is optimized at 20 µA load current which is also the minimum

load of this design. The test bench for the system level converter is shown in

Fig. 4.1. The simulated waveforms of the output voltage and the clock signals

are shown in Fig. 4.2. The inductor current, output current and switching node

waveforms can be observed in Fig. 4.3

The steady-state waveforms are shown in Figs. 4.4 and 4.5. In steady-state,

〈VOUT 〉 settles to 1.814 V and the the output voltage ripple (∆VOUT ) is 33.36 mV.

The constant on-time of the PMOS switch is 94.5 ns. The switching fre-

quency is 184.8 kHz. The total time of operation is 156 ns, whereas, the switching

period is 5.41 µs. Therefore, it can be said that the converter operates in deep

DCM. Results from the simulations are summarized in Table 4.1.

The role of the ZCD in the converter can be observed in Fig. 4.6. As soon

as VSW reaches zero, ZCD OUT goes high, which turns OFF the NMOS switch.

It can also be observed from the same figure that the inductor current also reaches

zero when VSW reaches zero.

The efficiency of the converter can be calculated from

η ≡ POUT

PIN

(4.1)

78

L= 100 µHRDCR= 3 Ω

RESR= 2 Ω

C = 3.3 nF

90 kΩ

3.6 V

+-

1.8 V

VIN

VIN

_C

L

I BIA

S_

CM

P2

I BIA

S_

Del

ayli

ne

VREF

VSW

VOUT

VS

S

VS

S_

CL

IB

IAS

_C

MP

1

IB

IAS

_C

MP

1_

AD

D

5 MΩ

35 MΩ

60

2 MΩ

+- +

-

+-

+-

0 V 0 V

3.6 V

RL

OA

DBuck_Converter_

System_Level

Figure 4.1: Test bench for system level buck converter

Table 4.1: Simulation results at 20 µA load current

Parameter Value

〈VOUT 〉 1.81 V

∆VOUT 33.46 mV

fSW 184.8 kHz

∆iL 1.38 mA

POUT from simulations is 36.4 µW and the input power is 42.8 µW. The efficiency

of the converter at 20 µA load is 85%.

79

0

0.5

1

1.5

2

(V)

VOUT

−1

0

1

2

3

4

(V)

Output Voltage and Gate signals

PMOS Gate

−1

0

1

2

3

4

(V)

Time 10(µs)/Div

NMOS Gate

Figure 4.2: Output voltage and gate signals at 20 µA load current.

The input power includes all of the losses in the converter. In particular,

PIN = POUT + PLOSS (4.2)

80

0

0.5

1

1.5

x 10−3

Inductor Current

−1

0

1

2

3

4

(V)

Time 10(µs)/Div

Inductor Current and Switching Node

Switching Node

Figure 4.3: Inductor current and switching node at 20 µA load current.

The total power loss in the converter is 6.4 µW of which 3.4 µW are the losses in

the controller (PLoss,Controller) and the other 3 µW losses are due to the switches

(PLoss,Switches). The total quiescent current in the controller is 830 nA.

81

1.78

1.79

1.8

1.81

1.82

1.83

1.84

1.85

(V)

VOUT

0

0.5

1

1.5

2

2.5

3

3.5

(V)

Time 10(µs)/Div

Output Voltage and Gate signals

PMOS GateNMOS Gate

Figure 4.4: Output voltage and gate signals in steady-state at 20 µA load current.

4.2 Simulation Results Across the Load Range

The maximum load current for this converter is 200 µA. The waveforms of

output voltage, gate signals, inductor current and output current equal to 200 µA

82

can be observed in Figs. 4.8 and 4.9. The switching frequency has scaled up

with the load current and the observed fSW at 200 µA load is 1.76 MHz which is

one order of magnitude more than the switching frequency at 20 µA. The output

voltage ripple is 27 mV which is less than the voltage ripple at 20 µA.

POUT measured in simulations is 360.4 µW. The input power is 400.9 µW.

The efficiency of the converter (POUT/PIN) at 200 µA load is 87.1 %.

