Download as pdf or txt
Download as pdf or txt
You are on page 1of 172

Radio Frequency Switch-mode Power

Amplifiers and Synchronous Rectifiers


for Wireless Applications
by

Sadegh Abbasian

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

THE COLLEGE OF GRADUATE STUDIES

(Electrical Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA


(Okanagan)
October 2015
c Sadegh Abbasian, 2015

Abstract
This thesis focuses on identifying and evaluating device, circuit, and system level issues
that affect the power efficiency of class-D and class-F switch-mode amplifiers, and class-
F synchronous rectifiers. The amplifier and rectifier circuits are used to implement pulse
encoded switch-mode power amplifier systems.
A detailed power efficiency analysis of current mode class-D amplifiers is presented for
variable duty cycle pulse trains. A device model with current saturation in the switch is
introduced and gives insight into how to select an appropriate load line for variable duty cycle
switching conditions. Other new results include the effect of capacitive switching losses which
are usually neglected in current mode amplifiers. The analytical results are compared with
simulation results and confirm that the model can provide good predictions of power efficiency
for a more general class of pulse encoded signals.
Class-F amplifiers are also investigated in this work. The work investigates how input
harmonic matching impedances at the gate affect amplifier power efficiency. Second harmonic
matching is very important and desensitizes the circuit to nonlinear capacitances in the device.
Third harmonic input terminations are much less significant. A comparison of voltage and
current mode circuits is also made and the current mode is better in terms of maximizing
power efficiency. The work is supported by experimental results.
Class-F amplifier circuits are reconfigured into synchronous rectifiers using the theory
of time-reversal duality. Time-reversal duality is usually applied in the context of lossless
circuits and a discussion of how loss impacts the circuit duals is presented. The rectifier dual
always has slightly higher power efficiency and insights into why this occurs are described.
Experimental results are shown for voltage and current mode class-F rectifiers as well as a
wideband current mode class-F rectifier.
The thesis concludes with the analysis and experimental results for an energy recycling
switch-mode power amplifier. A signal splitting network is implemented at the output of the
amplifier and out-of-band power is rectified to enhance the power efficiency of the amplifier.
Experimental results confirm that energy recycling can increase power efficiency. Concluding
remarks based on this research are summarized in the context of how best to use these circuits
for implementing high efficiency amplifiers and rectifiers for wireless applications.

ii
Preface
Some of the research results presented in this thesis have been published before in con-
ference and journal articles. My co-author for these publications was Dr. Thomas Johnson,
my research supervisor, and the relations between the published work and this thesis are
summarized below.

A part of Chapter 2 has been published as a conference paper.

∗ S. Abbasian and T. Johnson, “RF current mode class-D power amplifiers under periodic
and non-periodic switching conditions,” in IEEE International Symposium on Circuits
and Systems (ISCAS), May 2013, pp. 610-613.

Part of Chapter 3 has been published as a journal paper.

∗ S. Abbasian and T. Johnson, “Effect of second and third harmonic input impedances
in a class-F amplifier,” Progress In Electromagnetics Research C, vol. 56, pp. 39-53,
2015.

Parts of Chapter 4 have been published as a journal paper and a conference paper.

∗ S. Abbasian and T. Johnson, “High efficiency GaN HEMT class-F synchronous rectifier
for wireless applications,” IEICE Electronics Express, vol. 12, no. 1, pp. 1-11, 2015.

∗ S. Abbasian and T. Johnson, “High efficiency and high power GaN HEMT inverse
class-F synchronous rectifier for wireless power applications,” in European Microwave
Conference (EuMC), Paris, France, Sep. 2015, pp. 1-3.

iii
Table of Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv

Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

List of Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvi

Chapter 1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Architecture of Switch-Mode Power Amplifier Systems . . . . . . . . . . . . . . 6
1.2.1 Bandpass Sigma-delta Modulation . . . . . . . . . . . . . . . . . . . . . 7
1.2.2 Pulse Position Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Switch-mode Power Amplifier Circuits . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.1 Class-D Power Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.2 Class-E Power Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3.3 Class-F Power Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.3.4 Summary of Amplifier Classes Based on Harmonic Termination Impedances 20
1.4 Power Efficiency and Device Technology . . . . . . . . . . . . . . . . . . . . . . 21
1.5 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.5.1 Class-D Power Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.5.2 Class-F Power Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.5.3 RF Synchronous Rectifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.6 Research Goals and Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.6.1 Predicting the Power Efficiency of CMCD Power Amplifiers for Time
Encoded Input Signals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.6.2 RF Switch-mode Power Amplifiers with Energy Recycling . . . . . . . . 30
1.6.3 RF Rectifier Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.6.4 Class-F and Class-F−1 Power Amplifiers . . . . . . . . . . . . . . . . . . 32
1.6.5 Supporting Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

iv
TABLE OF CONTENTS

1.7 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

Chapter 2: Power Efficiency Analysis of RF Current Mode Class-D Amplifiers 34


2.1 The Current Mode Class-D RF Amplifier . . . . . . . . . . . . . . . . . . . . . 35
2.2 Device Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.2.1 Level 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.2.2 Level 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.2.3 Level 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.3 Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching . . . . 43
2.3.1 Selecting the DC Supply Voltage . . . . . . . . . . . . . . . . . . . . . . 43
2.3.2 Load Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.3.3 Power Loss Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.3.4 Selecting a Device Load Line . . . . . . . . . . . . . . . . . . . . . . . . 47
2.3.5 Current Saturation Model . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2.3.6 Analytical versus Simulated Results . . . . . . . . . . . . . . . . . . . . 51
2.4 Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals . . . . 54
2.4.1 Time Encoded Input Signal . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.4.2 Power Efficiency for 1 T Periodic Signals . . . . . . . . . . . . . . . . . . 54
2.4.3 N T Periodic Signals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.4.4 Power Efficiency Analysis for a 2T Periodic Signal . . . . . . . . . . . . 57
2.4.5 Power Efficiency for 6T Periodic Signals . . . . . . . . . . . . . . . . . . 58
2.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

Chapter 3: Class-F RF Power Amplifiers . . . . . . . . . . . . . . . . . . . . . 61


3.1 Level 3 Device Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.1.1 Bare Die Device Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.2 Class-F Amplifier Simulation Experiments . . . . . . . . . . . . . . . . . . . . . 70
3.2.1 Class-F Amplifier Design . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2.2 Harmonic Input Impedances and Sensitivity to Device Capacitances . . 73
3.3 Class-F Amplifier Experimental Results . . . . . . . . . . . . . . . . . . . . . . 76
3.3.1 Physical Circuit Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.3.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.4 Inverse Class-F Power Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4.1 Design Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.5 Wideband Inverse Class-F Power Amplifier . . . . . . . . . . . . . . . . . . . . 86
3.5.1 Design Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.5.2 Distributed Matching Networks . . . . . . . . . . . . . . . . . . . . . . . 93
3.5.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3.6 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

Chapter 4: Class-F RF Synchronous Rectifiers . . . . . . . . . . . . . . . . . . 99


4.1 The Principle of Time Reversal Duality . . . . . . . . . . . . . . . . . . . . . . 99
4.2 Definitions of Equivalence for Amplifier and Rectifier Duals . . . . . . . . . . . 102
4.3 High Efficiency GaN HEMT Class-F Synchronous Rectifier . . . . . . . . . . . 105
4.3.1 Rectifier Test Bench and Efficiency Definitions . . . . . . . . . . . . . . 105
4.3.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

v
TABLE OF CONTENTS

4.3.3 Class-F Amplifier and Rectifier Power Efficiency Analysis . . . . . . . . 111


4.3.4 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.4 High Efficiency GaN HEMT Inverse Class-F Synchronous Rectifier . . . . . . . 122
4.5 High Efficiency GaN HEMT Wideband Inverse Class-F Synchronous Rectifier . 126
4.6 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

Chapter 5: Switch-mode Power Amplifier with Energy Recycling . . . . . . . 129


5.1 Energy Recycling in Outphasing Power Amplifiers . . . . . . . . . . . . . . . . 129
5.2 Energy Recycling in RF Switch-mode Amplifiers . . . . . . . . . . . . . . . . . 129
5.3 Spectral Shaping to Enhance Energy Recycling Efficiency . . . . . . . . . . . . 131
5.4 Analysis of Power Efficiency Enhancement using Energy Recycling . . . . . . . 134
5.5 Experimental Implementation of a Switch-mode Power Amplifier with Energy
Recycling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.6 Discussion and Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . 140

Chapter 6: Conclusions and Future Work . . . . . . . . . . . . . . . . . . . . . 141


6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Appendix A: Measurement Results for Another Class-F PA . . . . . . . . . . . . . . 154

vi
List of Tables

Table 1.1 Coefficient values for the noise shaping filter HRF (s) . . . . . . . . . . . 9
Table 1.2 Some recently published results for class-D power amplifiers. . . . . . . . 24
Table 1.3 Some recently published results for class-F family PAs . . . . . . . . . . 26
Table 1.4 Some recently published results for RF synchronous rectifiers . . . . . . 28

Table 2.1 Summary of level 2 model values for the Cree CGH60015D die. . . . . . 41
Table 2.2 CMCD Design Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Table 2.3 Duty cycles for generating signals with a period of 6T. . . . . . . . . . . 59

Table 3.1 Level 3 model values for the Cree GaN HEMT (CGH60015D) in the
off-state bias condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Table 3.2 Summary of device capacitances for a Cree GaN HEMT (CGH60015D). 66
Table 3.3 Class-F amplifier designs with input harmonic termination networks. . . 71
Table 3.4 IMN transmission line lengths for a device model with linear capaci-
tances. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Table 3.5 Summary of simulation results for a device model with nonlinear capac-
itances. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Table 3.6 Source and load pull harmonic impedances for the class-F amplifier. . . 78
Table 3.7 Transmission line lengths for load and source matching networks. . . . . 79
Table 3.8 Source and load pull harmonic impedances for the class-F−1 amplifier. . 84
Table 3.9 Microstrip transmission line lengths for the class-F−1 amplifier. . . . . . 85
Table 3.10 Wideband class-F/family amplifier designs comparison. . . . . . . . . . . 87
Table 3.11 Results of the load/source pull simulations for ZLopt and ZSopt . . . . . . 91
Table 3.12 Admittance parameters and extracted values for a third order network. . 91
Table 3.13 Admittance parameters for low-pass network and extracted values for
the final band-pass structure corresponding to the input matching network. 93
Table 3.14 Microstrip transmission line lengths and widths for load and source
matching networks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

Table 4.1 Time reversal relations for circuit components. . . . . . . . . . . . . . . 101


Table 4.2 Some recently published results for RF synchronous rectifiers . . . . . . 105
Table 4.3 Comparison of class-F amplifier and rectifier experimental results. . . . 108
Table 4.4 Comparison of class-F amplifier and rectifier circuit duals. . . . . . . . . 120
Table 4.5 Comparison of class-F−1 amplifier and rectifier experimental results. . . 123

vii
List of Figures

Figure 1.1 RF switch-mode power amplifier architecture with energy recycling.


The figure also serves as a roadmap for this thesis. . . . . . . . . . . . . 2
Figure 1.2 A class-AB power amplifier. . . . . . . . . . . . . . . . . . . . . . . . . 3
Figure 1.3 Efficiency as a function of conduction angle for conventional power
amplifiers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Figure 1.4 Efficiency and output power as a function of input power for a class-AB
amplifier with a conduction angle of 244◦ (θ = 1.36π). . . . . . . . . . . 5
Figure 1.5 Block diagram of a SMPA with encoder and reconstruction filter. . . . 6
Figure 1.6 Block diagram of a bandpass sigma-delta modulator. . . . . . . . . . . 7
Figure 1.7 Power spectrum of a bandpass sigma-delta modulator. . . . . . . . . . . 7
Figure 1.8 A bandpass sigma-delta modulator pulse train in the time domain. . . 8
Figure 1.9 Block diagram of pulse position modulator. . . . . . . . . . . . . . . . . 9
Figure 1.10 A noise shaped PPM pulse train in the time domain. . . . . . . . . . . 9
Figure 1.11 A noise shaped PPM pulse train in the frequency domain. . . . . . . . 10
Figure 1.12 Schematic of a VMCD power amplifier. . . . . . . . . . . . . . . . . . . 11
Figure 1.13 VMCD amplifier voltage and current waveforms for a periodic drive
signal with a duty cycle of 30%. . . . . . . . . . . . . . . . . . . . . . . 12
Figure 1.14 Schematic of a CMCD. . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Figure 1.15 CMCD current and voltage waveforms for a periodic input pulse train
with a duty cycle of 30%. . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Figure 1.16 Schematic of a class-E power amplifier. . . . . . . . . . . . . . . . . . . 15
Figure 1.17 Class-E voltage and current waveforms for a 30% duty cycle pulse train. 16
Figure 1.18 Normalized voltage across the switch in a class-E PA for three different
duty cycles: 30% (circle), 50% (star) and 70% (square). . . . . . . . . 17
Figure 1.19 A class-F power amplifier circuit. . . . . . . . . . . . . . . . . . . . . . 18
Figure 1.20 Class-F power amplifier voltage and current waveforms. . . . . . . . . . 19
Figure 1.21 Amplifier classes in terms of harmonic load impedances. . . . . . . . . . 20
Figure 1.22 Band gap energy and saturated velocity for Si, GaAs and GaN. . . . . 21
Figure 1.23 GaN HEMT technology: packaged device from Cree (left) and MMIC
(right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

Figure 2.1 A CMCD circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35


Figure 2.2 Typical DC IV characteristics for a GaN device. . . . . . . . . . . . . 36
Figure 2.3 Schematic for current mode class-D power amplifier with a level 1 device
model (ADS schematic). . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Figure 2.4 Level 1 CMCD current and voltage waveforms for a 30% duty cycle
periodic drive signal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

viii
LIST OF FIGURES

Figure 2.5 Transition time (τ = 0.15T ) for a CMCD with Cree CGH60015D tran-
sistors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Figure 2.6 Level 2 device model for a CMCD amplifier. . . . . . . . . . . . . . . . 42
Figure 2.7 S-parameters for the on and off state for a Cree CGH60010D device. . 42
Figure 2.8 Signal with period T and variable duty cycle (α). . . . . . . . . . . . . 43
Figure 2.9 Device current waveforms at the drain terminal of the switching device.
The ADS simulation results are for a Cree large signal device model. . 45
Figure 2.10 Overlap of drain current and drain voltage waveforms in a CMCD am-
plifier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Figure 2.11 DC IV operating region for a CMCD amplifier including margin for
duty cycle variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Figure 2.12 Efficiency and output power of a CMCD as a function of load resistance
(α = 0.5). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Figure 2.13 Efficiency and output power of a CMCD as a function of load resistance
(α = 0.3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Figure 2.14 DC device current as a function of sin(απ) (α is duty cycle). . . . . . 51
Figure 2.15 Losses in the CMCD amplifier as a function of duty cycle. The overlap
period τ is 0.1 T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 2.16 Drain efficiency of CMCD amplifier as a function of duty cycle. . . . . 53
Figure 2.17 Signal with period 1T. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Figure 2.18 Comparison of power efficiency for a CMCD amplifier with a 1 T peri-
odic signal and SDM non-periodic signal. . . . . . . . . . . . . . . . . 55
Figure 2.19 Drain efficiency as a function of modulator drive level for SDM and
PPM encoders. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Figure 2.20 Signal with period 2T. . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Figure 2.21 A 6T signal with a zero mean DC component. . . . . . . . . . . . . . . 57
Figure 2.22 Drain efficiency of CMCD amplifier as a function of duty cycle when
driven with a 2T periodic signal. . . . . . . . . . . . . . . . . . . . . . . 58
Figure 2.23 CMCD amplifier power efficiency for periodic (1T , 2T , and 6T ) and
non-periodic pulse trains (SDM and PPM). . . . . . . . . . . . . . . . . 59

Figure 3.1 Level 3 equivalent circuit model for GaN HEMT Cree CGH60015D
[reproduced courtesy of The Electromagnetics Academy]. . . . . . . . . 62
Figure 3.2 Equivalent circuit model for off-state bias conditions. . . . . . . . . . . 63
Figure 3.3 Z-parameters for the level 3 device model (symbols) and for the large
signal device model (solid lines) for the off-state bias condition. . . . . 64
Figure 3.4 Y -parameters for the level 3 device model (symbols) and for the large
signal device model (solid lines) for the off-state bias condition. . . . . 65
Figure 3.5 Extracted intrinsic device capacitances for the Cree GaN HEMT (CGH60015D):
(a) drain-source capacitance, (b) gate-source capacitance, (c) gate-
drain capacitance, and (d) gate-source capacitance versus gate-source
voltage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Figure 3.6 Device model for packaged die [reproduced courtesy of The Electro-
magnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Figure 3.7 Comparison of Y11 parameters for the level 3 model including the pack-
age (symbols) and the GaN HEMT Cree large signal model for the
CGH40010F (solid line). The device bias conditions are in the off-state. 68

ix
LIST OF FIGURES

Figure 3.8 Comparison of Y12 parameters for the level 3 model including the pack-
age (symbols) and the GaN HEMT Cree large signal model for the
CGH40010F (solid line). The device bias conditions are in the off-state. 69
Figure 3.9 Comparison of Y22 parameters for the level 3 model including the pack-
age (symbols) and the GaN HEMT Cree large signal model for the
CGH40010F (solid line). The device bias conditions are in the off-state. 69
Figure 3.10 Schematic for the class-F PA [reproduced courtesy of The Electromag-
netics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Figure 3.11 Output matching network (OMN) structure [reproduced courtesy of
The Electromagnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . 73
Figure 3.12 Spectrum for the case where Cgd = 0 pF: drain voltage (left) and gate
voltage (right) [reproduced courtesy of The Electromagnetics Academy]. 73
Figure 3.13 Spectrum for the case where Cgd = 0.36 pF: drain voltage (left) and
gate voltage (right) [reproduced courtesy of The Electromagnetics Academy]. 74
Figure 3.14 Input matching network circuits: (a) Design 1, (b) Design 2 and (c)
Design 3 [reproduced courtesy of The Electromagnetics Academy]. . . . 74
Figure 3.15 Simulated drain efficiency as a function of second harmonic level for
a device model with linear capacitances [reproduced courtesy of The
Electromagnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . 77
Figure 3.16 Schematic of the class-F power amplifier with output and input match-
ing circuits and bias networks [reproduced courtesy of The Electromag-
netics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Figure 3.17 Simulated drain voltage and drain current waveforms (left) and gate
voltage and drain current waveforms (right)[reproduced courtesy of The
Electromagnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . 79
Figure 3.18 Photograph of the 10 W class-F power amplifier. . . . . . . . . . . . . . 80
Figure 3.19 The class-F amplifier test bench. . . . . . . . . . . . . . . . . . . . . . . 80
Figure 3.20 Measured and simulated drain efficiency and output power as a func-
tion of input power for a CW test signal [reproduced courtesy of The
Electromagnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . 81
Figure 3.21 Measured and simulated drain efficiency and output power as a func-
tion of frequency for a CW test signal [reproduced courtesy of The
Electromagnetics Academy]. . . . . . . . . . . . . . . . . . . . . . . . . 82
Figure 3.22 Measured output spectrums for a WCDMA signal at three different
output power levels: (a) 35.1 dBm (b) 33.4 dBm and (c) 31.6 dBm
[reproduced courtesy of The Electromagnetics Academy]. . . . . . . . . 82
Figure 3.23 Measured drain efficiency and ACLR as a function of output power
for a WCDMA signal [reproduced courtesy of The Electromagnetics
Academy]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Figure 3.24 Schematic of the class-F−1 PA with output and input matching circuits
and bias networks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Figure 3.25 Simulated drain voltage (solid) and drain current (dash) waveforms for
the class-F−1 power amplifier. The waveforms are shown for the drain
terminal of the packaged device. . . . . . . . . . . . . . . . . . . . . . . 86
Figure 3.26 Photograph of the class-F−1 power amplifier. . . . . . . . . . . . . . . . 86
Figure 3.27 Drain efficiency and output power as a function of input power for the
fabricated class-F−1 PA. . . . . . . . . . . . . . . . . . . . . . . . . . . 87

x
LIST OF FIGURES

Figure 3.28 Different steps for design of a lumped element matching network: (a)
low-pass network with normalized admittances gj ; (b) bandpass match-
ing network; (c) Norton transformation to increase output impedance. . 90
Figure 3.29 Impedance and frequency scaled lumped element lowpass network for
synthesizing a wideband output match. . . . . . . . . . . . . . . . . . . 91
Figure 3.30 Lumped element output network after applying a lowpass to bandpass
transformation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Figure 3.31 Impedance transformed output network (top) and the corresponding
Norton transformation to create the impedance transformation (bottom). 92
Figure 3.32 Bandpass output matching network with an impedance transformer to
match the output to 50 Ω. . . . . . . . . . . . . . . . . . . . . . . . . . 92
Figure 3.33 Bandpass input matching network by Norton transformation (nT =
1.1079). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Figure 3.34 Dividing the capacitance C1 into three parallel capacitances to reform
the output structure as a distributed network. . . . . . . . . . . . . . . 94
Figure 3.35 Equivalent transmission line circuit for a shunt resonator. . . . . . . . . 94
Figure 3.36 Distributed output matching network. . . . . . . . . . . . . . . . . . . . 94
Figure 3.37 Distributed input matching network. . . . . . . . . . . . . . . . . . . . 94
Figure 3.38 The fundamental frequency impedances of the input and output match-
ing networks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Figure 3.39 The second harmonic impedances of the input and output matching
networks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Figure 3.40 Wide frequency range sweep of the impedances of the output matching
network. Fundamental, second harmonic and third harmonic frequency
ranges are shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Figure 3.41 Photograph of the wideband class-F−1 power amplifier. . . . . . . . . . 97
Figure 3.42 Measured and simulated drain efficiency of the wideband class-F−1 PA
as a function of frequency for a CW test signal. . . . . . . . . . . . . . 97

Figure 4.1 (a) Network N and its current and voltage, (b) network Nvc , a voltage
and current dual of N , and (c) network Ntr , a time reversal dual of N . 100
Figure 4.2 The direction of energy flow in a network and its TR dual. . . . . . . . 101
Figure 4.3 Block diagrams of (a) a power amplifier, (b) synchronous rectifier dual,
and (c) synchronous rectifier with feedback to provide gate drive. . . . 103
Figure 4.4 Dynamic load lines for: (a) a class-F amplifier (b) a class-F rectifier. . . 104
Figure 4.5 The class-F rectifier test bench. . . . . . . . . . . . . . . . . . . . . . . 106
Figure 4.6 Measured RF to DC conversion efficiency and output DC power as a
function of load resistance for a class-F rectifier. . . . . . . . . . . . . . 109
Figure 4.7 Measured power efficiency and output DC power as a function of RF
input power for a class-F rectifier. . . . . . . . . . . . . . . . . . . . . . 109
Figure 4.8 Measured power efficiency and output DC power as a function of fre-
quency for a class-F rectifier. . . . . . . . . . . . . . . . . . . . . . . . . 110
Figure 4.9 Class-F rectifier power efficiency with and without mismatch loss. . . . 110
Figure 4.10 A class-F power amplifier with a series quarterwave transmission line. . 111
Figure 4.11 Drain voltage and drain current waveforms for: (a) a class-F power
amplifier and (b) a class-F rectifier. . . . . . . . . . . . . . . . . . . . . 112

xi
LIST OF FIGURES

Figure 4.12 Drain voltage and drain current waveforms with overlap loss for the
class-F amplifier and rectifier duals. . . . . . . . . . . . . . . . . . . . . 115
Figure 4.13 Estimated drain efficiency of class-F PA and rectifier as a function of
output capacitance. Ron for both the amplifier and rectifier are 2.2 Ω. . 119
Figure 4.14 Predicted losses in a class-F power amplifier as a function of output
capacitance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Figure 4.15 Predicted losses in a class-F rectifier as a function of output capacitance.121
Figure 4.16 Test bed for the class-F−1 rectifier. A class-F amplifier is used as a
high power RF input source. . . . . . . . . . . . . . . . . . . . . . . . . 122
Figure 4.17 Measured RF to DC conversion efficiency and output DC power versus
load resistance for the rectifier. The measurements conditions are for
an input RF source power of 40.76 dBm at a frequency of 910 MHz. . . 124
Figure 4.18 Measured power efficiency and output power as a function of frequency
for the rectifier. The measurements conditions are for an input RF
source power of 40.76 dBm. . . . . . . . . . . . . . . . . . . . . . . . . . 124
Figure 4.19 Power efficiency comparison of a class-F and class-F−1 synchronous
rectifier. Experimental results are shown. . . . . . . . . . . . . . . . . . 125
Figure 4.20 Measured drain efficiency as a function of frequency for the wideband
class-F−1 rectifier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Figure 4.21 Measured drain efficiency as a function of input power for the wideband
class-F−1 rectifier at frequencies of 650 MHz, 850 MHz and 1050 MHz. 127
Figure 4.22 Measured drain efficiency as a function of input power for class-F, class-
F−1 and wideband class-F−1 synchronous rectifiers. . . . . . . . . . . . 127

Figure 5.1 Outphasing amplifiers: (a) reactive signal combining and (b) isolated
signal combining with energy recycling. . . . . . . . . . . . . . . . . . . 130
Figure 5.2 Switch-mode power amplifiers (a) with reactive output filter and (b)
with energy recycling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Figure 5.3 Block diagram of a noise shaped PPM encoder with dither (top) and
example input and output waveforms (bottom). . . . . . . . . . . . . . 132
Figure 5.4 Power spectrum of a noise shaped PPM signal without out-of-band
spectral shaping (top) and with spectral shaping (bottom). . . . . . . . 133
Figure 5.5 Block diagram of a power amplifier with energy recycling. . . . . . . . . 134
Figure 5.6 Examples of amplifier power efficiency with energy recycling. . . . . . . 137
Figure 5.7 Test bed with a class F amplifier and a class-F−1 rectifier to recover
out-of-band energy. The system implements the block diagram shown
in Figure 5.2(b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Figure 5.8 Measured in-band and recovered power as a function of the coding
efficiency for encoder of the noise shaped PPM modulator. . . . . . . . 138
Figure 5.9 Measured drain efficiencies with and without energy recycling. . . . . . 139

Figure A.1 Measured drain efficiency as a function of input power for a CW test
signal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Figure A.2 Measured drain efficiency as a function of frequency for a CW test signal.155
Figure A.3 Measured output spectrums for a WCDMA signal at three different
output power levels: (a) 34.2 dBm (b) 32.4 dBm and (c) 30.5 dBm . . 155

xii
LIST OF FIGURES

Figure A.4 Measured drain efficiency and ACLR as a function of output power for
a WCDMA signal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

xiii
Acknowledgements
First of all, I would like to offer my sincere gratitude to my supervisor, Dr. Thomas
Johnson who has inspired me to continue my work in this field, and thanks to him for his
patient guidance and generous support throughout my studies.
I would like to thank Dr. Stephen O’Leary and Dr. Wilson Eberle for their support during
the past few years as members of my supervisory committee. I would also like to thank all the
professors in the electrical engineering department who contributed to my learning through
lectures and classes.
I would like to appreciate Dr. Fadhel Ghannouchi from the university of Calgary and Dr.
Homayoun Najjaran as members of my examining Committee.
I offer my gratitude to Dr. Andrew Labun with whom I spent a great time doing research
with in the summer of 2010. I also would like to thank my colleagues especially Dr. Ali
Tirdad and Mr. Saimoom Ferdous for their support throughout the work.
I offer my best and special thanks to my parents and family who have supported me with
their unconditional love throughout my years of education.

xiv
Dedication
I dedicate this thesis to my parents and family who have supported me throughout my
years of education.

xv
List of Acronyms
ADS Advanced Design System

ACLR Adjacent Channel Leakage Ratio

CMCD Current-Mode Class D

CE Codding Efficiency

CW Continuous Wave

GaN Gallium Nitride

HEMT High Electron Mobility Transistor

HSPA Hard Switched Power Amplifier

JFOM Johnson’s Figure of Merit

MMIC Monolithic Microwave Integrated Circuit

MSE Mean Square Error

PA Power Amplifier

PPM Pulse Position Modulation

PWM Pulse Width Modulation

RF Radio Frequency

SDM Sigma Delta Modulation

SMPA Switch-Mode Power Amplifier

TL Transmission Line

TR Time Reversal

VMCD Voltage-Mode Class D

ZVS Zero Voltage Switching

xvi
Chapter 1

Introduction
This thesis is about the design and implementation of high efficiency amplifiers and rec-
tifiers for wireless applications. The work is motivated by interest to improve the power
efficiency of transmit amplifiers in mobile devices like smartphones and basestations. The
transmit power amplifier (PA) is a large signal stage in the transceiver and the power con-
sumption of the amplifier circuit is a significant portion of the total power consumed by the
equipment. Reduced energy consumption improves battery utilization in mobile devices and
decreases electric utility costs for operating basestations.
The motivation to improve power efficiency in high power radio frequency (RF) amplifiers
has led to a shift from analog modes of amplification to digital modes of amplification. The
allure of the ‘digital’ amplifier is based on the concept that if an amplifying device is operated
as a switch instead of as an analog amplifier, then the power dissipation in the amplifying
device is significantly reduced. If device dissipation is reduced, then the overall amplifier
power efficiency is increased because more of the DC supply power is converted to RF power.
An amplifier that is designed to operate the amplifying device as a switch, is called a
switch-mode amplifier. Examples of switch-mode amplifier circuits are class-D and class-E.
The term ‘switch-mode’ can also be extended to circuits where the waveforms are designed to
minimize overlap losses, and in the limit of no overlap loss, the circuit operation is equivalent
to a switch. Class-F is an example of circuits which use waveform shaping to minimize device
dissipation.
In theory, very high power efficiencies are possible in switch-mode amplifiers providing
losses are minimal. Unfortunately, at high frequencies, ideal switching is not realizable with
current device technology, and losses can be very significant. Therefore, switch-mode power
amplifier circuits in the GHz frequency range are very challenging to implement. The designer
is forced to carefully evaluate circuits and seek to understand what limits performance and
determine how to overcome these limitations.
The goal of this research is to gain insight into ways to improve the power efficiency of
RF switch-mode amplifiers. The work can be divided into four main topics. The first topic
is an analytic study of power losses in class-D amplifiers for arbitrary duty cycle pulse trains.
The second topic is related to a study of the relationship between input harmonic impedance
and power efficiency for class-F amplifiers. The class-F work also includes experimental mea-
surements for three different types of class-F amplifiers and conclusions are made on the best
circuit topology for the highest power efficiency. The third topic is the analysis and imple-
mentation of a new switch-mode power amplifier system that employs energy recycling as an
efficiency enhancement technique. The experimental implementation of the energy recycling
amplifier led to a fourth topic which was the design of high efficiency RF rectifiers required
to convert RF power into DC power. The rectifier designs are closely linked to the design of
switch-mode power amplifiers and a design methodology based on the theory of time-reversal
duality has been used. An overview of the thesis is given in Figure 1.1.

1
Energy Recovery Class-F/Family
System Synchronous
Rectifiers
Chapter 5 Chapter 4

Bandstop Power
Spectrum

VDD RF to DC
Conversion

Out-of-band Power
A Signal
Encoder SMPA
RF Separator In-band Power
Modulated
Signal
Chapter 1. Introduction

Power Spectrum Bandpass Power


at node A Spectrum

PPM/SDM
with Chapter 2
Class-D−1
Spectral Shaping
Chapter 5 Class-F/Family Chapter 3

2
Figure 1.1: RF switch-mode power amplifier architecture with energy recycling. The figure also serves as a roadmap for this thesis.
1.1. Background

1.1 Background
The quest to design power efficient amplifiers has a long history. Two examples are the
Doherty amplifier [1] and the Chireix outphasing amplifier [2] both which were patented in
the early 1930’s. As a testament to this early work, the theory remains in widespread use
today and work continues to improve these types of amplifiers.
A study of power amplifiers nearly always begins with classical (conventional) transcon-
ductance amplifiers which are defined by the current conduction angle in the amplifying
device. These amplifiers are called class-A (100% current conduction), class-AB (typically
around 70% current conduction), class-B (50% current conduction) and class-C (less than
50% current conduction). An example of a transconductance amplifier circuit is shown in
Figure 1.2. The current conduction in the transistor is dependent on the gate bias and the
load line for the device. The relationships between conduction angle, output power, and
power efficiency are shown in Figure 1.3 [3]. The class-AB operating point is commonly used
in wireless communication applications as it provides a good balance between power efficiency
and distortion.
CB
vd (θ)
Vout
d
id (θ)
LDD RL
L0 C0
Vg M1
IDC
VDD

Figure 1.2: A class-AB power amplifier.

One of the drawbacks of the class-AB amplifier is that power efficiency is amplitude
dependent. The amplitude dependence is most commonly shown as a relationship between
input power and power efficiency. An example for a class-AB amplifier is shown in Figure 1.4.
The plot shows that maximum efficiency is obtained at maximum power which corresponds
to the maximum input amplitude to the amplifier. Power efficiency decreases as the input
amplitude (input power) is reduced. The term back-off refers to how much the input power
is backed off from the peak output power which the amplifier can deliver. For the class-
AB amplifier shown in Figure 1.4, the peak efficiency at the 1 dB compression point (0 dB
back-off) is 63%, while the power efficiency at 6 dB back-off is 38%, much less than peak
efficiency.
The amplitude response shown in Figure 1.4 is an example of the how the amplifier would
respond to an unmodulated signal. An unmodulated signal, also called a continuous wave
(CW) signal, is a sinusoidal signal with a constant frequency (fc ), constant amplitude (A)
and phase (φo ):

scw (t) = A cos(2πfc t + φo ). (1.1)

Although useful for characterizing amplifiers, a CW signal does not carry information, and in
a communication system application, modulation is added to the carrier. More generally, any

3
1.1. Background

Figure 1.3: Efficiency as a function of conduction angle for conventional power amplifiers.

communication signal which can be transmitted through a physical medium can be expressed
as

s(t) = r(t) cos[2πfc t + φ(t)] where r(t) ≥ 0. (1.2)

The signal envelope, r(t), adds information to the carrier by amplitude modulation (AM),
while the phase term, φ(t), adds information by phase modulation (PM).
Both the AM and PM components in a communication signal can be problematic for power
amplifiers. Since power efficiency in an amplifier is amplitude dependent, the average power
efficiency of the amplifier depends on the statistical distribution of the envelope variation
in the signal. Most signals have a peak amplitude that occurs infrequently and the average
envelope amplitude is typically much smaller than the peak amplitude. The measure most
commonly used to quantify amplitude variation in the signal is called the peak to average
power ratio (PAPR):
 
peak signal power
PAPR (dB) = 10 log10 . (1.3)
average signal power

Many common wireless communication signals have a PAPR in the range of 6-10 dB.
The implication of signals with high PAPR is that, on average, the power amplifier operates
in a back-off state approximately equal to the PAPR of the signal. Therefore, when a class-
AB amplifier is used to amplify a signal with a 6 dB PAPR, the average efficiency of the
amplifier is approximately equal to the efficiency at 6 dB back-off. For the class-AB amplifier
example shown in Figure 1.4, the average power efficiency of the amplifier for a 6 dB PAPR
signal would therefore be approximately 38%, much less than peak efficiency which is 63%.
Although a more accurate estimate of the average efficiency is obtained by integrating the
CW response characteristics over the amplitude probability distribution of the input signal,

4
1.1. Background

70 44
Designed Class−AB PA
Using Cree CGH60015D
60 42

Output Power (dBm)


Ideal Class−AB PA
1 dB
Efficiency (%) 50 40
1 dB
Compression Point
40 38

30 36

6 dB Back−off Mode
20 34
5 6 7 8 9 10 11 12 13 14 15 16
Input Power (dBm)

Figure 1.4: Efficiency and output power as a function of input power for a class-AB amplifier
with a conduction angle of 244◦ (θ = 1.36π).

the estimate using PAPR is a very useful approximation. Consequently, the average power
efficiency of a class-AB amplifier is much less than the peak power efficiency when amplifying
typical communication signals.
Another important characteristic of power amplifiers is distortion. Distortion is created by
nonlinear amplitude and phase characteristics in the amplifier. Under large signal conditions,
the amplifier saturates leading to amplitude compression. Under small amplitude conditions,
the amplifier may enter cut-off depending on the bias point of the amplifying device. Any
deviation from linear amplitude characteristics results in distortion that appears in the output
signal. Distortion can also be generated from phase distortion that may be both amplitude
and frequency dependent.
Because conventional transconductance amplifiers have amplitude dependent power effi-
ciency characteristics, the key to implementing a high efficiency amplifier is to devise circuits
whose power efficiency has reduced sensitivity to amplitude variation. There are different
approaches to this problem. One method is to implement parallel signal paths which work
together to reduce amplitude sensitivity. Examples of this method include Doherty [1] and
Chireix outphasing [2] techniques. Another approach is to create signalling and circuits that
maintain a saturated operating point in the amplifier. Examples of these methods include
envelope tracking [4] and switch-mode power amplifier techniques [5]. In this research, the
focus is on the latter method and new analytic and experimental results are presented for
switch-mode power amplifiers (SMPAs).

