+ All Categories
Home > Documents > high efficiency linear power amplifiers

high efficiency linear power amplifiers

Date post: 11-Feb-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
189
HIGH EFFICIENCY LINEAR POWER AMPLIFIERS: ANALYSIS, LINEARIZATION AND IMPLEMENTATION A DISSERTATION SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Mehdi F. Soltan June 2004
Transcript

HIGH EFFICIENCY LINEAR POWER AMPLIFIERS: ANALYSIS, LINEARIZATION AND

IMPLEMENTATION

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Mehdi F. Soltan June 2004

ii

© Copyright by Mehdi F. Soltan 2004 All Rights Reserved

iii

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in

scope and quality, as a dissertation for the degree of Doctor of Philosophy.

Donald C. Cox (Principal Adviser)

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in

scope and quality, as a dissertation for the degree of Doctor of Philosophy.

Bruce A. Wooley

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in

scope and quality, as a dissertation for the degree of Doctor of Philosophy.

Thomas H. Lee

Approved for the University Committee on Graduate Studies.

iv

Abstract Power amplifiers are key components in wireless transceivers. Their function is to

amplify the signal and generate the required RF power that allows transmission of the

signal over the appropriate range. Linear power amplifiers are needed in many modern

wireless systems that deploy non-constant envelope modulations. Examples of such

modulations are seen in CDMA and OFDM systems.

A common disadvantage of linear power amplifiers compared to their nonlinear

counterparts is their substantially lower power efficiency. Since the power amplifier is

the component that consumes most of the power in a transmitter, this lower power

efficiency directly translates into lower talk time in portable systems. Therefore,

improving efficiency of linear power amplifiers is a major objective. Finding fast and

systematic nonlinearity analysis methods and tools as well as linearization techniques that

promise higher efficiency are among the major challenges along this way. An additional

challenging requirement in portable systems is the need for smaller and lower cost power

amplifier devices. This work addresses these challenges and provides improved methods,

techniques and devices.

v

Acknowledgments I am indebted to so many people who have helped me in various stages of my life and

throughout my education. Completion of this work at Stanford would not have been

possible without all that support. It is beyond my capabilities to really articulate an

acknowledgment that can covey my most sincere and deepest gratitude to all those

people, whether explicitly mentioned here or not.

First I thank Professor Donald C. Cox, my primary academic advisor. Over the past

several years he has not only provided me with the best of the vision, advice and

encouragement that a graduate student might ever receive from his Ph.D. advisor, but

also has most kindly supported and advised me in almost every aspect of my life. I will

remain deeply indebted to him.

It is my pleasure to particularly thank Professors Bruce Wooley and Thomas Lee for all

their help during my stay at Stanford and for carefully reading a draft of this entire

dissertation. I also thank Professor David Leeson for kindly joining my orals committee

and Professor Abbas El Gamal who served as the chair of my orals committee.

I like to thank all my former teachers and professors as well as all my friends from whom

I have learned in the past.

I am most grateful to my family that has never let me feel that I am alone in my

endeavors. Without persistent support and encouragements from my parents and my wife

this work would never have finished. I also extend my thanks to my dear daughter,

Nazanin, who I hope can soon read this.

vi

Contents Abstract .............................................................................................................................. iv Acknowledgments............................................................................................................... v List of Tables....................................................................................................................viii List of Figures .................................................................................................................... ix Chapter 1: Introduction ....................................................................................................... 1

1.1 Basics of power amplifiers and classes ..................................................................... 3 1.2 Basics of nonlinearity................................................................................................ 9 1.3 Basics of linearization ............................................................................................. 12 1.4 Basics of signal and modulation dynamic range and effects on efficiency ............ 17

Chapter 2: More on PA nonlinearity and its effects on modulated signals....................... 19

2.1 Quasistatic versus dynamic nonlinearity................................................................. 21 2.2 A very fast method for simulating quasistatic nonlinearity effects......................... 26 2.3 Effects of the shape of the gain compression on ACPR ......................................... 38 2.4 A search for optimum gain compression shapes..................................................... 47

Chapter 3: PA linearization from an efficiency and complexity point of view ................ 49

3.1 More on EER and newer ER techniques................................................................. 50 3.1.1 Pulse deletion modulation ................................................................................ 53

3.1.1.1 In band vs. out of band linearity trade off ................................................. 56 3.1.1.2 Signal and modulation dynamic range effects .......................................... 60 3.1.1.3 Jitter in the signal constellations ............................................................... 65 3.1.1.4 Implementation methods for classes C and D amplifiers.......................... 73

3.1.1.4.1 Feedback methods .............................................................................. 79 3.1.1.4.2 Other methods .................................................................................... 82 3.1.1.4.3 Synchronous vs. asynchronous PDM................................................. 85

3.1.2 Pulse width modulation.................................................................................... 88 3.1.3 Pulse width and amplitude modulation (PW&AM)......................................... 92

3.1.3.1 EER with envelope feedback through bias control ................................. 100 3.1.3.2 Bias controlled envelope feedback without envelope elimination.......... 102 3.1.3.3 Envelope detector dynamic range and integrator offset effects .............. 108 3.1.3.4 Bandwidth effects and dynamic nonlinearity.......................................... 116 3.1.3.5 Potential for bias-controlled ER with envelope pre-distortion................ 123

3.2 Linearization by design of shape of gain compression curve ............................... 124 3.2.1 A note on self-biasing .................................................................................... 125 3.2.2 Cascade of self-biased nonlinear stages ......................................................... 128

Chapter 4: PA module constructions and matching methods ......................................... 139

4.1 Traditional output matching network topologies and implementations................ 141 4.2 A new approach to eliminate off chip discrete passives ....................................... 144

4.2.1 HP matching networks ................................................................................... 146 4.2.2 LPHP and LPLP networks ............................................................................. 150

vii

Chapter 5: Conclusion..................................................................................................... 159 5.1 Summary ............................................................................................................... 159 5.2 Future research directions ..................................................................................... 161

5.2.1 Integrated dynamic load adjustment .............................................................. 161 5.2.2 Sharing output devices for multiple bands..................................................... 163 5.2.3 CMOS PAs & breakdown.............................................................................. 164 5.2.4 Device layout.................................................................................................. 165 5.2.5 Dynamic interaction of thermal and electrical behaviors............................... 167

Bibliography.................................................................................................................... 169

viii

List of Tables Table 1-1: Published articles indexed under “Power Amplifiers” ...................................... 1 Table 1-2 Class A PA.......................................................................................................... 4 Table 1-3: Class B and Class C PAs ................................................................................... 5 Table 1-4: Class D and Class E PAs ................................................................................... 6 Table 1-5: Class F PA ......................................................................................................... 7 Table 1-6: Summary of a few basic PA measures for classes A-F ..................................... 7 Table 1-7: Linearization by back-off ................................................................................ 12 Table 1-8: Predistortion and feedforward linearization .................................................... 13 Table 1-9: Feedback (polar and cartesian) linearization ................................................... 14 Table 1-10: LINC method ................................................................................................. 15 Table 1-11: Envelope elimination and restoration ............................................................ 16

ix

List of Figures Figure 1-1: Sample output versus input power curve ......................................................... 3 Figure 1-2: Sample power added efficiency for different input powers ............................. 4 Figure 1-3: P1dB and IP3.................................................................................................... 9 Figure 1-4: AM to PM conversion .................................................................................... 10 Figure 1-5: Nonlinear capacitors are a source of AM-PM................................................ 10 Figure 1-6: Spectral regrowth ........................................................................................... 11 Figure 1-7: Probability distribution of signal average power in CDMA systems [1] ....... 18 Figure 2-1: Contributions of m and m& to spectrum of s& ............................................... 22 Figure 2-2: Compression curve, magnitude ...................................................................... 28 Figure 2-3: Compression curve, phase.............................................................................. 28 Figure 2-4: Power gain...................................................................................................... 29 Figure 2-5: Sample CDMA signal magnitude................................................................... 29 Figure 2-6: Undistorted input signal trajectory ................................................................. 30 Figure 2-7: Distorted output signal trajectory ................................................................... 30 Figure 2-8: Spectral regrowth ........................................................................................... 31 Figure 2-9: Amplitude probability distribution, CCDF .................................................... 31 Figure 2-10: ACPR for various power outputs ................................................................. 32 Figure 2-11: Comparison of measured and simulated ACPRs ......................................... 34 Figure 2-12: Transfer characteristic with 5% fluctuation ................................................. 36 Figure 2-13: Effect of magnitude error on ACPR............................................................. 36 Figure 2-14: Linear amplifier with sharp saturation ......................................................... 38 Figure 2-15: Gain and ACPR in a sharply saturated amplifier ......................................... 39 Figure 2-16: Effect of gain compression on ACPR .......................................................... 40 Figure 2-17: Effect of gain expansion on ACPR .............................................................. 41 Figure 2-18: ACPR of an expansive/compressive gain profile......................................... 41 Figure 2-19: ACPR of a compressive/expansive gain profile........................................... 42 Figure 2-20: Effect of gain magnitude profile on ACPR.................................................. 43 Figure 2-21: Effect of gain magnitude profile on ACPR.................................................. 43 Figure 2-22: Effect of gain magnitude profile on ACPR.................................................. 44 Figure 2-23: Effect of AM-PM on ACPR......................................................................... 45 Figure 2-24: ACPR of different phase profiles ................................................................. 46 Figure 2-25: ACPR in presence of large AM-PM............................................................. 46 Figure 3-1: Envelope elimination and restoration............................................................. 50 Figure 3-2: EER using a delta modulated Class-D switching power supply [32]............. 51 Figure 3-3: Envelope and phase modulated signals generated separately by DSP [1]. .... 51 Figure 3-4: Full power with no pulses dropped ................................................................ 53 Figure 3-5: 2 out of 8 pulses dropped ............................................................................... 53 Figure 3-6: 3 out of 8 pulses dropped ............................................................................... 53 Figure 3-7: Sample output filter frequency response ........................................................ 56 Figure 3-8: Case a ............................................................................................................. 57 Figure 3-9: Case b ............................................................................................................. 57 Figure 3-10: Spectrum of the signals for cases a and b..................................................... 57

x

Figure 3-11: PDM output spectrum .................................................................................. 58 Figure 3-12: Output spectrum with better filtering ........................................................... 59 Figure 3-13: Impact pulse dropping ratio on spectrum..................................................... 60 Figure 3-14: Impact pulse dropping ratio on spectrum..................................................... 61 Figure 3-15: Spur levels with PDM based power level adjustment.................................. 62 Figure 3-16: AM modulation with PDM........................................................................... 63 Figure 3-17: AM modulation with PDM........................................................................... 63 Figure 3-18: AM modulation with PDM........................................................................... 63 Figure 3-19: AM modulation with PDM........................................................................... 63 Figure 3-20: AM modulation with PDM........................................................................... 63 Figure 3-21: AM modulation with PDM........................................................................... 63 Figure 3-22: Digital modulation with PDM...................................................................... 65 Figure 3-23: Digital modulation with PDM...................................................................... 65 Figure 3-24: Filtered output tracks input........................................................................... 65 Figure 3-25: Digital modulation with PDM...................................................................... 65 Figure 3-26: I-Q trajectory ................................................................................................ 66 Figure 3-27: I-Q trajectory ................................................................................................ 66 Figure 3-28: Signal constellation ...................................................................................... 66 Figure 3-29: Signal constellation ...................................................................................... 66 Figure 3-30: PDM induced jitter ....................................................................................... 67 Figure 3-31: Time/phase jitter for the modulation ............................................................ 68 Figure 3-32: Output constellation for OSR=64................................................................. 69 Figure 3-33: Output constellation for OSR=16................................................................. 69 Figure 3-34: PDM jitter for OSR=64 ................................................................................ 70 Figure 3-35: PDM jitter for OSR=16 ................................................................................ 70 Figure 3-36: Phase jitter for various carrier/modulation frequencies ............................... 71 Figure 3-37: Modulation timing curve .............................................................................. 71 Figure 3-38: Modulation timing curve .............................................................................. 71 Figure 3-39: Simulation vs. fast estimates on phase jitter................................................. 72 Figure 3-40: PDM on Class C amplifier ........................................................................... 73 Figure 3-41: PDM on Class D amplifier ........................................................................... 75 Figure 3-42: Waveforms at full power.............................................................................. 75 Figure 3-43: Waveforms at 6dB back off.......................................................................... 75 Figure 3-44: Normalized efficiency for PDM and Class A .............................................. 77 Figure 3-45: Asynchronous envelope delta modulation ................................................... 79 Figure 3-46: Synchronous envelope delta modulation...................................................... 80 Figure 3-47: Lookup table based synchronous PDM........................................................ 82 Figure 3-48: Synchronous PDM using envelope sigma-delta........................................... 83 Figure 3-49: Magnitude of spurious in asynchronous PDM............................................. 86 Figure 3-50: Frequency of spurious in asynchronous PDM ............................................. 86 Figure 3-51: Constellation jitter in asynchronous PDM ................................................... 87 Figure 3-52: Modulating Class F amplifier with PWM .................................................... 88 Figure 3-53: Magnitude vs. duty cycle in PWM............................................................... 89 Figure 3-54: Waveforms in a PWM Class F amplifier ..................................................... 89 Figure 3-55: PWM only creates narrow band harmonics ................................................. 90 Figure 3-56: PW&AM on a Class C amplifier.................................................................. 92

xi

Figure 3-57: PW&AM efficiency for various maximum conduction angles.................... 94 Figure 3-58: 30% conduction angle profile....................................................................... 96 Figure 3-59: Efficiency for the 30% conduction angle profile ......................................... 96 Figure 3-60: 50% conduction angle profile....................................................................... 97 Figure 3-61: Efficiency for the 50% conduction angle profile ......................................... 97 Figure 3-62: 100% conduction angle profile..................................................................... 98 Figure 3-63: Efficiency for the 100% conduction angle profile ....................................... 98 Figure 3-64: Output magnitude as a nonlinear function of base bias................................ 99 Figure 3-65: EER with envelope feedback through bias control .................................... 100 Figure 3-66: Output power controlled by supply voltage ............................................... 102 Figure 3-67: Output power controlled by base bias voltage ........................................... 103 Figure 3-68 AM-PM in supply and base bias controlled amplifiers ............................... 103 Figure 3-69: Effect of limiter on the drain efficiency of the output stage ...................... 106 Figure 3-70: Envelope feedback PW&AM without limiter ............................................ 107 Figure 3-71: Waveform for an envelope detector with dead-zone.................................. 108 Figure 3-72: Effect of dead-zone on the spectrum of detected envelope........................ 109 Figure 3-73: Mismatch in envelope detectors ................................................................. 110 Figure 3-74: Error from envelope detector mismatch..................................................... 111 Figure 3-75: Spectrum of error form mismatch .............................................................. 111 Figure 3-76: Dead-zone effect in a closed loop system .................................................. 112 Figure 3-77: Dead-zone effect in a closed loop system .................................................. 113 Figure 3-78: Dynamic range of a PW&AM system with envelope feedback................. 114 Figure 3-79: Closed loop system tracks slower modulations.......................................... 116 Figure 3-80: Closed loop system fails to track fast modulations .................................... 117 Figure 3-81: Frequency response of a closed loop PW&AM system............................. 118 Figure 3-82: Loop gain effect on envelope stability ....................................................... 119 Figure 3-83: In-channel linearity and stability................................................................ 120 Figure 3-84: Out-of-channel linearity and stability......................................................... 121 Figure 3-85: Biasing an amplifier stage .......................................................................... 125 Figure 3-86: Effect of load impedance on self-biased DC current ................................. 126 Figure 3-87: Effect of load impedance on gain and power output.................................. 127 Figure 3-88: Self-biasing for various quiescent currents ................................................ 128 Figure 3-89: Shape of gain .............................................................................................. 129 Figure 3-90: Phase variation ........................................................................................... 130 Figure 3-91: Impact of the second stage on current of the driver stage .......................... 130 Figure 3-92: Power added efficiency of the two stage amplifier .................................... 131 Figure 3-93: Power added efficiency of the two stage amplifier .................................... 132 Figure 3-94: Power added efficiency of the two stage amplifier .................................... 132 Figure 3-95: Linearity of the two stage amplifier ........................................................... 133 Figure 3-96: Measured gain ............................................................................................ 135 Figure 3-97: Measured ACPR......................................................................................... 135 Figure 3-98: Measured power added efficiency.............................................................. 136 Figure 4-1: A two stage low pass matching network ...................................................... 141 Figure 4-2: A traditional PA module............................................................................... 142 Figure 4-3: Low pass matching....................................................................................... 147 Figure 4-4: High pass matching ...................................................................................... 147

xii

Figure 4-5: PA module in a lead frame package with no discrete passives .................... 148 Figure 4-6: Measure power added efficiency.................................................................. 149 Figure 4-7: Measured linearity........................................................................................ 149 Figure 4-8: Low pass-low pass matching........................................................................ 150 Figure 4-9: Low pass-low pass matching with one discrete capacitor............................ 151 Figure 4-10: Low pass-high pass matching..................................................................... 152 Figure 4-11: Implementation of LPHP matching............................................................ 152 Figure 4-12: High pass-high pass matching.................................................................... 153 Figure 4-13: Implementation of HPHP matching ........................................................... 153 Figure 4-14: A grossly power imbalanced PA................................................................ 155 Figure 4-15: A PA with some power imbalance ............................................................. 155 Figure 5-1: Power added efficiency with load adjustment.............................................. 162 Figure 5-2: A compact layout for a power transistor [86]............................................... 166

1

Chapter 1: Introduction

Power amplifiers are used in many different applications including the majority of

wireless and radio communications equipment, wireless and cable TV broadcast systems,

cable and other wired transmission systems, optical driver amplifiers, audio systems and

Radars. Depending on the application, frequencies range from audio frequencies to

millimeter wave frequencies. Amplifier power ranges from a few milliwatts to several

megawatts are used for different applications. Over many years the knowledge of related

technologies and the design of these amplifiers has been developed. Table 1-1 represents

interesting sample statistics on the most directly related publications indexed by IEEE

alone.

Publication Year Before 1980 1980 to 1990 1990 to 1998 1998 to 2/2004

Number of

Articles

195 397 1682 1854

Table 1-1: Published articles indexed under “Power Amplifiers”

in IEEE and affiliated publications

While some knowledge has been developed that is common to a wide variety of

applications, increasing specialization has led to the creation of a large number of

techniques and technologies that are useful only in very specific power amplifiers.

In this dissertation our main focus is on applications for cellular phone handsets and

portable wireless devices, with an emphasis on linear power amplification. CDMA,

TDMA and OFDM based handsets (or other portable systems) are among these related

applications. The carrier frequencies in these applications range from several hundred

megahertz to a few gigahertz, while peak powers of 100 milliwatts to a few watts are

most common. The most well known common objectives in amplifier designs for these

applications are: cost and size reduction; efficiency and talk time improvement; and

2

meeting linearity, gain, stability, robustness, reliability and other related requirements.

Many of the specific techniques, technologies and tradeoffs discussed in this dissertation

are directly useful in other applications as well. Therefore, while our main focus is linear

amplifiers, we briefly stretch further into some nonlinear applications such as GSM

handsets to highlight the more general applicability of the techniques.

Based on this focus, this dissertation is organized in five chapters. In the remainder of

this chapter, we will very briefly review some related background material. The objective

is to allow those new to the field to quickly gain a good overview of power amplifiers

and related technologies. One of the primary needs of researchers and designers working

on linear amplifiers is to quickly assess, analyze, design and optimize amplifier

nonlinearities based on their specific objectives. Chapter 2 focuses on a methodology and

a tool developed to address this need. In Chapter 3 we will present four non-traditional

categories of linearization methods. The common objective among those schemes is to

enhance the talk time of the portable wireless equipment using such amplifier systems.

Each of these is analyzed from the efficiency behavior and complexity viewpoints. In

cases where nontraditional effects or artifacts come to play analysis is provided to

enhance our understanding. Chapter 4 covers implementation issues related to power

amplifier modules. That chapter covers some newer topologies for output impedance

transformation networks that enable manufacturing of smaller and lower cost modules

followed by a brief discussion on some of the most important, but infrequently cited,

additional concerns in power amplifier design. We conclude in Chapter 5 by

summarizing the main points discussed in the first four chapters, followed by an

identification of main areas for future research on power amplifiers for handsets and

portable wireless systems.

3

1.1 Basics of power amplifiers and classes RF and microwave power amplifiers, PA’s, are devices that amplify the input RF or

microwave signals and deliver much higher power at the output. Power gain, defined as

the ratio of the output RF power to input RF power, is therefore a primary performance

measure. The power amplifier can also be considered a device that converts DC power

provided from the supply into RF power at the output. One of the most critical

performance measures of a PA is the efficiency of this conversion process. Drain or

collector efficiency is defined as the ratio of RF output power to the DC power provided

from the supply. Power added efficiency is another measure that is the ratio of the output

RF power, less input RF power, to the total power into the device (DC+RF).

total

inout

PPPPAE −== η Eq. 1.1-1

The power amplifier dissipates a major portion of the total power in many portable

systems such as phone handsets. Therefore, the efficiency of that amplifier is the most

important factor affecting the talk or operation time. Figure 1-1 and Figure 1-2 show

sample behaviors of a typical power amplifier. A rather good general review of these

basic concepts on RF power amplifiers can be found in [1].

Output Compression curveOutput Compression Curve

Pin

Pout

Output Compression curveOutput Compression Curve

Pin

Pout

Figure 1-1: Sample output versus input power curve

4

Power Added Efficiency

Pin

PAE

Power Added Efficiency

Pin

PAE

Figure 1-2: Sample power added efficiency for different input powers

In order to allow better analysis and understanding of their behaviors, power amplifiers

are grouped into different classes of operation. The classification is primarily based on

voltage and current waveforms, and hence based on operating conditions and topologies

[1, 2]. Tables 1-2 to 1-5 list the best known classes of PAs with some basic pictures of

waveforms and topologies.

Class Information: Class A Conduction angle is 360deg (100%) Circuit Topology

Waveforms

VDD

IDC

Vd

id

Ploss

t

t

t

VDD

IDC

Vd

id

Ploss

t

t

t

VDD

IDC

Vd

id

Ploss

t

t

t

Ploss is the power dissipated in the device

Table 1-2 Class A PA

RL

VDD

Vin

Vd

5

Class Information: Class B Conduction angle is 180deg (50%)

Circuit Topology

Waveforms

Class Information:

Class C Conduction angle is less than 180deg (less than 50%) Circuit Topology

Efficiency and Power Capability

as functions of Conduction angle

Waveforms are similar to those shown for Class B

Table 1-3: Class B and Class C PAs

ABC AB

Power Capabilityη 100%

78.5%

50%

0.125 0.134

68.1%

AB C A B 0

100% 100%50% 50% 0 0

Conduction angle Conduction angle

RL

VDD

Vin

Vd

VDD

IDC

Vd

id

Ploss

RL

VDD

Vin

Vd

6

Class Information: Class D No time overlap between transistor’s drain source voltage and drain current Circuit Topology

Waveforms

Class Information:

Class E No time overlap between transistor’s drain source voltage and drain current First derivative of VDS is zero at the moment device turns on Circuit Topology

Waveforms

Table 1-4: Class D and Class E PAs

RL

VDD

Vin

Vd

C2 L

C1

Vd1

id2

2VDD

id1 RL VDD

Vin

-Vin

id1

id2

Vd1

7

Class Information: Class F No time overlap between transistor’s drain source voltage and drain current Circuit Topology

Waveforms

Table 1-5: Class F PA

Class

Efficiency η %

Normalized RF Power

)2/(2max,

RVddPo

Normalized vd max

VddVd max

Normalized id max

DC

d

Ii max

Power capability

max,max,

max,

dd

o

iVP

A 50 1 2 2 0.125 B 78.5 1 2 3.14 (=π) 0.125

C 86 (θ=71°)

1 2 3.9 0.11

D 100 1.624 (=16/π2)

2 1.57 (=π/2)

0.318 (=1/π)

E 100 1.154 (=4/(1+π2/4))

3.6 2.86 0.098

F 100 1.624 (=16/π2)

2 3.14 (=π) 0.159 (=1/2π)

Table 1-6: Summary of a few basic PA measures for classes A-F

id1

2VDD

Vd

LRVDDπ

8

RL Vin

vd

λ/4 @ ω0

VDD

8

Aside from efficiency, there are other measures that can be used to benchmark power

amplifiers. Normalized RF power is a measure that can provide a rough estimate of the

load impedance required for delivering certain power in an ideal case. Normalized

maximum theoretical voltage and currents for different classes can help to determine the

required device voltage breakdown and current handling capabilities. Power capability is

another measure that can provide a basis to compare different classes from the required

device size point of view. Table 1-6 summarizes definitions and optimum theoretical

values for these measures for Classes A to F.