The total power loss in the converter is 40.5 µW of which 20.9 µW are the

losses in the controller (PLoss,Controller) and the remaining 19.6 µW losses are due

to the switches (PLoss,Switches).

Simulation results across the entire load current range are summarized in

Table 4.2. When the load current changes, the inductor current ripple, average

output voltage and on-time of the PMOS switch remain constant. However, pa-

rameters such as switching frequency, output voltage ripple and efficiency of the

converter change with the load.

Table 4.2: Simulated Performance Across Load Current Range

Load Current fSW ∆VOUT η

20 µA 184.8 kHz 33.36 mV 85.1 %

40 µA 368.7 kHz 32.45 mV 86.28 %

100 µA 1 MHz 30.3 mV 86.33 %

150 µA 1.36 MHz 27.95 mV 86.85 %

200 µA 1.76 MHz 26.47 mV 87.1 %

It can be observed from Table 4.2 that the switching frequency scales pro-

portionally with the load, efficiency increases with the load, whereas the output

voltage ripple reduces with the load.

83

4.3 Load Transient Response

The load transient testbench is shown in Fig. 4.10, in which a load re-

sistance of 90 kΩ is fixed and an additional resistance in series with an NMOS

switch is added in parallel to the 90 kΩ resistor. When the switch is OFF, the

load current is 20 µA and when the switch turns ON, additional load current of

180 µA is added to increase the load current from 20 µA to 200 µA. The switch

is controlled by a pulse signal. The rise and fall times of the load current as it

changes between 20 µA and 200 µA are 10 ns.

The output voltage response to the load transient is shown in Fig. 4.11. It

can be observed from the figure that whenever the load current (IOUT ) changes, the

switching frequency of the converter changes. The change in switching frequency

can be observed in the output voltage and gate signal waveforms.

A low to high load transient is shown in Fig. 4.12. From the waveform, it

can be observed that the load current changes from 20 µA to 200 µA. The rise

time is 10 ns. However, the converter does not react to fast changes in the load

current and will not be able to source the increased load current.

As soon as VOUT goes below VREF , the PMOS switch turns ON to charge

VOUT . As discussed previously, the decrease in the load resistance will cause the

switching frequency to increase. Because of the increased switching frequency,

the output capacitor charges more frequently to bring back VOUT to the desired

value. The immediate turn-ON of the PMOS switch when VOUT goes below VREF

prevents undershoot. As such the converter immediately transitions to steady-

state at the higher load current.

The output voltage behavior when a high to low load transient (200 µA

to 20 µA) occurs is shown in Fig. 4.13. When the high-low load transient occurs,

the OFF-time of the converter (i.e., the duration of the converter when both

84

switches are OFF) is increased to bring back VOUT to the desired value. There

is no overshoot in the output voltage. A new steady-state is reached within one

cycle.

How good the converter regulates the average output voltage to a steady-

state changes in load current is known as load regulation. Load regulation is

defined as the ratio of change in the steady-state output voltage to the steady-

state change in load current. Load regulation is expressed as

Looksterrible−−Loadregulation =∆ 〈VOUT 〉∆ 〈IOUT 〉

=VOUT,MAX − VOUT,MIN

IOUT,MAX − IOUT,MIN

(4.3)

For a low-to-high load current change, the steady-state 〈VOUT 〉 at 20 µA is 1.814 V

and the steady-state 〈VOUT 〉 at 200 µA is 1.801 V. The load regulation is

1.814 V − 1.801 V

200 µA− 20 µA= 72 V/A (4.4)

4.4 Line Transient Response

The testbench for line transient is shown in Fig. 4.14, where a pulse signal

is given as an input supply source. The rise and fall times of the pulse signal are

10 ns each. The input varies between 3.0 V and 3.6 V in 10 ns.

The line transient response can be observed in Fig. 4.15.