5
1.2. Architecture of Switch-Mode Power Amplifier Systems

1.2 Architecture of Switch-Mode Power Amplifier Systems


In a switch mode power amplifier (SMPA), the active device is used as a switch instead
of a linearly controlled current source. When the switch is on, the voltage across the switch
is low and current is high, and when the switch is open, the voltage is high and the current is
low (ideally zero). The switching action theoretically leads to 100% efficiency if the switches
are ideal, because the dissipation in the device, equal to the product of the current times
voltage, is zero. However, practical devices, especially at high frequencies, have capacitance,
inductance, and finite on and off state resistances that all lead to dissipation in the device
which degrades power efficiency.
Assuming that a high efficiency switch-mode amplifier can be implemented, the drawback
of switch-mode operation is that a modulated signal with an AM signal envelope cannot
be directly amplified because the amplitude states created by the switching action quantize
the output amplitude. For a switch-mode amplifier with two amplitude states, the output
amplitude is binary, and the only information which can be conveyed to the output signal
is the timing of level crossings. Therefore, if the high efficiency operation of a switch-mode
amplifier is to be utilized in a wireless communication application, additional circuit blocks
must be added to the amplifier system to map the modulated input signal into a pulse train
and to reconstruct the original modulated source after amplifying the pulse train. A block
diagram of the amplifier system is shown in Figure 1.5.

Figure 1.5: Block diagram of a SMPA with encoder and reconstruction filter.

Signal reconstruction in a switch-mode power amplifier is constrained by the types of


circuit elements that can be used in a high power radio frequency output stage. Almost
universally, signal reconstruction is implemented with a bandpass filter, and this in turn
imposes design constraints on the type of signal mapping which can be used to implement the
pulse encoder. By quantizing the modulated signal to binary amplitude levels, a large amount
of quantization noise is added by the pulse encoder. The quantization noise is spread over a
very wide bandwidth and signal encoders implement methods to shape the noise spectrum and
create a narrow region of high signal to noise ratio (SNR) where the source signal is placed.
Examples of compatible source encoding techniques include bandpass sigma-delta modulation
(SDM) [6, 7] and noise shaped pulse position modulation (PPM) [7, 8]. More generally, SDM
and PPM are examples of a larger signal set called time encoded signals which are a class of
signals that convey information in the timing of the zero-crossings (amplitude transitions). In
the following sections, a brief summary of SDM and PPM pulse encoding methods are given.

6
1.2. Architecture of Switch-Mode Power Amplifier Systems

1.2.1 Bandpass Sigma-delta Modulation


Bandpass sigma-delta modulation has been proposed by many researchers as one way of
implementing a pulse encoder for switch-mode power amplifiers [7, 6]. A block diagram of a
bandpass SDM is shown in Figure 1.6. The input to the encoder is an RF modulated source
signal and the output is a quantized two level pulse train. The quantization process generates
significant quantization noise that is shaped by a loop filter HRF (s). The noise shaping filter
creates a noise well where the RF input signal spectrum is placed. The noise well is called the
in-band spectrum and the broadband noise is called the out-of-band spectrum. An example
of the output spectrum from a SDM encoder is shown in Figure 1.7 and the corresponding
time domain signal is shown in Figure 1.8.

Noise Clock (fs ) Sampling


Shaping Filter Quantizer

s(t) p(t)
HRF (s)

Figure 1.6: Block diagram of a bandpass sigma-delta modulator.

Modulator Output Power Spectrum; 0.25Tc PA Output


20 20

0 0

-20 -20
Power [dBm]

Power [dBm]
-40 -40

-60 -60

-80 -80

-100 -100
600 700 800 900 1000 1100 1200 1300 1400 600 700 800 900
Frequency [MHz] Frequ

Figure 1.7: Power spectrum of a bandpass sigma-delta modulator.

The quantizer in a bandpass SDM is triggered by a sampling clock with a frequency fs .


The sampling clock is typically selected to be at least above the Nyquist rate of the carrier
frequency, which means the complex envelope is oversampled. For example, a 1 GHz wideband
code division multiple access (WCDMA) modulated input signal with a bandwidth of 10 MHz
sampled by a 3.4 GHz clock, has an envelope over sample ratio of 170 and a carrier oversample
ratio of 1.7.1 The signal to noise ratio of the reconstructed load signal which is determined by
1
In SDM theory, oversample ratio is usually defined relative to the Nyquist sample rate. For example,

7
1.2. Architecture of Switch-Mode Power Amplifier Systems

p(t): Sigma-delta pulse trains


0.8

0.6

0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1

1000 1001 1002 1003 1004 1005 1006


Time (ns)

Figure 1.8: A bandpass sigma-delta modulator pulse train in the time domain.

the envelope oversample ratio. Therefore, a high envelope oversample ratio is required; the
carrier oversample ratio can also affect the signal to noise ratio but in a less predictable way
[6].
Because the timing of the level crossings in a bandpass sigma-delta modulator are triggered
by a clock, the pulse widths in the output are integer multiples of the clock period, Ts , where
Ts = 1/fs . Therefore the minimum pulse width is constrained to Ts which is beneficial since
the current mode class-D power amplifier (CMCD) has a bandwidth limitation and cannot
amplify very narrow pulses. On the other hand, long pulses are possible, but occur very
infrequently. Examples of pulse distributions can be found in the literature [9].
For this research project, a fourth order bandpass SDM is used. The fourth order transfer
function for the noise shaping filter HRF (s) is
P3
bn sn
HRF (s) = P4n=0 (1.4)
n
n=0 an s

and the coefficients are shown in Table 1.1. The carrier oversample ratio for the quantizer is
always 3.4 times the carrier frequency of the input source signal. The quantizer amplitude
levels are normalized to ±1 V and a full scale input amplitude is defined as an amplitude of
1 V. The modulator is implemented in Matlab/Simulink and data files are generated for the
pulse trains. The data files can be used for both simulation and for experimental work where
the files are downloaded to an arbitrary waveform generator.

1.2.2 Pulse Position Modulation


Noise shaped pulse position modulation (PPM) [10] is another encoding method that
can be used for switch-mode power amplifiers. Unlike bandpass SDM which generates level
transitions that are synchronized with a clock, PPM amplitude changes are asynchronous and
can occur at any time. A block diagram of a noise shaped PPM encoder is shown Figure 1.9.
3.4 GHz/(2 × 1 GHz) = 1.7.

8
1.2. Architecture of Switch-Mode Power Amplifier Systems

Table 1.1: Coefficient values for the noise shaping filter HRF (s)

n bn an
0 7.3874 × 1037 1.5584 × 1039
1 2.0201 × 1027 8.6858 × 1026
2 1.8850 × 1018 7.8962 × 1019
3 2.6147 × 107 2.2 × 107
4 - 1

In the noise shaped PPM encoder, the amplitude and width of pulses are constant and the
position (timing) of pulse edges are variable and dependent on the input source signal s(t).
Similar to bandpass SDM, the spectrum of PPM is broadband, and the feedback loop shapes
the in-band noise to ensure the source signal is encoded with a high signal to noise ratio.
Examples of a noise shaped PPM signal in the time domain and frequency domain are shown
in Figures 1.10 and 1.11, respectively.
Noise Pulse
Shaping Generator
s(t) p(t)
HRF (s)
Tp

Figure 1.9: Block diagram of pulse position modulator.

1
p(t): Pulse−position pulse trains

0.8

0.6

0.4

0.2

−0.2

−0.4

−0.6

−0.8

−1

1000 1001 1002 1003 1004 1005 1006


Time (ns)

Figure 1.10: A noise shaped PPM pulse train in the time domain.

The noise shaped PPM encoder is implemented in a Matlab/Simulink model. The noise

9
1.3. Switch-mode Power Amplifier Circuits

10

−10

−20
Power [dBm]
−30

−40

−50

−60

−70
600 700 800 900 1000 1100 1200 1300 1400
Frequency [MHz]

Figure 1.11: A noise shaped PPM pulse train in the frequency domain.

shaping filter HRF (s) is identical to the bandpass SDM filter whose coefficients were given
in Table 1.1. The pulse width, Tp , is set to be equal to half the period of the input carrier
frequency. This leads to an efficient encoder with high coding efficiency. Data files are
generated from the Matlab models and used for simulation and the files are downloaded to
an arbitrary waveform generator for experimental work.

1.3 Switch-mode Power Amplifier Circuits


The high efficiency amplifier circuit topologies which are relevant to this work are class-D,
class-E and class-F. Class-D and class-E are called switch-mode classes because the gate of
the device is switched by the input signal. Class-F is also frequently lumped into the switch-
mode category, although it does not necessary require a two level input signal to switch the
amplifying device. Class-F originated from the design of harmonic tuning in the output circuit
rather than from a concept where the input signal switches the device. Within class-D and
class-F, the circuit designs can be subdivided into two types of circuits. Circuits which switch
voltage are called voltage mode (VM) circuits and circuits which switch current are called
current mode (CM) circuits. The terms ‘voltage switched’ and ‘current switched’ are most
widely applied to class-D amplifiers. Within the context of class-F amplifiers, the term inverse
class-F, also written at class-F−1 , is more widely used to distinguish current switching from
voltage switching which is simply written as class-F. A short overview of the basic operation
of these circuits is presented next.

10
1.3. Switch-mode Power Amplifier Circuits

1.3.1 Class-D Power Amplifiers


1.3.1.1 Voltage Mode Class-D
A voltage mode class-D power amplifier (VMCD) is shown in Figure 1.12. The circuit
consists of two active devices in a cascade configuration. The common junction between the
devices is connected to a series output filter to reconstruct a sinusoidal load signal from the
pulse train. The gate drive signals, Vin1 and Vin2 , are two antiphase pulse trains which control
the state of the switches (devices).

VDD

IDC PDC

LDD

CRF
VD1

M1
Vin1 Bandpass Filter
Pout

VD2 Lt Ct
RL VL
M2
Vin2

Figure 1.12: Schematic of a VMCD power amplifier.

Figure 1.13 shows current and voltage waveforms for the VMCD amplifier when the input
pulse train is a periodic pulse train with a duty cycle of 30%. The first row shows the gate
drive waveforms, the second row shows the drain-source voltages across each switch, and the
third row shows the current through each switch. The drain voltage waveforms are similar
to the gate voltage waveforms except for distortion arising from switch resistance. Since the
voltage waveform follows the input pulse train, the circuit is called a voltage mode class-D
amplifier. The current through the switches are a portion of a sinewave. The two switch
currents sum to provide a sinusoidal load current.

11
1.3. Switch-mode Power Amplifier Circuits

VD1 (V)

VD2 (V)

Figure 1.13: VMCD amplifier voltage and current waveforms for a periodic drive signal with
a duty cycle of 30%.

12
1.3. Switch-mode Power Amplifier Circuits

1.3.1.2 Current Mode Class-D


Figure 1.14 shows a current mode class-D power amplifier. Similar to a VMCD amplifier,
the input pulse trains, Vin1 and Vin2 , are two antiphase signals that control the state of the
switches. When a device is turned on, the voltage across the switch is zero and all the DC
current, IDC , provided by the supply goes through the switch. When the same device is
turned off, there is no current through the switch and the voltage across the switch is a
portion of a sinusoidal wave. The switch current is similar to a square wave and the voltage
across the device is a portion of a sinusoidal wave. In other words, the CMCD amplifier can
be considered as the voltage-current dual of the VMCD amplifier. The reconstruction filter is
a shunt filter in a CMCD circuit which is also the dual of the series resonator in the VMCD
circuit.
VDD

IDC PDC

LDD LDD
Lt
RL
VD1 VD2

Ct
Vg1 Vg2
M1 M2
Vin1 Vin2

Figure 1.14: Schematic of a CMCD.

One of the main advantages of the CMCD circuit in Figure 1.14 compared to the VMCD
circuit in Figure 1.12 is that the gate drive signals in CMCD are ground referenced, while
the gate drive signal for the upper transistor in the VMCD circuit, M1 , requires a bootstrap
drive circuit. This feature of CMCD amplifiers makes it more attractive for experimental
work [11, 12].
Current and voltage waveforms for an example of a CMCD amplifier are shown in Fig-
ure 1.15. In the first row, the gate drive waveforms are shown. The input signal is a periodic
square wave with a duty cycle of 30%. In the second row, the current through the switches is
shown. Clearly the current follows the gate waveform and current is switched in the circuit.
In the third row, the drain-source voltage across each switch is shown. The drain voltage
waveforms are a gated sinewave and the differential voltage, VD1 − VD2 , is a sinusoidal signal.
The differential drain voltage is the same as the voltage across the load resistor and the shunt
resonator circulates harmonic current between the switches.

13
1.3. Switch-mode Power Amplifier Circuits

VD1 (V)

VD2 (V)
V

Figure 1.15: CMCD current and voltage waveforms for a periodic input pulse train with a
duty cycle of 30%.

14
1.3. Switch-mode Power Amplifier Circuits

1.3.2 Class-E Power Amplifiers


Several years after the class-D circuit topology was introduced, Sokal reported the first
class-E amplifier in 1975 [13]. The novelty in the class-E circuit topology is that, unlike
class-D which requires two switches, the class-E amplifier requires only one switch. A class-E
amplifier circuit is shown in Figure 1.16. The switch M1 is shunted by a capacitor CP and the
load is connected through a series resonator. When the switch is on, current flows through
the switch and the voltage across the shunt capacitance CP is low. When the switch is open,
the capacitor provides current to the load and the voltage across the capacitor changes.
VDD

IDC PDC

LDD
Series
Resonator
Vdrain Pout
LX Lt Ct
isw ic RL VL
M1 CP
Vin

Figure 1.16: Schematic of a class-E power amplifier.

Example waveforms for a class-E amplifier are shown in Figure 1.17. The gate drive signal
is shown at the top of the figure and the waveform is a square wave pulse train with a duty
cycle of 30%. The load current is sinusoidal because the series resonator filters the non-
sinusoidal drain voltage waveform, and the sinusoidal current is alternately sourced/sunk by
the switch, M1 , or the shunt capacitor, CP . The current into the switch and the current into
the capacitor are shown, and the two currents sum to equal the sinusoidal load current. The
voltage waveform is more difficult to understand and requires analysis [3]. The key features
of the voltage waveform are that the voltage is zero when the current is switched between
the switch and the capacitor. By proper choice of capacitance CP and inductance LX , the
first derivative of the voltage waveform can also be zero at the switching instances. In this
way, the voltage waveform smoothly approaches zero at each switching instant and the circuit
implements a zero-voltage switching and zero-derivative switching condition. This is the key
feature of the class-E amplifier. Therefore, the circuit is attractive because it has a single
switch and very high efficiency because of the zero-voltage switching condition. The shunt
capacitance CP can also be partitioned between the intrinsic output capacitance of the device
and an external capacitance such that the sum is equal to CP .
The primary disadvantage of class-E is the peak voltage generated across the switch. The
peak voltage depends on the duty cycle of the input signal and the variation in peak voltage
is illustrated in Figure 1.18. The peak drain voltage is normalized to VDD in this figure and
ranges from 2.7 for a 50% duty cycle to 4.8 for a 30% duty cycle. The variation in peak
voltage as a function of duty cycle is distinctly different from class-D where peak voltage is
independent of duty cycle. The variation in peak voltage is even more problematic for non-

15
1.3. Switch-mode Power Amplifier Circuits

Figure 1.17: Class-E voltage and current waveforms for a 30% duty cycle pulse train.

16
1.3. Switch-mode Power Amplifier Circuits

periodic pulse trains generated by SDM or PPM pulse encoders where peak voltages can easily
be five times larger than VDD . Therefore, although class-E has attractive features, voltage
peaking limits its application in switch-mode power amplifiers and this amplifier topology will
not be analyzed further in this work. Literature references to class-E amplifiers will be made
later in the context of designing RF rectifiers.

5
Normalized drain voltage
4

0
200.0 200.5 201.0 201.5 202.0
Time (ns)

Figure 1.18: Normalized voltage across the switch in a class-E PA for three different duty
cycles: 30% (circle), 50% (star) and 70% (square).

1.3.3 Class-F Power Amplifiers


In class-D and class-E amplifier circuits, the device is operated as a switch and the switched
waveforms at the drain node of the devices are the result of both a switched gate drive signal
as well as an output resonator. In class-F, the principle idea is to shape the drain signal
waveforms by controlling the harmonic impedance of the output network such that overlap
between the current and voltage waveforms are minimized. In practical class-F amplifiers,
harmonic impedances up to the third harmonic are commonly controlled and in some designs
even higher harmonic order impedance terminations are implemented to maximize power
efficiency. Although output harmonic impedances presented to the drain terminal of the device
are very important in class-F circuits, harmonic impedances at the gate (input) terminal of
the device can also significantly affect the power efficiency of the amplifier. The importance
of input harmonic matching is a topic that is investigated further in this work.
A class-F amplifier circuit with harmonic control up to the fifth harmonic is shown in Fig-
ure 1.19 and example waveforms are shown in Figure 1.20. The input signal has a fundamental
frequency fo . The amplifying device is usually biased near a class-B operating point and the
current through the device conducts for half a cycle. The device current waveform is ideally a
half sinusoid and has a fundamental frequency component and even harmonics. The output
matching circuit is designed to short the even harmonics in the current waveform and present
an open circuit impedance for odd harmonic frequencies. With open circuit impedances at

17
1.3. Switch-mode Power Amplifier Circuits

odd harmonics, the voltage waveform across the switch is shaped to be a rectangular square
wave. Ideally, the overlap of the current and voltage waveforms across the device is small
which then leads to low dissipation and high power efficiency. Since the voltage is a square
wave in class-F, the circuit switches voltage.
VDD

IDC PDC

LDD
Third Harmonic Fifth Harmonic
C3 C5
Vdrain Pout

L3 L5
C0
M L0 RL VL
Vin
f0

Figure 1.19: A class-F power amplifier circuit.

Class-F amplifier circuits can also be designed to switch current and the current switched
dual is called inverse class-F or class-F−1 . In a current switched amplifier, the odd harmonics
are shorted at the drain node and the even harmonics are open. Under these conditions, the
current is switched and the voltage is a half sinusoid. Class-F amplifiers are explored much
more extensively in Chapter 3.

18
1.3. Switch-mode Power Amplifier Circuits

Vinput (V)
0

-1

-2
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Time (ns)

60
50
40
Vdrain (V)

30
20
10
0
-10
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Time (ns)

500
400
Idrain (mA)

300
200
100
0
-100
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Time (ns)

15
10
5
VLoad (V)

0
-5
-10
-15
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Time (ns)

Figure 1.20: Class-F power amplifier voltage and current waveforms.

19
1.3. Switch-mode Power Amplifier Circuits

1.3.4 Summary of Amplifier Classes Based on Harmonic Termination


Impedances
The concept of harmonic impedances in class-F can be applied more broadly to other
amplifier classes and this provides a unified way to see the relationships between different
amplifier classes. Every amplifier has an output matching circuit that provides a fundamen-
tal frequency match. The output matching circuit also presents the device with harmonic
impedances that may be either explicitly controlled as in class-F, or implicitly controlled as in
class-E. By considering the relative impedance of the odd and even harmonics, the different
amplifier classes can be identified in a diagram as shown in Figure 1.21. This type of visu-
alization for the amplifier classes was first presented by Raab in 2001 [14]. In the diagram,
the x-axis shows the relative magnitude of the even harmonic impedance (reactance) and the
y-axis shows the relative magnitude for the odd harmonic impedance presented to the drain
terminal of the amplifying device. Mid-scale on each axis is the relative load line resistance
presented to the device at the fundamental frequency.
Starting with class-F, the output matching network should have high impedance at odd
harmonics and low impedance for even harmonics. This places class-F in the upper left hand
corner of the diagram. Inverse class-F is in the lower righthand corner and requires high
impedance at even harmonics and low impedance at odd harmonics. Voltage switched class-
D has an output series resonator that in theory presents an open circuit impedance at all
harmonics and therefore class-D is in the upper right. Conversely, inverse class-D has an anti-
resonant parallel resonant circuit which shorts all harmonic frequency components placing
class-D−1 in the lower lefthand corner. Class-E has a specially designed output network
impedance that leads to zero voltage switching and consequently the harmonic impedances
are neither shorted nor open. Class-E harmonic impedances lie within the middle of the figure.

|Zn=3,5,...|
F D

R1 E

D−1 F −1
0
0 R1 |Zk=2,4,...|
Figure 1.21: Amplifier classes in terms of harmonic load impedances.

20
1.4. Power Efficiency and Device Technology

1.4 Power Efficiency and Device Technology


Of the different types of device technologies which are available for wireless communication
applications, Gallium Nitride (GaN) device technology is now used in many power amplifier
designs [8, 15, 16, 17, 18, 19, 20, 21]. GaN device technology has the following features:

• A large band-gap energy leads to high electric field breakdown potentials [22]; for in-
stance, the breakdown voltage of a Cree CGH40010 transistor is 84 V [23].

• GaN has a higher carrier saturation velocity compared to other technologies [22].

• The thermal dissipation of GaN devices is high. Combined with high breakdown volt-
ages, GaN devices have higher power densities (W/mm of gate width) compared to
other device technologies such as silicon LDMOS (laterally diffused metal oxide semi-
conductor).

Figure 1.22 shows the band-gap energy of three different materials versus saturation veloc-
ity. From this figure it shows that GaN offers much better high power and high frequency
possibilities compared to Gallium Arsenide (GaAs) and Silicon (Si) device technologies. For
a better comparison of different semiconductor materials, Johnson’s figure of merit (JFOM)
was proposed [24]. It uses the breakdown voltage and saturated electron drift velocity to
define a value for the high-frequency handling capability of a certain semiconductor material.
JFOM is expressed as Vsat Ec /(2π) where Vsat is the saturated electron velocity and Ec is the
critical breakdown field. For example, the JFOM for GaN is 27 times higher than that of
silicon, about 15 times that of GaAs, and about 1.4 times that of Silicon Carbide (SiC).

4.5

4
GaN
3.5

3
Eg (eV)

2.5

2
GaAs
1.5
Si
1

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Vsat (cm/s) 7
x 10

Figure 1.22: Band gap energy and saturated velocity for Si, GaAs and GaN.

GaN device technology is available in both discrete high power devices as well as mono-
lithic microwave integrated circuit (MMIC) processes. A picture of a discrete GaN device and
a MMIC GaN circuit are shown in Figure 1.23. In this research project, power amplifiers are

21
1.5. Literature Review

Figure 1.23: GaN HEMT technology: packaged device from Cree (left) and MMIC (right).

built using discrete GaN devices and a 10 W device available from Cree; model CGH40010,
is used. For modeling work, the bare die, model CGH60015D, is used along with a package
model. Although Cree provides comprehensive large signal device models for these compo-
nents, they are black-box models, and equivalent circuit models are developed to provide
further insight into device losses and to predict power efficiency of different amplifier circuits.

1.5 Literature Review


A short summary of relevant published work which provides context for this research is
given in the next sections.

1.5.1 Class-D Power Amplifiers


Of the three amplifier classes (class-D, class-E and class-F), the class-D circuit topology
is well suited for amplifying pulse encoded signals. Providing the devices are driven with a
broadband driver, the amplifier is broadband, and the circuit is fundamentally robust enough
to amplify signals with a range of different duty cycles. The output resonant circuit can also
be used as a reconstruction filter, although higher order filter structures are usually required
to sufficiently attenuate out-of-band noise. These features have motivated significant research
interest in the design of RF class-D power amplifiers for pulse encoded signals.

1.5.1.1 Voltage Mode Class-D Amplifiers


The origin of the class-D amplifier dates back just over fifty years ago, when Baxandall
reported the first experimental results in 1959 [25]. The circuit was designed as a way of
generating high power sinusoidal signals for RF transmitters. During the next two decades,
a number of different high frequency class-D amplifiers were designed and implemented for
various applications. Variants of the original class-D amplifier which used complementary
switches then evolved into transformer coupled voltage and current switched configurations
[26, 27]. Experimental circuits optimized for high power amplifiers which use periodic signal
source to generate sinusoidal load signals continue to the present day [28, 29].
The first reported work associated with class-D power amplifiers that amplify pulse en-
coded signals was described by Raab in 1973 [30]. In his circuit, an RF PWM (pulse width

22
1.5. Literature Review

modulated) signal was amplified by a class-D amplifier in the MHz frequency range. The first
simulated results for a class-D power amplifier with sigma-delta modulation was published in
[31]. In 2008, Johnson et al. [6] published the analysis of a VMCD with bandpass sigma-delta
modulation and he introduced the concept of coding efficiency to generalize the analysis of
power efficiency to include pulse encoded signals. The loss mechanisms of a voltage mode
class-D amplifier were formulated in [32] and supported by experimental results for a CMOS
VMCD amplifier in 2007. Recently, other experimental work with time-encoded signals in-
cluding SDM, PPM and PWM for a GaN VMCD amplifier have been reported [8].

1.5.1.2 Current Mode Class-D Amplifiers


In 2001, researchers at USCD published a paper on a CMCD power amplifier circuit that
used a differential tank with two inductors providing DC to the circuit [33]. The circuit
was shown earlier in Figure 1.14 and most CMCD work since 2001 has built on this circuit
topology. The analytic work associated with CMCD has focused primarily on conduction
losses and inductive switching losses for period periodic signals with 50% duty cycles [34]. In
[35], equations are derived for variable duty cycle switching conditions under the assumption
of independent loss mechanisms; however, the device model is based on a switch and effects
of current saturation are not analyzed.
Different device technologies have been used to implement CMCD circuits with most
work using either GaN devices for high power amplifiers and CMOS technology of low power
amplifiers. A CMOS design reported in 2011 [12] describes a fully integrated CMOS inverse
class-D amplifier but the experimental results are limited to periodic pulse trains with a fifty
percent duty cycle.
Several research groups in Germany have conducted experimental work to realize CMCD
amplifiers and evaluate the amplifier performance with pulse encoded signals [15, 16, 17]. They
have also evaluated a derivative of the CMCD amplifier called class-S which includes diodes
in series with the switches to prevent an off-state switch from turning on when the duty cycle
is not 50%. The experimental work shows that the CMCS amplifier is not any better than
CMCD, and the general consensus is that the additional losses in the diode are offset by gains
in preventing off-state switches from turning on under variable duty cycle switching conditions.
Also, their work has primarily focused on designing and implementing experimental circuits
and evaluating the circuit performance with different types of pulse encoders including SDM.
The work has not focused as much on detailed analysis and prediction of power efficiency. A
summary of recent work related to CMCD and CMCS amplifiers are shown in Table 1.2.

23
Table 1.2: Some recently published results for class-D power amplifiers.
Number Topology Technology Input signal Model Year Author Ref.
1∗ Class D (CM) GaN Sigma-delta Yes 2013 Abbasian et al [36]
2 Class D (CM) LDMOS RF PWM No 2012 Foad et al [37]
3 Class D (CM) GaN Periodic No 2012 Park et al [38]
4 Class D (CM) CMOS Periodic Yes 2012 Chowdhury et al [39]
5 Class D (CM) GaAs Periodic Yes 2011 Kamper et al [40]
6 Class D (CM) GaN Periodic No 2011 Aflaki et al [41]
7 Class D (CM) LDMOS PWM No 2010 Schuberth et al [42]
8 Class D (CM) GaN Periodic No 2009 Aflaki et al [43]
9 Class D (CM) LDMOS Periodic Yes 2009 Brackle et al [35]
10 Class D (CM) LDMOS Periodic No 2006 Nemati et al [44]
11 Class D (CM) LDMOS Periodic No 2005 Kim et al [45]
12 Class D (VM) GaN Sigma-delta No 2012 Wentzel et al [8]
13 Class D (VM) GaN Periodic No 2011 Lin et al [29]
1.5. Literature Review

14 Class D (VM) CMOS Sigma-delta Yes 2007 Hung et al [32]


15 Class D (VM) CMOS Sigma-delta Yes 2006 Johnson et al [6]
16 Class S GaN Sigma-delta No 2011 Wentzel et al [15]
17 Class S GaN Sigma-delta No 2010 Wentzel et al [16]
18 Class S GaN Sigma-delta No 2009 Wentzel et al [17]
19 Class S GaN Periodic No 2010 Samulak et al [18]
20 Class S GaN Periodic No 2009 Samulak et al [46]
21 Class S GaN Sigma-delta No 2008 Leberer et al [47]

24
1.5. Literature Review

1.5.2 Class-F Power Amplifiers


The first description of a class-F amplifier was presented by Tyler in 1958 [48]. A few
years later, the first application of a high efficiency class-F power amplifier was reported by
Snider [49] when he constructed and tested a 46 W, 250 MHz power amplifier for generating
high power CW signals. The theory of class-F expanded in the mid-90’s by Raab in several
papers where he systematically evaluated the shape of class-F waveforms related to the output
termination impedances for a finite number of harmonics [50, 51, 52].
Experimental and theoretical work to optimize class-F amplifier performance has continued
and reports of high power amplifiers with peak efficiencies greater than 80% in the GHz
frequency range are common [19, 53, 54, 55]. Although peak efficiency is high, the power
efficiency of these amplifiers reduces in a similar way to class-AB amplifiers as output power
is backed off from peak power. Also, the class-B bias point that is used in class-F leads to high
distortion levels. On the other hand, if the input signal to the amplifier can be conditioned
to maintain operation in a highly saturated state, then class-F amplifiers offer a promising
future for wireless applications. The harmonically tuned impedance networks inherent in the
class-F amplifier design methodology are not likely to be optimal for continuous spectrum
pulse encoded signals, so methods of implementing broadband class-F amplifiers are of great
interest. Broadband amplifiers also have applications in multi-band wireless communication
systems. A summary of recent work in terms of class-F amplifiers is given in Table 1.3.

25
Table 1.3: Some recently published results for class-F family PAs
Number Topology Technology Power Efficiency Frequency Year Author Ref.
1∗ Class F GaN 40.5 dBm 78.8% 0.99 GHz 2015 Abbasian et al [56]
2 Class F GaN 40.7 dBm 80.1% 2.1 GHz 2014 Hwang et al [57]
3 Class F−1 GaN 42.2 dBm 83.9% 1.9 GHz 2014 Kim et al [58]
4 Class F GaN 40 dBm 82% 3.1 GHz 2013 Chen et al [59]
5 Class F/F−1 GaN 41 dBm 73.5% 2.65 GHz 2012 Moon et al [19]
6 Class F GaN 30.4 dBm 69% 4 GHz 2012 Zomorodian et al [20]
7 Class F GaN 10.5 W 74% 0.55-1.1 GHz 2011 Carruba et al [21]
8 Class F GaAs 20 dBm 83% 0.9 GHz 2011 Carruba et al [53]
9 Class F GaN 26.24 dBm 62.5% 0.9 GHz 2010 Osori et al [60]
10 Class F GaN 37 dBm 79% 1 GHz 2008 Aflaki et al [61]
11 Class F CMOS 21.8 dBm 43.9% 2.4 GHz 2006 Huang et al [54]
12 class-F−1 GaN 39.9 dBm 76.7% 3.5 GHz 2012 Dong et al [62]
13 class-F−1 GaN 47.2 dBm 69.4% 3.54 GHz 2011 Kim et al [63]
1.5. Literature Review

14 class-F−1 GaN 41.4 dBm 73.5% 2.5 GHz 2009 Wu et al [64]


15 class-F−1 GaN 46 dBm 60.8% 2.35 GHz 2009 Tanany et al [65]
16 class-F−1 GaN 22.7 dBm 74% 1 GHz 2006 Woo et al [66]
17 class-F−1 LDMOS 41.2 dBm 74% 1 GHz 2006 Ouyahia et al [55]
18 Class FE GaN 2.36 W 86.8% 61.44 MHz 2010 You et al [67]

26
1.5. Literature Review

1.5.3 RF Synchronous Rectifiers


Interest in RF to DC rectification has grown recently with growing interest in wireless
power systems. Although wireless power is an active research area, significant contributions
began in the early 1960’s with the pioneering work of Brown [68]. He worked on high power
microwave rectifying antennas, called rectennas, to remotely power a helicopter. Microwave
power was also being considered as a way to distribute power on the moon and to transmit
power from orbiting solar arrays to earth. These systems require very high power microwave
amplifiers and highly directional antenna to collimate high power fields.
Recent work has focused on much lower power level applications with interest to remotely
power or remotely recharge portable devices and sensors. For these applications, the work
can be broadly classified into two categories: 1) diode rectifiers and 2) synchronous rectifiers
based on switching transistors [69]. RF diode rectifier circuits [70, 71, 72] are simpler to
implement and avoid the design issues associated with gate drive circuitry which is required
in a synchronous rectifier. On the other hand, synchronous rectifiers can also exploit advances
in device technology including high efficiency GaN power devices.
With clever insight, a paper in 2012 was presented by Ruiz et al. [73] where he showed the
implementation of class-E RF rectifier based on reconfiguring a class-E power amplifier circuit.
The RF rectifier delivered 50 mW of DC power with an efficiency of 83% at a frequency of 900
MHz. The insight for the circuit came from work presented by Hamill [74] in the early 1990’s
where he showed how the theory of time-reversal duality could be applied to relate amplifier
and rectifier circuit topologies. The 2012 work has led other researchers to investigate the
design of synchronous RF rectifiers based on converting switch-mode amplifiers circuits into
rectifiers. Other important references related to synchronous rectifiers are shown in Table 1.4.
In this thesis, class-F amplifiers are reconfigured into class-F rectifiers and new contribu-
tions have been made in terms of both understanding the impact of loss in applying time-
reversal duality as well as the implementation of high power rectifiers and wideband rectifiers.

27
Table 1.4: Some recently published results for RF synchronous rectifiers
Num. Type Device f (GHz) PDC η (%) Year Author Ref.
1∗ class-F GaN HEMT 0.985 8.7 W 81.3 2015 Abbasian et al [75]
synch.
2∗ class-F−1 GaN HEMT 0.91 10.15 W 85 2015 Abbasian et al [76]
synch.
3∗ class-D CMOS 2.45 2.5 mW 31.7 2015 Dehghani et al [77]
synch.
4 class-C GaN HEMT 10.1 1.67 W 64.4 2014 Litchfield et al [78]
self-synch. MMIC
5 class-C GaN HEMT 10.1 3.18 W 63.9 2014 Litchfield et al [78]
self-synch. MMIC
6 class-E E-PHEMT 0.9 35 mW 88 2014 Ruiz et al [79]
self-synch.
7 class-E E-PHEMT 2.45 15 mW 77 2014 Ruiz et al [79]
self-synch.
8 class-E GaAs PHEMT 2.45 5.7 mW 77 2013 Ishikawa et al [80]
1.5. Literature Review

self-synch.
9 class-F GaAs PHEMT 5.8 5.5 mW 68 2013 Ishikawa et al [81]
self-synch.
10 class-F−1 GaN HEMT 2.14 8.5 W 85 2012 Roberg et al [82]
self-synch.
11 class-E E-PHEMT 0.9 50 mW 83 2012 Ruiz et al [73]
synch.
12 class-F E-PHEMT 0.9 12.1 mW 85.4 2004 Gómez et al [83]
self-synch.