More on PA classes can be found in [3, 4, 5 and 6]. One has to keep in mind though that,

due to many known and unknown practical factors, optimum theoretical values can not be

achieved in reality. Indeed, even determining the class of operation in many practical

amplifiers is not easy and sometimes becomes impossible. In many cases amplifiers can

be modeled as one class at certain power output levels, and as another class in a different

power output range. However, this classification approach is very useful in providing

insights and starting points for design tradeoffs in power amplifiers. In addition to the

theoretical class-based design, some of the best known measurement based device

modeling and design approaches are based on load-pull and source-pull measurements

[2, 7].

9

1.2 Basics of nonlinearity Power amplifiers are nonlinear systems because the large signal behavior of the

semiconductor devices is nonlinear. Power output saturation, as shown in Figures 1-1 and

1-3, is one nonlinear behavior that occurs in every amplifier.

Among the measures for nonlinearity are P1dB and IP3. Figure 1-3 illustrates those

measures on a typical amplifier power out versus power in curve. IP3 is the extrapolated

intercept of the desired linear output with the 3rd order intermodulation with two-tone

excitation. P1dB is where the output drops 1dB below the desired linear output. These

measures are independent of input signal modulation schemes.

1dB

IP1dB IIP3

3rd order inter-mod term

Desired linear output

1dB

IP1dB IIP3

3rd order inter-mod term

Desired linear output

Figure 1-3: P1dB and IP3

In addition to the amplitude (AM) nonlinearity, power amplifiers usually also exhibit

some amplitude to phase conversion behavior (AM-PM) as shown in Figure 1-4 for a

sinusoidal input. AM-(to)-PM refers to the creation of phase modulation (PM) observed

at the output of a nonlinear amplifier when amplifying amplitude modulated (AM)

10

signals. AM-PM is often the result of voltage dependent capacitors (e.g. junction

capacitors). Figure 1-5 illustrates this mechanism.

Pha

se (

Deg

)

AM-PM curve

Pin (dBm)

Pha

se (

Deg

)

AM-PM curveAM-PM curve

Pin (dBm) Figure 1-4: AM to PM conversion

Iin=Asin(ωt) R C

Vout(t)

Iin=Asin(ωt) R C

Vout(t)

)(tan))]([(tan)(

)]([)]([

))(()()(

1

1

ωϕωϕ

ϕω

CRtBRCt

tBCCtVoutCC

ttCostBtVout

≈=

≈⇒=

+=

C = Average capacitor value at the fist harmonic

ϕ =Average phase at the fist harmonic

Figure 1-5: Nonlinear capacitors are a source of AM-PM

Amplitude dependent amplifier nonlinearity interacting with modulated carriers with

variable envelopes causes an effect called spectral regrowth. A modulated carrier can be

considered as a large set of tones squeezed into a particular frequency band around the

carrier. Amplifier nonlinearity creates intermodulation products among these tones. The

11

net effect is that additional unwanted spectral components are created. The creation of

these additional components is called spectral regrowth. Figure 1-6 shows a typical

spectrum of a nonlinear amplifier output that has suffered spectral regrowth.

Po

PHiPLo

Po

PHiPLo

Figure 1-6: Spectral regrowth

o

LO

o

HI

PP

LO

PP

HI

ACPR

ACPR

=

=

Eq. 1.2-1

Several measures of system nonlinearity are defined based on the amount of regrowth

related artifacts that are created for particular modulations. Adjacent channel power ratio

(ACPR) is the most often used parameter for characterizing CDMA amplifiers. ACPR is

defined as the ratio of unwanted power created in a specific channel within a specified

bandwidth at a specified frequency offset from the main carrier, divided by the power in

the desired modulation bandwidth. ACPR is illustrated in Figure 1-6. For example, in

PCS CDMA (IS-95) [34], ACPR is defined as the ratio of the power in a 30kHz

bandwidth at 1.25MHz offset from the center frequency, divided by the power in the

main modulation bandwidth. Unlike P1dB and IP3, ACPR and other spectral regrowth

based nonlinearity measures are heavily dependent on the modulating signal. As a result,

using specific signal characteristics and statistics is very important in simulating or

measuring ACPR in nonlinear systems.

12

1.3 Basics of linearization In order to improve the linearity of amplifiers, many linearization schemes have been

developed for different applications over many years. Tables 1-7 to 1-11 provide a listing

of the most well known linearization methods. There is extensive literature available on

most of these traditional linearization schemes and their more modern variations. Some

additional general information can be found in [8 and 9].

Back-off

Transmit less power to achieve higher linearity by avoiding the saturation nonlinearity.

Highlights: 1. Back-off is the simplest linearization

technique. 2. The required back off depends on

modulation type, AM-to-AM and AM-to-PM distortion levels.

3. Efficiency is sacrificed to achieve desired linearity

4. Low cost and no added complexity

Table 1-7: Linearization by back-off

13

Predistortion Compensate for amplifier gain and phase variation over power range by applying an inverse nonlinearity to the input signal prior to entering the amplifier.

ModulatorCorrection

GainGain

phasephase

Gain

phase

Baseband IN

RF out

Overall Response

ModulatorModulatorCorrectionCorrection

GainGain

phasephase

Gain

phase

Baseband IN

RF out

Overall Response

Highlights: 1. Look up table is perfectly suited for open loop predistortion. 2. Process and temperature variations as well as long term drifts usually mandate the

need for adaptation. 3. Adaptive predistortion systems can be complex. References: [9,10, 11 and 12]

Feedforward Correct the nonlinearity by subtracting an estimate of nonlinearity induced artifacts from the output.

PA+

Delay Σ

-

RFinPAPA

+Delay ΣΣ

-

RFin

PA

Auxiliary Amp

PAPA

Auxiliary Amp

-Delay

Attenuator

Error+ΣΣ

-Delay

Attenuator

Error

Highlights:

1. Requires a linear auxiliary PA to amplify the error signal. 2. Requires matching of the delays and attenuation/gain behaviors in order to function

correctly. 3. Efficiency is low due to losses in the output delay line, output power combiner and

the auxiliary amplifier. 4. Sensitivity to parameter variations over temperature and time usually mandates the

need for complex adaptive schemes. References: [9,13,14 and 15]

Table 1-8: Predistortion and feedforward linearization

14

Polar Feedback Using feedback from the output, detect magnitude and phase information and use it to correct the signal fed into the amplifier by comparing it with that of the desired input signal.

Σ

Filter

Envelope Detector

r + -ΣΣ

Filter

Envelope Detector

r + -

PAVGAVCOPD Filterθ

PAPAVGAVCOPD Filterθ

Highlights: 1. Two feedback loops: One for phase and one for magnitude. Stability is a concern. 2. Bandwidth of the envelope loop should be reasonably larger than the envelope

maximum frequency. 3. When AM-to-PM is manageable, phase feedback may be eliminated. 4. High performance envelope detection is required. Reference: [16] Cartesian Feedback Feedback a sample of the output, demodulate and detect in-phase and quadrature components and use them to correct the signals fed into the amplifier by comparing them with those of the desired signal.

-

-

Phase Adjustment

LO

PA

Σ

LPF

LPFQuadrature Modulator

Σ

I

Q

+

+

-

-

Phase Adjustment

LO

PAPA

ΣΣ

LPF

LPFQuadrature Modulator

ΣΣ

I

Q

+

+

Quadrature DemodulatorQuadrature Demodulator

Highlights: 1. Two feedback loops on I&Q. Stability is a concern. 2. I& Q phases have to be aligned 3. Loop bandwidth has to be reasonably higher than I & Q channel bandwidths. References: [17-21]

Table 1-9: Feedback (polar and cartesian) linearization

15

LINC (Linear amplification using nonlinear components.) Modulate the amplitude by combining two out-phased constant envelope amplified signals whose phase difference contains the amplitude information.

S1(t)

Signal Separator

S2(t)

S(t)

S1(t)

Signal Separator

S2(t)

S(t)

PA

PA

Σ+

+

K*S(t)

PAPA

PAPA

ΣΣ+

+

K*S(t)

)2/)((cos)(

))()(cos()())()(cos()(

))(cos()()(

12

1

Atbt

ttwtAtsttwtAts

twttbts

−=

−+=++=

+=

ααϕαϕ

ϕS1(t)

S(t)

S2(t)

α(t)

-α(t)

S1(t)

S(t)

S2(t)

α(t)

-α(t)

Highlights: 1. When fully matched hybrid combiners used, efficiency is proportional to the average

output power due to the loss in the hybrid. 2. Lossless combining on the other hand stresses the individual amplifiers reducing their

efficiency, and degrades the overall linearity. 3. Not efficient for modulations with large peak to average ratio 4. Difficult to implement the signal separator with analog components. References: [22-28]

Table 1-10: LINC method

16

Envelope Elimination and Restoration (EER) (Also known as Kahn technique) Amplify a constant envelope signal containing the phase information using an efficient nonlinear amplifier and superimpose the envelope variation by modulating the supply voltage of the amplifier.

PALimiter

RFin

RFout

PAPALimiter

RFin

RFout

Envelope Detector

Supply modulator

Envelope Detector

Supply modulator

Highlights: 1. Requires a very low loss, power efficient and fast supply modulator, such as class S

supply modulators or sigma delta DC-DC converters 2. Supply variations result in AM-to-PM distortion. 3. Delay mismatch between the envelope and RF phase paths can result in distortion. 4. Envelope feedback can be used to improve the linearity 5. A variation of this technique that uses amplifiers with additional back-off to

guarantee linearity is called envelope tracking References: [29-33]

Table 1-11: Envelope elimination and restoration

17

1.4 Basics of signal and modulation dynamic range and effects on efficiency As briefly noted in the previous section, linearity measures such as ACPR are dependent

on modulation characteristics such as dynamic range, statistical distribution and

modulation bandwidth and therefore vary for different modulation schemes. Measures

such as modulation peak to average and peak to minimum are used to convey further

insight into modulation dynamic range and how the signal amplitude fluctuates.

However, as will be seen in the next chapter, for precise measurement or estimation of

ACPR, signals with the exact modulation and having the correct cumulative distribution

function (CDF) of the amplitudes must be used. As a result, acquiring precise knowledge

of signal characteristics is a prerequisite to any amplifier optimization. Among the

systems requiring linear amplifiers are CDMA and OFDM based systems. A good review

of the signal characteristics in these kinds of systems can be found in [34 and 35]. In the

following chapters the effects of amplifier nonlinearity on modulated signals in general

and CDMA signals in particular will be discussed in further detail. Some related

information specific to OFDM systems can be found in [36].

In most modern wireless systems, power control schemes are used to reduce transmit

power when full power is not necessary in order to minimize unnecessary interference

and thus maximize the system capacity. For example, in CDMA systems, the average

power of the transmitted signals is varied over a range as large as 70dB [1 and 34].

The statistical distribution of the average power output varies as a function of where the

handset is functioning. This is illustrated in Figure 1-7 for urban and suburban areas. One

has to distinguish between modulation dynamic range, which can be defined as the peak

to average or peak to minimum signal power for a particular average power output, and

the dynamic range of the average power output such as the ones shown in Figure 1-7.

18

Figure 1-7: Probability distribution of signal average power in CDMA systems [1]

One practical implication of such large dynamic ranges is the need to characterize and

improve the efficiency of power amplifiers over wide power output ranges. This renders

single power efficiency data useless and therefore requires efficiency roll-off curves to

objectively compare various amplifiers and linearization schemes. This is done in

Chapter 3. More information on signal dynamic range can be found in [37]. Among the

main techniques developed to improve efficiency roll-off behavior in nonlinear amplifiers

are Doherty amplifiers [38] and stage bypassing [39]. In Doherty amplifiers an auxiliary

amplifier is used next to the main amplifier. At high power outputs the auxiliary amplifier

functions and causes a reduction in the load impedance seen by the main amplifier. This

allows the main amplifier to see higher load impedance at lower power outputs while

experiencing the correct impedance at higher power outputs. As a result Doherty

amplifiers show higher efficiencies at lower powers. Stage bypassing is based on the

concept that low power outputs can be delivered by driver stages. Therefore the output

stage devices in a PA can be turned off and bypassed to save power at lower power

outputs.

19

Chapter 2: More on PA nonlinearity and its effects on modulated signals

So far we have discussed the general effects caused by amplifier nonlinearity such as

spectral regrowth, harmonics, spurs, oscillatory behaviors, etc. We have also

discussed briefly the measures and metrics used for quantifying nonlinearity effects.

We have noted measures like IP3 and P1dB as general measures for the amplifiers.

We have also looked at modulation dependent measures such as ACPR, that are used

to assure compliance with standards with respect to unwanted signal emission that

occurs because of spectral regrowth.

One of the issues that designers face in the design of linear amplifiers for specific

digital modulation standards is how to simulate and predict the behavior of their

designs when amplifying modulated signals. We usually have to make sure that the

design robustly maintains linearity, e.g. the ACPR margin holds over a range of

frequencies, power outputs and temperatures. Also, in order to estimate ACPR

measures with acceptable variance, they have to be averaged over a few hundred

transmitted symbols, e.g. for CDMA this means an averaging time between 100µs to

1ms. The cell phone carrier frequencies are in the GHz range, e.g. 850 MHz for cell

band CDMA or 1.7- 1.9GHz for PCS CDMA. This means that if we want to estimate

ACPR using an FFT and running a regular transient simulation on our amplifier, it

would require us to obtain 4 to 40 million data points, depending on the desired

precision, for just one specific combination of frequency, temperature, power output,

circuit topology and bias level. With current computers and transient simulation

technologies, obtaining even one ACPR point for a specific case takes a prohibitively

long time. One feasible approach to avoid the problem is to use theoretical estimates

[40 and 41]. Unfortunately these theoretical methods can’t be easily adapted to

address various amplifier characteristics and hence do not provide a reliable and easy

to use tool for designers and researchers. One alternative solution is to use simulation

engines that support algorithms based on envelope transient simulation in conjunction

with harmonic balance simulation [42]. Those algorithms take advantage of the fact

20

that in most RF systems, the envelope frequency, or the modulation frequency, is

much smaller than the carrier frequency. Based on this assumption they have

combined harmonic balance and transient simulation algorithms to provide a solution

for these cases. While the approach provides useful results, the simulation time per

point ranges from several minutes to a few hours depending on the computer used,

circuit complexity, and how close to saturation the amplifier is driven. This limits the

usefulness of the approach only to providing reassurance for a few specific

conditions. It is almost impossible to use those tools for heavy exploration of the

trade offs or optimization in the design. In order to serve the unaddressed need for an

easy to use tool that could effectively predict nonlinearity effects of the linear power

amplifiers, an approach and a program were developed that allow very fast estimation

of effects such as ACPR and constellation jitter generated by amplifier nonlinearity.

The tool has enabled the development of an extremely interesting new linearization

technique that is discussed in chapter 3. This chapter will discuss the basics of the

approach, results from the program and results of a search to find optimum power

amplifier compression curve characteristics for specific modulations.

21

2.1 Quasistatic versus dynamic nonlinearity A modulated carrier can be represented as

( ) ( ))2cos()( ttftats c ϕπ +⋅= Eq. 2.1-1

or

( ) )(Re 2 tfj cetmts π⋅= Eq. 2.1-2

where

( )tjetatm ϕ⋅= )()( Eq. 2.1-3

is the complex modulation of a band limited signal with frequency content only at much

lower frequencies than the carrier frequency cf .

In a general case a power amplifier, in response to a signal ( )ts , will generate an output

signal ( )tu , that can be a nonlinear dynamic function of ( )ts with the dynamics

represented by derivatives of ( )ts .

),,,,( L&&&&&& ssssfu = Eq. 2.1-4

where

).2Re( 2 tfjc

cemfjms ππ+= && Eq. 2.1-5,

).)2(4Re( 22 tfjcc

cemfjmfjms πππ ++= &&&&& Eq. 2.1-6

etc.

The functional relationship, (.)f , depends on the amplifier circuit and operating

conditions.

22

Let us first consider a simple case where the modulation, )(tm , is a single tone signal at

the frequency of envf much smaller than cf .

)2cos()( tftm envπ= Eq. 2.1-7

)2sin(2)( tfftm envenv ππ=& Eq. 2.1-8

))2cos(2)2sin(2Re(

)2Re(2

2

tfjenvcenvenv

tfjc

c

c

etffjtff

emfjmsπ

π

πππππ

⋅+=

⋅+= && Eq. 2.1-9

cf2π

envf2π

Magnitude of Spectrum

Frequencyenvc ff +envc ff −

cf2π

envf2π

Magnitude of Spectrum

Frequencyenvc ff +envc ff −

component

componentm&

mcf2π

envf2π

Magnitude of Spectrum

Frequencyenvc ff +envc ff −

cf2π

envf2π

Magnitude of Spectrum

Frequencyenvc ff +envc ff −

component

componentm&

m

Figure 2-1: Contributions of m and m& to spectrum of s&

As observed in the spectrum of s& in Figure 2-1, the m& related component is suppressed

by a factor of envc ff / compared to the component due to m . When ( )tm is well behaved

and band limited, which is the case for practical modulation signals having maximum

frequency of envf , the same kind of suppression occurs on all spectral components of

m& compared to their equivalent in m . Since all the spectral content in the modulation

signal is at frequencies lower than envf , envc ff / serves as a lower bound for the

suppression factors of the spectral components in m& .

23

For example, this minimum suppression is more than 67dB in a PCS CDMA system. The

same suppression mechanism is in effect with the higher order time based derivatives of

the signal. Therefore with good precision the output can be estimated as

),(),)2Re(,2Re,(Re 2222 mfcgemfjemfjemfu tfjc

tfjc

tfj ccc =⋅⋅⋅⋅⋅≅ Lπππ ππ

Eq. 2.1-10

This conveys an approximate or quasistatic relationship between u and m .

Usually when the amplified signal is a narrowband modulated carrier as in the current

cellular phones and most other wireless systems, the output signal can also be represented

as

),,()(Re 2)( tmfceetu tnfjtjn

cn Ω+⋅⋅= ∑ πθα Eq. 2.1-11

where tnfjtjn

cn eet πθα 2)()( ⋅⋅ represents the signal around the thn harmonic and

),,( tmfcΩ represents other unwanted signal contributions including potential out of band

spurs, oscillations, subharmonics and noise. All of the in band components are included

in the tfjtj ceet πθα 2)(1

1)( ⋅⋅ term. Here in band refers to the spectral components whose

frequencies are close enough to the carrier that they fall in the same or immediately

adjacent application frequency band as that of the carrier, e.g. in CDMA systems spectral

components that are less than 50MHz away from the carrier for this purpose can be

considered in band. The remainder of the terms represent out of band contributions. The

phrase “in band nonlinearity” can be used to describe nonlinear behavior, specifications,

measures and signals that cause, contribute to, or otherwise refer to the tfjtj ceet πθα 2)(

11)( ⋅⋅ term. We can refer to all the other effects caused by nonlinearity as

“out of band nonlinearity.”

Thus, all the in band effects are governed by a relationship expressed as:

),()( 1)(

111 mfgetu c

tj ≈⋅= θα Eq. 2.1-12

24

The important message is that the dynamic effects of the circuit that shape the carrier

waveform or its harmonic content do not affect the modulation in a dynamic manner but

contribute to the shaping of a static relationship between input and output complex

modulation waveforms. The other important message is that the dynamic effects that

govern the behavior of the envelope or modulation do not affect the harmonic content

except through the magnitude and phase of the modulation. In summary the dynamics of

the envelope and carrier may be treated as decoupled.

In the above, which we will refer to as “quasistatic nonlinearity”, we have made two

basic assumptions:

1. The carrier frequency, fc , is much higher than maximum frequency of the

modulation, fenv.

2. The nonlinear signal path in the system is approximately the same for all the

spectral components. That means there is not a major decoupling or re-coupling

between the modulation and carrier signals when they travel through the

amplifier; they travel together.

The above two conditions together are sufficient but not always necessary to validate the

quasistatic nonlinear relationship between input and output of an amplifier.

An envelope detector with an envelope bandwidth much higher than the signal’s

maximum envelope frequency is a good example where we might still see a quasistatic

relationship between input and output even when the second condition is not met.

However when the second condition is not met, it can be a warning that we must be

careful about making the quasistatic assumption unless we can rule out the importance of

the modulation’s dynamics. We will revisit this matter later when discussing the EER

techniques. In EER systems the modulation and carrier signals are separated and

recombined and therefore they do not travel together. There we will provide examples of

systems in which the quasistatic assumption is not valid even though the condition 1 is

25

met. In general if the system’s bandwidth is narrow compared to the modulation, fenv, the

time dispersive behavior will violate condition 2.

A slightly different but essentially equivalent view of bandpass nonlinearities can be

found in [43]. There have been attempts in system level modeling of amplifiers with

dynamic nonlinearities, where the quasistatic assumption is not valid. The majority of

work has been focused on modeling of memory effects [44] or modeling and simulation

of weak wideband nonlinearities based on Volterra series [45]. Usefulness of such

approaches is usually limited to very specific cases. Fortunately, in most traditional single

carrier power amplifier circuits, where no elegant linearization techniques are used that

may involve narrow band feedback or dynamic adjustment based on narrow band

sampling of the signal or its envelope, the quasistatic assumption is valid. That covers all

of the handset PAs, even the ones with dynamic biasing schemes like the ones discussed

in section 3.2.

26

2.2 A very fast method for simulating quasistatic nonlinearity effects As we discussed in the last section, when the quasistatic nonlinearity assumption holds

true

)).(,())(,().( )(11

)(11

1 tjcc

tj etafgtmfgetu ϕθα =≈= Eq. 2.2-1

governs all the in band nonlinearity effects. Therefore, in band modulation dependent

measures like ACPR, signal constellation and jitter can be derived from an output signal

estimated using Eq.2.2-1. An amplifier’s quasistatic characteristic has been successfully

used for estimating nonlinearity effects on analog linear modulations such as SSB [46].

The function 1g is in fact the complex nonlinear compression curve of the amplifier

circuit at one specific carrier frequency and operating condition. In simulation it can be

estimated easily by running a single sweep analysis over input amplitude in a harmonic

balance simulation. Using transient simulation it can be calculated by estimating the

magnitude and phase of the FFT of the output signal at the carrier frequency, swept over

input power. These analyses are very simple and for typical power amplifier circuits they

take between several seconds to a few minutes depending on the computer used and the

amplitude sweep range.

Based on the above simulation approaches we can easily estimate and record the behavior

of 1g as a table of magnitudes and phases of the output versus magnitudes and phases of

the input. Since in most practical situations the gain magnitude is independent of the

input signal’s phase, and only the phase shift between input and output is a function of

the magnitude, the table simply becomes a table of output signal magnitudes and the

phase shift versus the input signal magnitude.

With such data, the most natural approach for estimating the output signal attributes is

first to estimate the output signal itself using a finely interpolated version of the lookup

table and the complex modulation signal )(tm , and then subsequently calculate all the

27

imperfection measures such as ACPR. For example, for PCS CDMA the right side

adjacent ACPR, defined as the ratio of signal power in an adjacent channel with an offset

of 1.25MHz from the carrier and bandwidth of 30kHz to the signal power in the main

channel, can be calculated from:

∫+

++

−+= kHzf

kHzf

kHzMHzf

kHzMHzf

c

c

c

c

dfuF

dfuFacprh 885

885

2

1525.1

1525.1

2

.)(

.)(1 Eq. 2.2-2

where 2)(uF is the estimated power spectrum of the amplifier output.

Based on the principle described above, a MATLAB program called ACP was developed

to perform the following:

1. Generate complex modulation signals for various types of modulations, with

further emphasis on various CDMA standards.

2. Read in the lookup table for the complex gain compression curve and interpolate

it to a very finely quantized lookup table to minimize errors.

3. Estimate a time based index number for each time sample of the input signal

4. Estimate the output signal )(11

1)()( tjettu θα ⋅= time samples by reading the

magnitude and adding the phase shift for the right index number from the table.

5. Perform FFTs and calculate discrete time equivalents of equations such as Eq.

2.2-2 for estimating parameters such as adjacent ACPR.