The line transient response when the input varies from 3.6 V to 3.0 V is

shown in Fig. 4.16

The line transient response when the input varies from 3.0 V to 3.6 V is

shown in Fig. 4.17

Line regulation is defined as the ratio of the steady-state change in output

to the steady-state change in input, which is expressed as

Lineregulation =∆VOUT,STEADY−STATE

∆VIN,STEADY−STATE

=VOUT,MAX − VOUT,MIN

VIN,MAX − VIN,MIN

(4.5)

85

For high to low line transient, i.e., when the input change from 3.6 V to

3.0 V, the steady-state VOUT at 3.6 V is 1.814 V and the steady-state VOUT at

3.0 V is 1.807 V. The line regulation is

1.814 V − 1.807 V

3.6 V − 3.0 V= 11.6 mV /V (4.6)

Similarly, for low to high line transient, i.e., when the input change from

3.0 V to 3.6 V, the steady-state VOUT at 3.6 V is 1.814 V and the steady-state

VOUT at 3.0 V is 1.807 V. The line regulation is

1.814 V − 1.807 V

3.6 V − 3.0 V= 11.6 mV /V (4.7)

86

0

0.5

1

1.5

x 10−3

(A)

Inductor Current

−1

0

1

2

3

4

(V)

Time 1(µs)/Div

Inductor Current and Switching Node

Switching Node

Figure 4.5: Inductor current and switching node in steady-state at 20 µA loadcurrent.

87

−1

−0.5

0

0.5

(V)

Switching NodeVSS

0

2

4

(V)

ZCD_OUT

0

2

4

(V)

NMOS Switch

0

0.5

1

1.5x 10

−3

(A)

Time 0.02(µs)/Div

Inductor Current

Figure 4.6: Detecting zero volts at node VSW at 20 µA load current.

88

−1

−0.5

0

0.5

1

1.5

2

2.5

3

3.5

4

(V)

Time 0.1(µs)/Div

Switching node

Figure 4.7: Ringing at the switching node in phase 3 at 20 µA load current.

89

1.78

1.79

1.8

1.81

1.82

(V)

VOUT

0

1

2

3

4

(V)

Time 0.2µs)/Div

PMOS GateNMOS Gate

Figure 4.8: Output voltage and gate signals in steady-state at a load current of200 µA.

90

−1

0

1

2

3

4

(V)

Switching node

0

0.5

1

1.5

x 10−3

(A)

Time 0.2(µs)/Div

Inductor current

Figure 4.9: Switching node and inductor current in steady-state at a load currentof 200 µA.

91

L= 100 µHRDCR= 3 Ω

RESR= 2 Ω

C = 3.3 nF

90 kΩ

3.6 V

+-

1.8 V

VIN

VIN

_C

L

I BIA

S_

CM

P2

I BIA

S_

Del

ayli

ne

VREF

VSW

VOUTV

SS

VS

S_

CL

IB

IAS

_C

MP

1

IB

IAS

_C

MP

1_

AD

D

5 MΩ

35

60 MΩ

2 MΩ

+- +

-

+-

+-

0 V 0 V

3.6 V

RL

OA

DBuck_Converter_

System_Level

+

10 kΩ

Figure 4.10: Load regulation testbench

92

0

1

2

x 10−4

(A)

Load Transient

IOUT

1.76

1.78

1.8

1.82

1.84

(V)

VREFVOUT

0

2

4

(V)

PMOS GateNMOS Gate

0

0.5

1

1.5x 10

−3

(A)

Time 50(µs)/Div

Inductor Current

Figure 4.11: Load transient response

93

0

1

2

x 10−4

(A)

Load Transient

IOUT

1.76

1.78

1.8

1.82

1.84

(V)

VREFVOUT

0

1

2

3

4

(V)

PMOS GateNMOS Gate

0

0.5

1

1.5x 10

−3

(A)

Time 2(µs)/Div

Inductor Current

Figure 4.12: 20 µA to 200 µA load transient

94

0

1

2

x 10−4

(A)

Load Transient

IOUT

1.76

1.78

1.8

1.82

1.84

(V)

VREFVOUT

0

1

2

3

4

(V)

PMOS GateNMOS Gate

0

0.5

1

1.5x 10

−3

(A)