28
1.6. Research Goals and Objectives

1.6 Research Goals and Objectives


The goals of this research are:

1. to improve analytic methods to predict the power efficiency of RF switch-mode power


amplifiers under general non-periodic switching conditions;

2. to build experimental prototypes of switch-mode power amplifiers to benchmark and


compare the performance of different circuit topologies;

3. to investigate and evaluate new switch-mode power amplifier circuit topologies that can
efficiently amplify pulse encoded signals.

The research goals are derived from the literature survey where the following observations
are made. First, there are many published experimental results that demonstrate high effi-
ciency operation for classes D, E and F; however, these results are almost always reported
under optimal switching conditions which correspond to a square wave drive signal with a
50% duty cycle. The power efficiency of switch-mode amplifiers degrades as the duty cycle
deviates from 50% and it is very important to understand how power efficiency changes as
the duty cycle (pulse width) changes in order to predict how well an amplifier will work for
wireless communication applications. Second, although there are many reports on different
switch-mode power amplifier circuits, it is not always that easy to compare results because
different devices are used in different circuit topologies and the objectives of the work have
not been to conduct comparative studies. In this work, amplifier circuits are analyzed and
evaluated using the same device to improve the consistency of experimental work and to draw
conclusions by comparing experimental results. Third, despite the high efficiency potential of
RF switch-mode power amplifier circuits, a competitive digital switch-mode power amplifier
has yet to be reported compared to conventional analog amplifier techniques such as Do-
herty [1]. Therefore, it still remains a challenge to realize high efficiency switch-mode power
amplifiers. This suggests that research must continue to evaluate new switch-mode power
amplifier architectures as a way to realize the full potential of high efficiency operation that
can theoretically be obtained by operating the device as a switch.
In pursuit of these research goals, specific research objectives are defined.

1.6.1 Predicting the Power Efficiency of CMCD Power Amplifiers for


Time Encoded Input Signals
As Table 1.2 shows, many publications have been written about CMCD; however, a careful
review of published work reveals the following limitations:

• Most authors have focused on building and measuring the performance of CMCD am-
plifier circuits to present experimental results without providing an analytic model to
predict and explain the experimental outcomes.

• Some researchers have presented analytical models; however, the models are usually
restricted to periodic input signals with fifty percent duty cycle.

• The peak current which the device can deliver in the saturation region is never consid-
ered. The device models are usually simple switches with an on-state resistance that
has no current limitations.

29
1.6. Research Goals and Objectives

• The output capacitance of the device is constant. However, most device capacitances
are nonlinear and capacitance can vary significantly under large signal conditions.

• Inductive switching losses associated with bond wires and the packaging are considered
to be a significant loss mechanisms in CMCD amplifiers, while capacitive switching losses
are usually neglected. Analytical results in Chapter 2 show that capacitive switching
losses can in fact be very significant under variable duty cycle switching conditions.

• Overlap losses have not been modeled; however, in practice rise time and fall times of
the voltage and current waveforms for the switches are significant.

Improving analytical models to predict power efficiency in CMCD amplifiers for general
switching conditions is a research objective. The analysis gives insight into the loss mech-
anisms as a function of the duty cycle as well as insight into how to appropriately select
the device load line considering current saturation in the switch. This work is described in
Chapter 2.

1.6.2 RF Switch-mode Power Amplifiers with Energy Recycling


As the experimental results of the published work show, it is difficult to obtain high effi-
ciency operation when amplifying pulse encoded signals. One of the reasons for this is that the
spectrum of the pulse encoded signals is broadband and continuous rather than harmonically
peaked. Also, the quantization noise is shaped to create a narrow noise well and consequently
a reconstruction filter with sharp attenuation characteristics is required to suppress out-of-
band noise. Highly selective filters have multiple resonators and are more complex than a
simple resonator. The filters are usually added after the amplifier circuit design and as a
result the out-of-band impedance presented to the device is rarely controlled in a predictable
way. If the required out-of-band impedance is violated, then the high efficiency operation of
the amplifier is compromised. As shown in Figure 1.21, all the amplifier classes can be related
to specific harmonic impedance conditions. Time encoded signals are challenging signals to
amplify because they have continuous power spectrums and are different from periodic signals
that have discrete harmonic frequency components.
A different approach to implementing RF switch-mode amplifiers was presented in [84]
where the output load circuit across the switch is broadband rather than harmonically tuned.
The load circuit was implemented by a complementary diplexer which has a matched broad-
band input port that splits into two complementary filter branches that separate the amplified
pulse train into an in-band signal and an out-of-band (quantization noise) signal. After signal
separation, the in-band signal couples to the antenna and the out-of-band signal can be recti-
fied. The rectified output power can provide an auxiliary supply which offsets the DC power
supplied to the amplifier. In this way, out-of-band power is recycled to improve the power
efficiency of the amplifier. A diagram of a switch-mode power amplifier with energy recycling
was shown in Figure 1.1. RF energy recycling has been reported for another power amplifier
architecture called linear amplification using nonlinear components (LINC), but the concept
is new in terms of its application to RF switch-mode power amplifiers.
Energy recycling is motivated by the observation that when high PAPR signals are encoded
into a pulse train, the coding efficiency of the pulse train is low, and large amounts of power
are generated in the out-of-band frequency spectrum. Therefore, two approaches to the design
of an efficient switch-mode power amplifier are: 1) to design a circuit that inherently operates

30
1.6. Research Goals and Objectives

with high efficiency even for pulse trains with broadband spectrums, or 2) design a high
efficiency broadband amplifier and use energy recycling to capture out-of-band energy and
use this to offset the DC supply to the amplifier. The second method is investigated in this
work. The concept of energy recycling is evaluated in detail and supported by experimental
results. The work is described in Chapter 5.

1.6.3 RF Rectifier Circuits


If energy recycling is to be implemented in a RF switch-mode power amplifier, a funda-
mental circuit block that must be designed is the RF rectifier. RF rectification has become
increasingly important especially in terms of research related to the development of wireless
power technology. In wireless power, RF signals are transmitted to a remote device, for ex-
ample a sensor, and the remote device rectifies incident RF power to autonomously power
the device or recharge a battery in the remote device. The research is motivated by the need
for remote systems that never need to be serviced and the devices can be embedded into
structures where access is difficult or impossible to replace batteries. Therefore, although
the motivation in this work is to design RF rectifiers for energy recycling in switch-mode
power amplifiers, the work contributes to a broader effort focused on implementing efficient
RF rectifiers for wireless power applications.
In a very interesting paper presented by Hamill in 1997, he showed how switch-mode power
amplifiers can be time reversed to operate as rectifiers. The work was motivated by the design
of power electronics circuits where he showed how time-reversal duality theory can be applied
to unify the design of inverters (amplifiers) and rectifiers. The application of time-reversal
duality to the design of RF rectifiers is quite recent with the first reported work in [73]. Since
this research began with a focus on RF switch-mode power amplifiers, the implementation of
synchronous RF rectifiers based on time-reversal duality is a natural extension.
Similar to switch-mode power amplifiers, the theory and analytic support for RF syn-
chronous rectifiers continues to evolve. This work contributes to the design of RF rectifiers
in terms of both analysis and experimental work by investigating the following points.

• When there is significant power loss, there is an ambiguity in how the amplifier and recti-
fier circuits should be compared in the context of time-reversal duality. Hamill’s original
work has been applied to lossless circuits, and the implications of loss are considered.

• The power efficiency of a class-F amplifier and the corresponding rectifier dual are
analyzed and compared experimentally. The results show that the rectifier has slightly
higher power efficiency and reasons for the difference are given.

• The power efficiency of RF rectifiers is dependent on the DC load impedance. The re-
lationship between power efficiency and load resistance is investigated both analytically
and experimentally.

• The class-F and class-F−1 synchronous rectifiers are compared to clarify which of these
configurations is better in terms of power efficiency and dynamic range.

• A wideband class-F−1 synchronous rectifiers is designed and implemented experimen-


tally. This is the first report of a wideband RF rectifier design based on time-reversal
theory applied to a wideband class-F−1 synchronous rectifier.

31
1.6. Research Goals and Objectives

The analysis and experimental results of different class-F RF rectifiers is described in


Chapter 4.

1.6.4 Class-F and Class-F−1 Power Amplifiers


In this research, class-F type amplifiers are used for both switch-mode amplifiers and
RF synchronous rectifiers. Therefore, a chapter in this work is dedicated to the design of
class-F amplifiers. In carrying out the detailed analysis of class-F circuits and reporting on
experimental results, effort has been made to contribute to a deeper understanding of these
circuits. Two contributions are made. First a detailed study of the importance of input
harmonic matching is made in terms of the relationship between input match and power
efficiency. The work builds on other published contributions and shows that second harmonic
input matching is very important, while third harmonic input matching provides diminishing
improvements in power efficiency. Second, unified experimental work is shown that provides
results for comparing the performance of both class-F and Class-F−1 . The conclusion is that
current switched Class-F−1 amplifiers are better than voltage switched class-F amplifiers in
terms of power efficiency and dynamic range. Chapter 3 describes the details of the class-F
amplifier design work and the results are used to compare with class-F rectifier circuits in
Chapter 4.

1.6.5 Supporting Publications


Some parts of this research work have been previously published in journals and confer-
ences as listed below.

1. S. Abbasian and T. Johnson, “RF current mode class-D power amplifiers under periodic
and non-periodic switching conditions,” in IEEE International Symposium on Circuits
and Systems (ISCAS), May 2013, pp. 610-613.

2. L. Xiao, S. Abbasian, and T. Johnson, “All-digital encoders for RF switch-mode power


amplifier applications,” in IEEE Wireless and Microwave Tech. Conf. (WAMICON),
Jun. 2014, pp. 1-6.

3. S. Dehghani, S. Abbasian and T. Johnson, “Tracking load to optimize power efficiency


in RF to DC rectifier circuits,” IEEE Wireless Power Transfer Conference (WPTC),
Boulder, Colorado, U.S.A., pp. 1-3, 2015.

4. S. Abbasian and T. Johnson, “High efficiency and high power GaN HEMT inverse
class-F synchronous rectifier for wireless power applications,” in European Microwave
Conference (EuMC), Paris, France, Sep. 2015.

5. S. Abbasian and T. Johnson, “High efficiency GaN HEMT class-F synchronous rectifier
for wireless applications,” IEICE Electronics Express, vol. 12, no. 1, pp. 1-11, 2015.

6. S. Abbasian and T. Johnson, “Effect of second and third harmonic input impedances
in a class-F amplifier,” Progress In Electromagnetics Research C, Vol. 56, pp. 39-53,
2015.

Also another journal article is under review and a second article is in preparation.

32
1.7. Thesis Outline

• S. Dehghani, S. Abbasian and T. Johnson, “Adjustable load with tracking loop to


improve RF rectifier efficiency under variable RF input power conditions” submitted to
IEEE Trans. Microwave Theory Tech., Jun. 2015.

• S. Abbasian, L. Xiao, and T. Johnson, “Energy recycling in RF power amplifiers,” in


preparation for IEEE Trans. on Circuits and Systems II.

1.7 Thesis Outline


A roadmap of this thesis was given in Figure 1.1. Chapter 2 describes analysis and
modeling of a CMCD power amplifier circuit with focus on predicting power efficiency for non-
periodic pulse trains. Chapter 3 describes the design of class-F and class-F−1 power amplifiers.
In Chapter 4, the class-F amplifiers in Chapter 3 are reconfigured as RF synchronous rectifiers
using the theory of time-reversal duality. Chapter 5 presents theory and experimental results
for a switch-mode power amplifier with energy recycling. Conclusions and future work are
presented in Chapter 6.

33
Chapter 2

Power Efficiency Analysis of RF


Current Mode Class-D Amplifiers
In this chapter, the power efficiency of RF current mode class-D (CMCD) amplifiers are
analyzed for a general class of time encoded signals. Time encoded signals are continuous-
time signals with only two amplitude levels. Consequently, a time encoder maps an input
source signal, such as a modulated carrier, into a binary amplitude pulse train where all the
source information is contained in the timing of the amplitude changes in the signal. In this
way, time encoding is similar to frequency modulation. Time encoded signals are suitable for
amplification in RF switch-mode amplifiers such as CMCD providing the encoders synthesize
signals which have a power spectrum that is compatible with signal reconstruction using an
output bandpass filter. Examples of time encoded signals which meet these requirement are
bandpass sigma-delta modulated signals and noise shaped pulse position modulation.
Although papers have been published which analyze the power efficiency of CMCD am-
plifiers, the input signal in most cases is assumed to be a periodic signal with a 50% duty
cycle [12, 34, 85, 86]. The 50% duty cycle switching condition in class-D is the optimal
switching condition that minimizes loss and maximizes output power; however, it does not
give insight into the amplifier power efficiency for a modulated source signal. In other work,
equations have been derived for variable duty cycle switching conditions under the assumption
of independent loss mechanisms [35] and for unconstrained peak current conditions that are
typically encountered in practical devices. Therefore, the limitations of existing work moti-
vates a more comprehensive power efficiency analysis of the CMCD amplifier under general
switching conditions that are typical for time encoded signals.
This chapter expands the analysis of power efficiency to include capacitive switching losses
associated with the effective output capacitance of the device under variable duty cycle switch-
ing conditions. Power efficiency equations are derived to model conduction losses, inductive
switching losses, capacitive switching losses, peak current limits (Imax ), and overlap losses
during switch transitions. The analysis under variable duty cycle switching conditions is
verified with a simulation in Keysight’s ADS software using large signal device models for a
CMCD amplifier with two 15 W Cree GaN power devices. Circuit simulations using time en-
coded signals are also compared with the analysis. Conclusions show that capacitive switching
losses normally assumed to be negligible in CMCD circuits can be significant under back-off
conditions from peak power. Also, the selection of the load line must consider the minimum
and maximum pulse widths of the switching signal to control power loss under variable duty
cycle switching conditions.
1
Parts of Chapter 2 have been published in an article. Reprinted with permission from IEEE [36].

34
2.1. The Current Mode Class-D RF Amplifier

2.1 The Current Mode Class-D RF Amplifier


A circuit diagram for a CMCD amplifier is shown in Figure 2.1. The circuit consists of
two devices, M1 and M2 , which are switched by a pair of complementary pulse trains, Vs1 and
Vs2 . The pulse trains can be either a periodic signal such as a square wave or more generally a
non-periodic pulse train which has an encoded source signal such as a modulated carrier. The
currents, isw1 and isw2 , form a pair of differential switched current signals. The differential
current is filtered by a parallel tank circuit (LT and CT ) which is anti-resonant at the carrier
(fundamental) frequency. Harmonic or out-of-band current components are shorted by the
tank and the fundamental frequency component or in-band current components are delivered
to the load.
vL

RL
Node A Node B
CT

LT

isw1 LDC1 LDC2 isw2


Rs1 IDC Rs2
M1 M2

Vs1 VDD Vs2

Figure 2.1: A CMCD circuit.

In this chapter, a CMCD amplifier design is analyzed for a circuit which uses Cree GaN
power devices for the switches. The devices are 15 W unpackaged die, model CGH60015D.
Although Cree provides a comprehensive large signal model for this device, the model is
proprietary and the details of the intrinsic devices are not available to the user. The Cree
model is like a black box with terminals for the gate, drain and source, and cannot be used
for analysis. Therefore, equivalent circuit models which capture the primary behaviour of the
devices are very useful to gain insight into the design of the CMCD amplifier and analyze the
contribution of different loss mechanisms to the overall power efficiency of the amplifier.

2.2 Device Models


Three levels of device models are identified to support analytical and simulation results
for CMCD amplifier circuits. The level 3 model described below is also used extensively in
later chapters related to the design of class-F amplifiers. The models are described in the
following sections.

35
2.2. Device Models

2.2.1 Level 1
The level 1 model is the most commonly used model in the analysis of CMCD amplifier
circuits. Many literature references can be found for this type of model and the associated
analysis of the class-D amplifier under periodic switching conditions [87, 41].
In the level 1 model, the device is modeled as a switch with an on-state resistance Ron
and an infinite off state resistance Rof f . The switch changes states instantaneously in accor-
dance with an amplitude transition in the gate drive pulse train. Power loss associated with
dissipation in the on-state is modeled and called conduction loss, PRon .
The equivalent on-state resistance for the device model is obtained from an approximation
of the DC IV characteristics for the device. In the on-state, the current is constrained to
operate in the linear region of the device at or below the knee point, the point which separates
the linear and saturation regions of the device IV curve. An example of the knee point and
the on-state resistance approximation are shown in Figure 2.2. As will be shown later, the
linear on-state resistance model can lead to analytic results that are less accurate especially
for cases where the input pulse train is periodic and the duty cycle is not 50%. A schematic
diagram of a CMCD amplifier using a level 1 device model is shown in Figure 2.3.

Figure 2.2: Typical DC IV characteristics for a GaN device.

The level 1 model can also include device capacitances and inductances. A fixed output
capacitance, Cout , models the total effective output capacitance of the device and the capac-
itance can be added to shunt the switch. Since device capacitances are nonlinear, the output
capacitance is the effective output capacitance at the drain node and includes contributions
from Cds and Cgd . In Figure 2.3, the effective device capacitances are modeled by C1 and C2 .

36
2.2. Device Models

Figure 2.3: Schematic for current mode class-D power amplifier with a level 1 device model (ADS schematic).

37
2.2. Device Models

As is well-known, one of the advantages of the CMCD amplifier is that there is zero-voltage
switching across the devices providing the gate pulse train in a periodic square wave with a
50% duty cycle. This means that with a 50% duty cycle there are ideally no switching losses
associated with the discharge of Cout . Further, the output capacitance of the two switches
shunts the differential tank circuit with an equivalent capacitance of Cout /2 and the tank
capacitance CT can be modified to compensate for the additional shunt capacitance.
When the duty cycle is not 50%, the zero-voltage switching is no longer valid and there
is stored charge in Cout that must be dissipated during the switching transition. An example
of the current and voltage waveforms at the drain terminals of the devices for a duty cycle of
30% is shown in Figure 2.4.2 The analysis of power loss associated with the discharge of Cout
for non-zero voltage switching conditions in CMCD is presented later in Section (2.3). The
power loss associated with the discharge of Cout is denoted as Pcap .
Another power loss mechanism that is commonly analyzed for CMCD circuits is the power
loss associated with series inductance in the switch, Ls . The inductance is primarily associated
with the drain inductance on the die and in the package. Because current is hard-switched
in the CMCD circuit, stored energy in the inductor is dissipated each time the switch opens.
The inductance also leads to voltage spiking when the switch changes state, because the
current cannot instantaneously change through the inductance. For a periodic gate signal
with a fundamental frequency fo , the corresponding power loss from the inductor, called Pind ,
2 f . The equation models the power loss in the two switches where the switched
is Ls IDC o
current has an amplitude of IDC , the total DC current into the CMCD circuit (see Figure 2.1).
Typically, inductive switching losses are low because the device and package design minimize
series inductance. The magnitude of inductive power loss relative to other loss mechanisms
is compared later in section (2.3.6).
In summary, the level 1 model is appropriate for independently calculating the power
loss associated with the on-state resistance (PRon ), capacitive switching losses (Pcap ), and
inductive switching losses (Pind ). Considering these power loss mechanisms, the overall drain
efficiency of the CMCD amplifier is
PL
η= (2.1)
PDC
where PL is the power delivered to the load resistor RL and the DC power is

PDC = PL + PRon + Pcap + Pind . (2.2)

2
This figure was shown earlier in Figure 1.15 and is repeated for convenience of reference.

38
2.2. Device Models

Figure 2.4: Level 1 CMCD current and voltage waveforms for a 30% duty cycle periodic drive
signal.

39
2.2. Device Models

2.2.2 Level 2
Although the level 1 model is used in many papers for analyzing CMCD circuits, it has
limitations especially under more general switching conditions including arbitrary duty cycle
periodic switching and non-periodic switching. Since the primary application of CMCD am-
plifiers in this research is to consider this amplifier as an efficient means of amplifying wireless
signals, it is necessary to consider how the CMCD works under general switching conditions
when the input signal is an encoded pulse train.
The level 2 model extends the analysis of the CMCD circuits and includes the following
features.

1. An important loss mechanism in switch-mode power amplifiers is the loss created from
non-ideal switching waveforms which have finite switching times. During the switching
interval, of duration τ , the current and voltage waveforms across the device overlap
which leads to power dissipation in the device. An example of the drain current and
drain voltage waveforms which include overlap loss are shown in Figure 2.5. In the level
1 model, switching is assumed to be instantaneous and overlap power loss is zero. In
level 2, overlap power loss, denoted as Pτ is included in the calculation of the overall
power efficiency of the amplifier. The overall DC power for the CMCD amplifier with
overlap loss is
PDC = PL + PRon + Pcap + Pind + Pτ . (2.3)
which can then be used in equation (2.1) to calculate the drain efficiency for the amplifier.

2. When the duty cycle for a periodic input signal is not 50%, then the average current
carried by each switch is no longer equal to IDC /2. This means that one switch carries
more current than the other. If a CMCD amplifier is designed using the assumption
of a 50% duty cycle condition which is often done, then current saturation in one of
the switches can appear if the load line is selected for maximum power. Consequently,
margin must be allocated in a CMCD design to consider the variation in pulse width
(duty cycle) to ensure that current saturation is avoided as this will significantly increase
the dissipation in the amplifier. An analysis with a level 1 model does not include any
current saturation limitations as the on-state resistance is an ideal resistor. In level
2, current saturation is modeled which constrains the peak current and models the
limitation of a practical device where the peak switch current cannot exceed Imax , the
maximum device current at saturation.

3. The effective output capacitance of the device is in general a nonlinear function of the
gate and drain voltages. Since often during a switching period the device is either in
the on-state or the off-state, the output capacitance model can be modified to include
an on-state capacitance Con and an off-state capacitance Cof f . A time averaged capac-
itance between the on and off state capacitances can then be used to analyze capacitive
switching losses under arbitrary duty cycle switching conditions.

The level 2 model is shown in Figure 2.6. The model is partitioned into two states, on-
state and off-state, and includes current saturation and different on-state and off-state switch
resistances and switch capacitances. The model also assumes that current and voltage overlap
for a state transition interval time of τ .

40
2.2. Device Models

2.0 70
60
1.5

Drain Current (A)

Drain Voltage (V)


50
1.0 40
30
0.5 20
10
0.0
0
-0.5 -10
200.1 200.6 201.1 201.4
Time (ns)

Figure 2.5: Transition time (τ = 0.15T ) for a CMCD with Cree CGH60015D transistors.

The model values used for the Cree CGH60015D are summarized in Table 2.1. The values
were extracted from the Cree large signal device model by measuring the S-parameters in the
on and off state. The on state corresponds to a gate voltage (Vgs,on ) of 0 V, the off state
corresponds to a gate voltage (Vgs,of f ) of -4 V, and the DC supply voltage VDD is 25 V.
S-parameters for a frequency range of 10 MHz to 2.5 GHz were fitted to the simplified model
to minimize the mean square error. A comparison of the S-parameters for the level 2 model
versus the S-parameters obtained for the Cree large signal model are shown in Figure 2.7. As
shown, the level 2 model matches well with the Cree model.

Table 2.1: Summary of level 2 model values for the Cree CGH60015D die.

Ron Con Cof f Ls


2Ω 1.8 pF 0.5 pF 250 pH

2.2.3 Level 3
Level 3 is a simulation model that is implemented to independently include and control
nonlinear device capacitances. The model includes a current generator that models the DC IV
characteristics, nonlinear device capacitances for Cgs , Cgd , and Cds , series inductance in the
drain, and gate resistance. The model is implemented in ADS and used for circuit simulations
to compare with analytic results. The simulation model is referenced much more in the next
chapter and additional details are given there.

41
2.2. Device Models

d
g
s
d d
ON state OFF state
Ld Ld
Vgs,on Vgs,of f
d’ d’
g g
Isat
Con Rof f Cof f
τ
Ron

s s

Figure 2.6: Level 2 device model for a CMCD amplifier.

+j1.0

+j0.5 +j2.0

S22 ON-State: Model


S22 ON-State: Device
S22 OFF-State: Model
+j0.2 +j5.0
S22 OFF-State: Device
0.2

0.5

1.0

2.0

5.0

0.0 

-j0.2 -j5.0

-j0.5 -j2.0

-j1.0

Figure 2.7: S-parameters for the on and off state for a Cree CGH60010D device.

42
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

2.3 Power Efficiency Analysis of a CMCD Amplifier with


Periodic Switching
The level 2 device model is used to analyze the power efficiency of a CMCD amplifier for
a periodic switching signal. For the analysis, the periodic signal has a period T and a duty
cycle α. An example gate drive waveform is shown in Figure 2.8. The analysis works through
the following steps. First, equations for the DC supply voltage and supply current are derived
for an arbitrary duty cycle. The analysis shows that the minimum DC current corresponds
to a duty cycle of 50% and increases for non-50% duty cycle conditions. The DC analysis
also provides guidelines on how to select an appropriate load resistance for an arbitrary duty
cycle.
The power loss associated with the discharge of the output capacitance is analyzed next.
The analysis shows that capacitive switching losses are not negligible for duty cycles which
vary significantly from 50%. Conduction, inductive and overlap loss mechanisms are also
analyzed. Finally, a comparison of analytic versus simulated results is made for each of the
loss mechanisms.
Amplitude

0 T 2T 3T Time

Figure 2.8: Signal with period T and variable duty cycle (α).

2.3.1 Selecting the DC Supply Voltage


The constraints on the DC supply voltage VDD for the CMCD amplifier are determined by
the drain-source breakdown voltage of the device, Vds,bd , and the maximum amplitude of the
load voltage, Vmax . For a periodic switching signal, the maximum load voltage corresponds
to a duty cycle of 50% and the voltage across the device (switch) is a half sinusoid. The
expected value of the drain voltage is equal to the DC supply voltage. Therefore, for a 50%
duty cycle
Vmax
VDD = (2.4)
π
with the constraint that Vmax must be less than the breakdown voltage. For the Cree
CGH60015D, the breakdown voltage is 84 V, and a DC supply voltage of 25 V is selected.
From (2.4), the peak amplitude of the load voltage is 78.5 V which leaves a margin of 5.5 V
relative to the breakdown voltage. The relative values of the DC supply voltage, the maximum
switch voltage, and the breakdown voltage are shown for a typical device in Figure 2.11.

43
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

2.3.2 Load Power


Let’s now consider the relations between the DC current, IDC , supplied to the CMCD
circuit and the peak amplitude of the load current, Iom , to estimate load power, PL . With
reference to the drain current waveforms in Figure 2.4, the amplitude of on-state switch
current is equal to IDC . Since the drain current waveforms are periodic square wave signals
with a duty cycle of α, they can be expanded into a Fourier series. The corresponding peak
amplitude of the load current is then equal to
2
Iom = IDC sin(απ). (2.5)
π
Once the peak output load current is found, an expression for the load power can be written
as
1 2
PL = RL Iom (2.6)
2
Equations (2.5) and (2.6) can also be combined to show the relationship between duty cycle
and load power:
2 2
PL = RL IDC [sin(απ)]2 . (2.7)
π2
From the last equation, it is clear that load power is a function of the load resistance, the
current switched by the transistors, and the duty cycle of the periodic signal. Load power is
maximized for a 50% duty cycle and power reduces as the duty cycle moves away from 50%.
By changing duty cycle, load power changes, and therefore duty cycle can be used to control
the amplitude of the load signal.

2.3.3 Power Loss Mechanisms


The level 2 device model includes output capacitance, inductance, and finite switching
time, and the current waveforms at the device terminals are modified compared to the ideal
class D switched current waveforms. An example of a typical drain current waveform including
additional losses is shown in Figure 2.9. When switch 1 (M1 ) is on, the output capacitance of
the device is essentially shorted by the low on resistance of the device (see Figure 2.6 for the
level 2 device model). Therefore, the on-state device capacitance, Con , is negligible and loss
is primarily associated with switching current through the series lead inductance Ls . On the
other hand, when M1 is on, M2 is in the off-state, and the off-state capacitance, Cof f , shunts
the drain-source terminals of the switch. The voltage across Cof f must be charged to track
the voltage at node B (see Figure 2.1). The energy required to charge Cof f must be supplied
by M1 and this leads to an additional current component that is superimposed on the drain
current flowing into switch 1. The drain current waveform through switch 1 including Cof f
is given by
isw1 (θ) = I1 (θ) − IRof f (θ) − ICof f (θ) (2.8)
where θ = ωt and
IDC (θ/τ ) 0<θ<τ



 IDC
 τ < θ < 2πα + τ
I1 (θ) = 2πα + 2τ − θ (2.9)
 IDC 2πα + τ < θ < 2πα + 2τ


 τ
0 2πα + 2τ < θ < 2π

44
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

Vom
IRof f (θ) = sin(θ) (2.10)
Rof f
4 sin(τ /2)
ICof f (θ) ≈ ωCof f Rp IDC sin(απ) cos(θ). (2.11)
π τ
In these equations, Rp = RL ||Rof f ; however, Rof f is usually much greater than RL , thus
Rp ≈ RL . Also, the current through Cof f is calculated from the derivative of the voltage
across the switch with respect to time.

Figure 2.9: Device current waveforms at the drain terminal of the switching device. The ADS
simulation results are for a Cree large signal device model.

The peak amplitude of the load current, Iom , which flows through the load resistance RL
is found by calculating the amplitude of the fundamental frequency component of the drain
current given in (2.8). From a Fourier series analysis of the pulse train, the peak amplitude
of the load current is
p
Iom = A2 + B 2 (2.12)
where
2 Rp sin(τ /2)
A = IDC [sin(D) − sin(τ /2)] − ICof f (2.13)
π RL τ
2 Rp sin(τ /2)
B = IDC [cos(τ /2) − cos(D)] (2.14)
π RL τ
and D = 2πα + 3τ /2. The current component ICof f associated with charging Cof f is
small relative to the amplitude of the load current. Therefore, the expression for Iom can be
simplified to
4 sin(τ /2)
Iom = IDC sin(απ + τ /2) . (2.15)
π τ

45
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

As a check on equation (2.15), consider the limit as τ → 0. In the limit, overlap reduces to
zero and the current waveform simplifies to the ideal CMCD waveform where
2
Iom = IDC sin(απ). (2.16)
π
When overlap loss is not negligible, as in a practical CMCD design, the peak load current is
reduced by the finite rise and fall times of the switched current waveform.
The power losses associated with on-state conduction loss and the charging of the off-state
switch capacitance are
 Z π 
1 2
PRon = 2 × Ron (IDC − ICof f (θ)) dθ
2π 0
2 1
= Ron IDC + Ron max(ICof f )2 . (2.17)
2
There is also power loss associated with non-zero voltage switching for duty cycles which
are not 50%. During an off to on state transition, the output capacitance of the off state
switch must be discharged. The capacitance is nonlinear, and as an estimate of the power loss
during the transition, the average drain source capacitance is used: Cds = (Cof f + Con )/2.
The corresponding power loss is given by [35]
 2
πVDD cos(απ) 1
Pcap = 2fo Cds . (2.18)
sin(απ) 1 + 4 fo Cds RL cos2 (απ)
Overlap loss during a switch transition is defined as the power loss associated with the
cross-over of the drain current and drain voltage waveforms. Using a linear approximation
for the amplitude transitions as shown in Figure 2.10, the overlap loss is estimated as

Z 2π 
τ −θ
   
1 θ
Pτ = 2× IDC U0 dθ
2π τ τ
 Z π0 
τ −θ
  
1 θ
= 2× IDC U0 dθ
π 0 τ τ
 Z π
τ θ − θ2

1
= 2× IDC U0 dθ
π 0 τ2
IDC
= U0 τ. (2.19)

where U0 = Vom cos(απ) τ and Vom is the peak amplitude of the load voltage. The peak
amplitude of the load voltage is equal to Iom RL , and using the approximation for Iom given
in (2.16), Vom ≈ (2/π) RL IDC sin(απ). Therefore, the overlap power loss can be expressed
more compactly as
τ
Pτ ≈ cot(απ) PL . (2.20)
3
The last power loss mechanism considered in this analysis is the power loss associated with
the discharge of energy stored in the series inductance, Ls . The fundamental frequency of the
switching waveform is fo and the on-state current is IDC . The stored energy is discharged
twice per cycle because there are two switches. Therefore, the inductive switching loss is
2
Pind = Ls IDC fo . (2.21)

46
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

Vom

U0
Idc

2απ 2απ + 2τ
θ
τ

Figure 2.10: Overlap of drain current and drain voltage waveforms in a CMCD amplifier.