6. Perform steps 3,4 and 5 for each output power to derive the characteristic curves

The following figures are sample results from the program. Figures 2-2 and 2-3 are the

magnitude and phase of the complex compression curve of the amplifier, i.e., the function

1g of the Eq. 2.2-1, for a particular carrier frequency.

28

Figure 2-2: Compression curve, magnitude

Figure 2-3: Compression curve, phase

29

Figure 2-4: Power gain

Figure 2-5: Sample CDMA signal magnitude

30

Figure 2-6: Undistorted input signal trajectory

Figure 2-7: Distorted output signal trajectory

31

Spectrum around carrier frequencySpectrum around carrier frequency

Figure 2-8: Spectral regrowth

Figure 2-9: Amplitude probability distribution, CCDF

32

Figure 2-10: ACPR for various power outputs

Figure 2-4 shows the calculated gain compression curve. Figure 2-5 shows a sample of

the envelope of a CDMA modulated signal. Figure 2-6 and Figure 2-7 show the signal

I&Q trajectories at the input and the output of the amplifier, respectively. The skew and

jitter introduced by the nonlinear amplifier are readily seen. Figure 2-8 shows the spectral

regrowth caused by the nonlinearity in the amplifier. Figure 2-9 demonstrates the impact

of the amplifier’s gain compression by comparing the cumulative probability distribution

function, CCDF, of the power of the output with that of the input for a CDMA modulated

signal.

Figure 2-10 contains the estimates of the left and right side ACPR for adjacent and

alternate channels of the amplifier, for various power outputs. Alternate channel ACPR is

defined the same way as the adjacent ACPR, but with a larger channel offset from the

carrier, e.g. 1.98MHz in PCS systems. There is a slight difference between right side and

left side ACPRs. This is why in Figure 2-10 we see a set of two close curves for adjacent

channels and another set of two curves for alternate channels. The primary cause for this

33

asymmetry in the spectrum is the simultaneous interaction of amplitude and phase

nonlinearities with the signal. If only amplitude or phase nonlinearities were present, the

spectrum would have been symmetric and the right and left side ACPRs should have

been the same. More on distortion sideband asymmetries can be found in [47].

Deriving the curves in Figure 2-11 using the more traditional envelope transient

simulation approach on a fast machine can easily take several hours to a few days. Using

the approach described above takes several seconds to a few minutes to derive the

compression curve of the PA using simple harmonic balance simulation and less than 10

seconds subsequently to derive all the above information including the ACPR curves.

This approach enables comprehensive exploration when designing and optimizing linear

amplifiers which wouldn’t have been possible otherwise. For example the unexpected dip

around 27dBm on the adjacent channel power ratio in Figure 2-10, was first seen using

this simulation method, and following it up permitted power amplifiers to be developed

that use this feature to enhance the efficiency. We will discuss this further in the next

section and in section 3.2.

In order to apply a simulation technique like the one above with confidence, one of the

first concerns is to seek conformance of the simulation results with measurements or

other simulation techniques. In order to compare simulation results from our ACP

program and a traditional envelope transient simulator, a reference amplifier model on

HP ADS was used to estimate ACPR at one power output. The ACPR estimates from this

program and ADS were less than 1dB apart from each other. For the reference circuit it

took ADS 12 hours to calculate that single reference data point on ACPR. Using our

approach described here it took a minute to generate the compression curves with ADS

and a few seconds for ACP to estimate ACPR versus power output, including the

reference data point.

The conformance can also be checked by comparing results from measurements on an

amplifier and the estimated ACPR curve derived from a simulation using an estimate of

the compression curve in the ACP program. That has been done many times and the

34

results usually match very well. The inherent problem with this type of verification is that

all the simulation errors and artifacts from passive and active device modeling all the way

to the modeling of the couplings in the bond wires, etc., add up and there are also errors

from measurements. These errors can mask potential inherent errors in the simulation

technique. Therefore, although the results match well, it is not a firm verification of the

program because it may very well be the effect of many errors canceling each other.

For additional verification the compression curve of an amplifier was measured in the

lab. The result was fed to the program and the estimated ACPR curves were compared

with the ones measured for the high side and low side on the same amplifier. This way all

other modeling and simulation errors are excluded and the only sources of error are in the

measurement techniques and the simulation program itself.

.

Figure 2-11: Comparison of measured and simulated ACPRs

Given the inherent high sensitivity of ACPR measurement and estimation, as noted in

Figure 2-11 the results from simulation and measurements match very well. Particularly

35

the shape of the curves is predicted very closely. Some of the known contributing sources

of error in the above experiment are:

1. Phase information is not available from simple measurements and all the phase

nonlinearity effects are neglected.

2. Power meter nonlinearity.

3. Power error in the output of the signal source.

4. Absolute error in power measurement, e.g. in the above curve due to the high

slope of the curve at powers above 27dBm just 0.2dB error in power measurement

could lead to an error of more than 1.2dB in the ACPR reading.

5. In measurement of ACPR a fluctuation of +/- 1dB is normal.

6. Different results can be obtained due to windowing effects and filter length.

7. The lookup table is usually generated by interpolating data and creating a table of

several thousand entries from a typical 30 to 60 point measurement set. This can

introduce offset caused by quantization error.

In order to get a sense of the magnitude of the errors that can be caused as noted in items

2, 3 and 7 above, a simulation was run on a linear compression curve that had a

sinusoidal fluctuation error:

xxkmy ⋅⋅⋅+= ))sin(1( Eq. 2.2-3

with 100 times m being the percentage error introduced for each power.

36

Figure 2-12: Transfer characteristic with 5% fluctuation

Figure 2-13: Effect of magnitude error on ACPR

37

Figure 2-12 shows a sample compression curve for 10 percent fluctuation or +0.8dB/ -

0.9dB. Figure 2-13 shows the effect on ACPR estimate for different error percentages

ranging from 0.1% to 10%. In summary this result suggests that a 5% magnitude or

equivalently 0.4dB power measurement error can in turn introduce substantial error in the

estimate of the ACPRs below -50dB. On the other hand it indicates that we should not

worry much about power measurement errors less than 0.1dB. Therefore in using the

simulation, attention should be paid to creating the compression curve and the input to

the program, with fine enough steps to avoid excessive quantization error. About 50 or

more points per curve is usually adequate.

In summary, confidence is obtained that this method for estimating the ACPR can

provide very useful results when combined with an adequate simulation of the

compression curve. This is confirmed in actual use and is predicted by the theory for

quasistatic nonlinear amplifiers.

38

2.3 Effects of the shape of the gain compression on ACPR In this section we explore various potential scenarios for nonlinear gain compression and

discuss the resultant trends observed in the ACPR behavior. Throughout this section we

use the same CDMA modulated input signal that was used in section 2.2.

The first case, which can be considered the most intuitive one, is the case of a sharply

clipped or saturated otherwise perfectly linear amplifier with no phase or gain distortion.

20 25 30-200

-100

0

100

200

Pout (dBm)

Out

put p

hase

(deg

rees

)

-60 -40 -20 0 20-40

-20

0

20

40

Pin (dBm)

Pou

t (dB

m)

20 25 30-70

-60

-50

-40

-30

Pout (dBm)

AC

PR

(dBc

)

0 10 20 30 4021

22

23

24

25

26

27

Pout (dBm)

Gai

n (d

B)

Adjacent Channels

Alternate Channels

Figure 2-14: Linear amplifier with sharp saturation

Figure 2-14 shows the characteristics and calculated ACPR for such a sharply clipped

amplifier. In all the following we maintain the saturation or clipping level the same at

31.6dBm output or 12volts on a 50 ohm load. Whenever the phase characteristic is not

39

shown explicitly, the phase is constant as shown in Figure 2-14. Whenever the

magnitude or gain magnitude is not shown, it is the same as shown in Figure 2-14.

Figure 2-15: Gain and ACPR in a sharply saturated amplifier

Figure 2-15 shows the same case as Figure 2-14, but with further detail. There is a knee

in the adjacent and alternate ACPR curves at 27dBm output or equivalently at a back-off

of 4.6dB with respect to full saturated power. It is only after that point that the nonlinear

effect of clipping becomes significant. We also observe that -50dBc adjacent channel

power ratio can be obtained at a back-off of 3.2dB. A second and much sharper knee is

observed in the alternate channel power ratio at about 0.3-0.4 dB back-off. The highest

slope of the adjacent channel power ratio curve occurs at back-off levels between -3.2dB

and -1.6dB and is as high as 7dB per a 1dB change in output power.

40

In a second case, the effect of gain compression prior to saturation is explored. As shown

in Figure 2-16 for the three cases of a, d and e we observe that with increasing

compression the knee moves further down to lower power outputs, i.e. to higher back-off,

and overall the ACPR increases further in the region of compression.

Figure 2-16: Effect of gain compression on ACPR

The scenario of ACPR for an expansive gain behavior is similar to that for compressive

gain behavior as seen in Figure 2-17 for the cases of a, b and c. Figure 2-18 shows what

is expected when an expansive gain behavior is followed by a gain decrease and then

saturation thereafter.

41

Figure 2-17: Effect of gain expansion on ACPR

Figure 2-18: ACPR of an expansive/compressive gain profile

42

Figure 2-19: ACPR of a compressive/expansive gain profile

Figure 2-19 shows an alternative case with a compressive gain behavior that is followed

by a gain increase and then saturation thereafter.

The interesting observations are:

1. A local maximum of the adjacent ACPR occurs at a power level where the gain

vs. power slope is highest.

2. There is a local minimum of ACPR at around 27.5dBm where the gain behavior

switches from expansive to compressive, and the second derivative of gain vs.

Pout is therefore zero.

3. When the power output is close to saturation, clipping is the dominant nonlinear

effect that determines the ACPR.

43

Figure 2-20: Effect of gain magnitude profile on ACPR

Figure 2-21: Effect of gain magnitude profile on ACPR

44

Figure 2-22: Effect of gain magnitude profile on ACPR

Figures 2-20, 2-21 and 2-22 show three other examples. Though atypical for actual

amplifiers, they further help to provide a better sense of how the shape of the gain versus

power output can change the ACPR behavior.

Among the above scenarios, the case shown in Figure 2-18 has the most important

practical impact. We observe the local minima in ACPR curve for the first time in such

an exploratory simulation. That inspires the thought that it may be possible to have an

amplifier as a cascade of two nonlinear stages that creates the expansive/compressive

behavior by tuning the conduction angle or bias of the two stages to achieve much better

ACPR at lower currents and higher efficiencies. In chapter 3, we will discuss results of

further research along this direction. In summary this approach has led to the design of

cellular and PCS handset CDMA power amplifiers with measured efficiencies of over

40% for the first time.

So far all the cases discussed have been constant phase cases. The following cases

explore the effect of AM-PM conversion on the ACPR for a constant gain magnitude

45

sharply saturated amplifier with peak power of 31.6dBm (see Figure 2-14 gain versus

Pout curve).

As intuitively expected and shown in Figure 2-23, increased AM-PM conversion also

degrades ACPR performance. Figures 2-24 and 2-25 present additional cases that show

how AM-PM conversion can also affect the ACPR, in addition to its effect on

constellation jitter.

Figure 2-23: Effect of AM-PM on ACPR

46

Figure 2-24: ACPR of different phase profiles

Figure 2-25: ACPR in presence of large AM-PM

47

2.4 A search for optimum gain compression shapes For all the cases discussed in the previous section, the ACPR curve shown in Figures 2-

14 and 2-15 is a lower bound on what can be achieved for all power outputs. This

suggests that among saturated amplifiers with a fixed maximum output power level, the

best ACPR is achieved at all power outputs by a perfectly linear amplifier with abrupt

saturation at the maximum power. Although intuitively appealing, this hypothesis is not a

trivial one. For example, can ACPR resulting from the clipping of the high side of the

signal be reduced, for example, by means of some additional nonlinear distortion on the

low side of the signal? What prevents the low side distortion from creating some adjacent

channel signal components in opposite phase to the components created by the high side

so that they cancel out and lead to lower ACPR?

In order to address the above questions a random search was performed for gain

characteristics with a saturation level of 31.6dBm with the target to find a curve that

could deliver an ACPR better than that shown in Figure 2-14 for at least one power level.

No such gain characteristic was found.

Using a different approach, a MATLAB optimization code was used to search for a

compression curve vector with the constraint of saturation at 31.6dBm. In one scenario

the objective function was defined to minimize the ACPR at one specific power level.

This was tried for various power levels, with back-off levels as low as 1.5dB. Although

various compression curves were obtained with different levels of distortion away from

the target output power level, they were all close to a straight line near the target power

level and the derived minimum ACPR was the same as shown in Figure 2-14 at that

power level. In the second scenario the objective function was set to minimize the

average of ACPR for all power levels up to 31.6dBm. The resultant vector was a constant

gain with sharp saturation at 31.6dBm and the minimum obtained was the same as that of

Figure 2-14.

48

Although these observations cannot be considered as complete proof, they are practically

suggestive that among saturated amplifiers with a fixed maximum power level, the best

ACPR is achieved at all power levels by a perfectly linear amplifier with sharp saturation

at the maximum power as shown in Figure 2-14.

49

Chapter 3: PA linearization from an efficiency and complexity point of view Introduction In this chapter various linearization techniques will be considered. The goal of this

exploration has been to achieve higher efficiency for linear amplifiers while maintaining

a low to modest complexity in the linearized system. As a result all the techniques are

based on using nonlinear amplifier classes in combination with control methods that

generate an overall acceptable in band and out of band linear performance for particular

applications.

In section 3.1 we revisit the envelope elimination and restoration (EER) technique with a

newer look at the envelope restoration (ER) function. Traditionally envelope restoration

is performed through modulation of the supply voltage. This chapter in contrast covers

only methods that re-inject the envelope modulation in the signal path, and usually at the

base of the output power devices. As a result the need for a fast tracking efficient power

supply, e.g. a fast tracking DC-DC converter [32], is eliminated and the power devices

themselves and the filtering function of their output matching networks play a main role

in restoring or modulating the envelope. Pulse deletion modulation (PDM), Pulse width

modulation (PWM) and bias controlled ER will be discussed in further detail.

In section 3.2 we will present one of the lowest cost and most efficient methods of

linearization based on a cascade of two self-biased nonlinear stages. As mentioned in the

previous chapter the possibility of such a solution was first predicted by the method

presented in chapter 2. This inspired further research that led to the final development of

this technique.

50

3.1 More on EER and newer ER techniques Envelope elimination and restoration is an old technique for using nonlinear amplifiers

for amplification of modulated non-constant envelope signals [29]. The concept is that a

nonlinear amplifier can be used to amplify a signal containing the all phase information

and the magnitude or envelope information can be superimposed on the signal after the

RF amplification or in other words through a different signal path other than the main RF

amplification path. Figure 3-1 shows this general concept.

Figure 3-1: Envelope elimination and restoration

Traditional envelope elimination and restoration often referred to as the Kahn technique

[30 and 31] is based on modulating the supply voltage of the output stage of an amplifier

with the envelope signal to create a fully modulated amplified signal at the desired output

power. There are various solutions for the efficient supply modulation. Some common

approaches used in efficient tracking DC-DC converters are based on Pulse width

modulation, delta-modulation and Class-S amplification. Two alternative

implementations are shown in Figures 3-2 and 3-3.

51

Figure 3-2: EER using a delta modulated Class-D switching power supply [32].

Figure 3-3: Envelope and phase modulated signals generated separately by DSP [1].

Key concerns and tradeoffs in such systems are:

1. Delay between envelope and phase paths cause distortion such as jitter and

ACPR, and thus has to be maintained at an acceptably low level. This is a

common problem in EER implementations and for most of the currently used

single carrier systems can be accommodated.

2. The switching power supply output has to be filtered finely enough in order not to

introduce spurs in the output RF signal. Designing those filters without requiring

bulky external components is still an open challenge.

52

3. Amplifiers also show AM-PM behavior in response to supply voltage modulation

particularly when operated in deep saturation for higher efficiency. Therefore

some additional back off is required, at the cost of losing efficiency.

4. Linear envelope detection of modulated RF signals over a wide enough dynamic

range is extremely challenging. This favors solutions like the one in Figure 3-3

over those like the one shown in Figure 3-2. However, their implementation

requires a change to the base band chips, which poses a different set of

challenges.

5. Designers have to be particularly careful not to lose the higher efficiencies they

gain from operating the PA’s closer to saturation to the power they consume in

the limiter, envelope detectors, switching power supply and filter.

6. Size and complexity of the switching power supply reflect on the overall size and

cost of power amplifier modules in different ways depending on how the budgets

for the bill of material in the systems are allocated.

With the idea of addressing concerns mentioned in items 2, 3, 6 (and to some extent, 5) a

new alternative for the implementation of the EER systems was explored [48]. The main

concept is to eliminate the supply modulator and instead inject the envelope back through

the base or gate of the amplifier output stage and let the nonlinear amplifier stage perform

the act of combining or superimposing the envelope onto the modulation. In the

remainder of section 3.1 we explore a few alternatives for implementing the envelope

restoration function with this philosophy.

53

3.1.1 Pulse deletion modulation Consider a periodic signal that consists of multiple pulses per period of the same

magnitude and evenly distributed over time. Let’s assume a period can consist of

maximum N pulses and in one sequence only m out of N pulses are transmitted. If we

look at the peak or envelope of the signal and average it over one period, this value will

be linearly dependent on the ratio of m over N.

mean= 0.22x (8/8)mean= 0.22x (8/8)

Figure 3-4: Full power with no pulses dropped

The above Figure shows the full or saturation magnitude of the envelope, m=8 and N=8.

mean= 0.22x (6/8)mean= 0.22x (6/8)

Figure 3-5: 2 out of 8 pulses dropped

mean= 0.22x (5/8)mean= 0.22x (5/8)

Figure 3-6: 3 out of 8 pulses dropped

54

Figures 3-5 and 3-6 show the signal for m/N=6/8 and m/N=5/8 respectively. The question

is how this concept could be used for linear modulation or linearization of power

amplifiers?

One scenario can be a class C stage with the above pulse trains as its current waveforms.

The filters would reconstruct signals with various envelope amplitudes that are linearly

dependent on the mean of the current waveforms and as a result the reconstructed

envelope will be a linear function of the m/N or the non-dropped pulses to the total pulses

ratio. The interesting promise of the technique is that the efficiency should not degrade

much for lower powers because there would be fewer current pulses for lower powers.

The equivalent of this scenario with voltage transition pulses has promise for class D

amplifiers.

This suggests that the method could be used for linear reconstruction of the envelope at

the output of an EER system with promise of higher efficiency at lower powers.

However, like every other promising technique, this one also has issues that need to be

addressed. As a result its usage will be limited and it cannot deliver the full promise of

perfectly linear high efficiency amplification.

The first fact that cannot be overlooked is that even in the case of a perfectly linear

reconstruction of the envelope, the main source of nonlinearity would be a potential

phase modulation dependent on the amplitude of the incoming or reconstructed signal.

For cases like the one shown in Figure 3-2 this mandates difficult specifications on the

limiter circuit. Also, in the design of the output matching network of the amplifier, it

would be critical to be sure that in interaction with the power device no variable phase

shift dependent on the amplitude of the output is created. Otherwise the AM-PM

characteristic may negate all the benefits of the linear envelope reconstruction.

Another systematic artifact of this technique is the spurious emission that is created as a

result of pulse dropping. Therefore, in using this approach one ends ups trading off in-

band linearity with out of band spurs or out of band nonlinearity. Another artifact is the

55

jitter introduced into the constellation as a result of pulse deletion. The magnitudes of all

these effects are dependent on the signal’s amplitude relative to saturation and also the

dynamic range of the signal. In the remainder of this subsection these issues will be

discussed further.

56

3.1.1.1 In band vs. out of band linearity trade off In order to gain a basic understanding of the spectrum trade offs to be expected in a PDM

based envelope restoration system we look at a very simple case. Let’s assume that the

output stage is a class C amplifier driven by a sequence of pulses so that the output

current pulses are generally similar to Figures 3-4, 3-5 and 3-6. The full power is

delivered when 8 out of 8 pulses are maintained. For now let’s consider that the output

matching performs a transimpedance and filtering function essentially equivalent to that

of a resonant tank circuit with a Q of 10 as shown in Figure 3-7.

Figure 3-7: Sample output filter frequency response

Let’s consider two cases as shown in Figures 3-8 and 3-9. Case a as shown in Figure 3-8

is the scenario of full power transmission. In case b, as shown in Figure 3-9, four out of

eight pulses are deleted. If the shape of the pulses does not change as a result of pulse

deletion function, as assumed in this scenario, the expected envelope magnitude out of

57

case b should be exactly half of that of case a, or the power should be 6dB lower for the

main carrier.

mean= 0.22x (5/8) case amean= 0.22x (8/8) case amean= 0.22x (5/8) case amean= 0.22x (8/8) case a

Figure 3-8: Case a

mean= 0.22x (5/8) case bmean= 0.22x (4/8) case bmean= 0.22x (5/8) case bmean= 0.22x (4/8) case b

Figure 3-9: Case b

Figure 3-10: Spectrum of the signals for cases a and b

Figure 3-10 shows the spectrum of the output of the filter for the two cases and in fact

confirms that carrier signal level at b is exactly 6 dB below that of case a. Of course this

exactly linear relationship only holds true when the pulse shapes are the same for the two

58

cases. In practice maintaining the pulse shape becomes challenging if a large percentage

of the pulses has to be deleted at the carrier rate.

Figure 3-11: PDM output spectrum

Figure 3-11 shows the spectrum of the filtered output normalized with respect to the

carrier frequency component. This shows that in the ideal case where pulse shapes are not

affected, the harmonic content of the amplifier output should not change for various

amplitudes. However, the pulse deletion function introduces spurious emission.

In this example, since the full signal cycle is 8 carrier cycles, spurs occur at subharmonics

at multiples of one-eighth of the carrier frequency. In general, however, these spurs need

not be subharmonics of the carrier and this varies with the implementation method. If a

fixed code size look-up table approach is adopted for implementation, spurs will always

be at subharmonics. On the other hand, when the delta-sigma approach is adopted there

will not be a fixed code size and therefore spurs can move around as a function of the

amplitude. For fixed output amplitude in a delta-sigma based implementation without

addition of dither or randomization signal, this cyclic behavior recurs and spurs will be

59

created at subharmonics. Later on in this chapter we will briefly discuss various potential

implementation schemes.

Obviously if a narrower band or higher order filter is used with higher out of band

rejection overall spurious emission will be reduced further. An example is shown in

Figure 3-12 for the same pulse sequence as case a with a different filter.

Figure 3-12: Output spectrum with better filtering

We observe that, using PDM, theoretically perfect in-band envelope or amplitude

linearity can be achieved at the cost of generating spurs out of band. In-band nonlinearity

may be seen in practical PDM based systems because the RF pulse shape may not remain

constant when the PDM is applied or due to nonlinear phase behavior.

The out of band spur level is a function of the filter shape as well as the pulse deletion

rates, average cycle length and modulation depth and shape. The frequency at which

spurs are maximized is also a function of the modulation type, filter shape and pulse

deletion rate. Needless to say, the pulse deletion rate itself has to be determined based on

the desired amplitude in order to reconstruct the right average amplitude at the output.

60

3.1.1.2 Signal and modulation dynamic range effects As mentioned in the previous subsection, an inherent characteristic of the PDM technique

is that it generates out of band spurs. In this section we will take a closer look at the

relative magnitude of the spurs generated for various output powers.

Figure 3-13: Impact pulse dropping ratio on spectrum

Figures 3-13 and 3-14 show the normalized spectrum of the output of an idealized PDM

modulated class C amplifier for various pulse deletion rates ranging from no deletion in

case a to deletion of 7 out of 8 pulses in case h. The observation suggests that the relative

spur levels in dBc go up as the pulse deletion ratio increases. The total dynamic range

explored in this case is 18dB, i.e. the output power level in case h is 18dB lower than in

case a.

61

Figure 3-14: Impact pulse dropping ratio on spectrum

As briefly mentioned before for fixed code size scenarios like the ones shown in the

previous subsection and in Figures 3-13 and 3-14 the relative harmonic content is not

changed for various power levels. All the harmonics of the carrier are exact linear

functions of the output magnitude and pulse deletion rate. The upper bound for both

harmonic content and sub harmonic spurs is primarily determined by the output filter.

Figure 3-15 shows the four closest spurs to the carrier. Due to lower rejection of the filter

at these frequencies these spurs show up as the strongest. It is more clearly seen that the

spur level in dBc goes up for lower powers produced by a higher pulse deletion ratio.