Time 2(µs)/Div

Inductor Current

Figure 4.13: 200 µA to 20 µA load transient

95

L= 100 µHRDCR= 3 Ω

RESR= 2 Ω

C = 3.3 nF

90 kΩ+-

1.8 V

VIN

VIN

_C

L

I BIA

S_C

MP

2

I BIA

S_

Del

ayli

ne

VREF

VSW

VOUT

VS

S

VS

S_

CL

IB

IAS

_C

MP

1

IB

IAS

_C

MP

1_A

DD

5 MΩ

35 MΩ

60

2 MΩ+

-+-

0 V 0 V

RL

OA

DBuck_Converter_

System_Level

+

Figure 4.14: Testbench for line transient simulation

96

2.83

3.23.43.6

(A)

V_IN

1.76

1.781.8

1.821.84

(V)

V_OUT

0

2

4

(V)

PMOS GateNMOS Gate

0

0.5

1

1.5x 10

−3

(A)

Time 50(µs)/Div

Inductor Current

Figure 4.15: Line transient simulation response

97

2.8

3

3.2

3.4

3.6

(A)

V_IN

1.76

1.78

1.8

1.82

1.84

(V)

V_OUT

0

1

2

3

4

(V)

PMOS GateNMOS Gate

Figure 4.16: Line transient simulation response (3.6 V - 3.0 V)

98

2.8

3

3.2

3.4

3.6

(A)

V_IN

1.76

1.78

1.8

1.82

1.84

(V)

V_OUT

0

1

2

3

4

(V)

Time 5(µs)/Div

PMOS GateNMOS Gate

Figure 4.17: Line transient simulation response (3.0 V - 3.6 V)

99

Chapter 5

HARDWARE IMPLEMENTATION AND MEASURED RESULTS

This chapter discusses the hardware implementation of the design. Layout of the

complete design and the individual block layouts are shown. Layout techniques

to prevent mismatch in the layout are also discussed. Finally, test results of the

fabricated chip are presented.

5.1 Layout

The complete layout with the pad frame is shown in Fig. 5.1. The padframe

consists of 40 pads, i.e., 40 I/O pins, and occupies an area of 1.5 mm × 1.5 mm.

Every pad occupies a length of 212.3 µm on each side, which leaves an active area

for layout of 1075.4 µm × 1075.4 µm.

The area consumed by this design, excluding the bias resistors, is 395.16 µm

× 89.34 µm. 21 out of 40 pins are used. The major area is consumed by on-chip

bias resistors of 50 MΩ and 2 MΩ for the comparator. The resistor model used

is ”oprrpresx”. The 50 MΩ resistor consumed an area of 565.44 µm × 302.04 µm

, whereas, the 2 MΩ resistor occupied an area of 150.04 µm × 97.24 µm. The

inductor and capacitor are off-chip and are not laid out.

The layouts of major blocks are shown in this section and the area con-

sumed by all the major blocks is summarized in Table 5.1.

For laying out the power switches, a gate strapping technique is used.

The gate of the MOSFET, which is siliced poly has high resistivity. This high

100

Figure 5.1: Complete layout of the DCM buck converter with pad frame

resistivity will cause delays in the clock signals that drive the gate. To reduce the

gate resistance, a contact from poly to metal1 is made wherever there is a gate.

Metal1, which has far lower resistivity than poly, will reduce the parasitic gate

resistance.

The layout of the comparator is shown in Fig. 5.2. Transistors that make

up a current mirror are laid out using the common centroid technique for better

101

Table 5.1: Area occupied by major blocks

Module Size (W/L) m

Chip 1500µ/1500µ

PMOS Power Switch 10.3µ/12.8µ

NMOS Power Switch 10.6 µ/7.6µ

Comparator 64.17µ/24.17µ

Zero current detector 38.2µ/16.95µ

Current-starved delay line-1 48.56µ/18.39µ

Current-starved delay line-2 33.46µ/18.39µ

D flip-flop 37.8µ/11.6µ

Non-overlapping clock generator 65.75µ/28.34µ

SR latch 12µ/11.6µ

Gate driver for PMOS switch 5.3µ/15.59µ

Gate driver for NMOS switch 5.3µ/13.09µ

matching. The common centroid technique is also implemented in the transistors

of the differential amplifier input stage.