The sum of all the power losses and the load power must equal the DC power supplied to
the amplifier:

PDC = PL + PRon + Pcap + Pτ + Pind . (2.22)

Analytical expressions have been derived for each term in the preceding analysis, and therefore
the overall drain efficiency of the CMCD under periodic switching conditions can be found
using
PL PL
η = = . (2.23)
PDC PL + Pon + Pcap + Pτ + Pind

2.3.4 Selecting a Device Load Line


An important design step in any amplifier design is to select the appropriate load line for
the amplifying device. In power amplifier design, this almost always means maximizing load
power and maximizing power efficiency. The load line for the CMCD amplifier is discussed
next. The analysis begins by considering conduction loss, then comments are made about
how the load line is adjusted for other losses.
The total DC power supplied to the amplifier is equal to the sum of the load power and
the losses in the circuit. The switches in a CMCD amplifier switch a current IDC , and the
corresponding conduction power loss is Ron IDC 2 . The DC power supplied to the amplifier is

equal to VDD IDC . Therefore,


2
PDC = VDD IDC = PL + PRon = PL + Ron IDC . (2.24)

Using the expression for PL in equation (2.7), (2.24) can be rearranged to solve for IDC :

VDD
IDC = . (2.25)
2 2
RL sin (απ) + R on
π2
A few remarks are made about this equation. First, the equation shows that DC power
is a function of duty cycle. The DC current is lowest for a 50% duty cycle (α = 0.5),
and DC current increases for non-50% duty cycles. Second, although the equation shows a

47
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

relationship between the load resistance and the DC supply current, the maximum DC current
is constrained by the device characteristics. If the maximum saturation current for the device
in the on-state is Imax , then IDC ≤ Imax . In a level 1 model (see Figure 2.4), the device is
modeled as a simple switch without current saturation. If this circuit is simulated to evaluate
power efficiency as a function of duty cycle, incorrect conclusions can be made because there
is no current saturation function in the model. The level 2 model adds current saturation and
ensures that unrealistic device currents are not generated in the simulation for arbitrary duty
cycles.
An important consequence of the variation of the switch current (IDC ) as function of
duty cycle is that a CMCD design must consider the range over which duty cycle will vary.
The 50% duty cycle switching condition is a commonly used benchmark in CMCD amplifiers
because this corresponds to maximum power and maximum efficiency. On the other hand, a
design optimized at 50% duty cycle is not optimal for a design which must amplify signals
with variable duty cycle. Margin must be allocated to ensure that the duty cycle limits do not
create switch currents that exceeds Imax , otherwise the power efficiency of the amplifier will
be severely reduced because the switch current pushes past the knee and deep saturation can
occur. These points are illustrated in Figure 2.11. In this figure, a DC current (switch current)
called Io is defined to correspond to an operating point with 50% duty cycle. Evaluating
equation (2.25) for α = 0.5 gives

VDD
Io = . (2.26)
2
R L + Ron
π2
For variable duty cycle switching conditions, the 50% duty cycle switching current Io must
be less than Imax . As an example, suppose the duty cycle switching range in the input pulse
train is constrained to a range from 30-70%. To avoid deep saturation at the end points of
the duty cycle range, the load resistance is selected for the 30% duty cycle, or equivalently
the 70% duty cycle point, such that IDC is equal to Imax . If the duty cycle exceeds the design
range, then the device becomes deeply saturated. The limiting case for saturation is when
α = 0, or α = 1, which corresponds to one switch closed and one switch open. Under these
conditions, the on-state device is pinned at IDC = Imax and VDS = VDD , and all the DC
power is dissipated in the switch — no power is delivered to the load since the switches are
not switched.
The preceding discussion implicitly links the choice of the load resistance RL to a condition
that constrains IDC to be less than or equal to Imax for the limits of the duty cycle range in
the pulse train. Another way to visualize the relationship between RL , load power, and power
efficiency is shown in Figure 2.12. With reference to equation (2.7), load power depends on
the load resistance, the DC current supplied to the circuit, and the duty cycle. Additionally,
we have equation (2.25) which shows that IDC varies as a function of duty cycle. The two
equations can be combined to rewrite load power as
 2
2 VDD
PL = 2 RL 2 sin2 (α π) (2.27)
π π2
RL sin2 (α π) + Ron

If conduction losses were zero, then PL would be proportional to VDD 2 /R for a fixed duty
L
cycle α which is intuitively satisfying. Equation (2.27) is more general and includes conduction
losses, but it still leads to an inverse relation between load power and load resistance, as shown

48
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

IDS

Load
Line
Vgs,on
α = 0.3 α=0

Imax

I0
α = 0.5

Vgs,of f
0 VDS
0 VDD Vmax Vds,bd

Figure 2.11: DC IV operating region for a CMCD amplifier including margin for duty cycle
variation.

by the dashed green line in Figure 2.12. Note that the relations in the figure are plotted for a
specific value of duty cycle; in this figure, α = 0.5. The figure also includes the corresponding
power efficiency for the CMCD amplifier as load resistance varies. Using the expression for
PL and PDC it is easy to show that the drain efficiency is
1
η= . (2.28)
Ronπ2
1+ csc2 (απ)
2 RL
The solid blue trace in Figure 2.12 shows that power efficiency increases as load resistance
increases, albeit with diminishing gains in efficiency as RL gets large.
Although equation (2.27) shows how load power varies as function of load resistance, it
is derived from a level 1 device model and depends on the on-state resistance, Ron . A direct
application of the equation cannot be used for a practical device without coupling it with a
constraint on the maximum device current, Imax . In order to determine the range of load
resistance that can be used with the device, a second trace is added to the graph to add the
Imax constraint.
Returning to equation (2.7), the maximum value of IDC that is possible with a physical
device is Imax . If IDC = Imax , then PL = (2/π 2 )Imax
2 sin2 (απ) RL which gives a linear bound
for PL as a function of RL for a specific duty cycle. A plot of this bound is shown as the
dashed red trace in Figure 2.12. The constrained load resistance range is the segment of the
dashed green curve that lies to the right of the intersection of the two lines. The range is
highlighted by a thick black line. The point where the dashed red and dashed green lines
intersect is the resistance which maximizes load power, and for the data shown in the figure,
the optimum load resistance is 45 Ω. Therefore, the addition of the constraint that IDC ≤ Imax
with equation (2.27) leads to analytic results which are consistent with the level 2 model that
includes current saturation in the switch.
Capacitive and inductive switching losses would lead to further adjustment in the final
value of load resistance, but equation (2.27) provides a good choice to begin optimizing a
CMCD design. As a final note, in a practical design, equation (2.27) should be evaluated

49
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

100 100

80 80

Efficiency (%)

Power (W)
60 60

40 40

20 20

0 0
0 20 40 60 80 100
Load Resistance (Ω)

Figure 2.12: Efficiency and output power of a CMCD as a function of load resistance (α = 0.5).

for the endpoints of the duty cycle range which is in the pulse train, since IDC increases for
non-50% duty cycles. An example is given in Figure 2.13 for a 30% duty cycle, and compared
to the 50% duty cycle case, the optimum load resistance is increased to approximately 65 Ω.

100 100

80 80
Efficiency (%)

Power (W)

60 60

40 40

20 20

0 0
0 20 40 60 80 100
Load Resistance (Ω)

Figure 2.13: Efficiency and output power of a CMCD as a function of load resistance (α = 0.3).

50
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

2.3.5 Current Saturation Model


As discussed in the previous section, a variable duty cycle periodic pulse train leads to
a range of on-state switch currents that fall in the range from Io to Imax . Over this range,
the device is transitioning from the linear region to the saturation region and an improved
model can be used to predict how IDC changes as duty cycle changes. The transition region
is highlighted in Figure 2.11 and can be modeled by the following equation

IDC ≈ Imax − (Imax − Io ) exp(4 [1 − csc(απ)]). (2.29)

The exponential function is scaled such that it is consistent with the approximation that
e−4 ≈ 0, and the function provides a smooth transition from Io to Imax . An example of how
this function models the change in IDC as a function of duty cycle is shown in Figure 2.14.

Figure 2.14: DC device current as a function of sin(απ) (α is duty cycle).

2.3.6 Analytical versus Simulated Results


The analytical theory which has been developed for the CMCD amplifier is now used
to compare the predicted power efficiency of the amplifier with simulation results. For the
comparison, a CMCD amplifier is designed using the Cree CGH60015D GaN HEMT. The
schematic diagram of the circuit is the same as the circuit shown in Figure 2.1 and the
component values are given in Table 2.2.
A summary of the analytical results of different power loss mechanisms versus duty cycle
is shown in Figure 2.15. The analytical results are compared with simulation results obtained
using the Cree large signal device model. For the simulated results, load power and DC source
power can be measured but no further details on the breakdown of individual loss mechanisms
can be isolated from the simulation results. This is where analytical results provide additional
insight, and the relative magnitude of different power losses can be compared. Conduction
losses and inductive switching losses are relatively insensitive to changes in duty cycle. On

51
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

Table 2.2: CMCD Design Values

Switches M1 and M2 Cree CGH60015D


Driver resistance Rdrv 5Ω
Drain supply voltage VDD 25 V
Off-state gate voltage Vgs,of f -4 V
On-state gate voltage Vgs,on 0V
Saturation current Imax 2.3 A
Duty cycle range (α) 30-70%
50% duty cycle switch current Io 1.7 A
Equivalent on-state resistance Ron 2Ω
Load resistance RL 65 Ω
Fundamental switching frequency fo 1 GHz
Drain bias inductors LDC1 and LDC2 100 nH
Tank Q 5
Tank capacitance CT 12 pF
Tank inductance LT 2 nH
Overlap interval τ 0.1 T

the other hand, capacitive switching losses, which are often neglected in CMCD analyses,
becomes quite significant particularly as the duty cycle deviates from 50%. As shown, when
all the individual loss mechanisms and the load power are summed together, the predicted
DC power is very close to the simulated results. Note that in this figure, all the powers are
normalized to the analytic value for PDC . Some of the simulated values for DC power are
slightly higher than the analytic value, therefore, the normalized value for some simulated
values are slightly greater than 100%.
Other conclusions from the analytical and simulated results are that the conduction loss
is the dominant loss mechanism for duty cycles in the range of 30% to 70%, and for very
low duty cycles (or very high) in the range of 20-30% (70-80%) capacitive switching losses
are significant. From this we can conclude that if it were possible to design a general pulse
encoder for CMCD amplifiers, it would be desirable to constrain pulse duty cycles to be in
the range of 30-70% to maintain good power efficiency. This conclusion is supported by the
power efficiency versus duty cycle data shown in Figure 2.16 where we see power efficiency
ranges from 55% to 78% for a duty cycle range of 30-70%.

52
2.3. Power Efficiency Analysis of a CMCD Amplifier with Periodic Switching

110
100

Normalized to Total Losses (%)


90
80
70 PL
60 PRon
Pcapacitive
50
Pinductive
40 Poverlap
PDC
30
PL: Simulation
20 PDC: Simulation

10
0
0.1 0.2 0.3 0.4 0.5 0.6
Duty Cycle (α)

Figure 2.15: Losses in the CMCD amplifier as a function of duty cycle. The overlap period τ
is 0.1 T .

Figure 2.16: Drain efficiency of CMCD amplifier as a function of duty cycle.

53
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

2.4 Predicting CMCD Amplifier Power Efficiency for Pulse


Encoded Signals
2.4.1 Time Encoded Input Signal
When the input pulse train to a CMCD amplifier is a periodic pulse train with a duty cycle
α, the load signal is a sinusoid at the fundamental frequency of the pulse train and the peak
amplitude is determined by the duty cycle. Although the duty cycle can be adjusted to create
an amplitude varying load signal, for a general modulated RF carrier signal, the spectrum of
the pulse train needs to be shaped to ensure that a modulated load signal with low distortion
can be delivered to the load. Examples of pulse encoders that can generate suitable pulse
trains for amplification by a CMCD amplifier are bandpass sigma-delta modulation (SDM)
and noise shaped pulse position modulation (PPM). The time domain and frequency domain
properties of SDM and PPM were briefly reviewed in Chapter 1.
As a way to predict the power efficiency of a CMCD amplifier when the input signal is a
general pulse encoded signal, CMCD power efficiency is evaluated for more complex periodic
signals. The motivation to use longer periodic sequences is to mimic the type of level crossings
that would be non-periodic pulse trains such as SDM or PPM. In the following sections, the
power efficiency of a CMCD amplifier is analyzed for periodic signals of different lengths. The
analytical results are then compared with simulations of a pulse encoded CMCD amplifier
which has non-periodic switching.

2.4.2 Power Efficiency for 1 T Periodic Signals


Analytic results have been developed in earlier sections for predicting the power efficiency
of a CMCD amplifier when the input signal is a periodic signal with duty cycle α. The signal
has a period of T and example pulse train is shown in Figure 2.17. By adjusting the duty,
the load power changes and so does the power efficiency. An example of the power efficiency
versus load power characteristic for a signal with period of T is shown in Figure 2.18. The
same CMCD amplifier design described in section (2.1) is used to obtain these results. The
peak load power is approximately 30 W and corresponds to a 50% duty cycle. The power
efficiency of the CMCD amplifier is approximately 75% at peak load power. As the duty cycle
decreases, load power decreases, there is a corresponding reduction in power efficiency because
conduction, inductive, and overlap losses are approximately independent of load power, while
capacitive switching losses increase as load power decreases. At 3 dB back-off the power
efficiency has dropped to 27%.
Amplitude

α1

T Time

Figure 2.17: Signal with period 1T.

54
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

100
Periodic Signal
90 SDM Signal
80

70

Efficiency (%)
60

50

40

30

20

10

0
0 5 10 15 20 25 30 35
Power (W)

Figure 2.18: Comparison of power efficiency for a CMCD amplifier with a 1 T periodic signal
and SDM non-periodic signal.

The results for the periodic signal model are now compared with simulated results for
bandpass sigma-delta modulator pulse trains. In Figure 2.18, a second response is shown
for the modulator pulse train. The simulation results are obtained using the SDM encoder
described in section 1.2.1. The input source signal to the encoder is a sinewave, and the
amplitude of the sinewave is swept to change output load power. Three observations are
made when comparing the 1T periodic signal response with the SDM response. First, for the
power range where the two responses overlap (approximately 21 W to 28 W), the two responses
are nearly identical. Second, the 1T response corresponds to a duty cycle range of 20-50%,
and clearly this sequence length cannot generate a load power below 21 W. Longer sequences
are required to generate a larger range of load powers. Third, the peak load power with the
periodic signal is higher than SDM. The reason for this is that a square wave with a 50% duty
cycle has maximum power at the fundamental frequency and an SDM pulse encoder usually
overloads before reaching a periodic limit cycle. Therefore, the peak power with an SDM
modulator is less than the total available peak power which can be obtained with a periodic
signal. This limitation led to the development of the PPM noise shaped encoder described in
section 1.2.2 which has higher coding efficiency. A comparison of the SDM and PPM encoders
is shown in Figure 2.19. For this figure, the x-axis corresponds to the peak amplitude of the
source signal relative to the quantization amplitude in the encoder. As shown, the PPM
encoder can overload to a 50% duty cycle square wave where the fundamental frequency
component has an amplitude of 4/π ≈ 1.27, exceeding the quantizer amplitude which is
unity.

55
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

100
SDM
PPM
80

Efficiency (%)
60

40

20

0
0 0.2 0.4 0.6 0.8 1 1.2
Modulator Drive Relative to Full Scale

Figure 2.19: Drain efficiency as a function of modulator drive level for SDM and PPM en-
coders.

2.4.3 N T Periodic Signals


Up to this point, we have considered periodic signals with a duty cycle of α and period T
as shown in Figure 2.17. If a pulse train with period 2T is considered as shown in Figure 2.20,
then it can be described in terms of two duty cycles α1 and α2 . The corresponding load signal
in a CMCD amplifier is determined by the amplitude of the second harmonic in the 2T pulse
train. This process can in general be extended to N T where the pulse train is parameterized
in terms of N duty cycles.
Amplitude

α1 α2

Time
T 2T

Figure 2.20: Signal with period 2T.

One drawback of the 1T and 2T pulse trains in Figures 2.17 and 2.20 is that there is
a variable DC level that depends on the duty cycles. For bandpass source signals, there is
no DC component and pulse encoders typically synthesize pulse trains with zero mean DC
components. One way to systematically create zero mean pulse trains is to alternate adjacent
intervals in the pulse train with a duty cycle of 1 − α. In this way, the sequence length doubles
for N duty cycles. An example of a 6T signal with N = 3 is shown in Figure 2.21. For the
6T signal, the amplitude of the sixth harmonic determines the amplitude of the load signal

56
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

in the CMCD amplifier.


Amplitude
α1 1 − α1 α2 1 − α2 α3 1 − α3

T 2T 3T 4T 5T 6T Time

Figure 2.21: A 6T signal with a zero mean DC component.

2.4.4 Power Efficiency Analysis for a 2T Periodic Signal


Assuming zero-mean pulse trains, we now consider how to predict the power efficiency of
a CMCD amplifier if the input pulse train is a 2T signal. There are two duty cycles, and for a
zero-mean signal, let α1 = α and α2 = 1−α. The parallel resonant filter in the CMCD output
is tuned to extract the second harmonic of the waveform. From a Fourier series analysis, the
second harmonic of this pulse can be expressed as
 
2
Iom2T = IDC2T sin2 (απ). (2.30)
π

The output power is then


1 2 2 2
PL2T = RL Iom = 2 RL IDC sin4 (απ). (2.31)
2 2T
π 2T

Using (2.31) and (2.7), the ratio of the output power of the 2T signal relative to the output
power of a 1T signal is
PL2T
≈ sin2 (απ). (2.32)
PL1T

The equation is written as an approximation because it based on the assumption that the
DC current drawn by the amplifier is the same for the 1T and 2T signals. Simulation results
verify that this is a good assumption as will be shown shortly.
Using equation (2.32), an expression for the relative power efficiency of the CMCD ampli-
fier when amplifying 1T and 2T is

η2T (PL2T /PDC2T )


= ≈ sin2 (απ). (2.33)
η1T (PL1T /PDC1T )

Again, the assumption that the DC currents for the two cases are the same is used to conclude
that the DC power for the two cases is the same (recall that VDD is fixed in a CMCD amplifier).

57
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

What the power efficiency ratio enables is an extension to lower load power levels, because
the second harmonic amplitude in a 2T pulse train with duty cycle parameter α is smaller than
the fundamental frequency amplitude in a 1T pulse train with a duty cycle of α. Equation
(2.33) is used to extend the power efficiency analysis of CMCD amplifiers from 1T to 2T . An
example of analytical and simulated results for 2T signals is shown in Figure 2.22. As the
results show, there is good agreement between the analysis and simulation which confirms the
assumption that the DC current in the CMCD amplifier is approximately equal for both the
1T and 2T signals.

100
Calculated results
Simulated results
80
Efficiency (%)

60

40

20

0
0 0.2 0.4 0.6 0.8 1
Duty Cycle

Figure 2.22: Drain efficiency of CMCD amplifier as a function of duty cycle when driven with
a 2T periodic signal.

2.4.5 Power Efficiency for 6T Periodic Signals


Longer periodic sequences can be generated to increase the amplitude range of the load
signal. As an example, a zero mean 6T signal is considered next. There are six duty cycle
parameters, but after imposing the zero-mean constraint, there are three independent duty
cycles. The corresponding duty cycles in the 6T pulse train are α1 , 1 − α1 , α2 , 1 − α2 , α3
and 1 − α3 as shown in Figure 2.21. Different combinations of duty cycles, lead to different
sixth harmonic amplitude levels and twenty different sequences are tabulated in Table 2.3.
Circuit simulations of a CMCD power amplifier were run with all the different 6T sequences
in Table 2.3 and the corresponding power efficiency is shown Figure 2.23. The simulation
results are compared with SDM and PPM pulse trains as well as with 1T and 2T periodic
sequences described earlier. Note that all periodic sequences have been constrained to have
pulse widths that are at least 0.2T or greater. From the results, it is clear that simulations
results with periodic sequences are very consistent with results for more general pulse encoded
signals like SDM and PPM. From these results we conclude that analytic results obtained for
1T and 2T periodic signals can lead to good predictions of the performance of a CMCD
amplifier for more general pulse trains like SDM and PPM.

58
2.4. Predicting CMCD Amplifier Power Efficiency for Pulse Encoded Signals

Table 2.3: Duty cycles for generating signals with a period of 6T.
Sequence (1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
α1 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2
α2 0.2 0.2 0.2 0.2 0.3 0.3 0.3 0.4 0.4 0.5
α3 0.2 0.3 0.4 0.5 0.3 0.4 0.5 0.4 0.5 0.5
Input Amplitude 0.44 0.57 0.68 0.72 0.7 0.81 0.85 0.91 0.95 0.98
Sequence (11) (12) (13) (14) (15) (16) (17) (18) (19) (20)
α1 0.3 0.3 0.3 0.3 0.3 0.3 0.4 0.4 0.4 0.5
α2 0.3 0.3 0.3 0.4 0.4 0.5 0.4 0.4 0.5 0.5
α3 0.3 0.4 0.5 0.4 0.5 0.5 0.4 0.5 0.5 0.5
Input Amplitude 0.83 0.94 0.98 1.04 1.08 1.11 1.15 1.17 1.21 1.27

100

90 Solid: SDM

80 Dash: PPM
Square: 1T−Signal
70
Efficiency (%)

Circle: 2T−Signal
60 Triangular: 6T−Signal
50

40 2T − Signal
6T − Signal
30 α=0.2
α=0.2
20
1T − Signal
10

0
0 5 10 15 20 25 30 35
Power (W)

Figure 2.23: CMCD amplifier power efficiency for periodic (1T , 2T , and 6T ) and non-periodic
pulse trains (SDM and PPM).

In wireless communication applications, the input signal is a modulated carrier with a


large PAPR. Therefore, the average power point is backed off relative to peak power. With
reference to Figure 2.23, peak load power is approximately 30 W and the power efficiency
is about 75%. At 3 dB back-off, the output power is 15 W and the corresponding power
efficiency is reduced to 28%. This shows that there is a steep reduction in power efficiency
as soon as the output power is backed off. The main reason for the reduction in efficiency is
the range of duty cycles in the pulse train. If the duty cycle variation could be minimized,
then higher efficiency could be obtained. For example, with reference to Figure 2.22, a duty

59
2.5. Chapter Summary

cycle range from 35-65% would maintain power efficiency at 50% or higher. Therefore, these
results show the importance of the signal mapping stage in a RF switch-mode power amplifier
system.

2.5 Chapter Summary


Three different device model levels were introduced to analyze the power efficiency of a
CMCD power amplifier. The level 1 model is a simple switch and a model that is commonly
used in the literature. The level 2 model introduced in this work extends the level 1 model
to include current saturation limitations in the switch as well as overlap losses. The level 3
model goes further and includes nonlinear device capacitances, and the model will be used
more extensively in the next chapter.
The level 2 model was used to analyze the power efficiency of a CMCD amplifier under
variable duty cycle switching conditions. A thorough analysis was presented for 1T periodic
signals. The analytic results were compared with simulations of a CMCD amplifier using a
SDM pulse train. The two results are very similar which shows that the analysis models the
behaviour of the amplifier under more general non-periodic switching conditions. The analysis
was extended to 2T periodic signals. Again there is a good match between the predicted and
simulated power efficiency over a large range of duty cycles. Simulation results were then
made to compare the power efficiency of CMCD amplifiers with longer periodic sequences up
to 6T . As with 1T and 2T , the 6T results closely match the performance of SDM and PPM
pulse trains.
The power efficiency of the CMCD amplifier is very dependent on the duty cycle variation
in the pulse train. As the analytical and simulation results show, power efficiency reduces
quickly as the load power is reduced. The results show that signal mapping in the pulse
encoder and the design of the switch-mode power amplifier are closely linked. In the next
chapter, class-F amplifiers are designed and later in Chapter 5, the class-F amplifiers are
tested with pulse encoded signals.

60
Chapter 3

Class-F RF Power Amplifiers


Class-F amplifiers use harmonically tuned input and output matching networks to shape
the waveforms at the gate and drain of the amplifying device. The goal of harmonic wave-
shaping is to reduce the overlap of the drain current and drain voltage waveforms to obtain
a high efficiency amplifier. Similar to the class-D amplifier, class-F amplifiers can approach a
switch-mode type of operating condition providing the amplifiers are operated near saturation.
There is a large body of work related to class-F amplifier design and the contributions
to this work are the following. First, a systematic study of the effect of input harmonic
termination impedance is made for a class-F amplifier. The study uses a level 3 device
model and the relative importance of first, second, and third harmonic input impedances are
analyzed. The comparison includes third harmonic terminations at the input which is new.
Experimental results for a class-F amplifier including third harmonic networks for both the
input and the output are also shown. Second, an inverse class-F (class-F−1 ) amplifier design
is implemented using exactly the same device as the class-F amplifier, and the two amplifier
designs are compared. The comparison is useful as it is more difficult to find reference designs
in the literature which provide a direct compare between the voltage switched class-F amplifier
and the current switched class-F−1 . The experimental results clearly show that the current
switched class-F−1 circuit has higher power efficiency over a larger dynamic range. The class-
F amplifier designs are also very useful benchmarks for Chapter 4, where the amplifier are
reconfigured into RF rectifiers. The chapter concludes with the design of a wideband class-
F−1 . Although experimental results for wideband class-F−1 can be found in the literature, the
design presented here is reconfigured into a wideband rectifier in Chapter 4, and the wideband
rectifier work which builds on the wideband amplifier design is new work.

3.1 Level 3 Device Model


All the experimental circuits in this thesis use a packaged 10 W GaN HEMT (model
CGH40010) available from Cree. Although a comprehensive large signal model for the device
is available from Cree, it is difficult to use this model to gain insight into the underlying
mechanisms that generate harmonic frequency components at the gate terminal of the device.
Also, the black-box Cree model does not enable any way to tune device parameters and
evaluate the sensitivity of the circuit with respect to different device parameters. Therefore, a
level 3 device model is first constructed from a large signal device model for a Cree CGH60015D
die. After building a model of the die, the model is then extended to include package parasitics
since a packaged device is used for all the experimental work. The packaged device (model
CGH40010) uses the CGH60015D die which is wire bonded to the package leads. The methods
used to extract the equivalent circuit values for the level 3 model are reviewed next.
2
Parts of Chapter 3 have been published in two articles. Reprinted with permission from PIER and EuMC
[56, 76].

61
3.1. Level 3 Device Model

Lg rg Ld
rd
G’ Cgd D’
G D
Cgs I Cds

V Intrinsic Device
S’ Model

rs

Ls
S

Figure 3.1: Level 3 equivalent circuit model for GaN HEMT Cree CGH60015D [reproduced
courtesy of The Electromagnetics Academy].

3.1.1 Bare Die Device Model


A level 3 device model is shown in Figure 3.1. The equivalent circuit for the device is
partitioned into an intrinsic device model composed of a current source and nonlinear device
capacitances (Cgs , Cgd and Cds ), and extrinsic inductances and resistances associated with
the interconnect to the intrinsic device. The values for the components in the equivalent
circuit model are obtained from Z and Y -parameters for the large signal device model. The
extraction process is described next.
Let Z 0 be the Z parameters for the intrinsic device. The corresponding Z parameters for
the overall device model are [88]
   0 0 + r + jL ω

Z11 Z12 Z11 + (rg + rs ) + j(Lg + Ls )ω Z12 s s
= 0 + r + jL ω 0 + (r + r ) + j(L + L )ω . (3.1)
Z21 Z22 Z21 s s Z22 d s d s

Therefore, once the extrinsic interconnect values for the series inductances and resistances are
known, we can find the Z-parameters for the intrinsic device. After Z 0 is obtained, the matrix
can be inverted to find the admittance parameters Y 0 . The individual admittance parameters
can be matched to the Π capacitance network and lead to the following equations [89].

0 ]
−Im[Y12
Cgd = (3.2)
ω

0 +Y0 ]
Im[Y11 12
Cgs = (3.3)
ω

0 +Y0 ]
Im[Y22 12
Cds = (3.4)
ω

62
3.1. Level 3 Device Model

3.1.1.1 Extraction of the Parasitic Inductances and Resistances


At this point, we need to find a way to determine the extrinsic circuit component values.
An important observation is that the series inductances and resistances are independent of
the bias point of the device, while the nonlinear device capacitances in the intrinsic model are
bias dependent. Therefore, a bias point can be selected that simplifiers the extraction of the
extrinsic component values.
A good choice for the extraction, is the off-state bias condition where VGS = VGSOF F and
VDS = 0 V. Under these conditions, the transconductor in the intrinsic model is short circuited
and nodes S 0 and D0 are shorted, as shown in Figure 3.2. Under the off-state bias condition,
the real part of the Z-parameters is directly related to the resistances in the equivalent circuit
and we have the following relations:

 Re[Z11 ] = rg + rs
Re[Z22 ] = rd + rs (3.5)
Re[Z12 ] = Re[Z21 ] = rs

From these equations, all the resistances can be uniquely found from Z-parameters for the
off-state. Also note that the real part is frequency independent.
Lg rg Ld
rd
G’ Cgd D’
G D
Cgs Cds

Intrinsic Device
S’ Model

rs VDS = 0
shorts nodes S’ and D’
Ls
S

Figure 3.2: Equivalent circuit model for off-state bias conditions.

The Z-parameters for the off-state were obtained from the large signal Cree model and
the results are shown in Figure 3.3. The real and imaginary components of the Z-parameters
are shown by the solid lines for a frequency range of 0.5 to 10 GHz. The data clearly show
the real component is constant over frequency, consistent with equation (3.5). The equations
were used to calculate the equivalent circuit resistances and the values are optimized in a final
step after all component values are extracted for the level 3 model.
The imaginary component of the off-state Z-parameters is more entangled and contribu-
tions to reactance are made both from the series lead inductances as well as the intrinsic
devices capacitances. The entanglement is simplified by using Z-parameters at very high fre-
quencies where the inductance starts to dominate the response and the intrinsic capacitances
have low reactance. The inductive behaviour of the reactance is evident in the Z-parameter
measurements shown in Figure 3.3 where the slope at high frequencies is related to the in-

63
3.1. Level 3 Device Model

ductance. Clearly, in the limit as ω gets large the Z-parameters are approximated as

 Z11 ' (rg + rs ) + j(Lg + Ls )ω
Z22 ' (rd + rs ) + j(Ld + Ls )ω (3.6)
Z12 = Z21 ' rs + jLs ω

These approximations are used to estimate values for the extrinsic inductances.

Figure 3.3: Z-parameters for the level 3 device model (symbols) and for the large signal device
model (solid lines) for the off-state bias condition.

3.1.1.2 Off State Device Model


Once the extrinsic inductances and resistances are found, the Z-parameters of the off-state
can be de-embedded to extract the intrinsic Z 0 -parameters. The instrinsic Z 0 -parameters are
then inverted to find the admittance matrix Y 0 from which values for the devices capacitances
can be found using equations (3.2), (3.3) and (3.4). After calculating initial values for all the
model components, a final step to optimize the model values is made to provide the best match

64
3.1. Level 3 Device Model

between the level 3 model and the Cree model. The Y -parameters are particularly useful
for optimizing the final values because they are sensitive to all the device characteristics. A
comparison of the off-state Y -parameters for the level 3 model compared to the Cree model are
shown in Figure 3.4. As shown, there is good agreement between the two models. A summary
of model 3 equivalent circuit component values for the off-state is shown in Table 3.1.

Figure 3.4: Y -parameters for the level 3 device model (symbols) and for the large signal device
model (solid lines) for the off-state bias condition.

3.1.1.3 Extraction of Nonlinear Device Capacitances


The intrinsic device capacitances are nonlinear, having dependencies on the gate-source
voltage and the drain source voltage. Now that the extrinsic inductances and resistances are
known in the level 3 model, parametric sweeps of the de-embedded Y 0 -parameters can be used
to extract models of the device capacitances over a wide range of bias conditions. It should be
noted, that when VDS is no longer zero, the model includes a voltage controlled current source
which models the transconductance of the device. At low frequencies, the transconductance

65
3.1. Level 3 Device Model

Table 3.1: Level 3 model values for the Cree GaN HEMT (CGH60015D) in the off-state bias
condition.

Element rg rd rs Lg Ld Ls Cgs,of f Cds,of f Cgd,of f


(Ω) (Ω) (Ω) (pH) (pH) (pH) (pF) (pF) (pF)
Value 0.62 0.6 0.1 91.8 88.65 1 4.1 0.87 0.6

Table 3.2: Summary of device capacitances for a Cree GaN HEMT (CGH60015D).

Model Model Datasheet


Vgs =-5 V to 0 V , Vds =0 V to 60 V Vgs =-8 V and Vds =28 V Vgs =-8 V and Vds =28 V
Capacitance Min. Max. Linear Value Typical
Cds 0.87 pF 2.4 pF 0.92 pF 0.87 pF 0.9 pF
Cgd 0.1 pF 0.7 pF 0.36 pF 0.19 pF 0.2 pF
Cgs 4.14 pF 7.65 pF 6.17 pF 4.14 pF 4.1 pF

term is associated with the real part of the Y 0 -parameters and the imaginary part of the Y 0 -
parameters can be used to find the intrinsic device capacitance values similar to the procedure
used in the off-state.
The bias points which were used to extract the nonlinear device capacitances span a VGS
range from -8 to 0 V (-8, -4, -2, -1, 0 V) and a VDS range from 0 to 80 V (0, 2, 4, 6, 8, 10,
20, 28, 40, 50, 60 and 80 V). The best fit values were optimized for a frequency of 1 GHz
and the device capacitance characteristics are shown in Figures 3.5 (a)-3.5 (d). Notable
characteristics include significant variation in Cds for low drain voltages, a significant change
in Cgs as the gate voltage swings between on and off states, and nonlinear Cgd characteristics
that depend on both the gate and drain voltages. The variation in the device capacitances
over the operating range of the device are summarized in Table 3.2. The table also includes
the nominal device capacitances given on the Cree datasheet for the die [90].
In order to implement the level 3 model in a circuit simulator, nonlinear circuit elements
are required for the device capacitances and the transconductance to model the IV curves
of the device. In ADS, nonlinear capacitor models are available that can be linked to look
up tables. In this way, the instantaneous capacitance is dependent on the instantaneous
gate-source and drain source voltages in the circuit. The nonlinear transconductance, which
models the IV characteristics of the device, is implemented using a special component called a
symbolically defined device (SDD). Similar to the capacitors, the SDD block is controlled by
a look-up table that models the drain source current dependency on gate-source and drain-
source voltages.

3.1.1.4 Level 3 Model for Die and Package


A packaged die is used for experimental work and a level 3 model for the device including
a package is shown in Figure 3.6. It consists of the level 3 bare die model shown earlier in

66
3.1. Level 3 Device Model

2.5 8
Vgs= −2
Vgs= −2 V,
V, −1
−1 V,
V, 00 V
V

7.5
Drain−Source Capacitance (pF)

Gate−Source Capacitance (pF)


2 7

6.5
Decreasing order
Vgs=0 to −5 Vgs= −3
Vgs= −3 V
V
1.5 6

5.5

1 5
(b)
4.5
(a) Vgs= −8
Vgs= −8 V,
V, −4
−4 V
V

0.5 4
0 5 10 15 20 25 30 0 10 20 30 40 50 60 70 80
Vds (V) Vds (V)

0.8 8

0.7 7.5

Gate−Source Capacitance (pF)


Gate−Drain Capacitance (pF)

0.6 Decreasing order 7


Vgs= 0 to −5
0.5 6.5

0.4 6

0.3 5.5

0.2 5

0.1 4.5
(c) (d)
0 4
0 10 20 30 40 50 60 70 80 −8 −7 −6 −5 −4 −3 −2 −1 0
Vds (V) Gate−Source Voltage (V)

Figure 3.5: Extracted intrinsic device capacitances for the Cree GaN HEMT (CGH60015D):
(a) drain-source capacitance, (b) gate-source capacitance, (c) gate-drain capacitance, and (d)
gate-source capacitance versus gate-source voltage.

Figure 3.1 with additional lead inductances (Lpg , Lpd ) and capacitances (Cpg , Cpd ). Similar
to the method used to find extrinsic inductances, the package inductances can be found from
Z-parameters measured at high frequencies where the response is dominated by the package
inductance and the package capacitance approaches a low impedance. After finding estimates
of the package inductances, the package capacitances in the level 3 model are tuned to match
the Z-parameters of the Cree large signal model for the packaged device (CGH40010F). The
extracted values for the package model are found to be: Lpg = 0.7 nH , Lpd = 0.6 nH,
Cpg = 0.41 pF and Cpd = 0.4 pF.
As a confirmation of the level 3 model, the Y -parameters of the model level 3 model
are compared to the Cree model. The results are shown in Figures 3.7 through 3.9. Good
agreement is obtained and the level 3 model is used in the next section to study the effect of
input harmonic termination impedances in a class-F amplifier.

67
3.1. Level 3 Device Model

Lpg Lg rg Cgd rd Ld Lpd

Cpg Cgs Cds Cpd


V

rs
Bare Die
Ls

Figure 3.6: Device model for packaged die [reproduced courtesy of The Electromagnetics
Academy].

Figure 3.7: Comparison of Y11 parameters for the level 3 model including the package (sym-
bols) and the GaN HEMT Cree large signal model for the CGH40010F (solid line). The
device bias conditions are in the off-state.

68
3.1. Level 3 Device Model

Figure 3.8: Comparison of Y12 parameters for the level 3 model including the package (sym-
bols) and the GaN HEMT Cree large signal model for the CGH40010F (solid line). The
device bias conditions are in the off-state.

Figure 3.9: Comparison of Y22 parameters for the level 3 model including the package (sym-
bols) and the GaN HEMT Cree large signal model for the CGH40010F (solid line). The
device bias conditions are in the off-state.

69
3.2. Class-F Amplifier Simulation Experiments

3.2 Class-F Amplifier Simulation Experiments


Class-F amplifiers are inherently nonlinear and use harmonic impedance control in the
output circuit to shape the voltage waveform at the drain. In an ideal class-F design, the
device is cut-off for half the cycle similar to class B, and harmonics are created in the output
circuit because the drain current is a half sinusoid. Harmonic termination impedances in the
output circuit preserve the shape of the current waveform by shorting even harmonics. The
drain voltage waveform is shaped to reduce overlap with the current waveform by open circuit
termination impedances at odd harmonics.
There are however other mechanisms in a class-F amplifier that create additional harmonic
frequency components in the current and voltage waveforms. Examples include deviations
from the class-B bias point which means the current waveform is no longer a perfect half-
sine waveform, and nonlinear device capacitances [19] and nonlinear device transconductances
which create new harmonic frequency terms even for perfectly sinusoidal input signals. Of
these mechanisms, nonlinear gate-source capacitance Cgs and nonlinear gate-drain capacitance
Cgd can directly create harmonic components at the gate. There are also feedback mechanisms
including Cgd and source inductance which can couple drain harmonics to the gate. For
example, an imperfect second harmonic termination in the output match creates a second
harmonic voltage component that can be coupled to the gate through Cgd . The effect of
harmonic signal components at the gate is that the gate waveform is no longer sinusoidal even
if the input to the amplifier is a perfect sinusoid. Changes to the gate waveform, affect the
drain current waveform, and consequently the drain current waveform deviates from an ideal
half sinusoidal shape. Therefore harmonic termination impedances at the gate can be used
to shape both the gate waveform and the drain current waveform.
The relationship between power efficiency and input harmonic impedances for class-F
amplifiers has been studied by others [91, 92, 93, 94, 95, 96, 97] and a summary is presented
in Table 3.3. Most work listed in the table has focused on evaluating the effect of fundamental
and second harmonic impedance terminations at the input of the device. Also, in these papers,
the mechanisms that generate harmonics in the device have been less important, and the work
has focused on investigating harmonic termination impedances using harmonic load pull and
time domain techniques. Although these techniques are general and can be extended to all
harmonics, the experimental work presented in these papers has been limited to circuits with
input harmonic terminations for fundamental and second harmonic impedances.
This work extends the investigation of input harmonic impedance to include the effect
of third harmonic terminations on the power efficiency of a class-F amplifier. A systematic
comparison of the power efficiency for amplifier designs with fundamental, second harmonic,
and third harmonic terminations are made. Simulations with the level 3 device model are
used to investigate the relationship between device capacitances and the harmonic levels at
the gate. The impact of the gate to drain capacitance, Cgd , is particularly important as
it provides feedback from the output to the input. The feedback, combined with imperfect
output harmonic termination impedances, can lead to the injection of both odd and even
harmonics to the gate terminal. The model is also used to investigate the sensitivity of the
design to nonlinear device capacitances. A comparison is made between a model with linear
capacitances and a model with nonlinear capacitances.