62

Figure 3-15: Spur levels with PDM based power level adjustment

In Figure 3-15 expected spur levels are shown for various nonmodulated output powers

when the PDM technique with a code size of 8 is used for power level adjustment. Back

off here means relative power level in dB with respect to the peak power. When an

envelope modulation is applied the spur levels are usually better than predicted for the

lowest power level in the modulation and worse than predicted for the highest power.

63

Figure 3-16: AM modulation with PDM

Figure 3-17: AM modulation with PDM

Figure 3-18: AM modulation with PDM Figure 3-19: AM modulation with PDM

Figure 3-20: AM modulation with PDM Figure 3-21: AM modulation with PDM

64

Figures 3-16 to 3-21 show the expected pulse sequence and the spectrum of the output

filtered by the typical filter shown in Figure 3-7 for an AM modulation with modulation

index of m=0.9. In this case the assumption is that pulses are either included in full or

deleted in full. The simulation was performed based on a sigma-delta envelope

modulation implementation technique that is discussed later in this chapter. For

generating this sequence of full pulses the sigma-delta modulator clock has to be set at

the same frequency and in synchronization with carrier. Figures 3-16, 3-18 and 3-20

show the current pulse sequence of an ideal class C amplifier and the desired envelope in

time domain. Figures 3-17, 3-19 and 3-21 show the spectrum of the filtered output. As is

seen in Figure 3-21 close in spurs, or harmonics of the envelope can be pushed down

easily and this in band linearity is achieved at the cost of out of band spurs.

The dynamic range of a signal can be looked upon in two ways. For a modulated signal

with fixed average power like the one shown in Figures 3-16 to 3-21, the peak to

minimum of 19dB can be used as an indicator of the modulation dynamic range at a fixed

average power. In a power controlled system, however, the average power of the

transmitted signal can vary substantially over a large transmit power dynamic range.

From the above discussion we conclude that the PDM method can reasonably handle

high average power signals with large modulation dynamic range but produces excessive

spurs at lower average power outputs in cases with large transmit power dynamic range.

In summary, the operable dynamic range of PDM systems is dictated by the feasible filter

rejections and tolerable spur level. These requirements severely limit the potential

applications of this approach, as most RF systems have tight spurious emission

specifications.

65

3.1.1.3 Jitter in the signal constellations

In considering potential artifacts of the pulse deletion method as described, one of the

first questions that comes to mind is whether the pulse deletion process introduces jitter,

i.e. timing or phase error, in the modulated signals and if so what are the related tradeoffs.

A simulation was performed for a hypothetical system with a carrier frequency of

250MHz and a 977 kilo-samples per second 4π QPSK modulated data stream.

Figure 3-22: Digital modulation with PDM

Figure 3-23: Digital modulation with PDM

Figure 3-24: Filtered output tracks input

Figure 3-25: Digital modulation with PDM

Figures 3-22 represents the PDM based modulated output in time domain. Figures 3-23

and 3-25 show the spectrum of the said output signal. The input to the system is a square

66

root raised cosine filtered 4π QPSK modulated data stream. Figure 3-24 shows the input

when filtered by a second square root raised cosine filter and superimposed on the output

of the system down converted to base-band and filtered by the same kind of filter. As

shown in Figure 3-24 the base-band spectrum is preserved very well.

Figure 3-26: I-Q trajectory

Figure 3-27: I-Q trajectory

Figure 3-28: Signal constellation

Figure 3-29: Signal constellation

Figures 3-26 and 3-27 show the I-Q trajectories for the input and output respectively.

Figures 3-28 and 3-29 show the signal constellation for the input and output. The input

and output match well. Therefore the proposed system theoretically is capable of

restoring the envelope with only small channel artifacts. In detecting the constellation the

assumption is that the clock and carrier frequencies are recovered perfectly and without

67

error and the timing recovery circuit successfully works to produce the lowest achievable

jitter.

Output InputOutput Input

Figure 3-30: PDM induced jitter

Figure 3-30 shows the spread of the phase error in degrees over several symbols for the

constellations shown in Figures 3-28 and 3-29. The dots correspond to the output

generated through PDM of Figure 3-29 and the solid line is that of the input.

For this particular modulation Figure 3-31 shows the relationship of the phase error

caused by timing errors with the corresponding timing error. It suggests that the mean

error of 0.11 degrees observed in this case is equivalent to the jitter caused by a 3 to

4nsec average timing jitter. Knowing the fact that in this system the carrier pulse duration

is 4nsec, it triggers another question. Is there any relationship between pulse duration and

the timing or phase jitter in such PDM systems?

68

Figure 3-31: Time/phase jitter for the modulation

Figures 3-32 and 3-34 show the constellation and jitter for a hypothetical system that has

carrier and maximum pulse deletion frequency that are 64 times the symbol rate or

oversampling ratio (OSR)=64. Figures 3-33 and 3-35 show the data for another case with

OSR=16. It appears that at lower carrier to envelope ratios or OSRs more jitter is

introduced by the pulse deletion function in PDM systems, as shown in Figure 3-36.

A potentially simpler descriptive relationship between jitter and pulse deletion rates is

suggested by the relationship between time and phase errors for the kind of modulation

discussed. A simple hypothesis is that the jitter in the pulse deletion schemes may be

introduced only by the pulse deletion function itself. If the jitter can all be attributed to a

timing error on the order of the duration of a deleted carrier pulse or a multiple of it

within same order of magnitude, then we could expect:

symbol

carrier

c

symbol

symbol

error

TT

ff

TT =≈∆

Eq. 3.1-1

69

Figure 3-32: Output constellation for OSR=64

Figure 3-33: Output constellation for OSR=16

70

Figure 3-34: PDM jitter for OSR=64

Figure 3-35: PDM jitter for OSR=16

71

Figure 3-36: Phase jitter for various carrier/modulation frequencies

Figure 3-37: Modulation timing curve

Figure 3-38: Modulation timing curve

Then Figure 3-37 that shows the relationship between timing error and phase jitter for the

4/π QPSK modulation is useful for obtaining a rough estimate of the phase jitter

resulting from pulse deletion in such a PDM system.

72

Figure 3-39 further illustrates this idea. The data points correspond to the simulated phase

errors shown in Figure 3-36 and the solid lines represent the estimated phase error using

Eq.5.1-1 and Figure 3-37. As observed, phase jitter in the PDM systems can quickly be

estimated by looking up the phase error in Figures 3-37 and 3-38 for a timing error equal

to the ratio of modulation over the carrier frequency, csymbol ff / . Such estimates are

particularly more precise for larger ratios. Equivalent phase vs. timing error data such as

the one shown in Figures 3-37 and 3-38 is usually available for standard modulation

schemes and is independent of tradeoffs of the PDM method.

Figure 3-39: Simulation vs. fast estimates on phase jitter

It is important to remember that in this section only the jitter directly resulting from the

pulse deletion process in PDM systems was discussed. Any saturation or clipping effect,

envelope nonlinearity, AM-PM conversion or delay difference between envelope and

carrier signal paths are other factors that can cause additional jitter-like artifacts to appear

in the output signal of a PDM amplifier system.

73

3.1.1.4 Implementation methods for classes C and D amplifiers An implementation of the PDM based on a Class C output stage is illustrated in Figure 3-

40. The current pulses of the output device ideally will be similar to those shown in

Figures 3-4 to 3-6.

RL

PDM PulseGenerator &Bias Control

VDD

Constant EnvelopeRFin

Envelope in

Class C PA

Driver Stages

Figure 3-40: PDM on Class C amplifier

Drain efficiency for a Class C PDM modulated stage is

averagecc

out

Total

out

IVP

PP

⋅==η Eq. 3.1-2

The highest efficiency maxη is obtained when there is no pulse dropping as in Figure 3-4.

Since

22 )( undeletedaveout NIP ∝∝ Eq. 3.1-3

74

where undeletedN is the number of undeleted pulses per second, we have

deletedundeleted

undeleted

out

out

average

ave

out

out

NNN

PP

II

PP

+==⋅=

max

max

maxmaxηη

Eq. 3.1-4

which shows that the efficiency versus power output rolls off as the square root of power

and is linearly dependent on the ratio of undeleted pulses to the total possible pulses. For

comparison, in a Class A amplifier, efficiency rolls off as a linear function of the power.

Also, the maximum efficiency can be as high as 60-70% (theoretically 100%, but at

essentially zero output) in Class C, while in Class A the peak theoretical efficiency is

50%. Thus, there is a promise of increasing the peak and average drain efficiency

substantially if the combination of PDM with a Class C output stage is used. The side

issues however are:

1. The preceding efficiency estimates are valid only as long as the output stage

current pulse shape remains constant. In reality it changes as a result of pulse

deletion. Although this is a secondary effect at low carrier frequencies, it can

become the dominant mechanism in degrading the efficiency at GHz-range carrier

frequencies.

2. The power consumed by the PDM pulse generator and bias control circuit may

offset a good portion of the power added efficiency improvement arising from

increasing the drain efficiency of the output stage. For implementing a complete

pulse deletion scheme the bias driver of this circuit has to switch the output stage

on and off at the carrier frequency. Given the big size of the output device, the

solution may prove ineffective by requiring excessive power to perform the

switching function. As a compromise we can decouple the pulse deletion rate

from the carrier frequency and eliminate the need for synchronous high speed

switching of the carrier. However, this may lead to truncated current pulses in the

amplifier which in return degrades the efficiency and linearity of the whole

system.

3. Given the current size and cost trends in GaAs handset power amplifier modules,

the addition of the PDM unit would be considered excessively expensive.

75

However if the PA integration into CMOS transceivers really becomes the trend

in the future it opens up the way for implementation of equivalent systems.

PDM can also be used to linearize Class D and Class F amplifiers. Figure 3-41 shows a

general sketch for a Class D amplifier modulated by PDM.

RL

VDDRL

VDD

M 1

M 2

PDMPulse

Generator&

BiasControl

Constant EnvelopeRFin

Envelope in

Driver Stages Figure 3-41: PDM on Class D amplifier

Figure 3-42: Waveforms at full power

Figure 3-43: Waveforms at 6dB back off

Figure 3-42 shows the current and voltage waveforms of the two transistors M1 and M2

and the output current waveform at full power. Figure 3-43 shows the same signals when

a back off of 6dB is applied through PDM.

The theoretical limits for efficiencies of Class D and Class F implementations are 100%.

Even at lower powers, i.e. higher pulse deletion rates, the current and voltage signals of

the devices theoretically do not have any time overlap. Therefore drain efficiency

remains at 100% with the assumption of pulse shapes not changing as a result of PDM. In

76

practical cases inevitably there are switching or nonlinear losses as well as passive

component losses. The losses in the passive components scale linearly with power output.

outLossLinear PP ∝. Eq. 3.1-5

The switching losses are more complex. The switching loss is the integral of the product

of the current and voltage waveforms of devices over the overlap time of the two

waveforms. If the pulse shapes of the current and voltage do not change as a result of

PDM, then the net effect of the PDM is the scaling of the current pulse magnitude. The

switching loss at each cycle then has voltage pulses that scale linearly with the current

magnitude, i.e., with square root of output power. The average switching loss is also

scaled linearly with the pulse transmission rate, which itself is proportional to square root

of output power.

outoutPulsePerLoss PIE ∝∝.. Eq. 3.1-6

outoutSecondPerPulses PIN ∝∝.. Eq. 3.1-7

outPulsePerLossSecondPerPulsesLossSwitching PENP ∝⋅∝ ..... Eq. 3.1-8

Therefore we have

outLossSwitchingLossLinearLossTotal PPPP ∝+= ... Eq. 3.1-9

and hence for drain efficiency we have

.max.

ConstPP

PPP

LossTotalout

out

Total

out ==+

== ηη Eq. 3.1-10

The above relation points to an extremely attractive potential, the promise of maintaining

the high efficiency of Class D or Class F power amplifiers over a range of power output

77

while amplifying modulated signals that require linear amplification. However, at high

carrier frequencies and high deletion rates the pulse shapes are affected by the PDM and

further degradation in the efficiency is expected for lower power outputs.

We cannot forget that even in these scenarios a back-off is required from the peak

amplifier power output to eliminate the clipping effects that were discussed in Chapter 2.

Figure 3-44: Normalized efficiency for PDM and Class A

Figure 3-44 shows the normalized efficiency roll off curves for the cases discussed in this

section. In summary, the theoretical maximum efficiency for the Class C and Class D

PDM is 100%, where for Class A it is 50%. At a back-off of 3.6dB we see from

Figure 3-44 a normalized efficiency of 0.66 for Class A, 0.82 for Class C PDM and 1 for

Classes D and F. This means a maximum theoretical linear (i.e., ACPR better than

-50dBc) drain efficiency of 33%, 82% and 100% for Class A, Class C and Class D or F,

respectively, at 3.6dB back-off.

78

At 1.8GHz, even considering a practically achievable peak efficiency of 55% for class C

and Class D, linear drain efficiencies of 45% and 55% may be achievable for Class C and

Class D/F, respectively. There will be substantial loss of power added efficiency due to

the power consumed in the PDM circuitry at these frequencies, however. This

substantially reduces the potential attractiveness of the PDM solution for GHz

frequencies.

In the remainder of section 3.1.1 we will briefly look at a number of potential approaches

for implementing the PDM pulse generator shown in Figures 3-40 and 3-41.

79

3.1.1.4.1 Feedback methods So far we have considered various effects in the PDM based amplifier systems assuming

the availability of the needed pulse sequence. However we have not yet suggested how

the PDM pulse sequence is generated. If we can have output envelope feedback in the

system, the most intuitive approach would be to sample the output envelope and compare

it with the desired envelope. When the output envelope is smaller than the desired one,

the system should switch on and allow the RF pulses to go through. When the output

magnitude exceeds the desired envelope the system needs to switch off and not allow the

RF pulses to go through. This was in fact the first methodology we used to generate and

simulate potential PDM effects.

Load

EnvelopeDetector

+

- Comparator

PAOutput Stage

Attenuation

Constant EnvelopeRFin

Envelope in

Driver Stages

Switch

Bias Circuit

Coupler

Figure 3-45: Asynchronous envelope delta modulation

Figure 3-45 shows the general concept of such an implementation. Asynchronous

envelope delta modulation based PDM can be a reasonable descriptive name for the

function performed here. An alternative to the above implementation can be the

synchronous version of the above which by controlling the switching time and

synchronizing it with the timing of the carrier pulses assures that full RF pulses go

through as shown in Figure 3-46.

80

Load

EnvelopeDetector

+

- Comparator

PAOutput Stage

Attenuation

Constant EnvelopeRFin

Envelope in

Driver Stages

Switch

Bias Circuit

Coupler

in

outCLK

FF

Figure 3-46: Synchronous envelope delta modulation

Previously we noted that in order to suppress the out of band spurs they should be pushed

further away from the carrier which means that higher switching rates are more desirable

for the pulse deletion. In the feedback implementations, the actual switching rate is

mainly determined by the bandwidth of the feedback loop which, in turn, is limited by the

bandwidth of the envelope detector and the speed of the comparator in the presence of

offsets. The feedback loop bandwidth also directly affects the jitter introduced in the

signal. The higher the bandwidth, the lower the jitter would be.

Design of linear RF signal envelope detectors with large enough dynamic range is a

challenging task by itself. One of the additional requirements for the envelope detectors

shown in Figures 3-45 and 3-46 is that their detection bandwidth has to be sufficiently

larger than the envelope frequency and the loop gain bandwidth of the system to yield

good in-band linearity and stability. On the other hand, wide bandwidth in the envelope

detector allows all harmonics and spurious signals to pass. This causes a degradation of

the envelope detector linearity. A wideband envelope detector also passes too much

81

extraneous signal fluctuation into the comparator which is translated into a time-varying

modulation-dependent offset in the comparator. This shows up in the output as extra

spurious signals that can also appear to be extra out of band nonlinearity. A bandwidth

close to the geometric mean of the envelope and carrier frequencies seems to deliver

more reasonable results based on the simulations. We will discuss issues of envelope

detectors and loop bandwidth in some further detail in sections 3.1.3.3 and 3.1.3.4.

The main problem of all these feedback based implementations is the coupling of too

many otherwise unrelated system tradeoffs, e.g., in-band and out-of-band nonlinearity,

the bandwidth of the envelope detector, the speed of the comparator and the ratio of

carrier to envelope frequency. These are enough to suggest looking for an alternative that

eliminates the feedback.

82

3.1.1.4.2 Other methods We shall not forget that the main promise of the PDM technique as we noted in section

3.1.1.1 is its promise to deliver perfect in-band linear behavior so long as the right pulse

sequence is fed into the amplifier and the pulse shapes are not affected by the pulse

deletion. Feedback is not necessary to accomplish this. Therefore if the appropriate pulse

sequence can be generated without feedback from the amplifier output, the system will be

feedback free and the limitations mentioned in the previous section go away, including

the need for an RF coupler and a linear wideband envelope detector.

One potentially effective approach could be to use a lookup table to maintain the

appropriate pulse sequence, i.e., codes with preferably fixed length, for various desired

envelope amplitudes. These code sets could be optimized to minimize out of band

spurious or in-band quantization noise. The quantization noise will be introduced because

the system will act as a digital to analog envelope converter with a finite number of levels

or code words. Such an implementation is shown in Figure 3-47.

LoadPA Output Stage

Constant EnvelopeRFin

DigitalEnvelope in

Driver Stages

Switch

Bias Circuit

in

outCLK

FF

1

2

N

12

N

ShiftRegister'sCLK

1/N

CLK

FrequencyDivider

Register

MxNMemory

LookupTable

N

Figure 3-47: Lookup table based synchronous PDM

83

Looking back at the main characteristic switching signal that is used to create the PDM

pulse sequence, we realize that it’s a sequence of pulses that at any time has its windowed

average tracking the modulation signal. So what we need to create is a sequence of pulses

that maintains this feature. Sigma-delta modulation provides exactly the same

characteristic [49]. So an alternative for the implementation shown in Figure 3-47 could

be a system that has a sigma-delta modulator replacing the lookup table as shown in

Figure 3-48.

LoadPA Output Stage

Constant EnvelopeRFin

Envelope in

Driver Stages

Switch

Bias Circuit

in

outCLK

FF

Σ1

s

Quantizer

First Order Sigma-Delta Converter

CLK

Figure 3-48: Synchronous PDM using envelope sigma-delta

Figure 3-48 shows a potential implementation of a synchronous PDM system based on

envelope sigma-delta modulation. In fact, the same model was used to simulate and

generate the modulated signals previously used in this chapter for analysis purposes.

The output signal spectrum around the carrier is the convolution of the spectra of the

sigma-delta modulator output and that of the constant envelope phase modulated carrier.

As a result all the available knowledge about close-in and spurious emission in sigma-

delta modulators can readily be used in estimating the spectrum of the output of such

PDM systems. Higher order sigma-delta modulators can also be used. As a result of their

84

noise shaping behavior, they yield better in-band linearity at the cost of more out of band

spurs.

Although in the example of Figure 3-48 the clock of the sigma-delta modulator is

assumed to be in synchronization with the carrier pulses, this synchronization it is not a

necessary condition. In fact the sampling clock of the sigma delta modulator can be a

fraction of the carrier frequency in alternative synchronous implementations, or it can be

a totally uncorrelated clock at a different frequency in asynchronous implementations.

Synchronous and asynchronous applications of sigma-delta modulators in PDM systems

have been described in [48]. Potential for synchronous application of band-pass sigma-

delta modulators in the transmitters has also been mentioned in [28 and 50]. Sigma-delta

modulators have been suggested as supply modulators in traditional EER as well [51].

The primary drawback of the asynchronous implementations is that partially deleted

pulses will be created. That affects the phase relationship between current and voltage in

the power device and, by increasing the unwanted overlap, degrades the efficiency. Since

the percentage of the distorted pulse shapes will depend on the output power, the

dependency of the efficiency on the output power will also slightly change from the

predictions made earlier in this section. Unfortunately, we still do not have an analytical

derivation of the efficiency in such cases and so it has to be simulated case by case.

Partial pulse deletion also leads to increased spurious around the harmonics of the carrier.

The primary drawback of the synchronous versions on the other hand is the extra

complexity as well as the challenges of running the sigma-delta modulator at carrier

frequency or dividing the carrier and synchronizing the switch to operate in synchronism

with both the carrier and the sampling clock.

85

3.1.1.4.3 Synchronous vs. asynchronous PDM In our analysis of spurious and jitter in PDM systems so far, the assumption has been that

the pulses are either fully deleted or fully transmitted, which is valid for true synchronous

implementations. In discussion of the implementation methods it was seen that

asynchronous approaches may also be used and they may be more desired from

complexity point of view. It is therefore important to have a sense of the level of artifacts

created in the signal and the related tradeoffs in such asynchronous implementations. In

this section we will briefly look at spurious and jitter in asynchronous PDM systems.

Simulations were performed on the same platform used for section 3.1.1.3. The platform

consists of a MATLAB flow that generates the time samples of the phase modulated

carrier and the envelope signal. The envelope of the signal is fed to a sigma-delta

modulator. The time samples of the output of the sigma delta modulator are multiplied by

the time samples of the phase modulated carrier to generate a PDM sequence. The

sequence is subsequently fed into a behavioral model of the power amplifier and its

output filter. For simulating asynchronous PDM, the clock frequency and the sampling

times were not synchronized with the timing of the carrier pulses. The modulation,

carrier frequency, and output filter were kept the same as in section 3.1.1.3. The clock

frequency of the envelop sigma-delta modulator was changed, however, to obtain results

for various oversampling ratios.

As seen in Figure 3-49 the maximum spurious level goes down as the oversampling ratio

increases, because the spurs are pushed further away from the carrier frequency as the

oversampling ratio increases. This can be seen in Figure 3-50. As a result the spurs are

filtered more effectively by the PA output filter.

86

Figure 3-49: Magnitude of spurious in asynchronous PDM

Figure 3-50: Frequency of spurious in asynchronous PDM

87

The jitter introduced in the signal constellation is another factor to be considered.

As observed in Figure 3-51 the phase error decreases for higher oversampling ratios.

Comparing the synchronous and asynchronous cases for the same carrier and envelope

frequency for an oversampling ratio of 256, Figure 3-35 and Figure 3-51 show that the

phase error in the asynchronous case is higher than the synchronous version by only

about 30%. However Figure 3-51 suggests that the phase error remains reasonably small

for all oversampling ratios. Therefore the primary criterion in selection of the rates in

asynchronous PDM systems should be the spurious emission.

Figure 3-51: Constellation jitter in asynchronous PDM

88

3.1.2 Pulse width modulation The biggest problem of the PDM method at high frequencies is the spurious emission that

cannot be filtered effectively due to lack of small low loss filters with narrow enough

bandwidth. In search for alternatives, one can think of a pulsed system that can modulate

the output by adjusting the pulse width. The goal, as before, is to maintain higher

efficiencies and to modulate the envelope.

PWMDuty CycleModulator

VDD

Constant EnvelopeRFin

Envelope in

Class F PA

Driver Stages

RLcf@4λ

Figure 3-52: Modulating Class F amplifier with PWM

One way of doing so is to modulate the switching duty cycle of the output device in a

Class F power amplifier, as indicated in Figure 3-52. The output magnitude or envelope

is theoretically governed by the following relationship.

)sin(. dutycyclesPs Maxoutoutout ⋅⋅== π Eq. 3.1-11

89

Figure 3-53: Magnitude vs. duty cycle in PWM

Since the behavior predicted by Eq. 3.1-11 is not linear, as also seen in Figure 3-53, some

additional envelope linearization may be required. Envelope feedback or envelope pre-

distortion may be applied.

Figure 3-54: Waveforms in a PWM Class F amplifier

The nice feature of no overlap between voltage and current in the device in the Class F is

maintained for all duty cycles. This is seen, for example, for a 15% duty cycle sequence

90

in Figure 3-54. Therefore, theoretically the efficiency can be maintained constant at its

peak for all power levels.

Figure 3-55: PWM only creates narrow band harmonics

As seen in Figure 3-55 for an example PWM based system with a carrier frequency of

1GHz, in contrast to PDM techniques, this approach does not inherently generate out of

band spurs. It only creates narrow band modulated harmonics.

The practical issues that limit the effective usage of this approach for GHz range

amplifiers in the order of their importance are:

1. The need for generation or modulation of the duty cycle of very narrow pulses. As

an example, for a system with only 16dB dynamic range, duty cycles as low as

5% need to be handled. In a 1.8GHz system that means pulse width modulation

needs to be done on pulses as short as 27 picoseconds. Currently no simple

economical solution is available to perform this task.