The layout of the zero current detector is shown in Fig. 5.3. Transistors

that are in a current mirror configuration are laid out using common centroid

technique for better matching.

The layout of two current-starved delay line blocks are shown in Figs. 5.4

and 5.5. These circuits also consist of transistors in a current mirror configuration

which require the common centroid technique.

102

Figure 5.2: Comparator layout

Figure 5.3: Zero Current Detector layout

The other major blocks that are used are the D flip-flop, shown in Fig. 5.6,

and the non-overlapping clock generator, shown in Fig. 5.7.

Figure 5.4: Current-starved delay line

103

Figure 5.5: Current-starved delay line

Figure 5.6: D flip-flop

5.2 Measured Results at Minimum Load

The design was fabricated in IBM 7HV 0.18 µm process through MOSIS.

A QFP-44 package was used. For testing purpose, the chip was soldered on a

TQFP44 socket. The inductor and the capacitor used were surface mount and

were soldered on the TQFP44 socket adapter. The TQFP44 was further soldered

to the perforated board. All other external components, such as bias resistors and

bypass capacitors, had wire leads and were soldered on the perforated board.

The minimum load for this design is 20 µA. As such, a load resistor of

90 kΩ is used. A 100 µH inductor and a 3.3 nF capacitor are used. The input

104

Figure 5.7: Non-overlapping clock generator

voltage (VIN) of 3.6 V, the reset signal (EXT RESET), enable signal for the pad

buffers, and reference voltage (VREF ) are all taken from DC power sources.

The output voltage waveform can be observed from the Fig. 5.8. The

measured voltage ripple on the output is 61.45 mV. The switching frequency

measured was 100 kHz.

Figure 5.8: Output voltage ripple

105

The non-ovelapping clock outputs can be observed in Fig. 5.9. They show

large amounts of ringing. The measured pulse width is approximately 94 ns.

Figure 5.9: Output signals of non-overlapping clock generators

The output waveform along with both gate driver inputs are shown in

the Fig. 5.10. When phi1 is high, the PMOS switch is turned ON and when

NMOSGD,IN is low, the NMOS switch is turned ON. After the NMOS switch

turns OFF, that is, when both switches are turned OFF, the output voltage

slowly decreases. The PMOS is again turned ON when the output voltage goes

below the reference voltage.

The output of the ZCD and the inverted gate signal of the NMOS switch

are shown in Fig. 5.11. It can be observed from the figure that as soon as the

output of the ZCD goes high, the inverted gate signal of the NMOS switch goes

high, turning OFF the NMOS switch.

The combined current consumption of the input power stage and gate

drivers is measured as 13.02 µA. This leads to a power consumption of 46.87 µW

106

Figure 5.10: Output voltage and input signals to the gate drivers

Figure 5.11: Output of ZCD and the NMOS gate driver input

by the power stage and gate drivers. The current consumed by the controller

was measured as 1.97 µA. The calculated power consumption of the controller is

7.09 µW. As the controller, power stage and gate drivers all are powered from

107

the input battery source, the power consumption of the controller also should

be included in the input power. Therfore, the measured total input power is

53.964 µW.

The measured steady-state output voltage is 1.805 V and the output cur-

rent calculated is 19.95 µA. The calculated output power is 35.93 µW.

As the input power and output power are known, the efficiency can be

calculated from

η ≡ POUT

PIN

(5.1)

The calculated efficiency is 66.5 %.

The simulated results are compared with the measured results in Table 5.2.

Table 5.2: Comparision of simulated and measured results for a 20 µA load

Parameter Simulated Measured

< VOUT > 1.814 V 1.82 V

< IOUT > 20.1 µA 19.93 µA

∆VOUT 33.36 mV 61.45 mV

Switching frequency (fSW ) 184.8 kHz 110 kHz

Efficiency (η) 84 % 66.5 %

From Table. 5.2, it can be observed that the measured output voltage

ripple (∆VOUT ) is significantly higher than the desired. This is due to higher

delay observed in the hardware current-starved delay circuit. The current-starved

delay circuit determines the constant on-time of the converter. Because of higher

constant on-time, the PMOS switch is turned ON for a longer time and the output

is charged to a higher value.