70
3.2. Class-F Amplifier Simulation Experiments

Table 3.3: Class-F amplifier designs with input harmonic termination networks.
Ref. Type Device Input Harmonics Approach
[91] class-F GaN fo , 2fo harmonic tuning
−1
class-F
[92] class-F PHEMT fo , 2fo device level
class-B
[93] class-F AlGaAs/GaAsN fo , 2fo phase relationship
−1
class-F
[98] class-F GaAs MESFET fo , 2fo power balance
class-E
[95] class-F Power MESFET fo , 2fo load/source pull

[96] class-F Power MESFET fo , 2fo , 3fo phase relationship

[97] class-F PHEMT fo , 2fo , 3fo load/source pull

This class-F GaN HEMT fo , 2fo , 3fo device level


work

3.2.1 Class-F Amplifier Design


The complete level 3 simulation model with provisions up to third harmonic termination
impedances at the gate is shown in Figure 3.10. The amplifier is designed for a fundamental
frequency of 990 MHz. The drain bias (VDD ) is 30 V and the gate bias (VGG ) is -2.6 V. The
optimal load and source impedances at the fundamental frequency were found using source
and load pull [99] test benches in the simulator.
The transmission line elements TL101 and TL102 in Figure 3.10 implement the fun-
damental frequency input match, and transmission lines TL109 and TL110 implement the
fundamental frequency output match. The output matching circuit consists of three other
transmission lines, TL106, TL107, and TL108, which create a second harmonic short and
third harmonic open at the drain. The input matching circuit also has three transmission
lines, TL103, TL104 and TL105, which are used to experiment with different input harmonic
terminations at the gate.
Further optimization of the third harmonic output stub TL108 can be made to compen-
sate for the output reactance of the device at the third harmonic frequency. However, the
optimization of the output matching network with the level 3 device model is not the primary
focus here and we constrain the output network topology to be the same for all design studies
of the input matching network (see Figure 3.11). Instead, here we adjust the length of the
output harmonic stubs and vary Cgd to create different harmonic injection levels that are
fed back to the gate. In this way, the significance of the input harmonic terminations versus
power efficiency can be swept for different harmonic injection levels at the gate.

71
3.2. Class-F Amplifier Simulation Experiments

Var Var
Eqn VAR Eqn VAR
VAR_wavelengths VAR_bias V_DC
Lambda0=360 VGG=-2.6 SRC1
Lambda2=Lambda0/2 TLOC Rdamp=5 Vdc=VGG V
Lambda3=Lambda0/3 TL104
Z=Z0 Ohm
L
Var
VAR E=ELharmonic3
Eqn
C L3
VAR_network1 F=f GHz
C1 L=Lch nH
ELharmonic1=Lambda0/4 R=
C=Cb pF
ELharmonic2=Lambda2/4
ELharmonic3=Lambda3/4
ELinputmatching1=80.25

Ref
ELinputmatching2=47.3
Var TLIN TLIN
Eqn
TL102 TL105
VAR Z=Z0 Ohm Z=Z0 Ohm
VAR_signal E=ELinputmatching2 E=ELharmonic1 R
f=0.990 F=f GHz F=f GHz R3
R=Rdamp Ohm
gate
Pindbm=24.5

DC_Block
DC_Block1

NonlinC
Ref

Ref

L Cgd L
L1 R R L2
TLOC
P_1Tone
TL101 TLOC gate L=Lg pH R1
R=rg Ohm
R2
R=rd Ohm
L=Ld pH
drain
Z=Z0 Ohm TL103 R= R=
PORT1
E=ELinputmatching1 Z=Z0 Ohm
Num=5
F=f GHz E=ELharmonic2
Z=50 Ohm
P=dbmtow(Pindbm) F=f GHz
Freq=f GHz NonlinC NonlinC
Cgs Cds
SDD2P
HARMONIC BALANCE Var
VAR SDD2P1
Eqn
VAR_parameters I[1,0]=(_v1)*0
HarmonicBalance I[2,0]=ids Var
VAR
rg=0.6 Eqn
HB1 C[1]= VAR3
rd=0.6
Freq[1]=f GHz Cport[1]= vds=_v2
Lg=92
Order[1]=10 vgs=_v1
Ld=88

DataAccessComponent TLOC
DAC1 TL108
InterpMode=Linear Z=Z0 Ohm
DAC InterpDom=Rectangular V_DC
E=ELharmonic3
ExtrapMode=Interpolation Mode SRC2
F=f GHz
iVar1="VGS" Vdc=VDD V
iVal1=vgs C
iVar2="VDS" C2
Ref

L
iVal2=vds C=Cb pF
L4
Var
Eqn VAR L=Lch nH
VAR_IDS TLIN R=
TL106 TLIN
ids=file{DAC1, "IDS"}
Z=Z0 Ohm TL109
E=ELharmonic1 Z=Z0 Ohm
DC_Block
E=ELoutputmatching1
drain F=f GHz
F=f GHz
DC_Block2

R
R4
R=RL Ohm
Ref

Ref

Var
Eqn VAR
TLOC
VAR_network2 TLOC TL110
ELoutputmatching1=65.1 TL107 Z=Z0 Ohm
ELoutputmatching2=54.3 Z=Z0 Ohm E=ELoutputmatching2
RL=Z0 E=ELharmonic2 F=f GHz
Z0=50 F=f GHz

Figure 3.10: Schematic for the class-F PA [reproduced courtesy of The Electromagnetics
Academy].

72
3.2. Class-F Amplifier Simulation Experiments

TLout3
(λo /12)
ZOM N 50 Ω
TLout1 TLout4
(λo /4) TLout2
(λo /8) TLout5

Figure 3.11: Output matching network (OMN) structure [reproduced courtesy of The Elec-
tromagnetics Academy].

3.2.2 Harmonic Input Impedances and Sensitivity to Device Capacitances


For the first simulation experiment, the device model consists of fixed (linear) device
capacitance values. The fixed capacitance values are selected to be the expected value of the
nonlinear capacitances over the operating range of the circuit. The fixed capacitance values
are: Cgs = 6.17 pF, Cds = 0.92 pF and Cgd = 0.36 pF. When the device capacitances are
linear, the only mechanism for generating harmonics at the gate is feedback from the output
circuit to the input circuit through Cgd .
The effect of feedback through Cgd is illustrated by comparing the harmonic spectrums at
the drain and gate. In Figure 3.12, the spectrums are shown for the case where Cgd is zero.
The drain spectrum for the voltage has DC, fundamental and third harmonic components.
The second harmonic is zero because the harmonic is shorted in the output circuit. The cor-
responding gate spectrum has only a DC and fundamental frequency component as expected
because there is no feedback from the drain to gate. On the other hand, when Cgd is not zero,
harmonic components from the drain voltage are fed back to the gate. An example is shown
in Figure 3.13 for Cgd = 0.36 pF. The amount of harmonic feedback from the output to the
input of the device depends not only on the size of Cgd but also on the harmonic levels in the
output circuit.

Figure 3.12: Spectrum for the case where Cgd = 0 pF: drain voltage (left) and gate voltage
(right) [reproduced courtesy of The Electromagnetics Academy].

In the next set of simulation experiments, the device model with linear capacitances is used
to compare the power efficiency of three different class-F amplifier designs where each design

73
3.2. Class-F Amplifier Simulation Experiments

Figure 3.13: Spectrum for the case where Cgd = 0.36 pF: drain voltage (left) and gate voltage
(right) [reproduced courtesy of The Electromagnetics Academy].

has a different harmonic input matching network. The input matching circuits are shown
in Figure 3.14. In the first design (Design 1), the gate is matched only for the fundamental
frequency. In the second design (Design 2), the input matching circuit provides a fundamental
frequency match and a short at the second and third harmonics. In the third design (Design 3),
the input matching circuit provides a match at the fundamental frequency, a short at the

(a) in gate
TLin2

TLin1

TLin3
(λo /12)
(b) in gate
TLin2

TLin1 TLin4
(λo /8)

TLin3 TLin5
(λo /12) (λo /4)
(c) in gate
TLin2

TLin4
TLin1 (λo /8)

Figure 3.14: Input matching network circuits: (a) Design 1, (b) Design 2 and (c) Design 3
[reproduced courtesy of The Electromagnetics Academy].

74
3.2. Class-F Amplifier Simulation Experiments

Table 3.4: IMN transmission line lengths for a device model with linear capacitances.

Harmonic input impedances


Design T Lin1 T Lin2 T Lin3 T Lin4 T Lin5 Γin,2f o Γin,3f o
fo 2fo 3fo
1 ZS,opt - - 74.8◦ 30.9◦ - - - 1∠343◦ 1∠58.5◦
2 ZS,opt Short Short 77.4◦ 26.65◦ 30◦ 45◦ - 1∠180◦ 1∠180◦
3 ZS,opt Short Open 80.25◦ 47.3◦ 30◦ 45◦ 90◦ 1∠180◦ 1∠0◦

second harmonic and an open at the third harmonic. Table 3.4 summarizes the input matching
network (IMN) designs.
As shown in Figure 3.13, odd harmonics in the drain voltage are fed back to the gate
through Cgd . With linear device capacitances, and an ideal second harmonic short in the
output, there are no even harmonics at the gate. However, when device capacitances are non-
linear or when imperfect second harmonic terminations in the output circuit are considered,
there are even harmonic components at the gate as well. In the next simulation experiment,
the same device model with linear capacitances is used except the level of second harmonic
distortion at the gate is swept over a large range by tuning the length of TL107 to create an
imperfect second harmonic short in the output circuit; in other words, the second harmonic
phase at the drain is swept from 150 to 180 degrees. An imperfect second harmonic short
is created in practical class-F amplifier designs as soon as the frequency is shifted from the
design frequency.
The results of the second harmonic sweep for the three different input matching circuit
designs is summarized in Figure 3.15. It can be seen that the input matching circuit design
is significant when the second harmonic level at the drain is high — for example greater
than -20 dB relative to the fundamental frequency component. For a second harmonic level
of -11 dB, going from Design 1 (fundamental match only) to Design 2 the power efficiency
increases by 5.5%. A small but measurable improvement in power efficiency is obtained with
Design 3 when a third harmonic open is added to the gate matching network. If the second
harmonic level is low, for example less than -30 dB, the power efficiency of the different designs
are similar which is expected. From these results it is clear that incremental improvements
in power efficiency can be obtained with harmonic terminations at the gate and the most
significant improvement is gained by a second harmonic short with a smaller improvement
gained by adding a third harmonic open. As discussed earlier, the amount of harmonic
feedback from the drain to the gate depends on the size of Cgd . When there is no feedback,
the gate signal has only a fundamental frequency component and the performance of the three
designs are similar. On the other hand, as Cgd is increased, the significance of harmonic input
impedance becomes increasingly important.
In the next set of simulations, the results for a device model with linear capacitances
are compared with simulation results for a device model with nonlinear capacitances. With
reference to Figure 3.1, a level 3 device model is used with the nonlinear device capacitance
characteristics shown in Figure 3.5. Similar to the simulations for linear capacitances, the
effect of nonlinear devices capacitances is evaluated for three different amplifier designs, each
with a different input harmonic matching circuit as shown in Figure 3.14. The output match-

75
3.3. Class-F Amplifier Experimental Results

Table 3.5: Summary of simulation results for a device model with nonlinear capacitances.

Harmonic input impedances


Design T Lin1 T Lin2 T Lin3 T Lin4 T Lin5 ηd (%)
fo 2fo 3fo
1 ZS,opt - - 79.1◦ 31.9◦ - - - 77.5
◦ ◦ ◦ ◦
2 ZS,opt Short Short 80.0 21.9 30 45 - 83.2
◦ ◦ ◦ ◦ ◦
3 ZS,opt Short Open 81.9 47.4 30 45 90 83.7

ing circuit in these simulations has perfect harmonic terminations at the second and third
harmonic.
A comparison of the three different designs with nonlinear device capacitances is shown
in Table 3.5. The results are consistent with the results shown in Figure 3.15 for linear
capacitances in that the biggest improvement in power efficiency results from adding a second
harmonic short at the gate and a small improvement of less than 1% is obtained by adding
a third harmonic open at the gate. Also, the improvement in power efficiency going from
Design 1 to Design 2 with nonlinear device capacitances is larger than the result with linear
capacitances. For example, for linear device capacitances and a second harmonic level of
approximately -16.5 dB in Figure 3.15, the power efficiency of Design 1 is about 77.5%, similar
to the power efficiency of Design 1 with nonlinear capacitances (Table 3.5). If Design 2 values
are compared, the result for the linear capacitance model is about 79.5% or an increase of
2% compared to Design 1, while the result for the nonlinear capacitance model is 83.2%, an
increase of 5.7% compared to Design 1.
The device level modeling work also shows that the performance of an amplifier with linear
device capacitances has a power efficiency within a few percent of an amplifier with nonlinear
device capacitances. In other words, the harmonic injection by Cgd is very significant. Also,
the simulation experiments show that a second harmonic short at the input is very significant
in terms of improving power efficiency in a class-F amplifier. The second harmonic short at
the gate node desensitizes the design to second harmonic injection from both feedback through
Cgd as well as second harmonic components created by nonlinear device capacitances. A small
but slightly higher power efficiency can be obtained with an additional input third harmonic
termination.

3.3 Class-F Amplifier Experimental Results

A class-F amplifier with third harmonic input matching and output matching circuits was
built. The design uses a packaged Cree 10 W device and the level 3 model including the
package model was used for the preliminary design. Final optimization of the design used the
Cree large signal model.

3.3.1 Physical Circuit Design


The device package introduces parasitic inductance and capacitance that modifies the op-
timal harmonic termination impedances at the terminals of the device. The required harmonic

76
3.3. Class-F Amplifier Experimental Results

Second / First Harmonic Ratio at Gate (dB)


85 −20

80 −30
Efficiency (%)

75 −40

Design 3
70 −50
Design 2
Design 1

65 −60
−45 −40 −35 −30 −25 −20 −15 −10
Second / First Harmonic Ratio at Drain (dB)

Figure 3.15: Simulated drain efficiency as a function of second harmonic level for a device
model with linear capacitances [reproduced courtesy of The Electromagnetics Academy].

impedances at the terminal planes of the device were determined from a load pull test bench
in the simulator. The results are shown in Table 3.6 for a fundamental frequency of 990 MHZ
with a bias of VDD = 30 V and VGG = −2.6 V.
The harmonic matching networks for the class-F amplifier are shown in Figure 3.16 and
are designed using the impedance buffer methodology described in [100]. With reference
to Figure 3.16, the output matching network at reference plane A should provide optimal
load impedances of ZL (f0 ) at f0 , ZL (2f0 ) at 2f0 and ZL (3f0 ) at 3f0 . The corresponding
load reflection coefficients at these frequencies are given in Table 3.6. In the design, all
transmission lines have a characteristic impedance of 50 Ω except for T Lin5 and T Lout3 which
have a characteristic impedance of 36 Ω.
The synthesis of the output match begins with the second harmonic network. The trans-
mission line T Lout2 is 90◦ at the second harmonic frequency and creates a short at plane B.
Consequently the reflection coefficient looking to the right of plane B at the second harmonic
is 1∠180◦ . The addition of the series transmission line T Lout1 modifies the phase to provide
the match ΓL (2fo ) at plane A. The next step in the synthesis of the output circuit is to add
transmission lines T Lout3 and T Lout4 to create the third harmonic match. Transmission line
T Lout4 is 90◦ at the third harmonic frequency creates a short at reference plane C. The short
is transformed through transmission line T Lout3 to create the required reflection coefficient
ΓL (3fo ) at plane A. Also, the characteristic impedance of T Lout3 is reduced from 50 Ω to
36 Ω and selected to improve the bandwidth of the fundamental frequency match by trans-
forming the lower output impedance at the device terminal plane to a higher impedance close
to 50 Ω. The fundamental frequency output match is implemented with a double stub circuit
consisting of T Lout5 , T Lout6 , and T Lout7 . For this design the double stub circuit results in a

77
3.3. Class-F Amplifier Experimental Results

Table 3.6: Source and load pull harmonic impedances for the class-F amplifier.

Pin (dBm) Pout (dBm) P AE(%)


24.5 40.5 81.6
ΓL (f0 ) ΓL (2f0 ) ΓL (3f0 )
0.34∠142.67 1.0∠166 1.0∠329.38
ΓS (f0 ) ΓS (2f0 ) ΓS (3f0 )
0.44∠146.11 1.0∠170.2 1.0∠33.91

VGG VDD
ZF ZE C0 ZA ZB ZC C0

F LGG TLout4 C IDC LDD


TLin3 E A B
Cin (λo /12)
(λo /12) TLin5 TLin7 Cout
Rg
Vin TLin2 D TL TLout6 Vout
M1 out1 TLout3
TLin1 TLin4 TLin6 TLout2 TLout5 TLout7
(λo /8) ZD (λo /8)
Harmonic Input Matching Network Harmonic Output Matching Network

Figure 3.16: Schematic of the class-F power amplifier with output and input matching circuits
and bias networks [reproduced courtesy of The Electromagnetics Academy].

more compact fundamental frequency match than a single stub matching circuit.
A similar design methodology is used for the input matching circuit design with the
exception of a series resistor Rg added to improve the stability of the device. The stability of
the device is evaluated using Rollet’s stability factor [101]

2 Re(Z11 ) Re(Z22 ) − Re(Z12 Z21 )


k = . (3.7)
|Z12 Z21 |

A series gate resistance increases the real part of Z11 to improve stability. Since Rg also
reduces gain, the choice of Rg is a compromise between stability and gain. In this design, a
value of 2 Ω is selected.
The class-F circuit design in Figure 3.16 is transformed into a physical circuit design using
microstrip lines for the transmission line structures. The design was fabricated using copper
tape transmission lines on a 1.524 mm Rogers 4350 substrate with dielectric constant of 3.70.
The final design values for the matching networks are summarized in Table 3.7.
The simulated drain and gate waveforms for the final class-F amplifier design are shown
in Figure 3.17. For these simulations, the current and voltage are shown referenced to the
terminal planes of the packaged device. The shape of the waveforms are modified relative to
the waveforms at the intrinsic device plane of the device because of the package parasitics.

78
3.3. Class-F Amplifier Experimental Results

Table 3.7: Transmission line lengths for load and source matching networks.

T Lout1 T Lout2 T Lout3 T Lout4 T Lout5 T Lout6 T Lout7


in mm 2 22.7 48.9 15.2 21.9 28.8 30.6
◦ ◦ ◦ ◦ ◦ ◦
in degree 4.1 46.2 102.4 31 44.6 58.6 62.2◦
T Lin1 T Lin2 T Lin3 T Lin4 T Lin5 T Lin6 T Lin7
in mm 29.6 27.5 27 15.7 43.7 21.3 1.3
◦ ◦ ◦ ◦ ◦ ◦
in degree 60.2 56.1 54.9 32 92.3 43.3 2.7◦

Figure 3.17: Simulated drain voltage and drain current waveforms (left) and gate voltage and
drain current waveforms (right)[reproduced courtesy of The Electromagnetics Academy].

3.3.2 Experimental Results


The experimental class-F amplifier design is shown in Figure 3.18 and a picture of the
experimental test bed is shown in Figure 3.19. The amplifier design was tested with both
continuous wave (CW) and modulated signals. The results for each test signal are described
in the following subsections.

3.3.2.1 CW Performance
The power efficiency of the class-F amplifier for a CW input signal is shown in Figure 3.20.
The measured power efficiency reaches a maximum value of 78.8% at an output power of
40.5 dBm. At maximum efficiency, the quiescent drain current is 19% of the maximum DC
current. The figure also includes simulation results for the same test conditions. Although
good agreement between simulation and measurement results are obtained, the discrepancy
at high power may be related to the self-heating in the device. At low output power, the
differences between simulated and measured performance may be related to the switch mode
operation of the model [102, 103, 104].
The power efficiency and output power as function of frequency are shown in Figure 3.21.
Power efficiency is greater than 60% over a frequency range of approximately 120 MHz. The

79
3.3. Class-F Amplifier Experimental Results

Figure 3.18: Photograph of the 10 W class-F power amplifier.

Rectified Voltage
Power Meter

40 dB
Attenuator
Class-F PA

Driver

RF signal
generator

Cooler

Figure 3.19: The class-F amplifier test bench.

bandwidth is primarily limited by the impedance variation of the input harmonic stubs near
the fundamental frequency. Although a double stub input match is added to compensate
for the harmonic stub impedances at the fundamental frequency, the network is inherently
narrowband. This illustrates the trade-off between bandwidth and power efficiency which can
result by shaping the gate waveform with harmonic terminations at the input.

80
3.3. Class-F Amplifier Experimental Results

3.3.2.2 Modulated Performance


The power efficiency of the class-F amplifier for a CW input signal was shown in Fig-
ure 3.20. After characterizing the amplifier with CW test signals, the amplifier was tested
with a 5 MHz WCDMA test signal. The WCDMA signal had a 8.8 dB peak-to-average (PAR)
power ratio and was generated using a Tektronix AWG70002A arbitrary waveform generator.
Performance of the amplifier was measured without linearization. The measured output spec-
trums for the WCDMA signals are shown in Figure 3.22 with a resolution bandwidth (RBW)
of 30 KHz. A summary of the adjacent channel leakage ratio (ACLR) and power efficiency as
function of output power is shown in Figure 3.23.
In the figures, three different output power levels are identified for comparison. At point
(a), the average output power is 35.1 dBm. The CW saturated output power is approximately
40.5 dBm (see Figure 3.20); therefore the peak power of the modulated signal at point (a)
is compressed by about 3.4 dB. The corresponding ACLR is -33.1 dBc (decibels relative to
the carrier) and the power efficiency is 46.1%. With the addition of digital predistortion, im-
provements in linearity of 15 dB or more could be expected [105, 106]. The other two points,
(b) and (c), are measured under back-off conditions and as expected linearity improves at the
expense of power efficiency.

90 41
Solid: Measurement
Dash: Simulation

Output Power (dBm)


80 40
Efficiency (%)

70 39

60 38
20 21 22 23 24 25
Input Power (dBm)

Figure 3.20: Measured and simulated drain efficiency and output power as a function of input
power for a CW test signal [reproduced courtesy of The Electromagnetics Academy].

81
3.3. Class-F Amplifier Experimental Results

90 41
Solid: Measurement
Dash: Simulation
80 40

Output Power (dBm)


Efficiency (%)
70 39

60 38

50 37

40 36
900 950 1000 1050
Frequency (MHz)

Figure 3.21: Measured and simulated drain efficiency and output power as a function of
frequency for a CW test signal [reproduced courtesy of The Electromagnetics Academy].

−10
Power spectrum density (dB/RBW)

−20

−30

−40
a
b
−50 c

−60

−70
975 980 985 990 995 1000 1005
Frequency (MHz)

Figure 3.22: Measured output spectrums for a WCDMA signal at three different output
power levels: (a) 35.1 dBm (b) 33.4 dBm and (c) 31.6 dBm [reproduced courtesy of The
Electromagnetics Academy].

82
3.3. Class-F Amplifier Experimental Results

60 −30

50 (a) −32

40
(b) (a) −34
Efficiency (%)

ACLR (dBc)
(c)
30 −36

20
(b) −38

10 −40
(c)
0 −42
30 31 32 33 34 35 36
Output Power (dBm)

Figure 3.23: Measured drain efficiency and ACLR as a function of output power for a WCDMA
signal [reproduced courtesy of The Electromagnetics Academy].

83
3.4. Inverse Class-F Power Amplifier

Table 3.8: Source and load pull harmonic impedances for the class-F−1 amplifier.

Pin (dBm) Pout (dBm) P AE(%) ΓL (f0 )


21.5 41.2 83.6 0.53∠119.67
ΓL (2f0 ) ΓL (3f0 ) ΓS (f0 ) ΓS (2f0 )
1.0∠32.85 1.0∠190.51 0.91∠146.41 1.0∠175.73

3.4 Inverse Class-F Power Amplifier


Although a class-F power amplifier has high power efficiency, the current switched class-
F−1 amplifier can have even higher power efficiency because capacitive switching losses are
reduced. In the class-F amplifier, the waveform shaping creates a switched voltage signal,
while in the class-F−1 amplifier, the voltage waveform across the device is a half-sinusoid. The
current switched class-F−1 can also have a larger dynamic range than the voltage switched
class-F amplifier in terms of maintaining higher power efficiency under back-off conditions.
This feature is shown later in Chapter 4 where class-F and class-F−1 rectifiers are compared.
The main disadvantage of a class-F−1 amplifier compared to a class-F amplifier is the device
utilization is slightly lower which means the maximum available power that can be delivered
by a device is slightly higher in the class-F configuration.

3.4.1 Design Methodology


A current switched class-F−1 amplifier is designed for the same 10 W Cree GaN device
used in the class-F amplifier. Similar to the class-F design, harmonic matching networks are
implemented in the input and output matching circuits to control the gate and drain wave-
forms. A second harmonic input match and a third harmonic output match are implemented
as shown in Figure 3.24. The required harmonic impedances at the device planes were deter-
mined from load pull test benches using the Cree large signal device model. The load pull
results are summarized in Table 3.8 for a bias of VDD = 24 V and VGG = −2.8 V.
The synthesis of the output match begins with the third harmonic network. The trans-
mission line T Lout2 is 90◦ at the third harmonic frequency and creates a short at plane B.
Consequently the reflection coefficient looking to the right of plane B at the third harmonic
is 1∠180◦ . The addition of the series transmission line T Lout1 modifies the phase to provide
the match ΓL (3fo ) at plane A. The next step in the synthesis of the output circuit is to add
transmission lines T Lout3 and T Lout4 to create the second harmonic match. Transmission
line T Lout4 is 90◦ at the second harmonic frequency which creates a short at reference plane
C. The short is transformed through transmission line T Lout3 and ideally maps to an open
circuit at the device. However, the line length needs to be modified to compensate for the
impedance contributions of the third harmonic network and to create the required reflection
coefficient ΓL (2fo ) at plane A. The fundamental frequency output match is implemented with
a single stub circuit consisting of T Lout5 and T Lout6 .
For the input matching network, a second harmonic termination impedance of ΓS (2fo ) is
required at the gate reference plane (plane D in Figure 3.24). The second harmonic impedance
is created by transmission line T Lin3 , with a length of 90◦ at the second harmonic frequency,

84
3.4. Inverse Class-F Power Amplifier

and transmission line T Lin4 . The fundamental frequency input match is implemented with
a single stub circuit consisting of T Lin1 and T Lin2 . Finally, a series resistor (Rg ) of 3.5 Ω is
added to improve the stability of the device.
VGG VDD
C0 C0

TLin3 TLout2 IDC LDD


(λo /4) Cin (λo /12)
TLin4 TLout3 Cout
Rg
Vin TLin2 TL TLout4 TLout5 Vout
M1 out1
TLin1 (λo /8)
TLout6

E D A B C

Harmonic Input Matching Network Harmonic Output Matching Network

Figure 3.24: Schematic of the class-F−1 PA with output and input matching circuits and bias
networks.

3.4.2 Experimental Results


An experimental prototype of the class-F−1 amplifier was fabricated on a 0.762 mm Rogers
4350 substrate with a dielectric constant of 3.70. The transmission lines were implemented
in microstrip and the dimensions were optimized using numerical simulations in ADS. The
final design values for the matching networks are summarized in Table 3.9. The simulated
drain voltage and drain current waveforms for the final class-F−1 amplifier design are shown
in Figure 3.25.

Table 3.9: Microstrip transmission line lengths for the class-F−1 amplifier.

T Lout1 T Lout2 T Lout3 T Lout4 T Lout5 T Lout6 T Lin1 T Lin2 T Lin3 T Lin4
in mm 2 12.65 7.86 21 21.97 28.98 37.37 9.18 44.63 2.8
◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦
in degrees 4.1 25.6 15.9 45.5 44.5 58.7 75.7 18.6 90.6 5.6◦

A photograph of the class-F−1 power amplifier is shown in Figure 3.26. The performance
of the amplifier was measured in a test bed similar to the class-F amplifier test bed shown in
Figure 3.19. The power efficiency of the amplifier for a CW input signal is shown in Figure 3.27
and reaches a maximum value of 83% at an output power about 40 dBm. These results are
compared with the CW measurements for the class-F amplifier (see Figure 3.21) where the
peak power efficiency is 78.8% at an output power of 40.5 dBm. The results confirm that
the power efficiency of the class-F−1 is higher than the class-F amplifier. The results also
shows that the power utilization in the class-F amplifier is slightly higher than class-F−1 and
delivers 0.5 dB more power at peak efficiency.

85
3.5. Wideband Inverse Class-F Power Amplifier

Figure 3.25: Simulated drain voltage (solid) and drain current (dash) waveforms for the class-
F−1 power amplifier. The waveforms are shown for the drain terminal of the packaged device.

Inverse Class-F Power Amplifier

Output

Input

Figure 3.26: Photograph of the class-F−1 power amplifier.

3.5 Wideband Inverse Class-F Power Amplifier


The class-F and class-F−1 power amplifiers described in the previous sections are inher-
ently narrowband. The harmonic impedance matching networks are designed for a specific
frequency and the design process does not consider the synthesis of broadband matching
networks.
Interest in wideband power amplifiers is high because most modern wireless communication
systems such as long term evolution (LTE), wideband code division multiple access (WCDMA)
and world wide interoperability for microwave access (WiMax) have services on multiple
frequency bands. For example, the Rogers Communication LTE wireless network in Canada

86
3.5. Wideband Inverse Class-F Power Amplifier

100 10

90 9

Output Power (W)


Efficiency (%)
80 8

70 7

60 6
20 21 22 23 24 25
Input Power (dBm)

Figure 3.27: Drain efficiency and output power as a function of input power for the fabricated
class-F−1 PA.

uses frequency bands spanning a frequency range from 700 MHz to 2600 MHz. Therefore, high
efficiency power amplifiers are required to meet multi-band operation requirements. Another
application of wideband power amplifiers is in the design of wideband RF rectifiers for wireless
power and RF energy recycling circuits [107]. In this work, the RF rectifier application is the
primary motivation for designing a wideband class-F type amplifier.
Different techniques have been proposed to design wideband power amplifiers. These
techniques include amplifiers with lossy matching networks[108], feedback amplifiers [109,
110], traveling-wave amplifiers (TWA) [111, 112], continuous mode power amplifiers [113, 114]
and harmonically tuned methods [104]. Although lossy matching networks and feedback
techniques are established methods of implementing wideband designs they usually lead to low
power efficiency. Traveling wave amplifiers use a chain of device in a transmission line structure
to create wideband responses but the area, size, and cost and can be high. Continuous mode
amplifiers and harmonically tuned methods are more recent and examples of designs using
these techniques are summarized in Table 3.10. For this work, the harmonically tuned design
approach [104] is selected as it is retains the underlying characteristics of the amplifier topology
with the added goal of synthesizing wideband matching networks. These synthesis of wideband
networks for switch-mode amplifiers also has other applications such as outphasing amplifiers
where signals with broad bandwidth need to be amplified efficiently.

Table 3.10: Wideband class-F/family amplifier designs comparison.


Ref. BW (GHz), (%) Pout (W) ηd (%) Type
2008 [115] 0.8-4, 133 1-2 40-55 Class-F
2010 [104] 1.9-4.3, 78 10-15 55-72 Class-F−1
2011 [21] 0.55-0.92, 51 8.5-13 70-80 Class-F
2012 [116] 1.45-2.45, 51 11-16.8 70-81 Class-F
2014 [114] 1.6-2.8, 54.5 8.1-17.8 67.5-81.9 Class-F
This work 0.6-1.2, 70.7 7-8.83 62.5-79.2 Class-F−1

87
3.5. Wideband Inverse Class-F Power Amplifier

3.5.1 Design Methodology


Fano [117] first presented fundamental design equations for the synthesis of wideband
matching networks in 1950. The equations are transcendental and must be solved numerically.
Recently, Dawson [118] has presented analytical closed-form solutions for alternate equations
that can be used to synthesize wideband networks. Dawson’s method to synthesize wideband
lumped element matching circuits is used to design a wideband class-F−1 power amplifier.
After the lumped element circuit is designed, it is transformed into a distributed matching
network [104].
The steps in the synthesis of the matching circuits are as follows:

1. The network synthesis begins by specifying the bandwidth of the amplifier. If f1 and f2
are the lower and upper band edge frequencies, then the centre frequency is
p
f0 = f1 f2 , (3.8)

the bandwidth is

∆f = f2 − f1 , (3.9)

and the fractional bandwidth is


f2 − f1
FB = . (3.10)
f0
For this design: f1 is 650 MHz, f2 is 1150 MHz, f0 is 865 MHz, ∆f is 500 MHz, and
the fractional bandwidth is 58%.

2. Use load pull measurements at the fundamental frequency over the operating frequency
range to construct an equivalent output circuit of the device. The circuit consists of a
shunt resistance R0 and shunt capacitance Cout . The resistance is the best fit load line
and the capacitance is the best fit to the output capacitance of the device.

3. Based on the fractional bandwidth and the impedance transformation ratio from the
device load line to the output load resistance (RL ), select a lowpass filter prototype.
For this design, the fractional bandwidth is 58% and an odd order prototype with
n = 3 is selected. The π network configuration is also selected such that the first shunt
capacitance in the lowpass prototype can be associated with the output capacitance of
the device. See Figure 3.28(a).

4. Use equations in [118] to calculate the normalized admittance values for the lowpass
filter prototype. The values depend on the terminal resistances of the network, the
device capacitance, and the relative bandwidth of the network.

5. Frequency and impedance scale the lowpass prototype. After scaling, the input terminal
resistance is R0 at the device port and the first shunt element C1 is equivalent to the
device capacitance Cout . Note that the output port impedance is not 50 Ω because the
impedance scaling is done to match the input port resistance to the device load line. The
output port impedance is adjusted by creating in step 7 using a capacitive impedance
transformer.

88
3.5. Wideband Inverse Class-F Power Amplifier

6. The lowpass network is transformed into a bandpass network with a centre frequency
f0 . See Figure 3.28(b).

7. Use a Norton transformation in the filter to shift the output terminal impedance to 50 Ω.
See Figure 3.28(c). The transformation creates a capacitive impedance transformer
which can be adjusted to match the output port.

8. Convert lumped element resonators into equivalent transmission line resonators. The
final network consists of microstrip lines and one discrete capacitance.

9. Verify that the input port impedance at the second harmonic has high impedance and is
not a short circuit. The current switched class-F−1 amplifier has odd harmonic current
components and even harmonic voltage components. Therefore, the voltage requires an
open impedance at the second harmonic. Since the network is broadband, it is difficult
to predict and control the second harmonic impedance. However, based on work by
Saad [104], he shows that the power efficiency of the amplifier remains high providing
the second harmonic impedance has a reflection coefficient far away from a short circuit.