2. The need for envelope feedback or envelope pre-distortion to linearize the PWM

behavior imposes the need for additional undesired complexity.

3. In practice the efficiency of the PWM modulated Class F amplifier does not

remain high at low power since due to the short pulse cycle, the device voltage

91

waveform does not maintain a rectangular pulse shape. In fact, at high enough

power at GHz frequencies the voltage waveform is not rectangular for any pulse

width. Therefore, the efficiency advantage simply disappears.

A good practical application for this technique may be found in low to medium power HF

transmitters. PWM has also been used in audio power amplifiers [52]. An alternative use

for PWM has been suggested in DC-DC converters used for modulating voltage supply

of power amplifiers in EER systems [53]. In general, however, from the in-band

nonlinearity point of view, PWM is not a good approach when high dynamic range is

required because distortion free modulation of the duty cycle over a wide range is not

easily achieved.

92

3.1.3 Pulse width and amplitude modulation (PW&AM)

In discussion of the PWM based solutions, we noted that adjusting output power level by

tuning pulse width yields reasonable expectation of producing good efficiency over a

range of power levels, but we could not offer a simple and practically viable technique

that can perform the pulse width modulation for large enough dynamic ranges at GHz

carrier frequencies. That triggers a search for a simpler approximation of PWM.

Dynamic bias control is an old concept suggested for Class A amplifiers [54 and 55]. We

also know that in a Class C stage the conduction angle is set by the bias level of the stage.

Therefore adjusting the bias level in a Class C stage could be the simplest way to adjust

the RF pulse width. The beauty of it is that no high frequency or high speed functional

block is required other than the device itself. PWM implementation could not be any

simpler than that. Since every device has a turn on threshold, the conduction angle of the

amplifier changes by adjusting the quiescent voltage level for the same RF input level, as

shown in Figure 3-56.

Vth Vth

Constant EnvelopeRFin

Driver Stages

RL

VDD

Class C PA

Envelope Control in

Bias Control

idvg

Figure 3-56: PW&AM on a Class C amplifier

93

This is a very simple approximate implementation for PWM. However, in Class C, PWM

implementation efficiency decreases faster than in Class-F. In this particular

implementation shown in Figure 3-56, which can be called pulse width and amplitude

modulation or PW&AM, not only the conduction angle but also the current pulse

amplitude is adjusted. The drain efficiency will be

)/()/(

.)./(

.)/(

max.

2max.

maxmax.max.

max.2

max.

. dcdc

fundfund

Dissdcdc

ofundfund

Diss

o

iiii

PiiPii

PP ηη === Eq. 3.1-12

in which

outfund Pii ∝−⋅−

= )2sin2()cos1(2

max ϕϕϕπ Eq. 3.1-13

)cos(sin)cos1(

max ϕϕϕϕπ

−⋅−

= iidc Eq. 3.1-14

where maxi is the peak current value in the device’s current waveform and ϕ2 is the

conduction angle of the device. Equations 3.1-13 and 3.1-14 show the relationship

between the fundamental frequency component of the current and dc current in Class C

amplifiers with respect to the peak of the current pulses and the conduction angle. In the

amplifier shown in Figure 3-56, the peak current and conduction angles are changing as a

result of the bias control. The highest conduction angles and highest peak currents

simultaneously occur when transmitting the highest power.

In order to predict the efficiency roll off behavior, in addition to the equations 3.1-12 to

3.1-14 the relationship between peak current and conduction angle must be known. This

depends on the pulse shape at the base or gate of the transistor and the device’s nonlinear

transconductance. In order to obtain a rough estimate of the efficiency roll off behavior

for such systems, we make the assumption that the gate voltage signal is sinusoidal and

94

the device has a sharp turn on threshold. Three sample transconductance profiles of

exponential, square-law and linear have been used to allow investigating the sensitivity to

device behavior for bipolar as well as long and short channel CMOS devices.

Figure 3-57: PW&AM efficiency for various maximum conduction angles

Figure 3-57 shows the efficiency roll off curves in PW&AM operation with linear

transconductance and turn on threshold and compares it with a classical Class A amplifier

as well as a PDM modulated Class C with 30% (108 degrees) conduction angle. In this

section we use a percentage scale in presenting conduction angles, e.g. 100% refers to

360 degrees conduction. In PW&AM maximum conduction angle is used at the highest

power output. Maximum conduction angle can be set by design. This suggests that

although the PW&AM method cannot achieve the theoretically expected efficiencies of

PDM and PWM, it promises substantial efficiency improvement compared to more

traditional linear solutions such as Class A amplifiers. When maximum conduction per

cycle is maintained below 50%, it guarantees Class C operation at all power outputs and

therefore efficiency is better than Class B or Class AB. When maximum conduction per

95

cycle is higher than 50%, the amplifier’s class of operation can shift from Class C at

lower power outputs to Class B or Class AB at higher power outputs. By evaluating the

efficiency for various profiles of conduction angles versus current magnitudes, figures 3-

58 to 3-63 further illustrate the tradeoffs seen in PW&AM operation.

From the figures 3-58 to 3-63 it is noted that profiles with lower conduction angles

deliver higher efficiencies. When the peak conduction angles are smaller not only the

efficiency is higher, but also the efficiency roll-off curves are less sensitive to the device

nonlinearity shape. As seen in Figure 3-63, when the peak conduction angle becomes

100%, the efficiency roll-off curve is quite sensitive to the device behavior. For example,

at 4dB back-off, the efficiency ranges from 33% for the lower curve, to 47% for the

higher curve. Therefore, the accuracy of any simulation of the efficiency will be very

heavily dependent on the accuracy of the models for devices, as well as for matching

networks. On the other hand, when the peak conduction angle is maintained below 50%

(180 degrees) the predicted efficiency roll-off behavior is much less dependent on the

device characteristics. The peak efficiency itself, however, is heavily dependent on

device parasitics as well as on other losses in the system.

Although the lower peak conduction angles are more desirable from the efficiency and

sensitivity points of view, the lower gain and the need for negative supply voltage to bias

the device to allow low conduction angles are significant limiting factors. The major

limiting issue with the PW&AM technique noted so far is this need for negative supply

voltage. Without a negative supply, achieving conduction angles less than 50% (180

degrees) is not easy and controllability is very limited. However, generating an additional

negative supply voltage in cheap solutions like cell phone handsets is so undesirable that

it has already caused the handset PA industry to shift from MESFET technology to GaAs

HBTs.

96

Figure 3-58: 30% conduction angle profile

Figure 3-59: Efficiency for the 30% conduction angle profile

97

Figure 3-60: 50% conduction angle profile

Figure 3-61: Efficiency for the 50% conduction angle profile

98

Figure 3-62: 100% conduction angle profile

Figure 3-63: Efficiency for the 100% conduction angle profile

99

Another important point is that the transfer characteristic from the gate bias to output

envelope is nonlinear, as seen in Figure 3-64 for the example device. Therefore, for

systems intended for linear applications, there is a need for a scheme to linearize the

overall transfer characteristic from the desired envelope input to the envelope of the

output. Also, if the output device and the matching network are not carefully designed to

minimize the AM-PM effects it would be necessary to compensate for those effects as

well. We will discuss more about the AM-PM later in this chapter.

Figure 3-64: Output magnitude as a nonlinear function of base bias

100

3.1.3.1 EER with envelope feedback through bias control

As we noted in the previous section, PW&AM offers the promising potential for

providing better efficiency roll-off over power changes at the cost of very little additional

complexity in the bias circuit of the output device in an amplifier. This is in fact even

more promising for the systems deploying nonlinear modulations with power control. In

linear systems, however, there is an additional need for linearizing the overall

characteristic so that the restored output envelope remains a linear function of the desired

input envelope.

Constant EnvelopeRFin

Driver Stages

RL

VDD

Class C PA

EnvelopeControl

Bias Control

vg Coupler

EnvelopeDetector

+

-

Attenuation

Envelope in

ErrorSignal

Integrator

Envelope Detector/

Pre-distorter

Figure 3-65: EER with envelope feedback through bias control

101

Figure 3-65 shows a scheme based on envelope feedback that can be used to linearize a

PW&AM based amplifier. The envelope feedback and error integration assures that, as

long as the feedback loop is stable and the loop gain is high enough, the output envelope

tracks the desired envelope of the signal.

102

3.1.3.2 Bias controlled envelope feedback without envelope elimination In the traditional EER method, envelope elimination is recommended to avoid or

minimize AM-PM conversion. Since a constant envelope signal is fed to the gate, phase

variation occurs only due to supply voltage modulation. Phase response in most

traditional amplifiers changes as a function of supply voltage and is most sensitive when

the amplifiers are pushed very deeply into compression. In the approach we propose for

PW&AM, the phase response is dependent on base or gate bias and therefore some

AM-PM conversion is expected as well. If specific care is not taken to avoid this design

problem in both approaches, AM-PM can become a main limiting factor in an amplifier’s

linear performance.

Figure 3-66: Output power controlled by supply voltage

Figure 3-66 shows how the output power of a typical InGaP HBT amplifier can be

controlled by supply voltage adjustment, as is done in traditional EER systems. Figure

3-67 shows the output power of the same amplifier when, instead of the supply voltage,

the base bias voltage is adjusted as suggested in PW&AM.

103

Figure 3-67: Output power controlled by base bias voltage

Figure 3-68 AM-PM in supply and base bias controlled amplifiers

104

As noted in Figure 3-68, AM-PM conversion occurs in both cases. For a 20dB dynamic

range from 10dBm to 30dBm, phase variation is about 24 degrees and 25 degrees in the

base bias control and supply control cases respectively. This is typical behavior for

amplifiers not specifically optimized for supply modulation or bias modulation. The

phase to amplitude sensitivity is similar in both approaches. However, as intuitively

expected, the bias control method generates less AM-PM at low power output compared

to supply control. The reason is that supply controlled devices operate very close to

maximum output, i.e. with very little headroom, at low powers. On the other hand, the

bias controlled approach creates more AM-PM at high powers because the device is

being pushed to its current handling limit.

The preceding results are derived from simulating a typical amplifier driven by a constant

input power while the supply or base bias voltages are swept. Among the results in

Chapter 2, we observed that in order to achieve a 50dBc or better ACPR at all power

outputs in CDMA, the maximum phase variation preferably should be maintained below

5 degrees, and definitely below 10 degrees. This suggests a need for a design change in

order to apply these techniques to CDMA.

The amount of AM-PM conversion is heavily dependent on the matching network

topology and impedance seen at the gate and drain of the device at the carrier and the first

few harmonics, as well as the size and behavior of the power devices. Making general

statements about how to reduce the AM-PM is not straightforward. In a few CMOS

designs we observed that decreasing the effect of the device’s parasitic capacitors by

purposely adding parallel linear capacitors to the matching network next to the device

helps in reducing AM-PM. This is an intuitive result because the impedances at device

terminals become less sensitive to the parasitic capacitance of the device. It is those

voltage dependent capacitors that change in response to a change in amplitude or

operating condition and cause the change in the impedance phases and hence the phase of

the signal. On the other hand, allowing some back-off helps considerably, but costs in

efficiency. Matching networks that deliver impedances that look closer to those of a

105

parallel tuned circuit at the device terminals tend to cause less AM-PM in CMOS

implementations. However, as the impedance transformation ratio increases due to

headroom limitations caused by higher power or lower supply needs, it becomes harder to

achieve that similarity and therefore the AM-PM sensitivity increases. In cases where

these simple design tradeoffs cannot reduce the AM-PM sufficiently, an external phase

correction mechanism may become required, such as phase pre-distortion or feedback

correction in traditional polar feedback.

From the foregoing it is clear that the envelope elimination function does not eliminate

the AM-PM problem. Then what is the advantage of having a limiter to eliminate the

envelope? The drawbacks are clear: additional complexity and cost, more power

consumption and hardship in providing the right clipping behavior over large enough

dynamic ranges. One benefit is that the drain efficiency roll-off curve will have a slower

slope compared to scenarios with variable input RF amplitudes. When translated to

power added efficiency, however, the tradeoffs are more delicate since the power

consumed in the driver stages and the limiter is taken into account. In a limiter-free

design, the power consumed in the driver stage is lower and obviously no power is

consumed for the limiting function. However, the drain efficiency in the output stage is

lower compared to a design with a limiter.

An example is shown in Figure 3-69 based on simulation of the same typical amplifier

used earlier in this section. Therefore, in a good design the limiter based architecture can

yield a lower slope in efficiency roll-off at power outputs higher than some threshold.

This translates into higher efficiency for some higher power outputs, and a higher slope

and lower efficiency at lower power outputs when the power consumed in the driver

stages and limiter become a substantial portion of the total power consumed.

106

Figure 3-69: Effect of limiter on the drain efficiency of the output stage

The conclusion is that having a limiter and performing the envelope elimination function

is not necessarily required or beneficial in all solutions that apply envelope feedback. It

potentially helps the efficiency and simplifies the envelope linearization problem at

higher power outputs, but does not help much with AM-PM and creates an additional

tradeoff between high power and low power efficiencies. Judgment is therefore required

on a case by case basis as to whether or not a limiter helps based on the specifics of the

power device and the efficiency, linearity and dynamic range specifications.

A general alternative configuration without a limiter is shown in Figure 3-70. The key

additional critical design factors in this configuration are the output stage base bias versus

desired power characteristic, as well as the gain compression curve of the driver and their

interaction. These factors determine the nonlinearity of the equivalent open loop transfer

function from the input envelope to the output envelope. That knowledge in turn dictates

the minimum gain required in the feedback loop to suppress artifacts generated by

envelope nonlinearity. Hence, the feasibility of a solution depends on whether the closed

loop system can be stabilized with the required feedback gain.

107

RL

VDD

Class C PA

EnvelopeControl

Bias Control

vg Coupler

EnvelopeDetector

+

-

Attenuation

ErrorSignal

Integrator

Envelope Detector

RFin withModulated Envelope

Driver StagesApproximately linear

RF with Modulated Envelope

RL

VDD

Class C PA

EnvelopeControl

Bias Control

vg Coupler

EnvelopeDetector

+

-

Attenuation

ErrorSignal

Integrator

Envelope Detector

RFin withModulated Envelope

Driver StagesApproximately linear

RF with Modulated Envelope

Figure 3-70: Envelope feedback PW&AM without limiter

108

3.1.3.3 Envelope detector dynamic range and integrator offset effects Earlier we very briefly mentioned some of the challenges seen in feedback methods in

general and with the techniques deploying envelope detectors in particular. Now it is time

to take a closer look at these issues.

One of the most common nonlinear effects in envelope detectors is the turn-on threshold

or dead-zone effect. The quiescent operating condition of the rectifying diode or device

has to be set to minimize the dead-zone. The linearity performance of most envelope

detectors is usually worst at lower powers, since dead-zone related distortion becomes a

more substantial portion of the signal.

Figure 3-71: Waveform for an envelope detector with dead-zone

Figure 3-71 shows the time response of an envelope detector with dead-zone in response

to an AM modulated carrier as its input. For amplitudes below a threshold the detector

fails to track the envelope. In the frequency domain this translates into harmonics of the

envelope modulation frequency as seen in Figure 3-72. Any other nonlinearity in the

109

envelope detector, e.g. resulting from the nonlinear voltage vs. current of the diodes, will

contribute to further envelope harmonic content.

Figure 3-72: Effect of dead-zone on the spectrum of detected envelope

RF feed-through and RF harmonics are also common in the output of envelope detectors.

As the ratio of the bandwidth of the detector to the carrier frequency increases these

effects become more important. In feedback systems like the one shown in Figure 3-70,

the feed-through and harmonic RF content at the integrator can contribute to additional

amplitude dependent offset that degrades the overall system linearity. However, this is

usually a second-order source of nonlinearity in the closed loop system unless very

wideband envelope detectors are used.

Another problem in systems with two envelope detectors, like the one in Figure 3-70, is

that the artifacts generated by the envelope detector nonlinearity are generally not the

same in both detectors, either because of mismatch in the design or more importantly

because of different RF signal amplitude in the two detectors. In order for the artifacts to

cancel in a design with matched envelope detectors, the inputs to the detectors must have

the same amplitude. Unfortunately, the best detection performance usually is obtained in

110

power ranges different from the regular system’s input power ranges. Therefore,

implementations like Figure 3-70 may require pre-amplification of the system’s input

signal fed into the detector. This pre-amplification causes additional sensitivity to

operating frequency and can limit the operable signal dynamic range substantially.

Feeding the desired pre-distorted envelope signal directly from the base-band signal

processing is advantageous from this perspective, since the amplitude of the input to the

detector at the amplifier output can be adjusted by correct setting of the feedback

attenuation. However, such configurations eliminate the chance of canceling the

nonlinear artifacts of two envelope detectors in the subtraction.

In order to obtain information on the effects of mismatch in the envelope detectors on the

detected error signal we performed a few simple simulations using two essentially

identical detectors with turn-on thresholds of 200mV and 170mV driven by the same

modulated RF signal. Thus, the entire mismatch is modeled in a 30mV offset as shown in

Figure 3-73. Such an offset is typical in practical implementations and unfortunately

there is not a simple practical way to eliminate it for continuous time, non-sampled

envelope detectors.

Figure 3-73: Mismatch in envelope detectors

111

Figure 3-74: Error from envelope detector mismatch

Figure 3-75: Spectrum of error form mismatch

112

In Figures 3-74 and 3-75 are shown in the time and frequency domains one of the

envelope detector outputs and the difference of the two outputs, i.e., the envelope error.

Since the loop gain of the closed loop system attempts to zero the error signal which is

the input to the integrator, these errors are directly reflected as in band or spurious

artifacts on the power amplifier output signal spectrum. Any offset in the integrator is

reflected in the output in the same way. As observed in Figure 3-76, the envelope

detector turn-on threshold or dead-zone effect is reflected on the output signal of the

closed loop system as small bumps on the envelope as it crosses into or out of lower

magnitudes. In the frequency domain these bumps are seen as spurs generated at inter-

modulation frequencies of the carrier and multiple harmonics of the envelope as observed

in Figure 3-77.

Figure 3-76: Dead-zone effect in a closed loop system

113

Figure 3-77: Dead-zone effect in a closed loop system

In most systems, not only do envelope fluctuations exist due to modulation, but also a

usually larger variation of the signal amplitude occurs due to power adjustments.

Therefore, as explored before with other linearization methods, it is important to have a

clear knowledge of the linearity and dynamic range tradeoffs for the closed loop system.

Figure 3-78 contains simulation results for an example closed loop system deploying

envelope detectors with a turn on threshold of 100mV, input signal swing magnitudes

ranging up to 1.5V, an amplifier with 1dB compression point of 32dBm, and a feedback

network with 22dB voltage attenuation from the output to the envelope detector. The

system is fed with an AM modulated signal with modulation index of m=0.7, i.e., 15dB

peak to minimum, with the input power swept to see the dynamic range related effects.

114

Figure 3-78: Dynamic range of a PW&AM system with envelope feedback

As observed, at power outputs lower than 15dBm, the closed loop system cannot

reconstruct the in-band modulation dynamic shape and the modulation index is

suppressed. Substantial spurious emission is also created. This is a result of the fact that

at lower amplitudes the low side clipping of the signal in the envelope detector in the

dead-zone occurs for a high percentage of the time and therefore becomes the dominant

nonlinear effect in the system. Also at lower input and output powers, due to the dead-

zone, the loop gain is substantially reduced and therefore the loop fails to effectively

correct the distortions. In summary, unfortunately, the loop is not capable of correcting

the dead-zone effect. As the power output increases, the probability of the low side

clipping decreases and the system starts to correctly track the modulation index. As a

result, spurious emissions are reduced as less clipping occurs in the envelope detectors.

As the power increases, the envelope detector becomes more effective. Regrettably

amplifier saturation occurs soon after, substantially degrading the linearity by

115

suppressing the modulation index and increasing the spurs. Unfortunately, these power

and voltage ranges in handset amplifiers are very typical and as a result usage of

envelope detector based systems in handsets becomes severely limited for this same low

dynamic range problem. When higher voltages are available, various microwave diodes

make it possible to implement envelope detectors with better dynamic range. This means

that in systems with higher power output and higher supply voltages, the high side

clipping is shifted to the higher power output and the low side clipping is further shifted

to lower power output and as a result a wider dynamic range with reasonable linear

performance becomes achievable. This suggests that infrastructure amplifiers offer better

potential for envelope detector based feedback system implementations.

116

3.1.3.4 Bandwidth effects and dynamic nonlinearity We mentioned before that one of the main drawbacks of the linearization methods

deploying feedback is a range of problems that arise from narrow bandwidth in the

feedback loop, including loop stability issues. In this section we quantify some of these

effects for the example closed loop PW&AM system with envelope feedback considered

earlier in this chapter. In a well designed case the closed loop system tracks the desired

envelope modulation very well for low enough envelope frequencies or narrowband

modulations. Figure 3-79 shows how well the output tracks the AM modulated input

when the envelope modulation frequency is around 150 kHz.

Figure 3-79: Closed loop system tracks slower modulations

117

Figure 3-80: Closed loop system fails to track fast modulations

In contrast when the same system is driven by a signal with the same AM modulation

depth, but with the envelope frequency of 15MHz, as observed in Figure 3-80, the closed

loop system fails to track the envelope modulation appropriately. Figure 3-81 shows the

effect of envelope modulation frequency on modulation depth and spurious emission at

the output of the system. As observed this particular system tracks the envelope

modulation reasonably well and generates relatively low spurs for envelope modulation

frequencies or modulation bandwidths lower than 1MHz. For envelope modulation

frequencies higher than 4MHz the spur level increases substantially and the closed loop

system fails to track the modulation depth as well.

In this example, while remaining approximately valid for modulation bandwidths lower

than 1MHz, the quasistatic assumption, discussed in the previous chapter, clearly fails for

modulation bandwidths higher than 1MHz. This example shows that application of the

quasistatic nonlinearity assumption needs careful consideration in systems with envelope

feedback, because in these kinds of systems envelope and RF signals are governed by

118

vastly different dynamics. This causes the difference in performance between

narrowband and wideband modulations.

Figure 3-81: Frequency response of a closed loop PW&AM system

One of the main and common considerations in every feedback system is stability.

Systems with envelope feedback are usually stable in the sense that when no modulation

is applied no oscillation is observed. However, a more descriptive notion would be

envelope stability. Envelope stability can be defined by how well the output envelope

tracks that of the input. We refer to this stability as in-channel envelope stability (ICES).

It can also be defined as a function of how much spur is generated as a result of imperfect

envelope tracking. We will refer to this second definition as out-of-channel envelope

stability (OCES). Thinking in the context of envelope stability as described here, we

observe that the loop is envelope stable up to certain frequencies and the envelope

stability gets worse for higher modulation frequencies. The loop gain in this kind of

systems is nonlinearly dependent on the envelope amplitude as well as the envelope

frequency. The frequency dependence is caused by the integrator, filters in the envelope

119

detectors and the frequency response of the amplifier’s output filter as well as the

feedback sampling network that are all frequency selective. The peak loop gain is

amplitude dependent because of the saturation in the integrator, output envelope detector

and the amplifier itself. As a result while still remaining a very useful tool for gaining

intuition about the behavior of the system, the simple linear stability analysis methods

based on the loop phase and gain margin fail to accurately predict frequencies where the

envelope stability starts to degrade.

Figure 3-82: Loop gain effect on envelope stability

However one of the expected tradeoffs that remains valid is the fact that higher average

loop gain reduces spurs further at the envelope frequencies where the loop is envelope

stable. At the same time higher loop gain reduces the overall envelope stable closed loop

bandwidth, while increasing the spurious level at higher envelope frequencies where the

system is not envelope stable. This is noted in Figure 3-82.

120

To summarize our observations, we have noted that linearity of the closed loop PW&AM

systems with envelope feedback, which can be seen by observing error in tracking of the

modulation depth as an in-channel measure and maximum spurious level as an out-of-

channel measure, is dependent on the signal amplitude and modulation depth as well as

the envelope frequency. The amplitude dependence is primarily caused by saturation

effects in PA stages and integrator, the turn-on threshold in the envelope detectors and

error offset in the envelope detectors and the integrator. The frequency dependence is a

result of the frequency dependent modules, e.g., error integrator, amplifier output filter,

coupling and feedback network and the envelope detector filters.

Before concluding this section let’s observe the combined effect of these factors in a

closed loop system.