108

The higher ∆VOUT also causes the switching frequency (fSW ) to decrease.

The decrease in fSW can be observed in Table. 5.2.

It can also be observed that the efficiency of the converter in the hardware

is degraded compared to the simulations. This is due to the increased current

consumption in the controller. Because the delays observed in the hardware were

very high, the bias current in the current-starved delay circuit, zero-cross detector

and comparator were increased to reduce the delay. This is one of the major

causes for efficiency degradation in the hardware.

Another possible cause for efficiency degradation could be negative induc-

tor current. The imprecise turn-OFF voltage of the ZCD could cause the NMOS

switch to turn-OFF late, thereby permitting negative inductor current. The in-

ductor current waveform could not be observed from the hardware to prove this

case. However, higher current was consumed by the power stage and the gate

drivers in the hardware compared to simulations.

5.3 Peak Efficiency

Table 5.3 shows results at another load current (96 µA), where peak effi-

ciency of 71.2 % is achieved. This is the maximum current where the converter is

operated in proper DCM even though it has high output voltage ripple.

At higher currents (¿ 96 µA), the comparator delay increases and will cause

the converter to operate in Boundary Conduction Mode (BCM) for a few cycles

before going back to DCM. It is similar to operating in burst mode. When the

converter is in BCM for a few cycles, the output gets charged higher than necessary

and then slowly discharges to VREF . The result is significantly high ∆VOUT .

109

Table 5.3: Comparision of simulated and measured results at 96 µA load

Parameter Simulated Measured

< VOUT > 1.806 V 1.807 V

< IOUT > 96 µA 95.9 µA

∆VOUT 30.3 mV 58 mV

Switching frequency (fSW ) 1 MHz 450 kHz

Efficiency (η) 85.1 % 71.2 %

110

Chapter 6

CONCLUSIONS

6.1 Summary

Ultra-low power applications such as wearable devices have limitations on

their size and weight. The limitations on size and weight of these devices will also

impose limitations on the battery size used in these devices, which thereby limits

the amount of battery energy available. Therefore, it is important to efficiently

use the limited battery energy, in order to achieve longer battery life.

This work proposes a highly efficient DC-DC buck converter design for

ultra-low power applications in order to fulfill the above requirement. The design

was implemented in IBM 180-nm technology and fabricated with a QFP44 package

through MOSIS. The maximum efficiency achieved by the design in simulation is

87.1 % at 200 µA load with an input voltage of 3.6 V.

A ripple-based constant on-time control with DCM-enabling circuitry is

implemented. The DC-DC converter is operated in Discontinuous Conduction

Mode (DCM), which is favorable for light loads.

The control-loop is implemented with simple circuitry, which greatly re-

duces the area occupied by the converter. Various power reduction techniques

are implemented to reduce the power consumption in the controller. The ZCD

and comparator are the only analog circuits which are power hungry. However,

in this work, the ZCD is enabled by the NMOS gate signal so that it is turned

ON only for a short duration in each cycle, thereby reducing the average power

consumption in the ZCD. The PMOS to NMOS ratio in the digital circuits, which

111

is generally 2:1, is violated in order to reduce the power consumption in the digital

blocks. Instead, the PMOS to NMOS ratio in all digital circuits is maintained at

1:1.

6.2 Comparison with the State-of-the-Art

The simulation results of the proposed ultra-low power design are compared

with the other state-of-the art low power designs. The comparison results are

summarized in Table 6.1.

The proposed work used IBM 180 nm technology for the design. The design

converts an input DC voltage range of 2.8 V - 4.2 V to an output DC voltage of

1.8 V, where 2.8 V - 4.2 V is the voltage range of a li-ion battery. The load

currents, which are ultra-low, range from 20 µA - 200 µA. The output voltage

ripple (∆VOUT ) varies with the load current between 26.2 mV - 33.5 mV, where

the highest ripple is at 20 µA. ∆VOUT is < 2% across the entire load range. The

switching frequency ranges from 185 kHz at 20 µA to 1.7 MHz at 200 µA. The

inductor and capacitor values of 100 µH and 3.3 nF are fixed across the input

voltage range and output load current range. Another parameter that varies

across the load range is the power efficiency of the converter with 85.1 % at 20 µA

that increases as the load current increases. The peak efficiency of the converter

is 87.1 % at 200 µA load. The entire load range is operated with discontinuous

conduction mode.