10. The same procedure is repeated for the input matching network.
With this design procedure, a wideband F−1 amplifier with fundamental and second har-
monic matching is implemented.

3.5.1.1 Lumped Element Output Matching Network


From Dawson’s paper [118], the normalized admittances in the lowpass filter prototype
are found from the following equations:

Q = R0 C1 ω0 (3.11)

ω0 = 2πf0 (3.12)

g0 = 1 (3.13)

1
g1 = (3.14)
( Q1ω0 ) g0

1
gj = for j = 2 to n (3.15)
gj−1 (kj−1,j )2

1
gn+1 = (3.16)
D( Q1ω0 )gn

For these equations, n is the order of the lowpass prototype, and ki,j and D are given by
(long) equations in [118].
In Equation (3.11), R0 is the load line resistance and C1 is the device output capacitance.
Values for R0 and C1 can be found from load and source pull simulations at different fre-
quencies. The load pull at the fundamental frequency includes a second harmonic open and a

89
3.5. Wideband Inverse Class-F Power Amplifier

g0 g1 g2 g3 g4

1 L2 2

R0 C1 C3 RL
Load Device
Line Capacitance

(a)
Device Model

1 L2 2
C2

R0 C1 L1 C3 L3 RL

(b)
Device Model Impedance
Transformer
1 L2 2
C20
R0 C1 L1 C40 C30 L03 0
(n2 RL) RL
(n2T L3) T

(c) Norton

Figure 3.28: Different steps for design of a lumped element matching network: (a) low-
pass network with normalized admittances gj ; (b) bandpass matching network; (c) Norton
transformation to increase output impedance.

third harmonic short. Once the load pull data is extracted, mean values for R0 and C1 must
be found. The real part of the load pull data can be averaged to find R0 and a linear fit to
the imaginary part can be used to find C1 [119]. Another method is to directly use midband
load pull data [104]. The latter method is used here and the input and output impedances
are summarized in Table 3.11.
Focusing on the synthesis of the output match, the conjugate impedance of ZL,opt is used
to estimate the equivalent device model. The equivalent circuit values are R0 = 52.3 Ω and
C1 = 4.31 pF. Equations (3.11) through (3.16) are evaluated for n = 3 to give normalized
lowpass prototype filter values in Table 3.12. The normalized lowpass prototype admittances
gk are then impedance and frequency scaled to admittances gF ISk . The scaled values and
the corresponding lowpass element values are shown in Table 3.12 and Figure 3.29. In order
to prepare the lumped element network for conversion into a distributed network, the value
of C1 = 5.13 pF is increased to be slightly larger than Cout . Capacitance C1 is partitioned
later into Cout and a second shunt capacitance which is combined into a shunt resonator.

90
3.5. Wideband Inverse Class-F Power Amplifier

Table 3.11: Results of the load/source pull simulations for ZLopt and ZSopt .

Freq. (MHz) Pin (dBm) Pout (dBm) P AE(%) ZLopt (Ω) ZSopt (Ω)
900 24.5 40.3 81.5 19.9 + j25.4 3.5 + j18

Table 3.12: Admittance parameters and extracted values for a third order network.

g1 (f) g2 (f) g3 (f) g4 (f)


0.8445 0.9593 0.6722 0.9108
gF IS1 gF IS2 gF IS3 gF IS4
5.1383e-12 1.5977e-8 4.0898e-12 47.6534
R0 (Ω) C1 (pF) L2 (nH) C3 (pF)
52.32 5.13 15.97 4.09

g0 g1 g2 g3 g4
1 0.8445 0.9593 0.6722 0.9108
1 L2 2
15.97
R0 C1 C3 RL
52.3 5.13 4.09 47.6

Figure 3.29: Impedance and frequency scaled lumped element lowpass network for synthesiz-
ing a wideband output match.

After impedance and frequency scaling, the lowpass network is transformed into a band-
pass network using a center frequency of f0 and bandwidth ∆f . The lumped element bandpass
network is shown in Figure 3.30.
C2
2.12pF

1 L2 2
15.97nH
R0 C1 L1 C3 L3 RL
52.3Ω 5.13pF 6.59nH 4.08pF 8.28nH 47.6Ω

Figure 3.30: Lumped element output network after applying a lowpass to bandpass transfor-
mation.

91
3.5. Wideband Inverse Class-F Power Amplifier

As a final step in the lumped element synthesis procedure, the output resistance of 47.6 Ω
needs to be increased to 50 Ω. Through a Norton transformation applied to C2 and C3 , an
impedance transformer can be constructed with a transformation ratio of nT . With reference
to Figure 3.31(a), after using the capacitive network in Figure 3.31(b), the modified output
load resistance RL 0 is n2 R . For the output matching circuit an impedance transformation
T L
ratio (nT ) of 1.0243 is required. The transformed component values for the output matching
network are shown in Figure 3.32.

L3 RL n2T L3 n2T RL

ZL n2T ZL
(a)

Z1 nT Z 1 nT Z1 Z2
(nT − 1) Z1 + (1 − nT )Z2

nT Z 1
Zi Z2 Zo Zi n2T Zo

(b)

Figure 3.31: Impedance transformed output network (top) and the corresponding Norton
transformation to create the impedance transformation (bottom).

C20
L2 2.07pF
15.97nH

1 2

R0 C1 L1 C40 C30 L03 RL0

52.3Ω 5.13pF 6.59nH 0.05pF 3.85pF 8.7nH 50Ω

Norton

Figure 3.32: Bandpass output matching network with an impedance transformer to match
the output to 50 Ω.

A similar design procedure can be followed to constructed the input matching network.
For the input network, the optimal source impedance at the fundamental frequency is ZSopt =
3.5 + j18 Ω (see Table 3.11). The corresponding normalized lowpass prototype admittances
and the frequency and impedance scaled admittances are summarized in Table 3.13. The final
input matching network is shown in Figure 3.33.

92
3.5. Wideband Inverse Class-F Power Amplifier

Table 3.13: Admittance parameters for low-pass network and extracted values for the final
band-pass structure corresponding to the input matching network.

g1 g2 g3 g4 gF IS1 gF IS2 gF IS3 gF IS4


(f) (f) (f) (f)
3.1059 0.5503 2.0157 0.4239 1.0290e-11 1.6829e-8 6.6788e-12 40.7305
R0in C1in C20 C30 C40 L1 L2 L03
(Ω) (pF) (pF) (pF) (pF) (nH) (nH) (nH)
96.071 10.291 0.196 1.817 5.264 3.293 16.830 6.228

C20
L2 1.81pF
16.83nH

1 2

R0in C1 L1 C40 C30 L03 RL0

96Ω 10.29pF 3.29nH 0.2pF 5.26pF 6.22nH 50Ω

Figure 3.33: Bandpass input matching network by Norton transformation (nT = 1.1079).

3.5.2 Distributed Matching Networks


At this point, lumped element matching networks have been synthesized for the amplifier
and the remaining step is to convert the networks into distributed structures using transmis-
sion lines. The first step is to divide capacitance C1 into three parallel capacitances consisting
of Cout , C11 and C12 as shown in Figure 3.34. The lumped elements can then be grouped
with adjacent circuit elements to create circuit networks that can be replaced with equivalent
transmission line structures. There are two shunt resonators in the circuit: 1) C11 and L1
and 2) C30 and inductor L03 . The LC shunt resonators can be implemented as 90◦ short circuit
transmission lines as shown Figure 3.35 [120]. The π-network consisting of Cp 0
12 , L2 and C4
is equivalent to a short transmission line with characteristic impedance Zo = L2 /C40 [121].
The only lumped element component which cannot be incorporated into a transmission line
structure is C20 , but this can serve as a DC blocking capacitor which is required in the circuit.
The final output matching network is shown in Figure 3.36.
The same procedure can be used for the input matching circuit and the final transmission
line input matching circuit is shown in Figure 3.37.
Microstrip transmission lines are designed to implement the distributed matching net-
works. The substrate is a Rogers RO4350 dielectric with a thickness of 0.762 mm and a
dielectric constant of 3.70. The final microstrip dimensions for each transmission line are
shown in Table 3.14.
Before proceeding to the fabrication of an experimental prototype, it is important to verify
the matching network designs and ensure the networks provide the required impedances at
the fundamental, second and third harmonic frequencies. The matching network designs

93
3.5. Wideband Inverse Class-F Power Amplifier

L2 C20
15.97nH 2.07pF
1 2

R0 Cout C11 L1 C12 C40 C30 L03 RL0

52.3Ω 4.31pF 0.77pF 6.59nH 0.05pF 0.05pF 3.85pF 8.7nH 50Ω

Transistor T L1 T L2 T L3

Figure 3.34: Dividing the capacitance C1 into three parallel capacitances to reform the output
structure as a distributed network.

θ = ∠90◦
L C TL
Z0 = (π/4)ωL

Figure 3.35: Equivalent transmission line circuit for a shunt resonator.

TL2 C20
Drain
50 Ω

TL1 TL3

Figure 3.36: Distributed output matching network.


0
CS2 TLS2
Gate
50 Ω

TLS3 TLS1

Figure 3.37: Distributed input matching network.

are verified in a simulator and the device plane impedances seen looking into the matching
networks are measured. Beginning with the fundamental frequency, a comparison of the
network impedances versus the load pull impedances are shown for the input and output
matching networks in Figure 3.38. The load pull measurements correspond to the points
and the simulated results correspond to the contours. As shown, the networks synthesize
fundamental load impedances that are close to the load pull values.
Next, the second harmonic impedances are simulated. The results are shown in Fig-
ure 3.39. Since this is a current switched class F−1 amplifier, the second harmonic is ideally a

94
3.5. Wideband Inverse Class-F Power Amplifier

Table 3.14: Microstrip transmission line lengths and widths for load and source matching
networks.

T L1 T L2 T L3 T LS1 T LS2 T LS3


Length (mm) 20.57 21.59 46.17 27.33 21.97 44.57
Width (mm) 0.2 0.2 2.65 6.5 0.2 3.9

Figure 3.38: The fundamental frequency impedances of the input and output matching net-
works.

open circuit. Other references on wideband amplifier designs [104] show that power efficiency
is relatively insensitive to an exact second harmonic impedance providing the impedance falls
outside a region around a short circuit. The simulated results for this design show that the
second harmonic impedances fall on the open circuit side of the Smith chart well away from
short circuit impedances and therefore the second harmonic impedance is satisfactory.
As a final check on the matching networks, the third harmonic impedance is evaluated. The
results for the output matching network are shown in Figure 3.39. Over the third harmonic
frequency range, the network should present a short circuit and the contour lies in the short
circuit region of the Smith chart. The plot includes a wide frequency sweep range and includes
the fundamental and second harmonic frequency ranges as well. The harmonic impedance
data shown in Figure 3.40 is consistent with the design goal of implementing a wideband
class F−1 . A similar result is obtained for the input matching network.

95
3.5. Wideband Inverse Class-F Power Amplifier

Figure 3.39: The second harmonic impedances of the input and output matching networks.

Figure 3.40: Wide frequency range sweep of the impedances of the output matching network.
Fundamental, second harmonic and third harmonic frequency ranges are shown.

96
3.5. Wideband Inverse Class-F Power Amplifier

3.5.3 Experimental Results


A photograph of the wideband class F−1 power amplifier is shown in Figure 3.41. Similar
to all the other amplifier designs, a 10 W Cree GaN HEMT is used (model CGH40010F).

Figure 3.41: Photograph of the wideband class-F−1 power amplifier.

The simulated and measured power efficiency of the wideband class-F−1 amplifier for a CW
input signal at different frequencies are shown in Figure 3.42. The measured power efficiency
reaches a maximum value of 79.2% with an output power of 8.83 W. Power efficiency is greater
than 60% over a frequency range of approximately 600 MHz.

100
Measurement
90 Simulation

80

70
Efficiency (%)

60

50

40

30

20

10

0
500 600 700 800 900 1000 1100 1200 1300
Frequency (MHz)

Figure 3.42: Measured and simulated drain efficiency of the wideband class-F−1 PA as a
function of frequency for a CW test signal.

97
3.6. Chapter Summary

3.6 Chapter Summary


In this chapter, the design methodology, simulation results and experimental results were
shown for three different class-F power amplifiers. The designs were for a narrowband switched
voltage class-F amplifier, a narrowband switched current class-F−1 amplifier, and a wideband
class-F−1 amplifier. All the experimental results use a packaged Cree 10 W GaN HEMT
device which makes comparison of experimental results insightful.
In the voltage switched class-F amplifier, a comprehensive study of the harmonic input
matching circuit was made. A level 3 device model was used for this work, and from the study,
it is concluded that input matching for the second harmonic is very important while the third
harmonic has little effect on the overall power efficiency of the amplifier. An experimental
prototype of the amplifier was built which included third harmonic input and output matching
circuits which is the first work to report third harmonic input matching in a class-F amplifier.
The experimental results of the voltage switched class-F amplifier are compared with a
similar current switched class-F−1 amplifier. The results show that the current switched
amplifier has higher peak power efficiency, which is attributed to lower capacitive switching
losses compared to the voltage switched amplifier. Conversely, the current switched power
amplifier delivered slightly less power at peak efficiency, about 0.5 dB less, which shows that
the device utilization is slightly higher in a voltage switched topology. A wideband current
switched class-F−1 amplifier design was also built using a network synthesis technique that
implements a wideband fundamental frequency match. It was also shown that the wideband
network has a high second harmonic impedance and a low third harmonic impedance over the
bandwidth of the design.
In the next chapter, the theory of time-reversal duality is applied to reconfigure these
amplifier designs into RF rectifiers. The experimental results presented in this chapter for the
amplifier configuration will be used to compare with experimental results for the RF rectifier
configuration.

98
Chapter 4

Class-F RF Synchronous Rectifiers


This chapter focuses on the design and implementation of high efficiency and high power
GaN HEMT class-F and class-F−1 synchronous rectifiers. The work provides new bench
marks for high power RF synchronous rectifiers operating at a frequency of 1 GHz and power
levels of approximately 10 W. The chapter begins with a brief overview of the time reversal
duality principle which is used to convert switch-mode power amplifier circuits into equivalent
synchronous rectifier circuits. Time reversal duality concepts have been used for several
decades in power electronic circuit applications, but the application to RF circuits is much
more recent. There are also interesting questions raised when applying time reversal duality
to circuits with loss and the implication of loss in terms of constructing circuit duals is
investigated. The clarification of these subtle points has not been discussed in the literature
and the primary implication is related to the operating point of the device when comparing
amplifier and rectifier duals.
Applying the principle of time reversal duality to class-F switch-mode amplifiers, three
different RF synchronous rectifiers were designed using a 10 W GaN HEMT device from
Cree. These rectifiers are a narrowband class-F rectifier, a narrowband class-F−1 rectifier,
and a wideband class-F−1 rectifier. In all cases, the rectifier is compared to the amplifier dual
and the work provides new references for comparing circuit duals. Most literature references
describing implementations of RF rectifiers typically use the principle of duality to create
rectifiers without providing direct comparisons with the amplifier circuit under equivalent
measurement conditions.
Other highlights of the research work include a comparison of class-F and class-F−1 recti-
fiers under constant load conditions. The optimal load resistance which maximizes RF to DC
conversion efficiency is dependent on the RF input power to the rectifier. A comparison of the
class-F and class-F−1 rectifiers shows that the class-F−1 is more robust in terms of operation
with a fixed load compared to a class-F rectifier.

4.1 The Principle of Time Reversal Duality


The concepts of duality are commonly used in circuit theory. Examples include transfor-
mations from Thévenin to Norton equivalent circuits and the synthesis of equivalent networks
with series or parallel circuit elements. Another example is the duality between voltage and
current waveforms. This is illustrated by comparing (a) and (b) in Figure 4.1; these figures are
based on the work of Hamill [74]. The network N can be transformed into a network dual Nvc
where the current and voltage waveforms are duals. The duality between current and voltage
waveforms is commonly used in switch-mode power amplifiers designs; for example, class-F

amplifiers are designed to switch voltage, while class-F 1 are designed to switch current.
3
Parts of Chapter 4 have been published in two articles. Reprinted with permission from ELEX and EuMC
[75, 76].

99
4.1. The Principle of Time Reversal Duality

I V
N

V I
t t

(a)
Ivc Vvc
Nvc

Vvc Ivc
t t

(b)
Itr Vtr
Ntr

Vtr Itr
t t

(c)

Figure 4.1: (a) Network N and its current and voltage, (b) network Nvc , a voltage and current
dual of N , and (c) network Ntr , a time reversal dual of N .

Another duality was recognized by Hamill [74, 122] which he called time reversal duality.
The time reversal (TR) concept is illustrated by comparing the current-voltage relations for
network N in Figure 4.1(a) with the current-voltage relations for the TR dual, network Ntr in
Figure 4.1(c). In the TR dual, the voltage across the network is Vtr = V (−t) and the current
into the network is Itr = −I(−t). As a consequence of the sign change in the current, power
flow is reversed at the terminals of network Ntr relative to the original network N . More
generally, for an n-port network, power flow is reversed at all n terminals. An example of
a two port network is shown in Figure 4.2. If network N is an amplifier, the primary input
power is the DC drain supply P1 and the output is the amplified RF signal P2 . The TR dual,
network Ntr , is a rectifier circuit where RF input power P20 is converted into a DC power P10 .
A general method of constructing a TR dual can be established [74, 122] and a summary
of the circuit relations is shown in Table 4.1. Capacitor and inductors are assumed to be
lossless and the circuit elements are unchanged in a TR dual. Resistance on the other hand is

100
4.1. The Principle of Time Reversal Duality

P1 N P2 P10 Ntr P20

Figure 4.2: The direction of energy flow in a network and its TR dual.

Table 4.1: Time reversal relations for circuit components.


Original Network (N ) TR-dual Network (Ntr )
Voltage V (t) Vtr (t) = V (−t)
Current I(t) Itr (t) = − I(−t)
Power P (t) Ptr (t) = − P (−t)
Inductance L Ltr = L
Capacitance C Ctr = C
Resistance R Rtr = − R

dissipative and because power flow is reversed, the circuit dual must have negative resistance.
Obviously negative resistance is not physically realizable except with active circuits, and
therefore the most common application of TR duality is in circuits where dissipative losses
are small. Examples include power electronic circuits which usually operate with very high
efficiency. At microwave frequencies, dissipative losses are more significant and power loss can
be substantial. This then motivates the question: how should losses be handled in constructing
circuit duals? More will be said about this later.
The other implication of TR duality is that active devices must be bidirectional and
operate under time reversed conditions. For an active device like a MOSFET or HEMT, this
means that a device which operates in quadrant I of the IV plane in an amplifier circuit,
must operate in quadrant III in a rectifier circuit [73]. The IV symmetry is not perfect in
practical devices, but in theory they have symmetry because the designations of source and
drain nodes are made relative to the polarity of the drain supply. The quadrant I versus
quadrant III device operation is discussed in more detail in the next section.

101
4.2. Definitions of Equivalence for Amplifier and Rectifier Duals

4.2 Definitions of Equivalence for Amplifier and Rectifier


Duals
Using the principle of time reversal duality, a switch-mode power amplifier circuit can
be transformed into a synchronous rectifier circuit as shown in Figure 4.3. In the rectifier
mode, the output port of the amplifier becomes an input port and the DC supply port of the
amplifier is replaced by a DC load, RDC . If the amplifier and rectifier were 100% efficient,
then RDC would be the equal to VDD /IDC in the amplifier circuit. For lossy circuits, the value
VDD /IDC is an approximation and the value of RDC needs to be swept to find the optimal
load resistance for a specific input power.
When loss is present in the amplifier and rectifier, we need to re-evaluate how the two
circuits are compared in the context of time reversal duality and relate this to measurements
of power and efficiency. In the amplifier, a source power of PDC is required to deliver a RF
load power of Pout , and in the rectifier circuit, a RF source power of Pin is required to deliver
a DC load power of PDC 0 . Now consider two different test conditions. In the first case, the

rectifier efficiency is measured under the condition where the RF powers are matched: the
output power Pout of the amplifier is the same as the input power Pin of the rectifier. The
matched RF power condition falls out naturally from a test configuration where the amplifier
and rectifier are arranged as a series cascade. In this case, the RF output of the amplifier is
connected to the RF input of the rectifier dual and RF powers are equal (Pout = Pin ). The
disadvantage of this test condition is that the DC input power to the amplifier and the DC
output power from the rectifier differ by the product of the efficiencies of the amplifier and
the rectifier.
A second method is to measure the two circuits under conditions where the input source
powers are identical. In this case, the DC source power for the amplifier (PDC ) is equal to the
RF input power (Pin ) for the rectifier. Under conditions of equal source powers, the amplifier
and rectifier should have similar power efficiencies and the power delivered to the loads should
be similar. The consequence of matching source powers is that the RF input power to the
rectifier is scaled relative to the RF output power delivered by the amplifier. The advantage of
the second method is that the operating points of the devices in the amplifier and rectifier are
closer than in the first method where the losses in the amplifier and rectifier are accumulated.
In this work, we use method two as a benchmark for comparing the amplifier and rectifier
based on the goal of minimizing the difference in the operating points of the active device.
The difference between the two test conditions can also be illustrated in terms of the
dynamic IV characteristics of the amplifier and rectifier. With reference to Figure 4.4(a), the
dynamic IV curve is shown for the class F amplifier described in Chapter 3. The expected
value of the drain voltage is equal to VDD and the expected value of the drain current is equal
to IDC . The corresponding DC operating point is denoted as point A on the IV curve.
In the rectifier circuit, as shown in Figure 4.4(b), the on-state is in quadrant III instead
of quadrant I as in the amplifier circuit. The average DC current flow is out of the drain
terminal when the device is on; therefore the device must operate in quadrant III to deliver
negative drain current. Similar to the amplifier, the expected value of the drain voltage is
0
VDC and the average value of the drain current is IDC 0 . Since V 0 0
DC is positive and IDC is
negative, the corresponding DC operating point for the rectifier falls in quadrant IV.
We now consider the two test conditions described earlier. If the RF powers of the amplifier
and rectifier are matched, the corresponding DC operating point for the rectifier is at point B
shown in the inset of Fig. 4.4(b). On the other hand, if the input source powers of the amplifier

102
4.2. Definitions of Equivalence for Amplifier and Rectifier Duals

VGG VDD RDC = VDD /IDC

(a) IDC PDC

LGG LDD Pout


Output
Rs Matching
Input
Matching M1
RL VL
Vs

P0DC
VGG V0DC
(b) I0DC R0DC
Pin
LGG LDD
Output
Rs Matching
Input RL
Matching M1
VRFin
Vs

P0DC
VGG V0DC

(c) I0DC
R0DC
Sampler Pin
LGG LDD
Output
Input Matching
RL
Matching M1
VRFin

Phase
Shifter

Figure 4.3: Block diagrams of (a) a power amplifier, (b) synchronous rectifier dual, and (c)
synchronous rectifier with feedback to provide gate drive.

and rectifier are equal, then the corresponding rectifier operating point is point C. If circuit
losses were identical for the amplifier and rectifier, the DC operating point for the rectifier
would be point D which mirrors the amplifier DC operating point obtained by changing the
sign of IDC . A comparison of the operating points shows that point C is closer to point D
which is the operating point dual of the amplifier. Therefore, a comparison of the amplifier
and rectifier are made under the condition of equal input source power conditions rather than
equal RF power conditions.
As a final remark on the dynamic IV curves for the amplifier and rectifier, it is noted that

103
4.2. Definitions of Equivalence for Amplifier and Rectifier Duals

2 2

A A
1 1
Ids (A)

(A)
0 0

ds
I
−0.3
B
−1 −1 C D
−0.4

−2 −2 −0.5
(a) (b) 24 26 28

0 20 40 60 80 0 20 40 60 80
Vds (V) Vds (V)

Figure 4.4: Dynamic load lines for: (a) a class-F amplifier (b) a class-F rectifier.

the effective on resistance of the rectifier is lower than in the amplifier. In quadrant III, the
reverse biased drain supply flips the gate control of the device to depend on the drain supply.
Therefore, in quadrant III, the device effectively turns on more as the drain swings more
negative. This observation is consistent with measured IV device characteristics reported by
other researchers for quadrant III device behaviour [123]. From the DC device characteristics
we therefore conclude that the effective on resistance of the rectifier in quadrant III is expected
to be slightly less than the amplifier on resistance in quadrant I. Also, the loop area under the
dynamic IV curve for the amplifier is larger in quadrant I than for the rectifier which means
current/voltage overlap losses are slightly higher in the amplifier than the rectifier. These
factors lead to the hypothesis that the rectifier efficiency is expected to be higher than the
equivalent amplifier efficiency.
At this point we have established a duality between the amplifier in Figure 4.3(a) and the
synchronous rectifier in Figure 4.3(b). Although a rectifier dual has been constructed from
the amplifier, the circuit in (b) requires a separate input gate drive similar to the amplifier.
The separate gate drive is inconvenient for rectification purposes and instead a feedback path
which samples the RF input signal is usually used to generate the gate drive signal. This is
shown in circuit (c). Since the gate must be switched with the correct phase relative to the
drain voltage and drain current waveforms, a delay line or phase shift circuit is required. When
the gate drive is derived from the RF input signal, the circuit is called a synchronous rectifier
circuit [73]. Another variation of synchronous rectifiers is a self-synchronous rectifier where
the intrinsic device capacitance Cgd is used in conjunction with a gate termination impedance
to create the required feedback signal for switching the gate [78]. In all the circuits shown
in this work, a directional coupler is used to sample the RF input signal and a delay line is
used as a phase shift circuit. In this way, any amplifier can be converted into a synchronous
rectifier by generating the appropriate gate drive signal from the RF input.

104
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

Table 4.2: Some recently published results for RF synchronous rectifiers


Ref. Type Device f (GHz) PDC η (%)
[73] class-E E-PHEMT 0.9 50 mW 83
synch.
[78] class-C GaN HEMT 10.1 1.67 W 64.4
self-synch. MMIC
[78] class-C GaN HEMT 10.1 3.18 W 63.9
self-synch. MMIC
[82] class-F−1 GaN HEMT 2.14 8.5 W 85
self-synch.
[79] class-E E-PHEMT 0.9 35 mW 88
self-synch.
[79] class-E E-PHEMT 2.45 15 mW 77
self-synch.
[80] class-E GaAs PHEMT 2.45 5.7 mW 77
self-synch.
[81] class-F GaAs PHEMT 5.8 5.5 mW 68
self-synch.
[83] class-F E-PHEMT 0.9 12.1 mW 85.4
self-synch.
[124] class-AB GaN HEMT 9.9 4.11 W 52
self-synch. MMIC
This class-F GaN HEMT 0.985 8.7 W 81.3
work synch.

4.3 High Efficiency GaN HEMT Class-F Synchronous


Rectifier
As will be shown, the power efficiency of a rectifier circuit is similar to the power efficiency
of an amplifier circuit providing they are tested under equivalent conditions where the input
power to the circuits is the same. Therefore, a high efficiency rectifier design begins with a
high efficiency amplifier design.
In this section we present experimental results for the implementation of a 10 W syn-
chronous class-F rectifier. The rectifier design is derived from the class-F amplifier described
earlier in Chapter 3. Other researchers have reported on GaN class-F−1 rectifier designs or
low power class-F rectifiers using pHEMTs, and this appears to be the first work for a high
power class-F design. A comparison of this design with other published results is given in
Table 4.2. Experimental results are also shown to compare the performance of the amplifier
and the rectifier dual under equivalent input power conditions. In this way, conclusions can
be made about the relative power efficiencies of the dual circuits.

4.3.1 Rectifier Test Bench and Efficiency Definitions


For comparison with the rectifier, the class-F PA in Chapter 3 (and Appendix-A) was first
re-tested. The device was biased with a 27 V drain supply and a gate bias of -2.6 V. At a
frequency of 985 MHz and for a sinusoidal input signal of 24.5 dBm (282 mW), the amplifier
delivers 8.3 W to a 50 Ω load. Under these conditions, the drain current is 0.398 A and the

105
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

Rectified Voltage

Power Meters

B A
Couplers Class-F PA

Class-F Rectifier RF signal


generator

Figure 4.5: The class-F rectifier test bench.

corresponding drain efficiency is 77.5%. The equivalent Thévenin source impedance of the
drain supply, RDC , is 67 Ω and the DC supply provides a source power of 10.7 W.
After testing the amplifier circuit, the amplifier was reconfigured as a synchronous class-F
rectifier. A photograph of the rectifier test bench is shown in Figure 4.5. An explicit feedback
loop is added to the amplifier to provide a gate drive signal from the RF input port as shown
in Figure 4.3(c). A directional coupler (A) samples the RF input signal and a variable delay
line is used to adjust the phase of the sampled signal to synchronously switch the GaN power
device. The measured sampling level of coupler A is -18.7 dB. The second port of coupler A
is connected to a power meter and calibrated to measure the reflected power at the input
of the rectifier. Since the input RF power is high and cannot be delivered by standard test
equipment, another class-F power amplifier is used to generate the RF source signal. A second
directional coupler (B) is placed in series between the rectifier and amplifier to measure the
available RF input power (Pin ) delivered to the rectifier. The input power is calibrated to
measure power at the interface between the two couplers.
Two definitions of RF to DC conversion efficiency which have been used in the literature
to report results for RF rectifiers are [80, 81]
PDC
ηr = (4.1)
Pin − Pref
and [82]
PDC
ηr = . (4.2)
Pin
In these equations, Pin is the incident or available power from the RF input source and Pref
is the amount of power reflected back from the input port because of mismatch loss. The

106
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

difference in the power efficiency measures is that in (4.1), the power efficiency accounts for
mismatch loss at the input of the rectifier, while in (4.2) the power efficiency is burdened by
input mismatch loss. From a system perspective, equation (4.2) is preferred because it includes
mismatch loss which is inherently present in the design of the rectifier circuit. If mismatch loss
is reduced by improving the design, then the corresponding power efficiency will be improved.
On the other hand, the efficiency measure in equation (4.1) accounts for mismatch loss and
gives insight into the maximum available power efficiency which can be obtained provided the
input is perfectly matched. Both measures of power efficiency are used in the literature and it
is important to identify which power efficiency measure is used for comparing results. Unless
otherwise stated, equation (4.1) which includes mismatch loss is used in this work as it is
deemed to be more appropriate for practical circuits where the terminal interface impedances
are usually specified and circuits must inherently include matching.

4.3.2 Experimental Results


As a starting point for evaluating the rectifier, the initial conditions were configured to be
the dual of the class-F amplifier where the input source powers of the amplifier and rectifier
are equal. The RF input power was set to 10.7 W (40.3 dBm) at a frequency of 985 MHz,
the same as the DC source power for the amplifier. The DC load resistance, RDC 0 , and phase

shifter are adjusted to optimize the power efficiency of the rectifier. The corresponding DC
load power is 8.7 W and the RF to DC conversion efficiency is 81.3% for an optimal load
resistance of 58 Ω. These numbers are compared to the equivalent class-F amplifier under
identical input source power conditions which showed a slightly lower power efficiency of
77.5% with an equivalent DC source resistance (RDC ) of 67 Ω, slightly higher than RDC 0 in
the rectifier dual. This shows that even under equivalent source power conditions where the
DC power supplied to an amplifier is equal to the RF power supplied to a rectifier, there are
small differences attributed to device operation in quadrant I versus quadrant III.
The rectifier power efficiency of 81% was calculated using (4.2) which includes input mis-
match loss. The effect of mismatch loss on the overall rectifier power efficiency can be found
by comparing the efficiency without mismatch loss using equation (4.1). Under the stated
test conditions, the reflected RF power at the input port of the rectifier is 28.2 dBm and the
input reflection coefficient is -12.1 dB. The corresponding RF to DC power efficiency without
mismatch loss is 86.5%, about 5% better. A summary of the class-F amplifier and rectifier
circuit duals tested under identical source power conditions are shown in Table 4.3.
Other test results for the rectifier are shown in Figures 4.6 through 4.8. In Figure 4.6,
the measured power efficiency and DC load power are shown as a function of the DC load
resistance RDC0 . These measurements are made at a frequency of 985 MHz with a RF input

power of 10.7 W. As shown, the optimal load resistance (RDC 0 ) is approximately 58 Ω.

Figure 4.7 shows the power efficiency of the rectifier as a function of the RF input power.
It shows that power efficiency peaks for an input RF power of 10.7 W (40.3 dBm). Power
efficiency remains above 50% for input power above 34 dBm and the maximum power delivered
by the rectifier is 11.3 W at an efficiency of 78%. Compared to other published work, this
result appears to be the highest reported DC power which has been measured for a RF
synchronous class-F rectifier circuit.
The rectifier performance as a function of frequency is shown in Figure 4.8. Efficiency
and load power peak at a frequency of 985 MHz. The bandwidth of the rectifier is dependent
on the bandwidth of the original amplifier design. In this case, the multiharmonic matching

107
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

Table 4.3: Comparison of class-F amplifier and rectifier experimental results.

Parameter Amplifier Rectifier

Frequency 985 MHz 985 MHz


Gate bias -2.6 V -2.6 V
DC supply/load resistance RDC = 67 Ω 0
RDC = 58 Ω
Input source power PDC = 10.7 W 0 = 10.7 W
Pin
Load power Pout = 8.3 W 0
PDC = 8.7 W
Power efficiency ηa = 77.5% ηr = 81.3%
Power efficiency without mismatch loss - ηr = 86.5%

networks are narrowband and a power efficiency of 50% is maintained over a frequency range
of about 50 MHz.
The power efficiency measurements shown in Figures 4.7 and 4.8 include mismatch loss.
If mismatch loss were reduced by improving the input match, then power efficiency would
increase. Measurements of mismatch loss were made over frequency and the results are shown
in Figure 4.9. The results show that mismatch loss reduces power efficiency by approximately
5% at 985 MHz and the loss increases as the frequency deviates from center frequency of the
design. The mismatch loss is directly related to the bandwidth of the output match in the
amplifier and improvements in the bandwidth of the matching circuit will reduce mismatch
loss over frequency. This is not any different than an amplifier which also has reduced efficiency
when the load match deviates from the optimum match. Later, in Section 4.5, results for a
wideband class-F amplifier and rectifier are shown.

108
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

90 10

80 9

Output DC Power (W)


70 8
Efficiency (%)

60 7

50 6

40 5

30 4
10 20 30 40 50 60 70 80 90
Resistance (Ohm)

Figure 4.6: Measured RF to DC conversion efficiency and output DC power as a function of


load resistance for a class-F rectifier.

90 12
11
80 10
9 Output DC Power (W)
70 8
Efficiency (%)

7
60 6
5
50 4
3
40 2
1
30 0
34 35 36 37 38 39 40 41 42
PRF-in (dBm)

Figure 4.7: Measured power efficiency and output DC power as a function of RF input power
for a class-F rectifier.

109
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

90 10

80 9

70 8

DC Power (W)
Efficiency (%)

60 7

50 6

40 5

30 4
950 960 970 980 990 1000
Frequency (MHz)

Figure 4.8: Measured power efficiency and output DC power as a function of frequency for a
class-F rectifier.

100

90
Efficiency (%)

80

70

60

50 Without mismatch loss


With mismatch loss
40
970 975 980 985 990 995
Frequency (MHz)

Figure 4.9: Class-F rectifier power efficiency with and without mismatch loss.

110
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

4.3.3 Class-F Amplifier and Rectifier Power Efficiency Analysis


Further insight into the small differences between the power efficiency of the amplifier
and rectifier duals can be obtained through analysis. It will be shown that switch losses do
not lead to perfect duality and dissipation inherently modifies the voltage waveform in the
rectifier relative to the amplifier. Also, device operation in quadrant I versus III modifies the
overlap losses in the amplifier and rectifier.
For the analysis, consider the class-F amplifier shown in Figure 4.10 and the waveforms
shown in Figure 4.11(a). In class-F, the ideal drain voltage, vd,P A (θ), is a square wave wave-
form with a 50% duty cycle. The Fourier series for the waveform consists of a fundamental
frequency component at f0 and only odd harmonics. The ideal current waveform, id,P A (θ), is
a half sinusoidal signal and the Fourier series for the waveform has a fundamental frequency
component at f0 and only even harmonics.