Figure 3-83: In-channel linearity and stability

121

As far as the in-channel linearity is concerned, as seen in Figure 3-83, the envelope

detector dead-zone is the primary degrading factor. In systems with no dead-zone the

desired envelope is tracked very well for the in-channel-envelope stable envelope

frequencies, while the envelope skew substantially increases at higher frequencies where

the envelope stability degrades. Envelope detector turn-on threshold always is reflected

as a dead-zone in the system and causes envelope skew even at lower envelope

frequencies where the system is otherwise envelope stable. Any offset in the envelope

detector output and integrator input is reflected as an additional random error in the

envelope and is observed as a modulation skew or a change in modulation depth. In

systems with an envelope detector induced dead-zone, higher amplitude signals suffer

less from modulation skew, as intuitively expected. Also, as is observed in Figure 3-83,

the in-channel-envelope stability bandwidth of the closed loop system with dead-zone is

dependent on the amplitude. The in-channel envelope stable bandwidth is lower for the

lower signal amplitudes.

Figure 3-84: Out-of-channel linearity and stability

122

In Figure 3-84 we see the impact on the out of channel spurious generated in the closed

loop system. As noted for the high amplitudes, the presence of the dead-zone and offset

have minimal impact on spurious levels at the frequencies where the system is envelope

stable. However, the sensitivity of the spur levels to the amplitude of the signal is quite

high when the envelope detector induced dead-zone is present. This refers to the same

impact of the dead-zone on limiting the linear dynamic range as earlier observed in

Figure 3-78.

In conclusion, envelope detector dynamic range severely limits the achievable dynamic

range for the envelope feedback based systems in general, and for PW&AM systems with

envelope feedback in particular. On the other hand, the presence of the feedback loop

substantially limits the usable modulation bandwidth of the system. Therefore alternative

implementations of linear power amplifier systems that use no feedback and no envelope

detectors may be more desirable, when possible.

123

3.1.3.5 Potential for bias-controlled ER with envelope pre-distortion Based on what is observed so far, an implementation of the bias-controlled envelope

restoration without envelope feedback or envelope detectors, like the one shown in

Figure 3-56, promises higher bandwidth and higher linear dynamic range. Unfortunately

as noted earlier, in most of the amplifiers, the transfer characteristic from the output stage

bias to the output envelope magnitude is not linear. Therefore, an appropriately pre-

distorted envelope control signal needs to be fed into the system in order to produce the

desired output modulation.

Common challenges of pre-distortion based systems, including sensitivity to temperature

and manufacturing and aging related drifts, apply to envelope pre-distortion based

PW&AM systems as well. As a result these may necessitate adaptive adjustments of the

envelope pre-distortion table. The advantage compared to traditional pre-distortion is the

potentially better efficiency vs. power roll off curves offered by bias controlled PW&AM

amplifiers.

124

3.2 Linearization by design of shape of gain compression curve Earlier in section 2.3 it was observed that an expansive gain response followed by a

compressive one right before the saturation provides a local minimum on the ACPR

versus power curve at a reasonably low back off from the saturation. This was in the

discussion of Figure 2-17 in a study of the effect of various biasing conditions and their

corresponding shape of gain compression curves. This means that an acceptable linear

performance can be achieved at back off levels lower than those traditionally assumed to

be required. The possibility of operating closer to saturation would allow higher peak

linear efficiencies to be achieved.

Also, as we will soon discuss in further detail, most power devices experience some level

of self-biasing [56 and 57]. That is, as the power output increases the bias current level in

the device increases. One positive benefit from this self-biasing behavior is that the

quiescent current of the device can be set at a lower current for most of the lower power

outputs and as the power output increases the device current is boosted up to the higher

currents required [58]. This leads to efficiency roll off curves with lower slope, i.e. higher

efficiencies at lower powers in devices with higher self-biasing [59]. One associated

problem though is that too much self-biasing with much lower bias currents at lower

powers creates an expansive gain behavior which can have destructive impact on the

linearity if not properly compensated. In this section we will discuss these issues in

further detail.

125

3.2.1 A note on self-biasing The base or gate bias lines of output and driver stage devices in power amplifiers is

usually supplied from bias reference circuits that have relatively low output impedance.

As a result these can be considered as voltage references. In HBT implementations the

base current needs to be supplied by the bias circuit. The impedance of the bias circuit’s

output has to be low enough to provide the drive capability to deliver the base current

while setting the correct base voltage. Figure 3-85 shows a bias scheme that is very

commonly used in InGaP HBT designs.

Vcc

RF

Bias Vcc Vref

Matching Network

To Next Stage

Figure 3-85: Biasing an amplifier stage

When the output or driver transistors are driven with large RF signals the nonlinear

behavior of the device causes the device to function as a rectifier and generate the a

portion of the required DC current from the RF current pulses. This occurs when the

initially set quiescent current is not high enough. The amplification and rectification

function when the device amplifies larger RF signals causes the device to draw higher

DC currents from the Vcc and the base bias line to produce the output current pulses.

This self-biasing behavior allows the device to move automatically to operate at the DC

current levels needed to provide the needed current pulse amplitudes and waveforms.

126

Figure 3-86 shows this behavior as observed in the output stage of a typical power device

biased with a scheme essentially the same as the one shown in Figure 3-85.

Figure 3-86: Effect of load impedance on self-biased DC current

In figure 3-86 the fixed quiescent current is 60mA. As a result of the self-biasing, the DC

current substantially increases to provide the required current for delivering the output

power. When load impedance is lowered, higher currents are needed to deliver the same

power. The three curves represent the collector current of the same device with the same

input biasing and matching for three different load impedances on the collector.

Interestingly, the self-biasing mechanism automatically increases the current to deliver

the higher needed currents when the load impedance is lower.

The shape and magnitude of the high power part of the self-biasing curves in Figure 3-86

are functions of the device structure and the load impedance seen by the collector at the

carrier frequency and its harmonics. On the other hand at lower power outputs the bias

circuit determines the current. This behavior is very useful since it allows the device to

move to the operating condition needed to deliver the required power. However, it

complicates the amplifier design process since it couples the output match and biasing

127

design problems together. As long as the bias reference circuit has the capability of

delivering the maximum required base DC current, the tradeoffs can be made

successfully.

Before proceeding to explore how the tradeoffs in amplifiers with multiple self-biased

stages can be used to provide required linearity while improving the power efficiency, it

will be useful to briefly consider how the load impedance affects gain and power output

of self-biased amplifiers. As intuitively expected, when the primary saturation

mechanism is supply voltage limitation as is the traditional design in most amplifiers,

higher load impedance translates into higher gain and lower maximum output power.

This is shown in Figure 3-87.

Figure 3-87: Effect of load impedance on gain and power output

Another practical point that needs to be noted is that almost every power device

experiences some self-biasing unless particular care is taken to avoid it. At high levels of

self-biasing, the conduction angle in the device is quite high. Therefore, practical

implementations of most of the other techniques discussed earlier in this chapter that rely

on lower conduction angles in the devices, require careful design to minimize self-biasing

and that may dictate a need for negative bias supply voltages as well.

128

3.2.2 Cascade of self-biased nonlinear stages As discussed in the previous section at higher power outputs the DC current of a device

in the self-biased region is primarily determined by the load it sees and the power of the

signal. When the device is large enough the compression behavior and the peak power is

primarily determined by the load impedance. Gain is always dependent on the current and

load impedance. On the other hand at lower power outputs the current is set by the

biasing circuit. As seen in Figure 3-86, a ten-fold variation in the DC current can occur.

Since gain is dependent on the current, more self-biasing is expected to cause more gain

variation over a power range. Since current at high power outputs is relatively

independent of the quiescent current, biasing devices at lower quiescent currents creates

more self-biasing which can cause more gain variation over the power range.

Figure 3-88: Self-biasing for various quiescent currents

129

Figure 3-89: Shape of gain

Figure 3-88, represents simulated DC current versus power output of the output stage in a

typical two stage amplifier for quiescent currents ranging from about 20mA to 80mA. For

higher power outputs, e.g., higher than 15dBm in this particular design, the current of the

stage is predominantly determined by the self-biasing mechanism. Figures 3-89 and 3-90

show the magnitude and phase of the amplifier gain for the same set of currents shown in

Figure 3-88. For lower quiescent current we observe more self-biasing, more gain

variation, larger gain peaking before saturation, compression at slightly lower power

outputs, and more phase variation over power change at lower power outputs. At

sufficiently large quiescent current the low power gain can be made large enough so that

no gain expansion occurs.

130

Figure 3-90: Phase variation

Figure 3-91: Impact of the second stage on current of the driver stage

131

In these simulations the maximum power output of the device changes from 31dBm to

31.5dBm as the quiescent current increases. Usually a change in the self-biasing curve,

i.e. DC current versus power output characteristic, in the output stage causes a change in

the self-biasing curve of the driver stage as well. This occurs because the base impedance

of the output stage changes as the current changes in the output stage. Through the

interstage matching network this change is reflected onto the impedance seen at the

collector of the driver stage. This changes the self-biasing curve of the driver stage as

shown in Figure 3-91. This further illustrates the fact that different stages in power

amplifiers interact so it is very important to perform the design and optimization of

multistage amplifiers as a whole with careful consideration of these interactions.

Figure 3-92: Power added efficiency of the two stage amplifier

For a two stage amplifier, simulated power added efficiency roll-off curves for the above

seven quiescent current values for the output stage are seen in Figure 3-92, and with

further detail in Figure 3-93 and Figure 3-94. At high power outputs the power added

efficiency roll-off slope is -4.4% per each dB of back off. This is reflected as a 2%

increase in the case of IQ=20mA compared with IQ=80mA as a result of the 0.5dB change

132

in the peak power. In general, designs using lower quiescent currents show higher

efficiencies compared to their equivalent counterparts with higher quiescent currents.

Figure 3-93: Power added efficiency of the two stage amplifier

Figure 3-94: Power added efficiency of the two stage amplifier

133

Therefore in more modern designs where there is a need to take advantage of all the

incremental improvements that can be achieved at no cost, the tradeoffs of linearity and

efficiency related to quiescent current and self-biasing behavior also need to be

considered. In general, lower current translates into better efficiency. Power capability of

the device and the load design dictate the minimum current at high power output, taking

into account the load requirements for the linearity. Minimum usable quiescent current is

limited by the required linearity and gain at lower power outputs as well as the

requirement to guarantee that devices remain biased on over the required temperature

range and manufacturing related device variations.

Figure 3-95: Linearity of the two stage amplifier

134

Figure 3-95 shows the simulated linearity performance of the two stage reference

amplifier used in this section for different quiescent currents. The ACPR curves are

derived using the method discussed in Chapter 2. As seen in the case of IQ=80mA, too

high quiescent current turns the gain response into a fully compressive one and therefore

the performance at power outputs below 27dBm is not as good as for IQ=70mA, which is

the closest approximation to a clipped linear gain characteristic. In all the other cases the

expansive/compressive nonlinear gain characteristic causes the ACPR to have a local

minimum between 26 and 27dBm. This range has been chosen by design to provide the

required linearity of close to, or better than, -50dBc ACPR at or below 28dBm output,

while minimizing the back-off and delivering the highest achievable efficiencies at

28dBm, the peak linear power specification for most CDMA handset amplifiers.

At lower power outputs as also discussed in Chapter 2, ACPR curves have a local

maximum between 18dBm and 13dBm and then roll down again. This is mainly due to

the gain magnitude having a zero second derivative somewhere in that power range. Of

course, the phase characteristics also have an impact on the shape of these ACPR curves.

Although considered second order effects due to reasonably low variation of the phase

with power output, these effects also have been taken into account. The case of 40mA

delivers the highest efficiency among the ones with acceptable linearity of better than

-50dBc at power outputs lower than 28dBm. This case shows 1.8dB of gain peaking,

peak power of 31.3dBm, i.e. back-off of 3.3dB at 28dBm delivering -50dBc linearity,

with a peak linear efficiency of 42% at 28dBm. These simulations were used here as an

example design case to show how the tradeoffs of gain, power, efficiency and linearity

related to load design, self-biasing and gain compression curve design for ACPR can be

interactively incorporated in one simple design that delivers the required linearity while

improving the efficiency. Before taking advantage of these tradeoffs in the design, the

highest linear efficiencies for CDMA handset power amplifiers were only slightly over

36%.

135

Vcc=3.4, Vref1=Vref2=3.0, F=1.765GHz

25.00

26.00

27.00

28.00

29.00

30.00

31.00

32.00

0.00 4.00 8.00 12.00 16.00 20.00 24.00 28.00

Pout (dBm)

GA

IN (d

B)

+25C -30C +85C

Figure 3-96: Measured gain

Vcc=3.4, Vref1=Vref2=3.0, F=1.765GHz

-70.00

-65.00

-60.00

-55.00

-50.00

-45.00

-40.00

0.00 5.00 10.00 15.00 20.00 25.00 30.00Pout (dBm)

AC

PR (d

Bc)

+25C -30C +85C

Figure 3-97: Measured ACPR

136

Vcc=3.4, Vref1=Vref2=3.0, F=1.765GHz

0.00

5.00

10.00

15.00

20.00

25.00

30.00

35.00

40.00

45.00

50.00

15.00 17.00 19.00 21.00 23.00 25.00 27.00 29.00Pout (dBm)

PAE

%

+25C -30C +85C

Figure 3-98: Measured power added efficiency

Figures 3-96, 3-97 and 3-98 show measurement results on a sample two stage power

amplifier designed using the above approach on a 2µm InGaP HBT process. Figure 3-96,

Figure 3-97 and Figure 3-98 show gain, adjacent channel ACPR, and power added

efficiency respectively. The data were collected at three ambient temperatures of -30C,

+25C and +85C. The expansion and compression is seen on the gain curves. Also local

minima close to 28dBm on the ACPR curves are observed as expected. Efficiency at

28dBm ranges from 39% to 42%. The test signal was a CDMA modulated carrier at

1.765GHz.

One interesting observation is that for +85C where the gain variation is less than that seen

for +25C and -30C, the ACPR curve is also lower at power outputs less than 26dBm.

This is consistent with the prediction from Figures 3-89 and 3-95 that cases with less gain

peaking would have lower ACPR at lower power outputs.

137

It also needs to be noted that the numerical error in simulating ACPR is dependent on the

length of the data sequence used in the simulation among other factors. In most of the

simulations in this thesis in order to maintain the simulation times reasonable, the length

was set and a fixed sequence was used. The numerical error due to this concatenation

causes the program to estimate an ACPR of -59dBc for the original non-distorted

sequence. This is reflected in most of the simulated ACPR curves in this thesis as a noise

floor of -59dBc. Since the target linearity is -50dBc, this sequence length provides a

reasonable margin to account for the critical ACPR induced by the amplifier nonlinearity

while maintaining the simulation times reasonable and the need for memory at

manageable levels. This is the reason for the observation that while the measured ACPR

results are below -60dBc at low power outputs, the simulation results have a minimum

value of -59dBc.

138

139

Chapter 4: PA module constructions and matching methods

In cellular phone handset applications, it is customary to use internally matched power

amplifier modules (PAMs). These modules include the amplifier die, elements for input

matching and output impedance transformation to 50 ohms and in some applications

additional semiconductor dice for bias and power control or other signal conditioning

functions [60]. The most usual construction includes a multilayer substrate to which the

other elements are soldered or glued using a die attach, and a mold on top that covers and

protects the wire bonds [61]. The input impedance for the module is 50 ohms.

The amplifier performance is optimized based on the assumption that the module sees a

50 ohm nominal load. The robustness is specified for a range of VSWRs. This has

become the standard way of specifying and measuring handset power amplifier module

performance. The objectives are to have more standardized ways of testing and

benchmarking PAs, standardizing transformation networks on the phone printed circuit

board and ultimately allowing “drop in” of equivalent power amplifiers from different

manufacturers. Since the output impedance transformation network has to have minimum

loss in order to minimize the efficiency degradation, it is not possible to design a module

with an output impedance of 50 ohms. Doing so will simply impose an extra

unacceptable 50% power loss. So the 50 ohm output matching refers to an impedance

transformation network that allows delivering the power to a 50 ohm load while

presenting the right impedance to the device so that it delivers appropriate power and

performance.

It is traditionally assumed that building high performance PAMs is only possible by using

off-chip inductors and capacitors for the output matching network. Among the reasons

are need for a) high-Q elements, since the loss at the output matching network causes

substantial efficiency degradation, b) low tolerance of amplifiers to output matching

element variations, and c) the need to manually modify and tune the module to yield the

140

desired performance, since most of the designs still are fine tuned by trial and error. As a

result the majority of today’s handset PAMs include a number of off-chip passive

elements that are used to implement the output impedance transformation network.

Various multi chip module (MCM) manufacturing technologies are used to integrate such

off-chip passives into the PAM [62][63][64]. In this chapter we show how these off-chip

passives can be eliminated to reduce the size and cost of the PA modules [65]. Pros and

cons of a few alternative matching network topologies are discussed along with some

related experimental data.

141

4.1 Traditional output matching network topologies and implementations

In today’s handset applications, power amplifiers operate on a rechargeable battery and

the supply voltage ranges from near 4V at full charge to 3.2V on a nearly fully

discharged one. Thus, amplifiers have to deliver power while meeting linearity

specifications at supply voltages as low as 3.2V. The specification for peak power, i.e. no

back-off, ranges from about 32dBm in CDMA handsets to 35.5dBm in some of the

multimode GSM handsets. This requires the real part of the load impedance on the

collector of the output device to be in the range of 1.5 to 3 ohms. Output device sizes are

primarily determined by the need to drive such loads with the required power. The

imaginary part of the load at the collector is usually tuned to optimize for efficiency,

linearity or a mixture of the two for a specific device. It is the function of the output

matching network to transform the 50ohm load into the desired impedance in that range,

e.g. 3 ohms, at the collector of the device.

One of the most common and efficient topologies for the output stage impedance

transformation network in handset power amplifier modules is a two stage low pass

matching network as shown in Figure 4-1. Some of the common variations come with

additional series LC tanks in parallel with the collector that resonate at 2nd or other

harmonics and act as harmonic traps [66].

To Collector of theOutput Transistor

Bond Wires

Chip Module Phone PCBModule

Vcc

(Z ~ R+jX Ohm)

Figure 4-1: A two stage low pass matching network

142

The two stage design is attractive since the Q of each stage can be below 2 compared to

the Q’s above 4 that are required for single stage networks. Lower Q in the stages means

that more wideband impedance transformation is possible and also the networks and their

respective amplifiers can be designed to be less sensitive to component variations. This

promises better yield with typical component variations. Wider bandwidth is desired

particularly in multi-band PA applications.

Discrete Passives

Power Amplifier Chip

Bias Control Chip

Discrete Passives

Power Amplifier Chip

Bias Control Chip

Figure 4-2: A traditional PA module

As seen in Figure 4-1 such implementations require at least 6 passive elements either

implemented on or soldered onto the module’s substrate for each amplifier. A module

may include more than one amplifier to address the need for multiple frequency bands. In

143

topologies with additional harmonic traps this number grows to 7 or 8 depending on the

specifics of the implementation. Depending on the values of the inductors, some can be

implemented using traces on the substrate. In handset PA modules the capacitors are

usually standard components soldered onto the substrate. Figure 4-2 shows a picture of a

module that deploys a matching network topology and implementation like the one in

Figure 4-1.

144

4.2 A new approach to eliminate off chip discrete passives In the traditional matching method discussed in the previous section, a major portion of

the manufacturing cost is due to the extra passives and multi layer substrates large

enough to allow enough space for routing, mounting the PA die and soldering the

passives.

Module yield is another major cost related concern. Although the two stage matching

network design promises better yield from the perspective of sensitivity to component

variations, maintaining the module yield high enough while attempting to squeeze all the

extra passive components in a smaller area can be very challenging. One of the major

problems is the soldering yield when passives are squeezed too close together or to the

bond wires.

On the other hand smaller PA modules are more attractive since they save space on the

phone boards. At the same time the need to have multiple power amplifiers in a single

module to address the need for multiple standards or multiple bands imposes tighter

restrictions on the module size. Based on these concerns it is obvious that a major step in

improving the PA modules from the cost, manufacturability and size points of view is to

eliminate the off chip passives. In any alternative implementation approach the concern

about whether desirable performance and yield can be achieved remains a critical one.

The new alternative approach presented in the remainder of this chapter is based on a

simple and old concept. Every capacitor has to be implemented on the power amplifier

die. Every inductor has to be made of bond wires that go between the PA chip and the

package leads or the module’s substrate. We can call this the “no passives guideline”.

The bond wires quality factor is high enough to allow implementation of impedance

transformation networks with acceptable loss. Fortunately high-Q capacitors are also

available in InGaP processes that are commonly used in making handset power

amplifiers. This promises that the loss in the output matching networks implemented

145

using this concept should have losses comparable with their off-chip counterparts in the

modules and therefore reasonable efficiencies should be achievable.

From a designer’s perspective one of the attractions of this approach is that one does not

have to deal with all the unwanted self resonance effects of the discrete passives

anymore. Those self resonances in the discrete capacitors are dominant factors in the

shaping of the impedance at the harmonics of the carrier in handset PAs. Therefore

design of higher order networks for the harmonic shaping can become more challenging

when using capacitors with low self resonance frequencies. A disadvantage of the newly

proposed approach however is that tuning the design by changing capacitor values

becomes impossible. At the same time dependence of the PA performance on the values

and couplings of the output related bond wires becomes more significant. Therefore more

careful electromagnetic modeling of the bond wire values and their couplings can create

substantially better insight in this kind of design. In summary some newer design

optimization methodologies, different from the ones used for more traditional modules,

are required to follow the no passives guideline presented here. That includes the need to

explore newer matching topologies for the output match, as well as devising fine tuning

methodologies based on appropriate design of experiments.

From the design for manufacturability perspective designers must deal with some newer

tradeoffs in the proposed no passives method. It becomes particularly important for the

designers to revisit the practical tolerances in the packaging process and their impact on

the range of values of the bond wires and their couplings and the subsequent effect on

amplifier performance.

In the remainder of this chapter we will discuss in further detail results obtained from

some specific attempts to make power amplifier modules based on the no passives

guideline.

146

4.2.1 HP matching networks

The first thought that naturally comes to mind is to attempt implementing the same

proven matching network shown in Figure 4-1 based on the no passives guideline. Every

inductor has to be implemented using one or a few bond wires in series or in parallel.

Each end of an inductor that meets a capacitor needs to be bonded on the PA die. The

inductors that are connected to two capacitors need to have the two ends on the die.

Therefore they have to be formed using at least two bond wires in series one going out

from the die to a pin on the package and one coming back from the same or a connected

pin on the package to another bond pad on the die. For a matching topology like the one

in Figure 4-1 this calls for at least six critical bond wires that determine the shape and

value of the impedance at the collector and directly affect all the performance measures

of a PA. Any error, unexpected coupling or off tuning related to those inductors can ruin

the PA’s performance. Two capacitors are also required on the chip after the output

device which imposes additional restrictions on the layout of the PA chip’s output.

In summary implementing Figure 4-1 based on the no passives guideline sounds very

risky as a first attempt in checking the feasibility of the approach. In order to avoid the

unwanted extra risk in the first attempt it seems more logical to try a simpler topology. A

first natural step in simplifying the problem is to use a single stage matching network.

The available alternatives for a single stage match are the low pass and the high pass

matching networks. The low pass method calls for an extra DC blocking capacitor.

Figure 4-3 shows the circuit schematic and a sample sketch of a bonding diagram for the

low pass method.

147

50 OhmLoad

Vcc

To Collector of theOutput Transistor

Low Pass Method

DC BlockOut

To Ground throughBackside Via orBond Wire

Vcc

Out

Chip Module/Package

To Collector of theOutput Transistor

Figure 4-3: Low pass matching

The alternative high pass method is represented in Figure 4-4. The naming is rooted in

the fact that the single stage LC networks that perform impedance transformation in the

two cases are low pass and high pass networks. Compared to the low pass case, the high

pass alternative needs one less capacitor and fewer bond wires and there is no need to

take the RF signal out of the chip and bring it back. This promises a much simpler tuning

problem. When needed, harmonic traps can be added to both networks by adding a

capacitor on the die with one side connected to the collector of the output stage and the

other side connected to bond wires going to ground on the module or package.