In Table 6.1, three other works [2], [3], [13] on ultra-low power applications

are reported for comparison with the proposed work. Observing the output voltage

and load current range of these works, it can be inferred that these three works are

not limited only to ultra-low power. The applications of these other three works

range from ultra-low power to low-power, whereas the proposed work is only for

ultra-low power applications.

112

∆VOUT is not reported in any of these works. The work in [2] reports a

switching frequency (fSW ) of 32 Hz for ultra-low loads which is significantly low

compared to fSW of 185 kHz in the proposed design. Not only the proposed design

is faster than [2], but also has better regulation. fSW at 200 µA is not reported

in [2],but the maximum fSW is 2.5 MHz at 20 mA load current. The work in [3]

reports a fSW of 1.65 MHz at 200 µA load which similar to fSW of 1.7 MHz in

the proposed design.

The 100 µH inductor value chosen for this design is significantly large

compared to the inductor values of other works, but the size (3 mm x 3 mm x

1.3 mm) of the 100 µH inductor is within the typical range of inductor sizes that

are used. However, the inductor size is not reported by other works. The output

capacitor of this work is significantly lower compared to the other works.

Comparing the power efficiency, the proposed work reports a higher effi-

ciency across the entire load range compared to [2] and [3]. However, the work in

[13] achieves higher efficiency compared to the proposed design.

Because of their wide load current range, the works in [2] and [3] switched

the operating modes accordingly with the load current, whereas the proposed

design operates only in DCM across the entire load range.

6.3 Issues

The delay in the comparator is the major issue of this design, which limits

the maximum load current to 200 µA. As the load current increases, because of the

delay in the comparator, the converter begins to operate in Boundary Conduction

Mode (BCM) for a few cycles and then again to operate in DCM. Because of the

successive turn-ON of the PMOS switch in BCM, the output capacitor overcharges

and increases the output voltage ripple.

113

Table 6.1: Comparison with state-of-the-art topologies.

Work [2] 2017 [3] 2016 [13] This work

Technology 130 nm 180 nm NA 180 nm

VIN 2.2–3.3 V 0.55–1.0 V 2.0–5.5 V 2.8–4.2 V

VOUT 1.7 V 0.35–0.5 V 1.3–5.0 V 1.8 V

IOUT,MIN 10 µA 100 nA 15 µA 20 µA

IOUT,MAX 20 mA 20 mA 50 mA 200 µA

∆VOUT NR NR NR 26.2–33.5 mV

fSW,MIN 32 Hz NR NR 185 kHz

fSW,MAX 2.5 MHz 1.65 MHz NR 1.7 MHz

L/C 3 µH/3 µF 4.7 µH/ NR 10 µH/22 µF 100 µH/3.3 nF

η, 20 µA 75 % 79 % 90 % 85.1 %

η, 200 µA 81 % 84 % 92 % 87.1 %

Mode Triple-mode Dual mode NR DCM only

The delay in the comparator also limits the switching frequency of the

converter.

There were several issues with the hardware. First, the ringing at the

switching node caused latch-up, which damaged the PMOS switch. Also, probing

the switching node would increase the ringing because of the added capacitance

from the probe. Second, the measured output voltage ripple in the hardware is

higher because the delay of the constant on-time block in the hardware is higher.

114

Also, the delays of the comparator and the zero-current detector in the hardware

did not match with the simulation results.

6.4 Future Work

Ultra-low power applications like wearable devices always demand longer

battery life. With the increasing functionality, the power consumption also in-

creases, thereby degrading the battery run-time. This demands the DC-DC con-

verters to be more efficient.

Some ideas for the future work in ultra-low power DC-DC buck converter

are:

• The comparator used in this design can be replaced by an adaptively-biased

comparator, where the bias current for the comparator increases with the

load current. If implemented, the issue of operating in BCM for multiple

cycles will be solved, which helps in achieving a wider load current range.