VDD

IDC ZL (f 0)
LDD
CB
vd,P A (θ)
Vout
d
id,P A (θ) iC (θ) TL1
(λo /4) L0 C0 ZL
Vg M1
Cout
Output Matching Network

Figure 4.10: A class-F power amplifier with a series quarterwave transmission line.

The shape of the voltage and current waveforms in the circuit depends on the harmonic
impedances presented to the switching device at the drain node, d. The tank circuit is anti-
resonant at the fundamental frequency f0 which forces the fundamental frequency component
of the drain current to pass through the load, RL . If the Q of the tank circuit is sufficient
high (Q > 5), then the impedance of the tank is small (ideally a short) at all the harmonic
frequencies. The transmission line is a quarter wavelength long at the fundamental frequency,
and the transmission line transforms the harmonic short created by the tank circuit into a
short circuit at the drain node d for even harmonics, and an open circuit at node d for odd
harmonics. Together, these conditions create impedance conditions at node d that lead to the
class-F waveforms.
The series transmission line also serves as a matching circuit, and the characteristic
impedance of the line, Zo , can be used to transform the load resistance, RL , to a funda-
mental frequency load line resistance, Rf 0 , where Rf 0 = Re[ZL (fo )]. The load line resistance
Rf 0 is used later in the power efficiency equations and is usually obtained from a load pull
simulation of the device.
The class-F amplifier has three primary loss factors which are analyzed.

1. PRon : conduction losses due to finite switch resistance;

2. Pcap : capacitive switching losses resulting from the discharge of the voltage stored on

111
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

vd,P A and id,P A


Id,max Vd,max

0 π 2π θ
(a)
vd,Rec and id,Rec

Vd,max

0 π 2π θ

−Id,max
(b)

Figure 4.11: Drain voltage and drain current waveforms for: (a) a class-F power amplifier and
(b) a class-F rectifier.

the output capacitance of the device, Cout ;

3. Poverlap : losses created during switching transitions when the drain voltage and drain
current waveforms overlap.

Each of these loss mechanisms modifies the ideal drain and current waveforms in the class-
F amplifier, and non-ideal waveforms are analyzed from the superposition of the different loss
mechanisms. Therefore, each loss mechanism is assumed to be independent and the losses are
combined to estimate the overall power efficiency of the class-F amplifier. The correspond-
ing losses in the rectifier dual are also analyzed. In the rectifier dual, it will be shown that
conduction losses are slightly less than the amplifier, and unlike the amplifier, there are no
overlap losses in the rectifier.

112
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

A. Conduction Losses
If the switch has finite on resistance, Ron , then the ideal voltage waveform is modified by
the voltage drop across the device during the on-state. Since the current is a half sine pulse,
the corresponding voltage drop is a half sine pulse. The on-state conduction loss is evident
in the voltage waveform shown in Figure 4.11(a). Based on this figure, the class-F amplifier
voltage and current waveforms including conduction losses are

Ron Id,max sin(θ) 0≤θ<π
vd,P A (θ) = (4.3)
Vd,max π ≤ θ ≤ 2π

and


Id,max sin(θ) 0≤θ<π
id,P A (θ) = (4.4)
0 π ≤ θ ≤ 2π.

In these equations, Vd,max is the maximum voltage across the device and equal to two times
that of the DC supply voltage, VDD . The current, Id,max , is the peak amplitude of the current
flowing through the device, and θ is a normalized time variable where θ = 2 π f0 t. Using these
equations, it is easy to calculate the conduction loss associated with the half sine voltage drop:

Id,max 2
Z 2π  
1
PRon = Ron i2d,P A (θ) dθ = Ron (4.5)
2π 0 2
The conduction loss is dissipated in the switching device M1 .
Now consider the time reversal dual of the amplifier circuit to construct an equivalent
rectifier. According to TR theory (see Table 4.1), the rectifier drain voltage waveform vRec (t)
is equal to vP A (−t) and the rectifier drain current waveform iRec (t) is equal to −iP A (−t). The
corresponding waveforms are shown in Figure 4.11(b). The current waveform is time reversed
and flipped, while the voltage waveform is only time reversed.
For the on-state in the rectifier, an exact time reversed voltage waveform would include the
positive oriented half sinusoidal pulse shown by the dashed line in Figure 4.11(b). However,
positive voltage and negative current during the on-state would imply negative power which is
not physically present in the switch. Therefore, in a dissipative switch, the dual of the voltage
waveform is modified during the on-state and includes a negative sinusoidal pulse as shown
by the solid line in Figure 4.11(b). From this, we conclude that the rectifier dual with losses
is not an exact TR dual, and the voltage waveform is modified by considering the dissipation
associated with finite switch resistance.
The drain voltage and drain current waveforms for the class-F rectifier assuming an on-
state switch resistance, Ron0 , are:


Vd,max 0≤θ<π
vd,Rec (θ) = 0 I (4.6)
Ron d,max sin(θ) π ≤ θ ≤ 2π

and

0 0≤θ<π
id,Rec (θ) = (4.7)
Id,max sin(θ) π ≤ θ ≤ 2π.

113
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

The corresponding conduction loss in the rectifier is

Id,max 2
Z 2π  
0 1 0 2 0
PRon = Ron id,Rec (θ) dθ = Ron (4.8)
2π 0 2
where the prime terms denote measures for the rectifier. The conduction loss in the rectifier is
0 , corresponds to device operation in
similar to the amplifier except the on-state resistance, Ron
quadrant III instead of quadrant I, as in the amplifier. Since on-state resistances in quadrant
I and III are similar, conduction losses in the amplifier and rectifier duals are similar and
PR0 on ≈ PRon .

B. Capacitive Switching Losses


All physical switching devices have output capacitance and this is modeled in the circuit
as Cout (see Figure 4.10). When the switch M1 is off, the DC drain inductor LDD charges
Cout to Vd,max , and when the switch is on, the capacitor is discharged through M1 and the
energy stored in the capacitor is dissipated in the device creating a capacitive switching loss.
Therefore, the power loss is associated with the falling edge of the drain voltage waveform in
Figure 4.11(a).
The energy stored in Cout is 21 Cout Vd,max
2 . The discharge occurs with a frequency f0 ;
therefore the total power dissipated is
1 2
Pcap = Cout Vd,max fo . (4.9)
2
We now consider the capacitive switching loss in the rectifier dual. With reference to
the drain voltage waveform in the rectifier circuit shown in Figure 4.11(b), it is clear that
the capacitor Cout is charged to Vd,max and discharged to zero similar to the amplifier cir-
cuit. However, one constraint on the discharge waveform is that it must occur prior to the
switch turning on. During the discharge (falling edge of the voltage waveform), the voltage is
positive, and consequently, the discharge current must also be positive to create positive dis-
sipation otherwise negative power would be generated. Therefore, the total power dissipated
0 .
by discharging the capacitor Cout is the same as for the amplifier; hence, Pcap = Pcap

C. Overlap Power Loss


The third loss mechanism that is important to consider in the class-F amplifier and rectifier
is overlap loss. In order to analyze overlap loss, the ideal waveforms shown in Figure 4.11
need to be modified to include the rise and fall time of the voltage waveforms. The modified
waveforms are shown in Figure 4.12. For analysis, it is assumed that the rise and fall times
change linearly over a time interval τ .
At this point, before continuing with the analysis, it is insightful to consider a qualitative
comparison of overlap in the amplifier and the rectifier dual. In the amplifier, the drain
current and drain voltage are always in quadrant I of the IV plane for the device. This
means that overlap loss is primarily determined by the harmonic impedances in the matching
network and the gate waveform. The situation for the rectifier dual is different. When the
device is off, the current and voltage waveforms are in quadrant I, and when the device is
on, the current and voltage are in quadrant III. The implication of the on-state in quadrant
III means that the voltage waveform in the off-state has to fully discharge to zero before the
current can flow in quadrant III. Another way to describe the current and voltage constraints

114
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

Id,max

vd,P A and id,P A


Vd,max

0 π 2π θ
τ τ
(a)
vd,Rec and id,Rec

Vd,max

0 π 2π θ
τ τ

−Id,max (b)

Figure 4.12: Drain voltage and drain current waveforms with overlap loss for the class-F
amplifier and rectifier duals.

is that the dissipation must be positive since negative power cannot be generated in the device.
Therefore, the only voltage waveform that can co-exist with the negative on-state drain current
is a negative drain voltage waveform. The quadrant III constraints in the rectifier dual lead
to the modified voltage and current waveforms shown in Figure 4.12. The waveforms show
that unlike the amplifier, the rectifier dual does not have overlapping current and voltage
waveforms during the transition intervals. From this observation we conclude that there is no
overlap power loss in the rectifier dual.
In the class-F amplifier, the voltage and current waveforms overlap during the transition
intervals. Overlap loss during a switch transition is defined as the power loss associated with
the cross-over of the current and voltage waveforms as shown in Figure 4.12(a). Using a linear

115
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

approximation over a transition time of τ , the overlap loss is


Z 2π
1
Poverlap = vd (θ) id (θ) dθ
2π 0
Z τ
−Iτ
 
1 Vd,max
=2× θ (θ − τ ) dθ (4.10)
2π 0 τ τ
Vd,max Id,max
= τ sin(τ ).

The overlap power loss is consistent with the condition that the loss is zero when τ is zero.
Factors which affect the shape of the current and voltage waveforms will determine the
length of the transition interval and consequently the amount of overlap loss dissipated in
the amplifier. An important factor which affects waveshapes are the harmonic termination
impedances in the output and input matching networks. The transition time can also be af-
fected by the size of the output device capacitance, Cout , because the device has a maximum
available current to discharge the capacitance. However, it should be noted that the amount
of power dissipated to discharge the capacitance is in addition to the overlap loss calculated
above because the waveforms in Figure 4.12 do not include the current required to discharge
the capacitor. Therefore, the total power loss in a class-F amplifier includes conduction loss,
capacitive switching loss and overlap loss, while the total power loss in the rectifier dual only
includes conduction and capacitive switching losses. Based on these losses, we expect the
rectifier dual to have higher power efficiency compared to the amplifier.

D. Power Efficiency
Equations to predict the power efficiency of class-F amplifiers and the corresponding rec-
tifier TR dual are derived next using the three loss mechanisms. For the amplifier, the input
source power is the DC source power, PDC . The DC power source power must equal the sum
of the output power delivered to the load, Pout , and the power dissipated in the switching
device:
PDC = Pout + Ploss = Pout + PRon + Pcap + Poverlap . (4.11)
For the rectifier, the source power is the input RF power, Pin , and the source power must
equal the sum of the DC power delivered to the load PDC0 plus the power dissipated in the
rectifying device:
0 0 0
Pin = PDC + Ploss = PDC + PR0 on + Pcap
0
. (4.12)
Using these expressions, the drain efficiency of the amplifier is
Pout Pout
ηa = = (4.13)
PDC Pout + PRon + Pcap + Poverlap
and the RF to DC conversion power efficiency of the rectifier is
0
PDC 0
PDC
ηr = = 0 . (4.14)
Pin PDC + PR0 on + Pcap
0

Expressions for all the terms in the power efficiency equations have been found except for
0 . These terms are found from the Fourier series expansions of the waveforms in
Pout and PDC
Figure 4.11.

116
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

The amplitude of the fundamental frequency component of the drain voltage vd,P A (θ) is

2 Id,max
V f0 = Vd,max − Ron (4.15)
π 2
and the amplitude of the fundamental frequency component of the drain current id,P A (θ) is

Id,max
I f0 = . (4.16)
2
The fundamental frequency components are in-phase across the load and the total load power
is

Vd,max Id,max Ron Id,max 2


 
Vf o If o
Pout = = − . (4.17)
2 2π 2 2

The equation can also be written in terms of the fundamental frequency load line impedance
at the drain. Define Rf 0 as the ideal load line resistance at the drain assuming a perfect
switch (Ron = 0). Then,

Vf 0 4 Vd,max
Rf 0 = = (4.18)
If 0 Ron =0 π Id,max

and another expression for Pout is


 2
1 Id,max
Pout = (Rf 0 − Ron ) . (4.19)
2 2

Using expressions for Pout , PRon , Pcap , and Poverlap , the power efficiency of the class-F amplifier
is
1
ηa =
PRon Pcap Poverlap
1+ + +
Pout Pout Pout
1 (4.20)
= 2 .
2Ron 2
π Rf 0 Cout fo τ sin(τ ) Rf 0
1+ + +
Rf 0 − Ron 4 Rf 0 − Ron 3 Rf 0 − Ron
0
For the class-F rectifier, we need an expression for PDC to evaluate equation (4.14). Similar
to the amplifier, the DC terms of the Fourier series for the rectifier waveforms in Figure 4.11(b)
are

0 Vd,max Id,max
VDC = − Ron (4.21)
2 π
and

0 Id,max
IDC = . (4.22)
π
The DC output power of rectifier is therefore

117
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

 2
0 0 0 Vd,max Id,max 0 Id,max
PDC = VDC IDC = − Ron
2π π
(4.23)
Id,max 2
 2  
π 0
= Rf 0 − Ron .
8 π
0 , the power efficiency of the class-F rectifier can be expressed as
Using this expression for PDC
1
ηr =
PRon
0 Pcap0
1+ 0 + 0
PDC PDC
1 (4.24)
= 2 .
0
2 Ron π 2 Rf 0 Cout fo
1+ +
Rf 0 − π82 Ron
0 4 Rf 0 − π82 Ron
0

A comparison of the power efficiency equations for the class-F amplifier and rectifier shows
that the primary difference between the circuit duals is that the rectifier does not have overlap
loss. Therefore, we expect the power efficiency of the rectifier to be higher than the amplifier.

0
E. Expressions for RDC and RDC
Analytical expressions for the Thévenin equivalent resistance of the DC power supply in
the amplifier, RDC , and the DC load resistance in the rectifier RDC0 are derived next. Since
the power loss is different for the amplifier and rectifier, these two resistances are slightly
different.
For the class-F amplifier, the DC components of the Fourier series for the class-F waveforms
in Figure 4.11(a) are
Vd,max Id,max
VDC = + Ron (4.25)
2 π
and
Id,max
IDC = (4.26)
π
where the sign of IDC is chosen to be consistent with notation in Figure 4.3. The Thévenin
resistance of the DC supply is then
 
VDC π Vd,max
RDC = = + Ron
IDC 2 Id,max
(4.27)
π2
= Rf 0 + Ron .
8
The DC components for the rectifier waveforms were given earlier in equations (4.21) and
(4.22). Using these expressions,
0
0 VDC π Vd,max 0
RDC = 0 = − Ron
IDC 2 Id,max
(4.28)
π2 0
= Rf 0 − Ron .
8
Comparing this equation with the Thévenin equivalent resistance of the amplifier DC supply,
we expect that the optimal load resistance for the rectifier to be less than the amplifier.

118
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

4.3.4 Simulation Results


The analytical equations are verified by comparing predicted performance with simulation
results for the class-F power amplifier model in Figure 4.10. For the simulation, a level 2
device model is used without device capacitances. Instead, the total effective output device
capacitance is modelled as Cout in Figure 4.10, the same capacitance used to analyze capacitive
switching losses. The gate drive for the amplifier consists of a 24.5 dBm sinusoidal source at
a frequency fo of 985 MHz and a gate bias of -2.6 V. The drain supply voltage, VDD is 27 V
and the effective on resistance of the device for these bias conditions is estimated to be 2.2 Ω.
The resonant tank circuit is chosen to have a Q of 5 for a load resistance of 50 Ω.

100

95

90

85
Efficiency (%)

80

75

70

65
Analytical model for amplifier
60 Analytical model for rectifier
Simulation model for amplifier
55
Simulation model for rectifier
50
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Cout (pF)

Figure 4.13: Estimated drain efficiency of class-F PA and rectifier as a function of output
capacitance. Ron for both the amplifier and rectifier are 2.2 Ω.

Under these conditions, time-domain simulations were run for both a class-F amplifier and
the rectifier dual as Cout is swept over a range of 0 to 1.6 pF. A comparison of the simulated
and analytical results is shown in Figure 4.13. As Cout changes, the switching losses in the
amplifier change and the drain current/voltage overlap changes. The overlap interval, τ ,
was obtained from the simulated drain voltage and drain current waveforms and varied from
0.065 π to 0.15 π radians, as Cout varied from 0.2 to 1.6 pF, respectively.
Although a simplified device model is used to validate the analytical power efficiency rela-
tions, a good reference point for comparison with the experimental results shown in section 4.3
is to consider the case of Cout equal to 1 pF. The effective output capacitance of the Cree
CGH40010 device used in the experimental work is approximately 1 pF. The corresponding
simulated and analytical results are summarized in Table 4.4 and can be compared with the
experimental results in Table 4.3. Although not an exact match, the analytical results show
that the optimal DC load (RDC 0 ) for the rectifier is expected to be less than the equivalent

Thévenin resistance (RDC ) of the DC supply which is consistent with the experimental results.
The analytical and simulated results predict a slightly higher power efficiency for the rectifier

119
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

Table 4.4: Comparison of class-F amplifier and rectifier circuit duals.

Parameter Amplifier Rectifier

Frequency (fo ) 985 MHz 985 MHz


Peak drain voltage (Vd,max ) 52 V 52 V
Peak drain current (Id,max ) 1.3 A 1.3 A
Fundamental frequency load resistance (Rf 0 ) 50.8 50.8
Output device capacitance (Cout ) 1 pF 1 pF
Current/voltage overlap (τ ) 0.12 π rad 0
Gate bias -2.6 V -2.6 V
DC supply/load resistance RDC = 64.6 Ω 0
RDC = 60.3 Ω
Input source power PDC = 12 W Pin = 12 W
Load power Pout = 9.5 W 0
PDC = 9.8 W
Power efficiency ηa = 79% ηr = 81.7%
Conduction loss (analytic) 0.93 W 0.93 W
Capacitive switching loss (analytic) 1.33 W 1.33 W
Overlap loss (analytic) 0.49 W -
Power Efficiency (analytic) 79.6% 82.1%

(3% in this case) which is similar to the experimental results where the rectifier efficiency was
4.8% higher than the amplifier. Also, the analytic, simulation and experimental results for
power efficiency are all within 2% of each other which demonstrates good agreement between
theory and experiment.
One of the advantages of constructing an analytical model is that it provides a way to
explore the contribution of different loss mechanisms to the overall power efficiency of the
amplifier and rectifier duals. A breakdown of losses are shown in Figures 4.14 and 4.15
for the class-F amplifier and rectifier, respectively. Conduction losses are independent of
Cout and contribute a fixed loss to both the amplifier and rectifier. On the other hand,
capacitive switching losses increase as Cout increases for both the amplifier and rectifier. For
a capacitance of 1 pF, conduction losses reduce power efficiency by approximately 8%, while
switching losses reduce efficiency by about 10%. In the amplifier, there is an additional
power loss from the overlap of the drain voltage and current waveforms, and for a capacitance
of 1 pF, overlap loss reduces efficiency by approximately 3%. The figures also include the
simulation results which closely follow the analytical results confirming the theory which has
been developed to predict power efficiency in the class-F amplifier and rectifier duals.

120
4.3. High Efficiency GaN HEMT Class-F Synchronous Rectifier

100

95
Efficiency (%)
90

85

80 PRon
P +P
Ron cap
75
PRon + Pcap + Poverlap
Simulation Results
70
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Cout (pF)

Figure 4.14: Predicted losses in a class-F power amplifier as a function of output capacitance.

100

95
Efficiency (%)

90

85

80
PRon
75 PRon + Pcap
Simulation Results
70
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Cout (pF)

Figure 4.15: Predicted losses in a class-F rectifier as a function of output capacitance.

121
4.4. High Efficiency GaN HEMT Inverse Class-F Synchronous Rectifier

4.4 High Efficiency GaN HEMT Inverse Class-F


Synchronous Rectifier
The class-F synchronous rectifier is a voltage switched rectifier. The class-F rectifier can
also be implemented in an inverse configuration where the current is switched. In this section,
experimental results for a class-F−1 synchronous rectifier are presented for the time reversed
dual of the class-F−1 amplifier described in Chapter 3. The class-F−1 uses the same Cree
10 W GaN device used in the class-F rectifier design, and the performance of these two
rectifier designs are compared.
In Section (3.4) of Chapter (3) a 10 W class-F−1 amplifier was designed and implemented.
In this section, the amplifier is reconfigured as a high power GaN RF synchronous class-
F−1 rectifier by adding feedback from the output to the input. The rectifier is tested under
identical source power condition associated with the dual amplifier. The measurement results
for the class-F−1 rectifier are also compared with measurement results for the class-F rectifier
design presented in Section (4.3) in terms of performance and dynamic range.
The test bench for the class-F−1 rectifier is shown in Figure 4.16. Similar to the class-F
rectifier, a -18.7 dB input coupler (coupler A) and a phase shifter are used to create a gate
drive to synchronously switch the device. The other port of coupler A is connected to a power
meter to measure the reflected input power. The high power RF input signal (Pin ) from the
rectifier is generated by a 10 W class-F amplifier, the same amplifier used to implement the
class-F synchronous rectifier. A second coupler, coupler B, is inserted between the rectifier
and amplifier to measure the available RF input power (Pin ) delivered to the rectifier.

Rectified Voltage

Power Meters

B A
Class-F PA
Couplers

RF signal Inverse Class-F


generator Rectifier

Figure 4.16: Test bed for the class-F−1 rectifier. A class-F amplifier is used as a high power
RF input source.

As an initial test, the performance of the rectifier is compared with the performance of
the amplifier under equivalent test conditions. From Chapter 3, the amplifier had a power
efficiency of about 83% for a DC source power of 11.9 W. When the RF input power to the
rectifier has the same source power as the amplifier (11.9 W), the rectifier has an RF to DC

122
4.4. High Efficiency GaN HEMT Inverse Class-F Synchronous Rectifier

conversion efficiency of 85%, slightly higher than the amplifier. The results are consistent
with observations made for the class-F amplifier which also shows a slightly higher efficiency
for the rectifier configuration. The corresponding equivalent Thévenin resistance of the DC
supply for the amplifier, RDC , is 48.4 Ω compared to the optimum DC load resistance, RDC 0 ,

which was found to be 47 Ω. Again, the observation that the DC load resistance for the
rectifier is slightly less than the Thévenin resistance of the amplifier is consistent with the
class-F experiments. The input reflection coefficient under these test conditions is -14.25 dB
and if efficiency is calculated without mismatch loss, the efficiency of the rectifier is 88%.
A summary of the test results for the class-F−1 amplifier and rectifier duals are given in
Table 4.5.

Table 4.5: Comparison of class-F−1 amplifier and rectifier experimental results.

Parameter Amplifier Rectifier

Frequency 910 MHz 910 MHz


Gate bias -2.8 V -2.8 V
DC supply/load resistance RDC = 48.4 Ω 0
RDC = 47 Ω
Input source power PDC = 11.9 W Pin = 11.9 W (40.8 dBm)
Load power Pout = 9.8 W 0
PDC = 10.15 W
Power efficiency ηa = 83% ηr = 85%
Power efficiency without mismatch loss - ηr = 88%

Other test results for the rectifier are shown in Figures 4.17 and 4.18. In Figure 4.17, it is
seen that the optimal DC load resistance, RDC 0 , is about 47 Ω, while in Figure 4.18, it is seen

that peak efficiency is obtained at a frequency of 910 MHz. As with the class-F rectifier, the
bandwidth of the class-F−1 is fundamentally limited by the bandwidth of the multiharmonic
matching network in the amplifier. For this design, power efficiency remains above 70% over
an 80 MHz frequency range.
The most interesting experimental results for the class-F−1 rectifier relate to the relative
dynamic range of the rectifier compared to the class-F rectifier. These results are shown in
Figure 4.19. Two observations are made. First, the peak efficiency of the class-F−1 rectifier is
higher than the peak efficiency of the class-F rectifier for the same RF input power conditions.
This suggests the switching losses are lower in the class-F−1 circuit; this can be explained by
the difference between zero voltage switching in class-F−1 as opposed to hard switching in
class-F. The second observation is that the dynamic range of the class-F−1 rectifier is much
larger than the class-F rectifier. For example, for a minimum power efficiency of 60%, the class-
F rectifier has a 6 dB dynamic range compared to the class-F−1 rectifier which has a 16 dB
dynamic range, 10 dB higher than class-F. From this comparison, which uses identical devices
in two different circuit topologies, it shows that the class-F−1 RF rectifier has significantly
better overall performance compared to a class-F RF rectifier. Therefore, it is concluded that
a class-F−1 is the preferred circuit topology, a result which does not appear to be clearly

123
4.4. High Efficiency GaN HEMT Inverse Class-F Synchronous Rectifier

86 10.4
84 10.2
82 10

Output Power (W)


80 9.8
Efficiency (%)
78 9.6
76 9.4
74 9.2
72 9
70 8.8
68 8.6
66 8.4
20 25 30 35 40 45 50 55 60 65 70
RDC (Ω)

Figure 4.17: Measured RF to DC conversion efficiency and output DC power versus load
resistance for the rectifier. The measurements conditions are for an input RF source power
of 40.76 dBm at a frequency of 910 MHz.

90 11

80 10

70 9
Output Power (W)
Efficiency (%)

60 8

50 7

40 6

30 5

20 4

10 3
860 880 900 920 940 960 980 1000
Frequency (MHz)

Figure 4.18: Measured power efficiency and output power as a function of frequency for the
rectifier. The measurements conditions are for an input RF source power of 40.76 dBm.

presented in the existing literature.

124
4.4. High Efficiency GaN HEMT Inverse Class-F Synchronous Rectifier

90

80
Efficiency (%)

70

60

50

40
Class−F−1 Rectifier
Class−F Rectifier
30
20 25 30 35 40 45
Input Power (dBm)

Figure 4.19: Power efficiency comparison of a class-F and class-F−1 synchronous rectifier.
Experimental results are shown.

125
4.5. High Efficiency GaN HEMT Wideband Inverse Class-F Synchronous Rectifier

4.5 High Efficiency GaN HEMT Wideband Inverse Class-F


Synchronous Rectifier
The class-F and class-F−1 synchronous rectifier designs presented so far are inherently
narrowband because of the design of the multiharmonic matching networks in the amplifier.
In some applications, such as in energy recycling, the spectrum of the RF input signal can
be spread over a large bandwidth and wideband rectifier designs are of interest. In other
applications, such as wireless power and RF energy harvesting, wideband rectifiers are also
useful. In this section, experimental results for a wideband class-F−1 are presented.
The same test bench for the narrowband class-F−1 rectifier shown in Figure 4.16 is used
for measuring the performance of the wideband rectifier. Frequency errors introduced by the
bandlimited response of the power amplifier are compensated by measuring the RF input
power at each frequency using coupler B. The power meter connected to coupler B measures
the incident RF input power, while the power meter connected to coupler A measures the
reflected input power.
The power efficiency of the rectifier as a function of frequency is shown in Figure 4.20.
The results show that power efficiency remains above 60% over a bandwidth of approximately
500 MHz ranging from 600 MHz to 1150 MHz. The peak efficiency is 80.1% at a frequency of
650 MHz. For these measurements, the DC load resistance is 34 Ω and the RF input power
is 10 W. Also, in these measurements, the phase shifter has been tuned at each frequency to
maximize power efficiency.

100

90

80

70
Efficiency (%)

60

50

40

30

20

10

0
600 700 800 900 1000 1100
Frequency (MHz))

Figure 4.20: Measured drain efficiency as a function of frequency for the wideband class-F−1
rectifier.
The rectifier performance as a function of the RF input power is shown in Figure 4.21.
Results are shown for frequencies of 650 MHz, 850 MHz and 1050 MHz and compared with the
narrowband rectifier measurements in Figure 4.22. The wideband rectifier has good efficiency
and dynamic range over a wide frequency range, while the narrowband designs have good

126
4.5. High Efficiency GaN HEMT Wideband Inverse Class-F Synchronous Rectifier

performance at a specific frequency. Therefore, the bandwidth of the rectifier can be designed
to match the bandwidth of the RF input spectrum which is to be rectified.

100
Wideband: 650 MHz
90
Wideband: 850 MHz
80
Wideband: 1050 MHz
70
Efficiency (%)

60

50

40

30

20

10

0
20 25 30 35 40
Input Power (dBm)

Figure 4.21: Measured drain efficiency as a function of input power for the wideband class-F−1
rectifier at frequencies of 650 MHz, 850 MHz and 1050 MHz.

100
Class F−1 Rectifier
90 Class F Rectifier
Wideband: 650 MHz
80 Wideband: 850 MHz
Wideband: 1050 MHz
70
Efficiency (%)

60

50

40

30

20

10

0
20 25 30 35 40
Input Power (dBm)

Figure 4.22: Measured drain efficiency as a function of input power for class-F, class-F−1 and
wideband class-F−1 synchronous rectifiers.

127
4.6. Chapter Summary

4.6 Chapter Summary


The design of class-F RF synchronous rectifiers based on the application of time reversal
duality has been demonstrated. The implications of loss in the switch-mode amplifier proto-
type were considered and a set of equivalent conditions were proposed to evaluate the rectifier
dual. A direct comparison was then made between the amplifier and rectifier circuits using the
class-F amplifiers described in Chapter 3. The experimental results provide new benchmarks
for high power class-F synchronous rectifiers including a wideband rectifier design. The RF
synchronous rectifiers described in this chapter are used in an energy recycling switch-mode
power amplifier described in the next chapter.

128
Chapter 5

Switch-mode Power Amplifier with


Energy Recycling
5.1 Energy Recycling in Outphasing Power Amplifiers
The concept of energy recycling in power amplifiers was first introduced in 1999 for out-
phasing amplifiers [125]. In outphasing amplifiers, the modulated source signal is mapped to
two constant envelope signals, amplified, and after amplification, the signals are combined
to reconstruct the original modulated source signal. The primary motivation behind out-
phasing schemes is to exploit the high efficiency operation of amplifiers that can be obtained
when amplifying constant envelope signals. The signal mapping commonly used in outphasing
amplifiers is called LINC, linear amplification using nonlinear components.
Although the outphasing concept is conceptually attractive as a way to implement high
efficiency amplifiers, the practical difficulties of signal combining and reconstruction at the
output of the amplifier have been difficult to solve. Two approaches have been used in
implementing outphasing amplifiers. The first approach is to reactively combine the output
of the two branches. This method is called Chireix outphasing based on the original invention
of the outphasing amplifier which was described in 1935 [2]. By reactively combining the
amplifiers, there is no isolation between the two amplifier branches and the primary design
challenge is to maintain efficient combining over a large dynamic range to accommodate the
amplification of signals with high peak to average power ratios. A block diagram of a Chireix
outphasing amplifier is shown in Figure 5.1(a).
The second approach to outphasing that has been proposed uses an isolating signal com-
biner [125, 126]. In this method, the interaction between the two amplifier branches is mini-
mized and both amplifier branches can operate with very high efficiency. The disadvantage of
isolated combining is that significant power is dissipated in the combiner especially for signals
with high PAPR. Therefore, the overall power efficiency of the amplifier is reduced signifi-
cantly by the isolated combiner. As a way to offset the power loss in an isolated combiner,
energy recycling has been proposed. Rather than dumping the RF power in the combiner to
an internal load, the power is rectified and returned to the DC supply of the amplifier. A block
diagram of this method is shown in Figure 5.1(b). The published work on energy recycling
in outphasing amplifiers includes simulation work, measurements of RF to DC rectification
efficiency, and some implementations of complete systems [125, 126, 127, 128].

5.2 Energy Recycling in RF Switch-mode Amplifiers


In an analogous way, the two approaches which have been taken in exploring implementa-
tions of outphasing amplifiers can also be applied to switch-mode power amplifiers using pulse
modulation techniques. The reactive signal reconstruction approach has been the mainstay

129
5.2. Energy Recycling in RF Switch-mode Amplifiers

c1 (t)

si (t) Signal Constant Non-isolated Reactive so (t)


Envelope VDD combining Signal
Mapping
Modulated Signals (load modulation) Combiner
Source
Signal
c2 (t)
Saturated or
(a) Switch-mode PA’s

c1 (t)

si (t) Signal Constant Energy a(t) Isolated so (t)


Envelope VDD Signal
Mapping Recycling
Modulated Signals Combiner
Source
Signal
c2 (t)
Saturated or
(b) Switch-mode PA’s

Figure 5.1: Outphasing amplifiers: (a) reactive signal combining and (b) isolated signal com-
bining with energy recycling.

of switch-mode amplifier work where a filter (a reactive structure) is used to reconstruct the
output signal from the pulse modulated signal while trying to simultaneously create out-of-
band impedances that lead to high efficiency switching in the amplifying device. A block
diagram for this type of switch-mode power amplifier is shown in Figure 5.2(a). Similar to
Chireix outphasing, the reactive approach is in practice difficult to implement and much work
remains to be done to implement high efficiency switch-mode amplifiers with reactive signal
reconstruction.
A second approach to switch-mode power amplifiers is to employ energy recycling in the
amplifier by terminating the switch in a broadband load instead of a reactive out-of-band
load. The disadvantage of this approach is that power is now dissipated in the out-of-band
spectrum. As a way to recapture this power, energy recycling has been proposed as an
efficiency enhancement for this type of amplifier [129].
If energy recycling is to be implemented in a switch-mode power amplifier, the signal
reconstruction block must include signal separation to isolate the out-of-band power. One way
to implement signal separation is to use a complementary diplexer [129, 84]. The diplexer is a
three port filter structure where the input port is split into two complementary filter branches.
One branch is a bandpass filter that isolates the in-band signal spectrum that is transmitted
to the antenna, while the other filter branch isolates the out-of-band signal spectrum which
can then be rectified to recapture out-of-band power. The insertion loss of the diplexer is
critical to the overall performance of the amplifier and a stripline design reported in [84] has

130
5.3. Spectral Shaping to Enhance Energy Recycling Efficiency

a loss of approximately 2 dB in the in-band path and 0.8 dB in the out-of-band path.
Another way to implement signal separation is to use a circulator and a bandpass filter.
A block diagram of a switch-mode amplifier with this type of signal separation is shown in
Figure 5.2(b). The circulator is a non-reciprocal device and out-of-band power reflected by
the in-band bandpass filter is reflected back to the circulator and coupled to an isolated port.
The isolated port, port C in Figure 5.2(b), can then be connected to a RF rectifier to recover
power from the out-of-band spectrum. As with the complementary diplexer, the insertion
loss of the isolator is critical to the overall power efficiency of the architecture. Later, in
Section (5.5), experimental results are presented for a switch-mode power amplifier using a
circulator for signal separation.

VDD
Out-of-band reactive
BPF
termination
si (t) Pulse p(t) so (t)
Encoder
Modulated In-band
Source Signal Switch-mode power Output Signal
Signal Mapping PA Reconstruction
Filter
(a)

VDD Energy a(t)


Recycling Out-of-band
C power BPF
si (t) Pulse p(t) A B so (t)
Encoder
Modulated In-band
Source Signal Switch-mode Circulator power Output Signal
Signal Mapping PA Reconstruction
Filter
(b)

Figure 5.2: Switch-mode power amplifiers (a) with reactive output filter and (b) with energy
recycling.

5.3 Spectral Shaping to Enhance Energy Recycling Efficiency


If energy recycling is employed in a switch-mode amplifier, it is desirable to shape the out-
of-band spectrum to improve rectification efficiency. The out-of-band spectrum generated
by the encoding process usually has a very large bandwidth which is created by quantizing
the source signal. Implementing a high efficiency broadband RF rectifier is more challenging
than a narrowband rectifier, therefore spectral shaping to concentrate out-of-band power in a
reduced frequency range could improve rectification efficiency of out-of-band energy.