50 OhmLoad

Vcc

Bond Wire'sParasitic

High Pass Method

To Collector of theOutput Transistor

Out

Chip Module/Package

Vcc

Out

To Collector of theOutput Transistor

Gnd

Figure 4-4: High pass matching

148

Figure 4-5: PA module in a lead frame package with no discrete passives

As already noted when following the no passives guideline, matched power amplifier

modules can be built using standard lead frame based packaging techniques, such as the

one shown in Figure 4-5. Based on this approach several amplifiers have been designed,

built and tested. They successfully delivered their target performance for different

applications. As an example Figures 4-6 and 4-7 contain some measurement results from

a PA module implemented using the same approach for a wideband CDMA application.

Using this approach allows shrinking of the CDMA PA module sizes from 6mm x 6mm

to 4mm x 4mm and subsequently to 3mm x3mm in the newest generations.

149

Vcc=3.6, Vref=3.0, F=1.95GHz

0.00

5.00

10.00

15.00

20.00

25.00

30.00

35.00

40.00

45.00

50.00

15.00 17.00 19.00 21.00 23.00 25.00 27.00 29.00

Pout (dBm)

PAE

%

PAE +25C PAE -30C PAE +85C

Figure 4-6: Measure power added efficiency

Vcc=3.6, Vref=3.0, F=1.95GHz

-60.00

-50.00

-40.00

-30.00

-20.00

-10.00

0.00

15.00 17.00 19.00 21.00 23.00 25.00 27.00 29.00

Pout (dBm)

AC

LR (d

Bc)

ACLR1 +25C ACLR1 -30C ACLR1 +85C

Figure 4-7: Measured linearity

150

4.2.2 LPHP and LPLP networks As briefly mentioned in the previous section, two stage matching networks like the one in

Figure 4-1 can be designed to provide wider bandwidth for the impedance transformation

compared to the single stage cases discussed in the previous section. This is due to the

additional degree of freedom in the design provided by the second matching stage. In

some applications, such as some multiband and multistandard handsets where the PA

modules have to perform over wider bandwidths, single stage matching networks fail and

inevitably one has to consider the two stage matching networks. In this section we

discuss some results obtained from experiments in designing power amplifier modules

based on two stage output matching while attempting to follow the no passives guideline.

Following the same naming convention used in the previous section, the matching

network shown in Figure 4-1 can be referred to as a low pass-low pass (LPLP) network.

That topology can be implemented using on chip capacitors and bond wires as presented

in Figure 4-8.

To Ground throughBackside Vias orBond Wires

Vcc

Out

Chip Module/Package

To Collector of theOutput Transistor

LPLP Matching

Figure 4-8: Low pass-low pass matching

151

The main problem with such an implementation is the critical dependence of amplifier

performance on too many critical coupled bond wires. The coupling is high because of

the physical proximity that is dictated by the small size of the die’s output side. This

makes the task of tuning and optimization of the module much more complicated.

Another disadvantage of the implementation shown in Figure 4-8 is the need for three on

chip capacitors. Comparing this with the simple implementation of the single stage high

pass network in Figure 4-4 that has only one capacitor on the chip with only one critical

inductor implemented by a single bond wire or two parallel bond wires shows the

additional complexity of LPLP matching.

To reduce practical complexity of the implementation, one solution is to release the no

passives requirement to allow for only one capacitor on the module. As seen in

Figure 4-9, this approach reduces the number of bond wires and allows more space

between them and therefore reduces the couplings to a more manageable level. There is a

cost and space penalty for having the capacitor on the module but it is still much simpler

than traditional implementations requiring 6 passives or more on the module.

To Ground throughBackside Vias orBond Wires

Vcc

Out

Chip Module/Package

To Collector of theOutput Transistor

LPLP Matchingwith One Capacitoron the Module

Figure 4-9: Low pass-low pass matching with one discrete capacitor

152

Another alternative to the LPLP matching network is a two stage low pass-high pass

(LPHP) network consisting of a low pass and a high pass section as shown in

Figure 4-10.

50 OhmLoad

To Collector of theOutput Transistor

Bond Wire'sParasitic

Vcc

Out

LPHP Matching

Figure 4-10: Low pass-high pass matching

As observed in Figure 4-11, this topology can be fabricated following the no passives

guideline and the complexity is slightly lower than Figure 4-8 since it needs fewer bond

wires and one less capacitor on the chip.

To Ground throughBackside Vias orBond Wires

Vcc

Out

Chip Module/Package

To Collector of theOutput Transistor

LPHP Matching

Gnd

Figure 4-11: Implementation of LPHP matching

153

Other alternatives are high pass-low pass (HPLP) and high pass-high pass (HPHP)

networks. HPLP does not reduce the complexity of the problem any further. HPHP

networks are the simplest ones from the implementation point of view. The HPHP

network has only two or two sets of parallel critical bond wires as shown in Figure 4-12

and Figure 4-13.

50 OhmLoad

To Collector of theOutput Transistor

Bond Wire'sParasitic

Vcc

Out

HPHP Matching

Figure 4-12: High pass-high pass matching

Chip Module/Package

Vcc

Out

To Collector of theOutput Transistor

Gnd

Gnd

HPHP Matching

Figure 4-13: Implementation of HPHP matching

However, another important practical concern is the allowable harmonic level at the

output of a power amplifier. With HPHP output networks more care and additional

mechanisms are needed to maintain tolerable harmonic levels. Hence, in a specific set of

154

experiments performed to study feasibility and practical issues of two-stage output

matching networks for multiband power amplifiers intended for GSM applications which

have rather tight specs on harmonic levels, the experiments were narrowed to

accommodate only LPHP as in Figure 4-11 and LPLP with one capacitor on the module

as in Figure 4-9. We should remember that for each scenario of output matching, a new

chip has to be designed to include the correct capacitors and pin out for the output

matching network.

With careful design and enough effort in optimizing the solution through iterations both

of the topologies are promising. As expected, the simpler less coupled solution of

Figure 4-9 could be tuned to deliver the power output and efficiency with less effort. At

35dBm the efficiency was close to 55% after module tunings in the first attempt on the

chip. In the first IC run, the LPHP case delivered only 44% efficiency at around 33dBm

and was short on delivering the target power of 35dBm. Balanced and symmetric PA

layouts are common particularly at high frequencies [66]. However, in both of these

experiments in order to reduce the die size aggressively, no specific network was put in to

balance the load on the output device. This limits the balancing to whatever level is

provided by quasi-symmetric placements of wire bonds and other elements on the die and

module. This aggressive move created reasonably acceptable results in the LPLP case

where enough room was available to balance the layout more effectively while facing

fewer bond wire coupling related problems. On the other hand due to the space tightness

and coupling related problems, the more complex LPHP implementation had more

unsolvable asymmetry that degraded the amplifiers performance substantially.

This asymmetry effect was observed in thermal images of the amplifiers. Figure 4-14, the

LPHP case, shows much more loading asymmetry to the extent that the upper half of the

device is almost not functioning. Transistor temperature in the other half is as high as 189

degrees C. In comparison, as seen in Figure 4-15 for the LPLP case, transistor load is

balanced more effectively and most of the output transistor cells except for the ones close

to the two edges are functioning relatively well, allowing the hottest output device cells

to operate at cooler temperatures, as low as 102 degrees C.

155

This area is least effectiveTemperature= 167 °C

This area provides most of the power outputTemperature= 189 °C

This area is least effectiveTemperature= 167 °C

This area provides most of the power outputTemperature= 189 °C

Figure 4-14: A grossly power imbalanced PA

Most active cellsTemperature= 102 °C

Bonding is more symmetric compared to Figure 4-14

Most active cellsTemperature= 102 °C

Bonding is more symmetric compared to Figure 4-14

Figure 4-15: A PA with some power imbalance

156

Unfortunately, there is no known systematic relationship between the achievable design

symmetry and the topology of the output match or its bonding complexity. In fact, with

additional iterations in making the LPHP design more symmetric, it should be possible to

achieve the same performance levels as the other topology. However, dealing with

multiple interrelated complexity and tradeoff problems at the same time can distract a

designer from being able to address all issues and thus to achieve the desired optimal

performance in one attempt. The many interrelationships between design tradeoffs are the

primary reason that power amplifier design is a highly iterative process. In summary,

comparing various alternative topologies and implementations for the output match, it is

still not clear what systematic performance related disadvantages or advantages exist for

any of them, except from the harmonic content point of view. One expects better

performance results can be achieved with less effort on iterations and tuning for the ones

that look simpler.

Since we brought up the issue of asymmetry and its effect on the thermal and electrical

performance of the amplifiers, it may be appropriate to conclude this section with a few

more sentences on this topic. An alternative and perhaps more appropriate phrase for the

asymmetry mentioned above could be power imbalance which can be defined as variation

of the delivered RF power and or dissipated power over different transistor cells in the

output stage. When transistor cells are identical in construction, power imbalance can be

caused by source imbalance, i.e. variation in the impedance seen by the base of the

transistor cells, DC or bias circuit imbalance, i.e. variation in the quiescent voltage or

current at the base from one cell to the other, or load imbalance, i.e. the load impedance

seen at the collectors of the transistors.

Power imbalance among the output power transistor cells of an amplifier results in some

of the cells delivering a higher portion of the total power output, and thus operating at

higher temperatures. Some other cells deliver less power and operate cooler. This

effectively reduces the active transistor size. This power imbalance leads to higher loss in

the combined device and is reflected in the overall PA performance as reduced efficiency

and lower power handling capability by the amplifier. This effect is seen in the

157

experiment on the LPHP case in this section. Since with power imbalance, a portion of

the device operates at higher temperature, reliability of the device is reduced. This is

reflected in related measures, e.g. imbalanced devices may show shorter mean time to

failure.

Among the methods that help to reduce the power imbalance in larger devices are emitter

and base ballasting [68 and 69], although their primary objective is to avoid thermal

runaway [70, 71 and 72]. The primary mechanism is negative feedback. When power

output or temperature increases in one device, its current increases and causes additional

IR drop in the ballast resistors. This voltage drop in turn tends to reduce the operating

current. The net result of this mechanism is to reduce the current and power output in

that particular cell and thus to push the input impedance of the cell up and cause more of

the RF power input to be directed to other cells. Therefore the ballasts provide most help

in balancing when there are enough cool and low power adjacent cells available to

effectively take the load from the overloaded cells. In more modern amplifiers where cost

considerations further squeeze device sizes and the ballasting resistor values have to be

pushed to minimum acceptable values in the pursuit of higher power efficiency, the

balancing effect of the ballast resistors is decreased and they tend to provide mainly the

protection against thermal runaway. One of the areas open for further research is the

quantifying of the systematic balancing effects of ballast resistors in response to

temperature and power gradient over the power device cells.

The main method for reducing or avoiding power imbalance is to use balancing trees in

the layout between the amplifier stage RF input and the RF input to the bases of the cells,

between the DC bias network and the bias input to the bases of the cells, and between

collectors of the cells and the output load [66]. The cells are the same by construction, are

biased at the same levels (since they see the same impedance and voltage drop on the bias

network), and see the same load impedance on their collectors. They therefore present the

same large signal input impedance in the presence of the same RF signals. Therefore the

balanced tree in the RF input path sees the same impedance looking into the bases of the

cells and therefore delivers the same RF power to each cell. This way all the cells in the

158

output device are expected to have the same power output and loss which yields a perfect

power balancing condition. However, we made a hidden assumption that the temperature

of all the cells is the same. Cells that generate the same heat, i.e. same power loss, will

have the same temperature only when the space around them exhibits the same thermal

behavior, i.e. same thermal conductivity and boundary conditions.

The fixed temperature assumption is valid when in layout the cells are positioned in a

thermally balanced way, or on a thermal balancing tree that makes the junction to

ambient thermal resistances the same. In practice there usually remains some imbalance

at the corners of the output device, but in good layouts most of the cells experience

approximately the same thermal environment. In these scenarios the ballasts may come

into play to further equalize the power levels on adjacent devices that are closer to the

corners. One interesting question that remains open for further research is in amplifiers

driven by power controlled signals with fast changes in power levels. The question is

what are the dynamics of the change in the temperature and power balance over the

device and how does that affect the performance of the amplifier.

Before we conclude let’s look back at Figures 4-14 and 4-15. As mentioned before no

balancing tree was used on the output of these devices. The objective was to save chip

space and associated cost and see if it still has reasonable performance. Note that the

output current is high and therefore the output balancing trees are composed of wide

traces. Therefore they take considerable space on the die. While we fully paid the toll in

the LPHP case, we obtained reasonable, yet sub-optimal efficiency in the LPLP case.

159

Chapter 5: Conclusion Section 5.1 provides a summary of the previous three chapters. We conclude by briefly

discussing future research directions in section 5.2.

5.1 Summary We noticed the need for faster simulation methods and tools for estimating the effects of

power amplifier nonlinearities in the presence of non constant-envelope modulations. In

chapter 2 a very fast method was presented to address this need.

One of the main problems in linear amplifiers is their lower efficiency compared to the

nonlinear amplifiers. Also most of the modern handsets amplifiers have to operate over a

wide range of power levels due to the widely used power control schemes. Very low

efficiency at lower power is therefore an issue for most amplifiers. In chapter 3 PDM,

PWM, PW&AM and nonlinear cascading were presented as four alternative techniques to

the traditional linearization methods with the goal of addressing the need for higher

efficiencies in linear amplifiers both at high and low powers. The first two can

theoretically achieve much higher efficiencies when fast enough and efficient power

devices are available. Unavailability of such devices is the primary limitation on

effectiveness of these techniques at GHz carrier frequencies. PW&AM was presented as

a compromise solution with feasible implementations based on today’s devices. This

technique promises higher efficiencies for linear amplifiers.

In today’s handset applications the need for very low cost and small size in power

amplifiers severely limits the usefulness of any extra design complexity to improve the

performance. Nonlinear cascading was presented as an effective and yet most simple

possible solution to address the need for higher linear efficiency at high powers while

improving the efficiency at lower powers to some extent. As discussed in chapter 2, in

the ideal linear amplifier with saturation a back-off of 3.2 dB is needed to achieve -50dBc

of ACPR in a CDMA amplifier. In chapter 3 we observed measurement results of an

160

amplifier designed based on nonlinear cascading delivering -50dBc ACPR at 3.3dB

back-off. Thus, using today’s existing devices and the nonlinear cascading design

approach, the theoretically achievable linear performance at high powers can be

practically delivered and there is no room left for any additional improvement (at high

powers) through further linearization. Therefore, any further efficiency improvement at

high powers will result from improvements in devices or by the use of higher efficiency

classes of amplifiers. Although at high power, performance (i.e. efficiency while

delivering acceptable linearity) is under control relatively well, there is huge room left for

improvements in efficiency at lower powers as will be discussed in section 5.2.

As discussed in chapter 4 in order to minimize the size and cost of power amplifier

modules a very important step was to eliminate the need for off-chip passives in the

module. Along this line we observed well performing alternative implementations of

output impedance transformation networks based on on-chip capacitors and bond wires

as the inductors. Through this approach the goal of size and cost reduction was achieved.

161

5.2 Future research directions The most important needs for future research in the design of power amplifiers for

handsets must take into account today’s major issues and objectives. Size and cost

reduction in handset PAs is a very common and pressing objective. Integration of PAs for

multiple frequency bands of operation in single packages is another objective. Delivering

higher efficiencies at lower powers is a major performance related improvement that is

most desirable if it can be achieved at very minimal or no additional cost. Integration of

the PAs with the other chips in a transceiver, if possible, is another long term goal. Based

on these objectives we briefly introduce a few research directions that we believe can

have the most valuable impact on the handset power amplifier technology.

5.2.1 Integrated dynamic load adjustment As noted in Chapter 3 efficiency for most of the more practical classes of operation falls

rather sharply when the power output is reduced from the maximum. A question is

whether we can design the amplifier in such a way that its efficiency remains close to the

efficiency when the power output is close to its maximum. The answer is yes. The

maximum efficiency is usually achieved at maximum power output because at that power

the voltage swing across the collector or drain of the device is the largest and also

potentially the current and voltage waveform overlaps are minimized. In fact the output

impedance transformation network is traditionally designed to deliver the correct

impedance at the collector or drain to do so. This is why in designing amplifiers for

different maximum power outputs, different values are used for the elements in output

matching networks. This suggests that if we can switch between different matching

networks, or switch between or tune the component values for different power outputs

while operating an amplifier, the efficiency and linearity performance can be optimized

for different the power levels.

162

Figure 5-1: Power added efficiency with load adjustment

Figure 5-1 shows the hypothetical efficiency curves expected from such an amplifier

designed with three output matching configurations, one for power below 100mW, one

for the 100mW to 200mW range and one for the 200mW to 300mW range. Efficiency at

lower power outputs, such as 50mW, can be doubled using such a scheme compared to a

fixed matching scenario.

The concept of load switching or load tuning for efficiency improvement at low powers is

not new [73]. However in the known implementations so far, there are tunable elements

or switches needed in series or parallel with components in the matching network that can

not be readily integrated on the power amplifier die. Design or use of semiconductor

processes that allow such devices to be integrated on the PA chip, such as the process

suggested in [74], is one approach to solving the problem. Another direction for

exploration is to try to make those tuning or switching functions out of the same kind of

devices used in the majority of today’s handset amplifiers, HBTs to be more specific. A

potentially more valuable yet achievable approach is to use portions of the output power

device to perform the switching or tuning functions as well. That way a die size increase

can be avoided and low cost can be maintained.

163

5.2.2 Sharing output devices for multiple bands Multi-band multi-mode cell phones have been available for several years. In many

phones there is a separate PA module being used for each frequency band. Multi-band

PA modules are available for some applications, and whenever available they are the

more desirable choice. In today’s multi band PA modules there is a separate amplifier for

each frequency band when bands are widely separated. These amplifiers may be

implemented on multiple dies or may be put on a single die.

In handset power amplifier dies, the output stages are the biggest elements on the die and

they take from a third to more than half of the chip space. Therefore a major future step

in reducing the cost and size of multi-band multi-mode PAs would be the sharing of all or

portions of the output device units for different bands or modes of operation. The nature

of the sharing related problems is essentially the same as is expected in load switched

PAs. Therefore some of the solutions developed for simple on chip load switching may

very well find their ways into device sharing multi-band PAs and vice versa.

164

5.2.3 CMOS PAs & breakdown

Most of today’s commercial handset power amplifiers are made of HBT devices on

InGaP processes. The primary reasons are efficiency, reliability and single-supply

operation. InGaP processes are primarily used for power amplifiers and, as a result, these

limited-use specialized processes are 5 to 6 times more expensive than the most widely

used general purpose digital or even RF CMOS processes. Implementing digital or

analog signal conditioning blocks or switches in InGaP HBT processes is not usually

possible and in the few special possible cases is not trivial. This has dictated a need for an

additional CMOS chip in some of the more recent power amplifier modules that contain

more integrated signal conditioning functions such as power control [60]. These translate

into higher cost and bigger size for the power amplifier modules.

Two primary technical limitations have held the industry from being able to effectively

implement handset power amplifiers in CMOS. One is the achievable efficiency of 1 to 3

watt amplifiers in CMOS technology, and the common perception of the achievable

efficiency. The second and perhaps even more critical problem is the low breakdown

voltages of CMOS devices. In traditional handset applications, the power amplifier is

directly connected to the battery to minimize power waste. The phones have to function

while connected to chargers and in that mode the battery voltage can easily go above

4volts. Unfortunately most of today’s CMOS processes that can potentially deliver PA

performance at GHz frequencies have a breakdown of 3 volts or lower.

There has been considerable recent work on improving performance factors such as

efficiency and gain of watt range CMOS RF power amplifiers [75-79]. Results of further

research on design and configurations of single or combinations of devices on standard

processes that can act effectively as high speed power transistors while improving the

breakdown characteristics may lead to enabling technologies for commercial CMOS

PA’s [80-83]. Any future trend in reduction of battery voltages, usage of DC-DC

165

converters for the supply line of the power amplifiers, or specific breakdown avoiding

methodologies like regulators and other protecting circuitry that become active when the

battery voltage is high, e.g. when on charger, may be effective alternative ways to enable

and encourage the usage of CMOS PAs for handsets. However one has to remember that

the total additional savings resulting from using CMOS technology compared to the

InGaP processes is very small and therefore the available budget for potential additional

complexity for breakdown avoiding mechanisms is very limited.

5.2.4 Device layout Output power devices in power amplifiers are very large and occupy substantial chip

area. They usually consist of a large number of smaller transistor cells. There is not much

systematic knowledge available about the effects of relative placement and layout of

these cells on the thermal or electrical behavior of power amplifiers. Most of the existing

designs follow the rather straightforward parallel placement of the cells in a row or

multiple rows or columns [84]. Device size usually is chosen based on rules of thumb

derived from trial and error and based on engineering judgment on the size of fingers and

cells the output transistor layout is planned. From that point on most of the effort is

focused on squeezing the performance (i.e. desired combination of gain, efficiency and

linearity) out of that device through matching, biasing and other design steps. As a result

of this design approach there is no guarantee that better performance cannot be achieved

from power amplifiers built using existing processes.

Any positioning of device cells that can lead to cooler operation of the cells, lower series

resistance of traces or lower emitter or source inductance seen per cell can lead to better

performance. Spreading the cells apart can lead to cooler operation. On the other hand

most compact layouts, such as the constructions suggested in [85 and 86], are most

desirable from the cost point of view. There is no systematic knowledge available on how

to optimize a layout based on these objectives and constraints or on whether any

reasonable improvement, e.g. better efficiency or smaller size, over existing designs can

be achieved through such an optimization.

166

With a potential for increasing the efficiency and power of the devices in the same chip

area, or reducing the area for the same power and efficiencies, one newer set of proposed

candidates includes the compact layout shown in Figure 5-2.

Figure 5-2: A compact layout for a power transistor [86]

These kinds of layouts reduce the emitter inductance by directly connecting the emitter

contacts of the device to a wide sheet of metal going to ground through a connected via,

and therefore may improve gain and efficiency of the stage [86]. Since vias act to block

the heat conduction though the substrate, laying out the device around the via and leaving

the surroundings open promises better overall conduction to take the heat out of the

device and therefore promises cooler operation for the same power outputs compared to

the example of Figure 4-17 in which the vias are to some extent thermally blocking one

167

side of the device. This happens since the substrate itself is the best conductor of heat

found in a power amplifier chip. Vias are holes coated on the sides with very thin layers

of metal for electrically connecting the top and bottom metal plates through the

electrically non-conducting GaAs substrate, and are therefore a relatively nonconducting

area from the thermal point of view compared to the remainder of the substrate. While it

appears that better thermal and electrical performance in smaller die areas should be

achievable using these kinds of devices, quantification of the potential improvements is a

problem left open for further study.

In general, based on today’s knowledge the interactions among layout, thermal and

electrical issues are not effectively and easily quantified. As a result most of the efforts

on layout improvements start with speculative hypotheses that can only be checked by

experimental confirmation. Further systematic research on these interactions can lead to

design of more efficient or more compact power devices and amplifiers using existing

semiconductor processes.

5.2.5 Dynamic interaction of thermal and electrical behaviors There is fairly limited literature available on electro-thermal modeling of power

amplifiers [87-89]. Dynamic interaction of thermal and electrical behaviors and their

effect on the performance of power amplifiers in general and power controlled PA’s in

particular, is an area that is not well understood. Results of further studies in this area

may lead to further effective size (and potentially cost) reduction or power density and

efficiency improvement in power devices.

As mentioned in section 5.2.4 any knowledge obtained through study of these

interactions can lead to more compact device layouts capable of delivering higher

efficiencies at higher power densities. In 5.2.4 we proposed a heuristic search for finding

better devices. Here the objective should be developing a system approach for modeling

power amplifiers. Such models should incorporate device cell positions, thermal

resistances and time constants, power and loss distribution, thermal and electrical ballast

168

effects, and compression curve characteristics as well as the interactions among all these

factors.

More traditional device level modeling approaches attempt to incorporate some of the

specific thermal characteristics into the transistor model and leave the rest of the

modeling job to the designers and circuit simulator. Unfortunately circuit simulators are

not fast enough to allow exploratory simulation of transistor level amplifier models with

extensive behavioral model based electro-thermal cross coupling between elements. The

need for the simulations to cover multiple thermal and power control time constants

while delivering results over fine enough time steps that allow continuity and validity of

the electrical simulations is rendering the task even more difficult.

Therefore, the objective should be obtaining a very detailed yet manageable system level

modeling method for PAs. Provided that such a method is found, fast simulation

programs can be developed to enable systematic study and optimization of layout, load

and source and bias distribution, balancing, ballasting, and thermal response in the

presence of power controlled signals.