• Adaptive on-time technique helps in achieving lower output voltage ripple.

• The negative inductor current in the buck converter, which will introduce

power losses, can be avoided if the ZCD is fast and precise. A ZCD imple-

mented with a comparator may not be precise because of the mismatch in

the hardware. In the future, we need to make the reference voltage attached

to the ZCD an external input, so that, through calibration, the mismatch

can be canceled.

put together...

• The reference voltage for the zero-current detector should be made pro-

grammable so that the zero-current detector turns OFF the NMOS switch

at the right time.

115

• Techniques to reduce ringing at the switching node could be explored. Talk

about the resistor and the effect of it...

116

REFERENCES

REFERENCES

[1] Z. Bi and W. Xia, “Modeling and simulation of dual-mode dc/dc buck con-

verter,” in 2010 Second International Conference on Computer Modeling and

Simulation. IEEE, 2010, pp. 371–375.

[2] Y. J. Park, J. H. Park, H. J. Kim, H. Ryu, S. Kim, Y. Pu, K. C. Hwang,

Y. Yang, M. Lee, and K. Y. Lee, “A design of a 92.4% efficiency triple mode

control dc 8211;dc buck converter with low power retention mode and adap-

tive zero current detector for iot/wearable applications,” IEEE Transactions

on Power Electronics, pp. 6946–6960, 2017.

[3] P. H. Chen, C. S. Wu, and K. C. Lin, “A 50 nw-to-10 mw output power

tri-mode digital buck converter with self-tracking zero current detection for

photovoltaic energy harvesting,” IEEE Journal of Solid-State Circuits, pp.

523–532, Feb 2016.

[4] R.J.Baker, CMOS Circuit Design, Layout, and Simulation, 3rd ed. Wiley-

IEEE Press, 2010.

[5] W.-H. K. Hylas Y. H. Lam and D. Ma, “Loop gain analysis and development

of high-speed high-accuracy current sensors for switching converters,” in Cir-

cuits and Systems, 2004. ISCAS ’04. Proceedings of the 2004 International

Symposium. IEEE, 2004, pp. 828–831.

118

[6] Wearable technology and IoT wearable devices. [Online]. Available:

1.)http://www.businessinsider.com/wearable-technology-iot-devices-2016-8,

Accessed:July5th,2017

[7] Advances in sensor technology, integration and power manage-

ment are enabling the future of wearable technology. [Online].

Available: 1.)https://www.dialog-semiconductor.com/sites/default/files/

enabling-the-future-of-wearable-technology.pdf.Accessed:January26,2017

[8] K.-H. Chen, Power Management Techniques for Integrated Circuit Design,

1st ed. John Wiley Sons, Incorporated, 2016.

[9] A. Paidimarri and A. P. Chandrakasan, “10.8 a buck converter with 240pw

quiescent power, 922017 IEEE International Solid-State Circuits Conference

(ISSCC). IEEE, Feb 2017, pp. 192–193.

[10] J. C. Tsai, T. Y. Huang, W. W. Lai, and K. H. Chen, “Dual modulation

technique for high efficiency in high-switching buck converters over a wide

load range,” IEEE Transactions on Circuits and Systems I: Regular Papers,

pp. 1671–1680, 2011.

[11] Achieving High Efficiency under Micro-Watt Loads with Switching Buck

DC-DC Converters. [Online]. Available: 1.)https://pdfs.semanticscholar.

org/bff3/afc2b58fed5f03b4902a6dcae3cd98741d85.pdf.Accessed:June6,2017

[12] Power Management Reference Design for a Wearable Device with

Wireless Charging Using the bq51003 and bq25120. [Online]. Available:

1.)http://www.ti.com/lit/ug/slua748a/slua748a.pdf,Accessed:July5th,2017

119

[13] TPS6273x Programmable Output Voltage Ultra-Low Power Buck Converter

With Up to 50 mA / 200 mA Output Current. [Online]. Available:

1.)http://www.ti.com/lit/ds/symlink/tps62736.pdf,Accessed:July5th,2017

120