For example such a tool can lead to early prediction of potential hot spots in a design by

any designer. This effort can also lead to answers to the following questions: What are

the best layout, ballasting, and balancing methods from a static thermal design point of

view? In a power controlled system with wide dynamic range and continuous change in

average power, how much deviation in the performance is observed compared to

predictions by static models? Is the approach good at predicting signal dependent thermal

memory effects? If so, can those simulation methods be used to optimize design of

amplifiers and pre-distorters jointly to minimize the complications due to memory

effects? Availability of such a tool should lead to PAs with better overall performance.

169

Bibliography [1] Raab, F.H.; Asbeck, P.; Cripps, S.; Kenington, P.B.; Popovic, Z.B.; Pothecary, N.;

Sevic, J.F.; Sokal, N.O.; “Power amplifiers and transmitters for RF and microwave,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 50, Issue: 3, March 2002, Pages:814 – 826.

[2] Cripps, S. C.; RF Power Amplifiers for Wireless Communication. Norwood, MA:

Artech House, 1999. With more on load-pull [3] Raab, F.H.; “Class-E, Class-C, and Class-F power amplifiers based upon a finite

number of harmonics,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 49, Issue: 8, Aug. 2001, Pages: 1462 – 1468.

[4] Krauss, H. L.; Bostian, C. W.; Raab, F. H.; Solid State Radio Engineering. New

York: Wiley, 1980. [5] Raab, F.H.; “Switching transients in class-D RF power amplifiers,” HF Radio

Systems and Techniques, Seventh International Conference on (Conf. Publ. No. 441), 7-10 July 1997, Pages:190 – 194.

[6] Sokal, N.O.; “Class E high-efficiency power amplifiers, from HF to microwave,”

IEEE MTT-S Int. Microwave Symp. Dig., vol. 2, Baltimore, MD, June 7–11, 1998, Pages: 1109–1112.

[7] Colantonio, P.; Giannini, F.; Limiti, E.; Teppati, V.; “An Approach to Harmonic

Load– and Source–Pull Measurements for High-Efficiency PA Design,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 52, Issue: 1, Jan. 2004, Pages: 191 – 198.

[8] Katz, A.; “Linearization: reducing distortion in power amplifiers,” Microwave Magazine, IEEE, Volume: 2, Issue: 4, Dec. 2001, Pages: 37 – 49.

[9] Kenington, P.B.; High Linearity RF Amplifier Design. Norwood, MA: Artech

House, 2000. [10] Kusunoki, S.; Yamamoto, K.; Hatsugai, T.; Nagaoka, H.; Tagami, K.; Tominaga,

N.; Osawa, K.; Tanabe, K.; Sakurai, S.; Iida, T.; “Power-amplifier module with digital adaptive predistortion for cellular phones,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 50, Issue: 12, Dec. 2002, Pages: 2979– 2986.

170

[11] Jaehyok Yi; Youngoo Yang; Myungkyu Park; Wonwoo Kang; Bumman Kim; “Analog predistortion linearizer for high-power RF amplifiers,” Microwave Theory and Techniques, IEEE Transactions on , Volume: 48 , Issue: 12 , Dec. 2000, Pages:2709 – 2713.

[12] “Look-up table techniques for adaptive digital predistortion: a development and

comparison,” Muhonen, K.J.; Kavehrad, M.; Krishnamoorthy, R.; Vehicular Technology, IEEE Transactions on , Volume: 49 , Issue: 5 , Sept. 2000, Pages:1995 – 2002.

[13] Pothecary, N.; Feedforward Linear Power Amplifiers. Norwood, MA: Artech

House, 1999. [14] Parsons, K.J.; Kenington, P.B.; “The efficiency of a feedforward amplifier with

delay loss,” Vehicular Technology, IEEE Transactions on , Volume: 43 , Issue: 2, May 1994, Pages:407 – 412.

[15] Hsieh, C.-C.; “A feedforward S-band MIC amplifier system,” Solid-State

Circuits, IEEE Journal of , Volume: 11 , Issue: 2 , Apr 1976, Pages:271 – 278.

[16] Asbeck, P.; Fallesen, C.; “A polar linearisation system for RF power amplifiers,” Electronics, Circuits and Systems, 2000. ICECS 2000. The 7th IEEE International Conference on , Volume: 1 , 17-20 Dec. 2000, Pages:478 – 481.

[17] Johansson, M.; Mattsson, T.; “Transmitter linearization using Cartesian feedback

for linear TDMA modulation,” Vehicular Technology Conference, 1991. 'Gateway to the Future Technology in Motion', 41st IEEE, 19-22, May 1991, Pages: 439 – 444.

[18] Johansson, M.; “Linearization of RF power amplifiers using Cartesian feedback,”

Tech. Rep., Lund University.

[19] Boloorian, M.; McGeehan, J.P.; “Automatic removal of Cartesian feedback transmitter imperfections,” Communications, IEE Proceedings- , Volume: 144, Issue: 4 , Aug. 1997, Pages:281 – 288.

[20] Dawson, J.L.; Lee, T.H.; “Automatic phase alignment for a fully integrated

cartesian feedback power amplifier system,” Solid-State Circuits, IEEE Journal of, Volume: 38 , Issue: 12 , Dec 2003, Pages: 2269 – 2279.

[21] Carrara, F.; Scuderi, A.; Palmisano, G.; “Wide-bandwidth fully integrated

Cartesian feedback transmitter,” Custom Integrated Circuits Conference, 2003. Proceedings of the IEEE 2003, 21-24 Sept. 2003, Pages: 451 – 454.

[22] Cox, D.; “Linear Amplification with Nonlinear Components,” Communications,

IEEE Transactions on, Volume: 22, Issue: 12, Dec 1974, Pages: 1942 – 1945.

171

[23] Cox, D.; Leck, R.; “Component Signal Separation and Recombination for Linear Amplification with Nonlinear Components,” Communications, IEEE Transactions on, Volume: 23, Issue: 11, Nov 1975, Pages: 1281 – 1287.

[24] Cox, D.; Leck, R; “A VHF Implementation of a LINC Amplifier,”

Communications, IEEE Transactions on, Volume: 24, Issue: 9, Sep 1976, Pages: 1018 – 1022.

[25] Raab, F. H.; “Efficiency of outphasing power-amplifier systems,” IEEE Trans.

Communications, vol. COM-33, Oct. 1985, Pages: 1094–1099.

[26] Conradi, C.P; Johnston, R. H.; McRory, J. G.; “Evaluation of a lossless combiner in a LINC transmitter,” IEEE Canadian Electrical and Computer Engineering Conf., Proceedings of, Edmonton, AB, Canada, May 9–12, 1999, Pages: 105–109.

[27] Xuejun Zhang; Larson, L.E.; Asbeck, P.M.; Langridge, R.A.; “Analysis of power

recycling techniques for RF and microwave outphasing power amplifiers,” Circuits and Systems II: Analog and Digital Signal Processing, IEEE Transactions on, Volume: 49, Issue: 5, May 2002, Pages: 312 – 320.

[28] Asbeck, P.M.; Larson, L.E.; Galton, I.G.; “Synergistic design of DSP and power

amplifiers for wireless communications,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 49, Issue: 11, Nov. 2001, Pages: 2163 – 2169.

[29] Kahn, L.; “Comparison of linear single-sideband transmission with envelope

elimination and restoration single-sideband transmitters,” Proc. IRE, Dec. 1956, Pages: 1706–1712.

[30] Raab, F. H.; Sigmon, B. E.; Myers, R. G.; Jackson, and R. M.; “L-band

transmitter using Kahn EER technique,” IEEE Trans. Microwave Theory Tech., pt. 2, vol. 46, Dec. 1998, Pages: 2220–2225.

[31] Raab, F. H.; “Intermodulation distortion in Kahn-technique transmitters,” IEEE

Trans. Microwave Theory Tech., vol. 44, Dec. 1996, Pages: 2273–2278.

[32] Su, D.K.; McFarland, W.J.; “An IC for linearizing RF power amplifiers using envelope elimination and restoration,” Solid-State Circuits, IEEE Journal of, Volume: 33, Issue: 12 , Dec. 1998, Pages:2252 – 2258.

[33] Staudinger, J.; Gilsdorf, B.; Newman, D.; Norris, G.; Sadowniczak, G.; Sherman,

R.; Quach, T.; “High efficiency CDMA RF power amplifier using dynamic envelope tracking technique,” Microwave Symposium Digest., 2000 IEEE MTT-S International, Volume: 2, 11-16 June 2000, Pages:873 – 876.

[34] Groe, J. B.; Larson, L.E.; CDMA Mobile Radio Design. Norwood, MA: Artech

House, 2000.

172

[35] Van Nee, R.; R. Prasad, R.; OFDM for Wireless Multimedia Communications. Norwood, MA: Artech House, 2000.

[36] Costa, E.; Pupolin, S.; “M-QAM-OFDM system performance in the presence of a

nonlinear amplifier and phase noise,” Communications, IEEE Transactions on, Volume: 50 , Issue: 3 , March 2002, Pages:462 – 472

[37] Sahu, B.; Rincon-Mora, G.A.; “A High-Efficiency Linear RF Power Amplifier

With a Power-Tracking Dynamically Adaptive Buck-Boost Supply,” Microwave Theory and Techniques, IEEE Transactions on , Volume: 52 , Issue: 1 , Jan. 2004, Pages:112 – 120.

[38] Iwamoto, M.; Williams, A.; Pin-Fan Chen; Metzger, A.G.; Larson, L.E.; Asbeck,

P.M.; “An extended Doherty amplifier with high efficiency over a wide power range,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 49, Issue: 12, Dec. 2001, Pages: 2472 – 2479.

[39] Staudinger, J. “Applying switched gain stage concepts to improve efficiency and

linearity for mobile CDMA power amplification,” Microwave J., vol. 43, no. 9, Sept. 2000, Pages: 152–162.

[40] Zhou, G.T.; Kenney, J.S.; “Predicting spectral regrowth of nonlinear power

amplifiers,” Communications, IEEE Transactions on, Volume: 50, Issue: 5, May 2002, Pages:718 – 722.

[41] Gard, K.G.; Gutierrez, H.M.; Steer, M.B.; “Characterization of spectral regrowth

in microwave amplifiers based on the nonlinear transformation of a complex Gaussian process,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 47 , Issue: 7 , July 1999, Pages:1059 – 1069.

[42] Ngoya, E.; Larcheveque, R.; “Envelop transient analysis: a new method for the

transient and steady state analysis of microwave communication circuits and systems,” Microwave Symposium Digest, 1996., IEEE MTT-S International, Volume: 3 , 17-21 June 1996, Pages:1365 – 1368 vol.3.

[43] Kaye, A.; George, D.; Eric, M.; “Analysis and Compensation of Bandpass

Nonlinearities for Communications,” Communications, IEEE Transactions on, Volume: 20 , Issue: 5 , Oct 1972, Pages:965 – 972.

[44] Le Gallou, N.; Ngoya, E.; Buret, H.; Barataud, D.; Nebus, J.M.; “An improved

behavioral modeling technique for high power amplifiers with memory,” Microwave Symposium Digest, 2001 IEEE MTT-S International , Volume: 2, 20-25 May 2001, Pages:983 - 986 vol.2.

173

[45] Ngoya, E.; Le Gallou, N.; Nebus, J.M.; Buret, H.; Reig, P.; “Accurate RF and microwave system level modeling of wideband nonlinear circuits,” Microwave Symposium Digest., 2000 IEEE MTT-S International , Volume: 1 , 11-16 June 2000, Pages:79 - 82 vol.1.

[46] Saleh, A.; “Frequency-Independent and Frequency-Dependent Nonlinear Models

of TWT Amplifiers,” Communications, IEEE Transactions on, Volume: 29, Issue: 11 , Nov 1981, Pages:1715 – 1720.

[47] Borges de Carvalho, N.; Pedro, J.C.; “A comprehensive explanation of distortion

sideband asymmetries,” Microwave Theory and Techniques, IEEE Transactions on , Volume: 50 , Issue: 9 , Sept. 2002, Pages:2090 – 2101.

[48] Cox, D. C.; Soltan, M. F.; Lee, T. H.; “High-efficiency high-power amplifier,”

United States Patent 6,583,679. [49] Boser, B.E.; Wooley, B.A.; “Quantization error spectrum of sigma-delta

modulators,” Circuits and Systems, 1998, IEEE International Symposium on, 7-9 June 1988, Pages: 2331-2334 vol.3.

[50] Keyzer, J.; Hinrichs, J.; Metzger, A.; Iwamoto, M.; Galton, I.; Asbeck, P.;

“Digital generation of RF signals for wireless communications with band-pass delta-sigma modulation,” Microwave Symposium Digest, 2001 IEEE MTT-S International , Volume: 3 , 20-25 May 2001, Pages:2127 - 2130 vol.3.

[51] Yuanxun Wang; “An improved Kahn transmitter architecture based on delta-

sigma modulation,” Microwave Symposium Digest, 2003 IEEE MTT-S International , Volume: 2 , 8-13 June 2003, Pages:1327 - 1330 vol.2.

[52] Berkhout, M.; “An integrated 200-W class-D audio amplifier,” Solid-State

Circuits, IEEE Journal of , Volume: 38 , Issue: 7 , July 2003, Pages:1198 – 1206. [53] Hanington, G.; Chen, P.F.; Radisic, V.; Itoh, T.; Asbeck, P.M.; “Microwave

power amplifier efficiency improvement with a 10 MHz HBT DC-DC converter,” Microwave Symposium Digest, 1998 IEEE MTT-S International, Volume: 2 , 7-12 June 1998, Pages:589 - 592 vol.2.

[54] Saleh, A.A.M.; Cox, D.C.; “Improving the Power-Added Efficiency of FET

Amplifiers Operating with Varying-Envelope Signals,” Microwave Theory and Techniques, IEEE Transactions on , Volume: 83 , Issue:1, Jan 1983, Pages: 51-56.

[55] Kyounghoon Yang; Haddad, G.I.; East, J.R.; “High-efficiency class-A power

amplifiers with a dual-bias-control scheme,” Microwave Theory and Techniques, IEEE Transactions on , Volume: 47 , Issue: 8 , Aug. 1999, Pages:1426 – 1432.

174

[56] McRory, J.G.; Johnston, R.H.; “Self-bias for Class B bipolar transistors,” Electrical and Computer Engineering, 1996. Canadian Conference on, Volume: 2, 26-29 May 1996, Pages:855 - 858 vol.2.

[57] Wei, C.-J.; Sprinkle, S.; Hu, J.T.; Chung, H.-C.; Mitchell, B.; Dicarlo, P.; Bartle,

D.; “Waveform characterization and modeling of dynamic charge behavior of InGaP-GaAs HBTs,” Microwave Symposium Digest, 2001 IEEE MTT-S International , Volume: 2 , 20-25 May 2001, Pages:675 - 678 vol.2.

[58] Teeter, D.A.; East, J.R.; Haddad, G.I.; “Use of self-bias to improve power

saturation and intermodulation distortion in CW class B HBT operation,” Microwave and Guided Wave Letters, IEEE [see also IEEE Microwave and Wireless Components Letters] , Volume: 2 , Issue: 5 , May 1992, Pages:174–176.

[59] Sowlati, T.; Leenaerts, D.; “A 2.4 GHz 0.18 µm CMOS self-biased cascode

power amplifier with 23 dBm output power,” Solid-State Circuits Conference, 2002. Digest of Technical Papers. ISSCC. 2002 IEEE International, Volume: 1, 3-7 Feb. 2002, Pages:294 - 467 vol.1.

[60] Shuyun Zhang; Madic, J.; Bretchko, P.; Mokoro, J.; Shumovich, R.; McMorrow,

R.; “A novel power-amplifier module for quad-band wireless handset applications,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 51 , Issue: 11 , Nov. 2003, Pages:2203 – 2210.

[61] Makioka, S.; Tateoka, K.; Yuri, M.; Yoshikawa, N.; Kanazawa, K.; “A high

efficiency GaAs MCM power amplifier for 1.9 GHz digital cordless telephones,” Microwave and Millimeter-Wave Monolithic Circuits Symposium, 1994. Digest of Papers, IEEE 1994 , 22-25 May 1994, Pages:51 – 54.

[62] Kagaya, O.; Sekine, K.; Yamashita, K.; “Very compact thin power amplifier

module for gigahertz-band cellular systems,” Electronics Letters, Volume: 34, Issue: 3 , 5 Feb. 1998, Pages:244 – 245.

[63] Heo, D.; Sutono, A.; Chen, E.; Suh, Y.; Laskar, J.; “A 1.9-GHz DECT CMOS

power amplifier with fully integrated multilayer LTCC passives,” Microwave and Wireless Components Letters, IEEE [see also IEEE Microwave and Guided Wave Letters] , Volume: 11 , Issue: 6 , June 2001, Pages:249 – 251.

[64] Jos, R.; “Technology developments driving an evolution of cellular phone power

amplifiers to integrated RF front-end modules,” Solid-State Circuits, IEEE Journal of , Volume: 36 , Issue: 9 , Sept. 2001, Pages:1382 – 1389.

[65] Soltan, M. F.; Rategh, H. R.; “Internal impedance match in integrated circuits,”

United States Patent Application, Pub. 20040012457.

175

[66] Spirito, M.; de Vreede, L.C.N.; Nanver, L.K.; Weber, S.; Burghartz, J.N.; “Power amplifier PAE and ruggedness optimization by second-harmonic control,” Solid-State Circuits, IEEE Journal of , Volume: 38 , Issue: 9 , Sept. 2003, Pages:1575 – 1583.

[67] Tang, O.S.A.; Duh, K.H.G.; Liu, S.M.J.; Smith, P.M.; Kopp, W.F.; Rogers, T.J.;

Pritchard, D.J.; “A 560 mW, 21% power-added efficiency V-band MMIC power amplifier,” Gallium Arsenide Integrated Circuit (GaAs IC) Symposium, 1996. Technical Digest 1996, 18th Annual , 3-6 Nov. 1996, Pages:115 – 118.

[68] Gao, G.-B.; Unlu, M.S.; Morkoc, H.; Blackburn, D.L.; “Emitter ballasting resistor

design for, and current handling capability of AlGaAs/GaAs power heterojunction bipolar transistors,” Electron Devices, IEEE Transactions on , Volume: 38 , Issue: 2 , Feb. 1991, Pages:185 – 196.

[69] Liu, W.; Khatibzadeh, A.; Sweder, J.; Hin-Fai Chau; “The use of base ballasting

to prevent the collapse of current gain in AlGaAs/GaAs heterojunction bipolar transistors,” Electron Devices, IEEE Transactions on , Volume: 43 , Issue: 2, Feb. 1996, Pages:245 – 251.

[70] Adlerstein, M.G.; “Thermal stability of emitter ballasted HBT's,” Electron

Devices, IEEE Transactions on , Volume: 45 , Issue: 8 , Aug. 1998, Pages:1653 – 1655.

[71] Jaejune Jang; Kan, E.C.; Dutton, R.W.; Arnborg, T.; “Improved performance and

thermal stability of interdigitated power RF bipolar transistors with nonlinear base ballasting,” Bipolar/BiCMOS Circuits and Technology Meeting, 1997. Proceedings of the , 28-30 Sept. 1997, Pages:143 – 146.

[72] Chih-Hao Liao; Chien-Ping Lee; “Optimum design for a thermally stable

multifinger power transistor with temperature-dependent thermal conductivity,” Electron Devices, IEEE Transactions on , Volume: 49 , Issue: 5 , May 2002, Pages:909 – 915.

[73] Raab, F.H.; “High-efficiency linear amplification by dynamic load modulation,”

Microwave Symposium Digest, 2003 IEEE MTT-S International, Volume: 3 , 8-13 June 2003, Pages:1717 – 1720.

[74] Kumar, M.; Tan, Y.; Sin, J.K.O.; Cai, J.; “An SOI LDMOS/CMOS/BJT

technology for integrated power amplifiers used in wireless transceiver applications,” Electron Device Letters, IEEE , Volume: 22 , Issue: 3 , March 2001, Pages:136 – 138.

[75] Wang, W.; Zhang, Y.P.; “0.18-Micron CMOS Push-Pull Power Amplifier With

Antenna in IC Package,” Microwave and Wireless Components Letters, IEEE, Volume: 14 , Issue: 1 , Jan. 2004, Pages:13 – 15.

176

[76] Shirvani, A.; Su, D.K.; Wooley, B.A.; “A CMOS RF power amplifier with

parallel amplification for efficient power control,” Solid-State Circuits, IEEE Journal of , Volume: 37 , Issue: 6 , June 2002, Pages:684 – 693.

[77] Mertens, K.L.R.; Steyaert, M.S.J.; “A 700-MHz 1-W fully differential CMOS

class-E power amplifier,” Solid-State Circuits, IEEE Journal of, Volume: 37, Issue: 2 , Feb. 2002, Pages:137 – 141.

[78] Aoki, I.; Kee, S.D.; Rutledge, D.B.; Hajimiri, A.; “Fully integrated CMOS power

amplifier design using the distributed active-transformer architecture,” Solid-State Circuits, IEEE Journal of , Volume: 37 , Issue: 3 , March 2002, Pages:371 – 383.

[79] Fallesen, C.; Asbeck, P.; “A 1 W CMOS power amplifier for GSM-1800 with

55% PAE,” Microwave Symposium Digest, 2001 IEEE MTT-S International, Volume: 2 , 20-25 May 2001, Pages:911 – 914.

[80] Mendonca Santos, P.; Castro Simas, M.I.; Lanca, M.; Finco, S.; Behrens, F.H.;

“Breakdown voltage improvement of standard MOS technologies targeted at smart power,” Industry Applications Conference, 1995. Thirtieth IAS Annual Meeting, IAS '95., Conference Record of the 1995 IEEE , Volume: 2 , 8-12 Oct. 1995, Pages:937 - 945 vol.2

[81] Santos, P.M.; Casimiro, A.P.; Lanca, M.; Simas, M.C.; “High-voltage solutions in

standard CMOS,” Power Electronics Specialists Conference, 2001. PESC. 2001 IEEE 32nd Annual, Volume: 1, 17-21 June 2001, Pages: 371– 377.

[82] Mitros, J.C.; Chin-Yu Tsai; Shichijo, H.; Kunz, M.; Morton, A.; Goodpaster, D.;

Mosher, D.; Efland, T.R.; “High-voltage drain extended MOS transistors for 0.18-µm logic CMOS process,” Electron Devices, IEEE Transactions on , Volume: 48, Issue: 8 , Aug. 2001, Pages:1751 – 1755.

[83] Kuo, T.C.; Lusignan, B.; “A 1.5 W class-F RF power amplifier in 0.2 µm CMOS

technology,” Solid-State Circuits Conference, 2001. Digest of Technical Papers. ISSCC. 2001 IEEE International , 5-7 Feb. 2001, Pages:154 - 155, 442.

[84] Jackson, G.; Teeter, D.; Bradford, D.; Cobb, M.; “A scalable MMIC-compatible

power HBT,” Microwave Symposium Digest, 1995., IEEE MTT-S International, 16-20 May 1995, Pages:457 - 460 vol.2.

[85] Tanaka, K.; Mochizuki, K.; Yamada, H.; Takubo, C.; “GaInP/GaAs collector-up

tunnelling-collector heterojunction bipolar transistors with underneath via-hole structure,” Electronics Letters, Volume: 39, Issue: 3 , 6 Feb. 2003, Pages:326-327.

[86] Kiaei, A.; Soltan, M.F..; Rajaei, A.; Rategh, H.R.; “Compact layout for a

semiconductor device,” United States Patent Application, Pub.20040012457.

177

[87] Baureis, P.; “Electrothermal modeling of multi-emitter heterojunction-bipolar-transistors (HBTs),” Integrated Nonlinear Microwave and Millimeterwave Circuits, 1994. Third International Workshop on , 5-7 Oct. 1994, Pages:145 – 148.

[88] Snowden, C.M.; “Characterization and modelling of microwave and millimeter-

wave large-signal device-circuit interaction based on electro-thermal physical models,” Microwave Symposium Digest, 1997., IEEE MTT-S International, Volume: 1 , 8-13 June 1997, Pages:155 - 158 vol.1.

[89] Batty, W.; Panks, A.J.; Johnson, R.G.; Snowden, C.M.; “Electrothermal modeling

and measurement for spatial power combining at millimeter wavelengths,” Microwave Theory and Techniques, IEEE Transactions on, Volume: 47 , Issue: 12 , Dec. 1999, Pages:2574 – 2585.


Recommended