131
5.3. Spectral Shaping to Enhance Energy Recycling Efficiency

As an example of out-of-band spectral shaping, a noise shaped pulse position modulator


with an adaptive sinusoidal dither signal has been implemented. The block diagram of the
encoder is shown in Figure 5.3. The encoder consists of a negative feedback loop with a noise
shaping filter H(s) and a pulse generator. The pulse generator creates a pulse width equal to
the half the period of the carrier frequency of the source signal (Tc /2). The zero-crossings of
the error signal e(t) determine the delay (position) of the pulse and a rising edge triggers the
pulse generator. The pulse generator creates an amplitude quantized output pulse train and
the quantization noise is shaped by the noise shaping filter H(s). In this way, the loop is very
similar to a sigma-delta modulator except that the timing of pulse edges are asynchronous
rather than synchronous as in sigma-delta modulation.

Dither
Amplitude fdither
Control
d(t)
s(t) e(t) p(t)
H(s)
Modulated Encoded
Source Signal Noise T /2 Pulse Signal
Shaping Filter Generator

Amplitude p(t)
1
s(t)
0

-1 t/T
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2

Figure 5.3: Block diagram of a noise shaped PPM encoder with dither (top) and example
input and output waveforms (bottom).

The noise shaping loop includes a sinusoidal dither signal d(t) which is added to the error
signal e(t). The amplitude of the dither signal is controlled by the source signal envelope.
When the source envelope has a small amplitude, the out-of-band quantization noise has much
higher power than the signal and the dither amplitude is large. Conversely, when the source
signal envelope is at peak envelope power, then the dither amplitude reduces to zero because
most of the power in the output pulse train is signal power. The exact amplitude mapping
function used to control the dither amplitude depends on a compromise between signal to
noise ratio (SNR), frequency of the dither signal, and loop stability. Examples of the output
spectrums from the encoder with and without spectral shaping are shown in Figure 5.4.

132
5.3. Spectral Shaping to Enhance Energy Recycling Efficiency

Relative Power Spectral Density (dB)

10
without spectral shaping
0

−10

−20 signal
encoded in
noise well
−30

−40

−50

−60
850 900 950 1000 1050
Frequency (MHz)
Relative Power Spectral Density (dB)

10
with spectral shaping
0 adaptive dither
to improve
−10 out−of−band
rectification
−20 efficiency

−30

−40

−50

−60
850 900 950 1000 1050
Frequency (MHz)

Figure 5.4: Power spectrum of a noise shaped PPM signal without out-of-band spectral
shaping (top) and with spectral shaping (bottom).

133
5.4. Analysis of Power Efficiency Enhancement using Energy Recycling

5.4 Analysis of Power Efficiency Enhancement using Energy


Recycling
An RF power amplifier with energy recycling is shown in Figure 5.5. The block diagram
is sufficiently general to capture the primary concepts of recycling in both outphasing and
switch-mode power amplifiers. A modulated RF carrier s(t) is mapped to a set of signals ck (t)
with reduced amplitude variation; for example, constant envelope signals. The signal mapping
generates N input signals for the high efficiency amplifier block. For the cases considered here,
a switch-mode amplifier has N = 1 and LINC has N = 2.
0
PDC
ηer

Per Energy Padd G31


Recycling
PDC 3
ηd0
Pin Psig + Padd G31 Psig G21
Signal
Mapping
s(t) ck (t) N N 1 G21 2
RL
High Efficiency Signal Combiner /
Amplifier(s) Separator

Figure 5.5: Block diagram of a power amplifier with energy recycling.

The outputs of the amplifier are combined in a signal reconstruction block. The signal
reconstruction block also has signal separation and spectral power added by the signal map-
ping function can be isolated. An example of a signal combiner/separator in an outphasing
amplifier design is a 180 degree hybrid which provides sum and difference output signals.
For switch-mode amplifiers, the signal reconstruction and separation can be implemented by
either a complementary diplexer or the combination of a circulator and bandpass filter. The
power loss associated with signal reconstruction is modeled by the gain term G21 from port
1 to port 2, and power loss associated with signal separation to recover dissipated power is
modelled by the gain term G31 from port 1 to port 3. The output path gains are lossy and
consequently the path gains are less than unity: 0 < |G| ≤ 1.
An expression for the overall drain efficiency of the amplifier with energy recycling is
derived next. The native drain efficiency of the power amplifier is defined as the power
efficiency of the amplifier block when amplifying the signals ck (t) under a load condition
equivalent to the input impedance of the signal combining network at port 1. The output
power spectrum from the amplifier can be partitioned into the desired signal power (Psig ) and
power added by the signal mapping function (Padd ). Therefore, the native drain efficiency of
the amplifier is
Psig + Padd
ηd0 = (5.1)
PDC
where PDC is the DC input power to the amplifier.
After signal reconstruction, the total output power delivered to the load at port 2 is equal
to Psig G21 . The signal combiner also outputs residual power which is created by the signal
reconstruction process. The total residual power available at port 3 is Padd G31 . The residual

134
5.4. Analysis of Power Efficiency Enhancement using Energy Recycling

power is rectified to generate an auxiliary DC supply that is summed with the main supply.
The return of output RF power back to DC supply power implements an energy recycling
loop.
The DC power provided to the amplifier consists of an external source of power PDC 0 and
recycled power Per . The available recycling power depends on both the efficiency of the energy
recycling block (ηer ) as well as the loss G31 in the signal reconstruction block. Therefore,

Per = Padd G31 ηer . (5.2)

The overall power efficiency of the amplifier can then be written as


Psig G21 Psig G21
ηd = 0 = . (5.3)
PDC PDC − Per

The equation for overall power efficiency can be expanded to write it in terms of the native
power efficiency ηd0 . Using (5.2) in (5.3),
Psig G21 Padd + Psig
ηd = ×
PDC − Padd G31 ηer Padd + Psig
Padd + Psig Psig /(Padd + Psig ) G21
= ×
PDC 1 − (Padd /PDC ) G31 ηer
(5.4)
Psig /(Padd + Psig ) G21
= ηd0 ×
1 − [(Psig + Padd )/PDC ] [Padd /(Psig + Padd )] G31 ηer
Psig /(Padd + Psig ) G21 ηd0
= .
1 − [Padd /(Psig + Padd )] G31 ηer ηd0

A more concise expression for the overall power efficiency of the energy recycling amplifier
can be obtained by adopting the concept of coding efficiency. Coding efficiency is commonly
used as metric to evaluate how efficient the pulse encoder is in a switch-mode power amplifier.
However, coding efficiency can be applied more broadly to any class of signals that adds power
to the original source signal spectrum. The definition of coding efficiency is
in-band signal power Psig
CE = = (5.5)
total power Psig + Padd

and it gives a measure of how much signal power is added by the signal mapping block to
create a set of output signals. For LINC, the RF modulated source signal, s(t), is mapped
to two constant envelope signals, c1 (t) and c2 (t), and the signal mapping adds power to the
source signal. Therefore, coding efficiency can be applied to LINC in the same way as it is
used as a metric to quantify the efficiency of pulse encoders in switch-mode amplifiers.
Using the definition of coding efficiency and equation (5.5), we also get the relation Psig =
(Psig + Padd )CE. Therefore the overall power efficiency of an amplifier with energy recycling
is
ηd0 G21 CE
ηd = . (5.6)
1 − ηd0 ηer G31 (1 − CE)

Equation (5.6) is very useful for exploring the theoretical bounds of power efficiency for
different power amplifier scenarios. A number of examples are shown in Figure 5.6. As a
first example, consider the upper bound for a power amplifier with a native power efficiency,

135
5.5. Experimental Implementation of a Switch-mode Power Amplifier with Energy Recycling

ηd0 , of 80% and with perfect energy recovery and lossless signal separation.3 Perfect energy
recovery is defined as an RF to DC rectification efficiency, ηer , of 100%, and lossless signal
separation means G21 and G31 are both 0 dB. For this case, if the coding efficiency of the
constant envelope signal were 25%, then the corresponding power efficiency of the amplifier is
28% without energy recycling and 50% with energy recycling. Under these conditions, energy
recycling boosts power efficiency by 78% compared to an amplifier without energy recycling:
(50% - 28%)/28% x 100% = 78%. For this scenario, energy recycling clearly provides a
significant boost in the power efficiency of the amplifier.
For a second example, consider the same constant envelope power amplifier efficiency of
80% with -1 dB loss for the output path, G21 , and -0.5 dB for the G31 . The losses are
practical values that could be obtained with an optimized output network consisting of either
a circulator and bandpass filter or a complementary diplexer. The third example shown in the
figure corresponds to measured values that model the experimental test-bed described later
in the following section. A comparison of these three examples shows that it is significant to
minimize losses in the signal separation block to maximize the efficiency enhancement that is
obtained from energy recycling.

5.5 Experimental Implementation of a Switch-mode Power


Amplifier with Energy Recycling
An experimental test-bed was implemented to evaluate energy recycling in a switch-mode
power amplifier. A photo of the test-bed is shown in Figure 5.7. The system consists of
an arbitrary waveform generator to generate pulse encoded input signals, a class-F 10 W
amplifier, a circulator and bandpass filter for a signal reconstruction and signal separation,
and a wideband class-F−1 rectifier. The class-F 10 W amplifier design and the wideband
class-F−1 rectifier design were described earlier in Chapters 3 and 4, respectively. All the
pulse train waveforms were generated in Matlab using the noise shaped PPM pulse encoder
with out-of-band spectral shaping described earlier in section 5.3. The carrier frequency of
the source signal is centered at 930 MHz and the sinusoidal dither signal has a frequency of
986 MHz. The clock rate of the arbitrary generator is 24.25 Gs/sec. A two stage circulator
and a bandpass filter are used for signal separation and signal reconstruction. The bandpass
filter is a cavity filter with a center frequency of 930 MHz and bandwidth of 10 MHz. The
path gain from the input port (port 1) to the load port (port 2) is -1.12 dB (G21 ) and the
path gain from the input port (port 1) to the recycling power port (port 3) is -2.1 dB (G31 ).
The native power efficiencies of the amplifier and rectifier were first measured indepen-
dently of other components. For the pulse encoded waveforms the class-F amplifier power
efficiency (ηd0 ) is approximately 50%. The power efficiency does vary slightly depending on
the amplitude of the source signal and the amplitude of the dither signal. For the rectifier,
the peak RF to DC conversion efficiency (ηr ) is approximately 80% and the efficiency re-
duces as the input power is backed off. Since the amplitude of the dither signal in the pulse
train spectrum is inversely proportional to the amplitude of the source signal, we expect the
rectifier will have high power efficiency for low input amplitude source signals and reduced
efficiency for high amplitude input source signals. The dependence of rectification efficiency
on power is an important observation and an optimized energy recycling system would adjust
3
A saturated or switch-mode power amplifier with a drain efficiency of 80% is considered to be a very good
amplifier and in alignment with the best performance that is obtained from experimental amplifiers.

136
5.5. Experimental Implementation of a Switch-mode Power Amplifier with Energy Recycling

80

70

Overall Efficiency (%)


60

50

40

30

20

10

0
0 20 40 60 80 100
Coding Efficiency (%)
ηd0 = 80%; ηER = 100%; G21 = 0dB; G31 = 0dB
ηd0 = 80%; ηER = 0%; G21 = 0dB; G31 = 0dB
ηd0 = 80%; ηER = 80%; G21 = −1dB; G31 = −0.5dB
ηd0 = 80%; ηER = 0%; G21 = −1dB; G31 = −0.5dB
ηd0 = 50%; ηER = 80%; G21 = −1.12dB; G31 = −2.1dB
η = 50%; η = 0%; G = −1.12dB; G = −2.1dB
d0 ER 21 31

Figure 5.6: Examples of amplifier power efficiency with energy recycling.

peak efficiency to coincide with the average amplitude level of the source signal. However, in
this design, peak rectification efficiency is obtained at the lowest source amplitude level.
The test results for the energy recycling switch-mode power amplifier are shown in Fig-
ures 5.8 and 5.9. In Figure 5.8, the in-band power measured at the output of the filter is
shown as a function of the modulator drive level. The modulator drive level is expressed in
terms of coding efficiency which gives the ratio of the signal power relative to the total power
in the pulse train spectrum (see equation (5.5)). Therefore, low coding efficiency corresponds
to a low amplitude source signal and conversely high coding efficiency corresponds to a high
amplitude source signal. Since the encoded pulse train has constant amplitude, the total
power in the pulse train is constant and the therefore the sum of signal power (Psig ) and
added power Padd is equal to a constant. The figure shows the relative distribution of power
as the coding efficiency of the signal changes. From this graph, we expect energy recycling to
be most effective for low amplitude source signals.
In Figure 5.9, the power efficiency of the in-band signal component relative to the total
DC power supplied to the amplifier is shown as function of coding efficiency. Since the en-
coded signal is a constant power signal, the power efficiency decreases as the amplitude of

137
5.5. Experimental Implementation of a Switch-mode Power Amplifier with Energy Recycling

Powermeter
Voltmeter
Attenuator Phase shifter
Class-F PA Coupler
Driver
Rectifier

Circulator
BPF

RF arbitrary signal
generator

Figure 5.7: Test bed with a class F amplifier and a class-F−1 rectifier to recover out-of-band
energy. The system implements the block diagram shown in Figure 5.2(b).

7
In−band Power
Recovered Power
6

5
Power (W)

0
0 20 40 60 80 100
Coding Efficiency (%)

Figure 5.8: Measured in-band and recovered power as a function of the coding efficiency for
encoder of the noise shaped PPM modulator.

the source signal decreases. A second trace on the plot shows the in-band power efficiency
of the amplifier with the energy recovery system. At high input levels to the modulator,
the energy recovery system does not offer any benefit because the out-of-band dither power
is low. However, as the RF signal level drops, the energy recovery system can significantly

138
5.5. Experimental Implementation of a Switch-mode Power Amplifier with Energy Recycling

improve power efficiency relative to the power efficiency without energy recovery. As can be
seen in this figure, for a coding efficiency of 25%, the power efficiency without energy recovery
is 9.2%, while the corresponding power efficiency with energy recovery is 12.5%. From these
measurements the efficiency enhancement provided by the energy recovery system is 36%, a
large increase in efficiency. The results in Figure 5.9 include analytical results using the equa-
tions in Section 5.4. The analytical results have good agreement with experiment except for
high input signal levels. As mentioned earlier, the deviation is expected because rectification
efficiency is not constant and decreases as the power available at port 3 decreases. Signals
with high coding efficiency have low out-of-band power levels, and consequently the rectifica-
tion efficiency will be lower than pulse trains with low coding efficiency. The measurements
at high coding efficiency also show that power efficiency with and without energy recycling
converge which is also consistent with the reduction in rectification efficiency.

50
7
6
40 5
4
3
Efficiency (%)

30 8 10 12 14

20

10
In−band (Measured)
In−band+Recovered (Measured)
Overall efficiency: Analytical
0
0 20 40 60 80 100
Coding Efficiency (%)

Figure 5.9: Measured drain efficiencies with and without energy recycling.

139
5.6. Discussion and Chapter Summary

5.6 Discussion and Chapter Summary


The analytical and experimental results confirm that energy recycling can be used for
efficiency enhancement in switch-mode power amplifiers. However, the available hardware in
the test-bed has performance limitations and this is where the analytical results in Figure 5.6
are useful to project the overall power efficiency for better designs. For example, a significant
limitation in the experimental test-bed was the insertion loss of the signal separation and
reconstruction filters. The reconstruction filter has 2.1 dB loss and the out-of-band filter has
1.12 dB loss. An optimized filter design using cavity resonators could be much better and
losses may be as low as 1 dB for the reconstruction filter and 0.5 dB for the out-of-band filter.
Also, the class-F amplifier used in the test-bed had a native power efficiency of 50% when
amplifying pulse encoded signals and this could probably be improved.
In Figure 5.6, an example of an improved design with an amplifier efficiency of 80% and
reduced losses in the filter paths is shown. Relative to the experiments reported in this work,
the power efficiency for a coding efficiency of 25% would increase from 9.2% to 12.5% with
the improved design. Figure 5.6 also includes an upper bound on the power efficiency of an
energy recycling amplifier where filter losses are 0 dB and RF to DC rectification efficiency is
100%. This bound is useful to assess the limits of an energy recycling switch-mode amplifier.
It should also be noted that the analysis given in Section 5.4 is general and can also be
applied to outphasing amplifiers using LINC. Therefore, the analytical method can be used to
evaluate the benefits of energy recycling in any amplifier method that employs signal mapping
to reduce amplitude variation as a way of exploiting a high efficiency mode of operation in an
amplifier.

140
Chapter 6

Conclusions and Future Work


The research work in this thesis has been motivated by interest to improve the power
efficiency of RF switch-mode power amplifiers for wireless communication applications. If
power efficiency is to be improved, then it is essential to understand factors which affect
power efficiency and evaluate new methods to improve the design of RF switch-mode power
amplifiers. Conclusions from this research and recommendations for future work follow.

6.1 Conclusions
In Chapter 2, a detailed power efficiency analysis of CMCD amplifiers was presented. The
analysis includes a number of new contributions related to the predicting power efficiency
under variable duty cycle switching conditions. A level 2 device model was introduced to model
current saturation in the switch. Current saturations impose constraints on the maximum
switch current and considers limitations that are not observed when the device is modeled as
a simple switch (level 1 model). The implications of current saturation are important when
the duty cycle deviates significantly from 50% and an appropriate load line must be selected
to avoid deep saturation over the duty cycle operating range. Capacitive switching losses
were also shown to be very significant as the duty cycle deviates significantly from 50%. In
most literature references on CMCD analysis, capacitive switching losses are usually neglected
because the zero-voltage switching condition is assumed; an assumption that is valid only for
50% duty cycles. The analysis of power efficiency was extended to 2T signals and, similar
to the 1T analysis, there is good agreement between the analytical and simulation results.
The CMCD power efficiency analysis for periodic pulse trains was compared to simulation
results using both SDM and PPM pulse encoders. The comparison shows that analytical
results derived for variable duty cycle conditions provide good insight into predicting the
power efficiency of the CMCD amplifier for more general pulses such as SDM and PPM. In
terms of maximizing the power efficiency of CMCD amplifiers, the analysis clearly shows that
it is very important to constrain the range of duty cycle variation in the pulse train. Using
the circuit models for the Cree CGH60015D GaN HEMT, if duty cycles could be constrained
to a range from 35-65%, then power efficiency greater than 50% could be maintained.
In Chapter 3, class-F power amplifiers were studied. Class-F is another high efficiency
mode of operation that has received significant attention in the literature. Despite the large
body of work on class-F amplifiers, it is difficult to find comparative work which benchmarks
different circuit topologies. The research work was motivated by three questions: 1) How
significant is input harmonic matching in terms of power efficiency?; 2) Is class-F or inverse
class-F the better circuit topology?; and 3) What performance can be expected from using a
class-F amplifier in RF switch-mode power amplifier systems?
A systematic study of the sensitivity of power efficiency with respect to input harmonic
impedance termination was made for class-F amplifier. A device model was used to show how
harmonic injection through Cgd from the drain to gate is clearly evident and this also shows

141
6.1. Conclusions

how imperfect output terminations affect input harmonic levels. Second harmonic impedance
terminations were clearly shown to be very important, and further improvements in power
efficiency with a third harmonic match is incremental. An experimental class-F amplifier with
harmonic input matching including the third harmonic was designed and tested. A second
class-F amplifier based on the inverse current switched topology was designed and built to
compare with the voltage switched class-F design. A comparison of the two designs showed
that the inverse class-F amplifier has slightly higher power efficiency at the expense of a small
reduction in output power (0.5 dB). Under backed off power conditions, the inverse class-
F amplifier also shows better performance than the class-F amplifier; therefore, the current
switched inverse class-F design is the preferred choice in terms of power efficiency. A third
amplifier was designed, which was a wideband inverse class-F amplifier. The design was
built in anticipation of the RF rectifiers required for the energy recycling switch-mode power
amplifier which amplifies wideband pulse encoded signals. Experimental results were shown
and the wideband design has a minimum efficiency of 67% over a frequency range of 650 MHz
to 1150 MHz.
The class-F switch-mode amplifier designs described in Chapter 3 were reconfigured as
RF synchronous rectifiers in Chapter 4. The motivation for designing RF rectifiers is driven
by the need to implement circuits for the energy recycling amplifier described in Chapter 5.
In carrying out this work, new contributions were made to the design of RF synchronous
rectifiers. Experimental work on RF synchronous rectifiers based on time-reversal concepts is
very recent and began in 2012. After reviewing the design methodology, it was clear that the
principle of time reversal duality was useful for synthesizing rectifier duals from switch-mode
amplifiers, but the supporting analysis of the rectifier dual operating in the time-reversed mode
was less well developed. In this work, progress was made to understanding the operating mode
of the rectifier particularly in terms of how to interpret the dual when losses are significant.
Topologically, prior work has focused on inverting lossless amplifier circuits and the impact
of loss has not be studied. In addition to contributions in terms of the analysis of the RF
synchronous rectifier, new experimental benchmarks were established. An inverse class-F
rectifier design was recently presented in a conference paper that reports the highest power
efficiency for a high power RF rectifier [76]. The wideband inverse class-F rectifier is also new
experimental work. It should also be noted that RF synchronous rectifiers can be used in
other applications including wireless power and RF energy harvesting.
Chapter 5 brings together the work of Chapters 3 and 4 to implement a RF switch-mode
power amplifier with energy recycling. An analysis of energy recycling as a means of enhancing
the power efficiency of a switch-mode power amplifier is presented. The analysis is general
and to other amplifier architectures such as outphasing amplifiers using LINC. In switch-mode
amplifiers, out-of-band power is extracted in the signal reconstruction filter block and rectified
to provide DC power which can supplement the main DC supply for the amplifier. The work is
the first report experimental results for energy recycling in RF switch-mode power amplifiers.
The work included the implementation of a PPM pulse encoder with out-of-band dither to
improve rectification efficiency of out-of-band energy. Although the overall power efficiency
of the experimental amplifier was low, it does demonstrate how energy recycling can improve
power efficiency. In the experimental test bed power loss in the signal reconstruction filters
was a main source of loss and better designs could significantly improve results.

142
6.2. Future Work

6.2 Future Work


The methodology followed in this work has led to contributions that can benefit the design
of class-D amplifiers, class-F amplifiers, RF synchronous rectifiers and energy recycling. After
exploring different aspects of RF switch-mode power amplifiers, an observation is that the
overall power efficiency is still quite low especially when amplifying modulated source signals.
The goal of realizing a competitive high efficiency switch-mode power amplifier still requires
more innovation and research. So what should be pursued next?
Many of the results in this work illustrate the inter-relationships between signal mapping
and the power efficiency performance of the amplifier circuit. For example, if a pulse encoder
could be designed that constrains pulse duty cycles to a narrow range, then it may be possible
to get much better efficiency from CMCD amplifiers. Also, integrating the reconstruction
filter and the amplifier are essential to reduce output power losses as well as control output
impedances at the device plane.
For class-F amplifiers, the design methodology focuses on terminating harmonic impedances
and therefore a better pulse train spectrum for this type of amplifier would probably be a
harmonically shaped spectrum rather than the continuous spectrum generated by SDM and
noise shaped PPM. Therefore, an improved pulse encoder could possibly lead to higher ef-
ficiency with class-F amplifiers. The other objective in creating a harmonic pulse spectrum
would be to reduce the dissipation of out-of-band power. By shifting out-of-band power to
harmonic frequencies, less power should be dissipated in the out-of-band spectrum.
Wideband switch-mode amplifiers are naturally more suited to pulse trains like SDM and
PPM; however, they will create out-of-band power that will be dissipated. If wideband ampli-
fiers are pursued, then energy recycling would appear to be needed to offset the amplification
of out-of-band power.
The concept of RF switch-mode amplifiers rests on operating a device in a highly saturated
or switching state. When the device is operated with two states, there is a high cost in terms
of coding efficiency because signal power must be added to the source signal by the signal
mapping block. Is the ‘digital’ amplifier too costly? Perhaps a mixed mode amplifier with a
small range of signal amplitude variation would be better.
LINC and outphasing are close cousins to switch-mode amplifiers. Both amplifier archi-
tectures are based on operating the device with constant envelope signals. Recent progress
in terms of Chireix outphasing where reactive combining is used instead of isolated combin-
ing have been encouraging. For example, a research group in the Netherlands has reported
a drain power efficiency of 53.5% for a Chireix outphasing amplifier amplifying a WCDMA
signal [130]. There are still challenges with Chireix amplifiers such as distortion; however, it
has the advantage that no additional output filter is required for reconstructing the source
signal and instead the vector addition of the two branches creates the required cancellation
of added source power.
Finally, the work on RF synchronous rectifiers has many new possibilities in terms of
wireless power applications. As shown by results in this thesis, the dependence of rectification
efficiency on the DC load resistance can be significant and therefore the design of the DC load
circuit needs to be coupled with the design of the rectifier. In this way, the power efficiency
of the rectifier can be maximized over a larger dynamic range. Maintaining high rectification
efficiency over a large dynamic power range is important for both wireless power and energy
recycling applications.

143
Bibliography
[1] W. H. Doherty, “A new high efficiency power amplifier for modulated waves,” Proc.
IRE, vol. 24, no. 9, pp. 1163–1182, Sep. 1936. → pages 3, 5, 29

[2] H. Chireix, “High power outphasing modulation,” Proc. IRE, vol. 23, no. 11, pp. 1370–
1392, Nov. 1935. → pages 3, 5, 129

[3] S. C. Cripps, RF Power Amplifiers for Wireless Communications. Artech House, Inc.,
1999. → pages 3, 15

[4] D. Kimball, J. Jeong, C. Hsia, P. Draxler, S. Lanfranco, W. Nagy, K. Linthicum, L. Lar-


son, and P. Asbeck, “High-efficiency envelope-tracking W-CDMA base-station amplifier
using GaN HFETs,” IEEE Trans. Microwave Theory Tech., vol. 54, no. 11, pp. 3848–
3856, Nov. 2006. → pages 5

[5] T. Johnson and S. Stapleton, “RF class-D amplification with bandpass sigma-delta
modulator drive signals,” IEEE Trans. Circuits Syst. I, vol. 53, no. 12, pp. 2507–2520,
Dec. 2006. → pages 5

[6] ——, “Comparison of bandpass sigma-delta modulator coding efficiency with a periodic
signal model,” IEEE Trans. Circuits Syst. I, vol. 55, no. 11, pp. 3763–3775, Dec. 2008.
→ pages 6, 7, 8, 23, 24

[7] A. Wentzel, C. Meliani, and W. Heinrich, “Optimized coding scheme for class-S ampli-
fiers,” in Microwave Conference (EuMC), 2011 41st European, Oct. 2011, pp. 329 –332.
→ pages 6, 7

[8] A. Wentze1, C. Meliani, G. Fischer, and W. Heinrich, “An 8W GaN-based H-bridge


class-D PA for the 900 MHz band enabling ternary coding,” in International Microwave
Symposium (IMS), May 2012. → pages 6, 21, 23, 24

[9] T. Johnson, K. Mekechuk, and D. Kelly, “Noise shaped pulse position modulation for RF
switch-mode power amplifiers,” in European Microwave Conference (EuMC), Manch-
ester, UK, Oct. 9–14 2011, pp. 320 – 323. → pages 8

[10] T. Johnson, K. Mekechuk, D. Kelly, and J. Lu, “Asynchronous modulator for lineariza-
tion and switch-mode RF power amplifier applications,” in IEEE RFIC Symposium,
Boston, MA, Jun. 7–9, 2009, pp. 185–188. → pages 8

[11] H. Kobayashi, J. Hinrichs, and P. M. Asbeck, “Current mode class-D power amplifiers
for high efficiency RF applications,” IEEE Trans. Microwave Theory Tech., vol. 49,
no. 12, pp. 2480–2485, Jan. 2001. → pages 13

144
Bibliography

[12] D. Chowdhury, S. V. Thyagarajan, E. A. L. Ye, and A. M. Niknejad, “A fully-integrated


efficient CMOS inverse class-D power amplifier for digital polar transmitters,” in IEEE
Radio Frequency Integrated Circuits Symp., June 2011, pp. 1–4. → pages 13, 23, 34

[13] N. O. Sokal and A. D. Sokal, “Class E - A new class of high-efficiency tuned single-
ended switching power amplifiers,” IEEE J. Solid-State Circuits, vol. SC-10, no. 3, pp.
168–176, Jun. 1975. → pages 15

[14] F. Raab, “Class-E, class-C, and class-F power ampliers based upon a limite number of
harmonics,” IEEE Trans. Microwave Theory Tech., vol. 49, no. 8, pp. 1462–1468, Jun.
2001. → pages 20

[15] A. Wentzel, C. Meliani, and W. Heinrich, “A voltage mode class-S power amplifier for
the 450 MHz band,” Int. Journal of Micro. And Wireless Tech., vol. 3, pp. 311–318,
February 2011. → pages 21, 23, 24

[16] ——, “RF class S power amplifiers: State-of-the-art results and potential,” IEEE
MTT-S Symp. Dig., pp. 812–815, May 2010. → pages 21, 23, 24

[17] A. Wentzel, C. Meliani, E. Flucke, J.and Ersoy, and W. Heinrich, “Design and realiza-
tion of an output network for a GaN-HEMT current-mode class-S power amplifier at
450MHz,” in German Microwave Conference, March 2009, pp. 1–4. → pages 21, 23, 24

[18] A. Samulak, G. Fischer, and R. Weigel, “Design and simulation GaN based class-S PA
at 900 MHz,” in 18th Int. Conf. on Micro., Radar and Wireless Comm. (MIKON), Jun.
14–16 2010, pp. 1–4. → pages 21, 24

[19] J. Moon, S. Jee, J. Kim, J. Kim, and B.Kim, “Behaviors of class-F and class-FI ampli-
fiers,” IEEE Trans. Microwave Theory Tech., vol. 60, no. 6, pp. 1937–1951, Jun. 2012.
→ pages 21, 25, 26, 70

[20] V. Zomorrodian, U. Mishra, and R. York, “A high-efficiency class F MMIC power am-
plifier at 4.0 ghz using AlGaN/GaN HEMT technology,” in Compound Semiconductor
Integrated Circuit Symposium (CSIC), 2012. → pages 21, 26

[21] V. Carrubba, J. Lees, J. Benedikt, P. Tasker, and S. Cripps, “A novel highly efficient
broadband continuous class-F RFPA delivering 74% average efficiency for an octave
bandwidth,” in IEEE MTT/S International Microwave Symposium, May 2011, pp. 1–4.
→ pages 21, 26, 87

[22] R. J. Trew, “SiC and GaN transistors - Is there one winner for microwave power ap-
plications?” in Proceedings of the IEEE, vol. 90, Jun. 2002, p. 1032 1047. → pages
21

[23] “CGH40010 GaN HEMT,” Cree, Data Sheet, 2015. [Online]. Available: http:
//www.cree.com → pages 21

[24] E. Johnson, “Physical limitations on frequency and power parameters of transistors,”


RCA Review, vol. 26, pp. 163 – 177, 1965. → pages 21

[25] P. J. Baxandall, “Transistor sinewave oscillators,” Proc. Inst. Elec. Eng., vol. 106, part
B, suppl. 16, pp. 748–758, May 1959. → pages 22

145
Bibliography

[26] H. L. Krauss, C. W. Bostian, and F. H. Raab, Solid-State Radio Engineering. John


Wiley & Sons, Inc., 1980. → pages 22

[27] F. H. Raab, “Class-D power amplifier load impedance for maximum efficiency,” in Proc.
RF Technology Expo, Anaheim, CA, Jan. 1985, pp. 287–295. → pages 22

[28] S. El-Hamamsy, “Design of high-efficiency RF class-D power amplifier,” IEEE Trans.


Power Electron., vol. 9, no. 3, pp. 297–308, May 1994. → pages 22

[29] S. Lin and A. E. Fathy, “Development of a wideband highly efficient GaN VMCD
VHF/UHF power amplifier,” Progress In Electromagnetics Research C, vol. 19, pp.
135–147, December 2011. → pages 22, 24

[30] F. H. Raab, “Radio frequency pulsewidth modulation,” IEEE Trans. Commun., vol. 21,
pp. 958–966, Aug. 1973. → pages 22

[31] A. Jayaraman, P. F. Chen, G. Hanington, L. Larson, and P. Asbeck, “Linear high-


efficiency microwave power amplifiers using bandpass delta-sigma modulators,” IEEE
Microw. Guided Wave Lett., vol. 8, no. 3, pp. 121–123, Mar. 1998. → pages 23

[32] Tsai-Pi Hung, J. Rode, L. E. Larson, and P. M. Asbeck, “Design of H-bridge class-D
power amplifiers for digital pulse modulation transmitters,” IEEE Trans. Microwave
Theory Tech., vol. 55, no. 12, Part 2, pp. 2845–2855, Dec. 2007. → pages 23, 24

[33] S. D. Kee, I. Aoki, and D. Rutledge, “7-MHz, 1.1-kW Demonstration of the


new E/F2,odd switching amplifier class,” in IEEE MTT/S International Microwave
Symposium, 2001. → pages 23

[34] Tsai-Pi Hung, A. G. Metzger, P. J. Zampardi, M. Iwamoto, and P. M. Asbeck, “Design


of high-efficiency current-mode class-D amplifiers for wireless handsets,” IEEE Trans.
Microwave Theory Tech., vol. 53, no. 1, pp. 144–151, Jan. 2005. → pages 23, 34

[35] A. Brackle, S. Heck, M. Berroth, and I. Dettmann, “Efficiency of current-mode Class-D


amplifiers with delta-sigma modulated drive signals,” in German Microwave Conf., Mar.
16–18, 2009, pp. 1–4. → pages 23, 24, 34, 46

[36] S. Abbasian and T. Johnson, “RF current mode class-D power amplifiers under periodic
and non-periodic switching conditions.” in International Symposium on Circuits and
Systems (ISCAS), May 2013, pp. 610 –613. → pages 24, 34

[37] F. Malekzadeh, R. Mahmoudi, M. Sarkeshi, and A. Roermund, “Fine tuning of switching


frequency for minimal distortion in high frequency PWM systems,” in IEEE MTT/S
International Microwave Symposium, June 2011, pp. 1–4. → pages 24

[38] J.-C. Park, C.-S. Yoo, W. Kang, D. Kim, J.-G. Yook, and W. S. Lee, “GaN HEMT
based high-efficiency current-mode class-D amplifier using chip-on-board technique,”
Microwave and Optical Technology Letters, vol. 54, no. 2, pp. 358–362, Feb. 2012. →
pages 24

[39] D. Chowdhury, S. V. Thyagarajan, E. A. L. Ye, and A. M. Niknejad, “A fully-integrated


efficient CMOS inverse class-D power amplifier for digital polar transmitters,” IEEE J.
Solid-State Circuits, vol. 47, no. 5, pp. 1113 – 1122, May 2012. → pages 24

146
Bibliography

[40] M. Kamper, G. Ulbricht, R. Weigel, and G. Fischer, “Comparison of class a and class d rf
amplifier operation with focus on reverse intermodulation and efficiency performance,”
in Microwave Conference (EuMC), 2011 41st European, Oct. 2011, pp. 272 – 275. →
pages 24

[41] P. Aflaki and M. Helaoui, “Effect of the class of switching-mode power amplifiers on the
efficiency of band-pass delta-sigma architectures,” in Microwave Conference (EuMC),
2011 41st European, Oct. 2011, pp. 312–315. → pages 24, 36

[42] C. Schuberth, P. Singerl, M. E. Gadringer, H. Arthaber, A. Wiesbauer, and G. Magerl,


“Highly efficient switched-mode transmitter using a current mode class-D RF amplifier,”
International Journal of RF and Microwave Computer-Aided Engineering, vol. 20, no. 4,
pp. 446–457, Jul. 2012. → pages 24

[43] P. Aflaki, R. Negra, and F. Ghannouchi, “Highly efficient switched-mode transmitter us-
ing a current mode class-D RF amplifier,” IET Microwaves, Antennas and Propagation,
vol. 3, no. 6, pp. 997 – 1006, Jul. 2009. → pages 24

[44] H. M. Nemati, C. Fager, and H. Zirath, “High efficiency LDMOS current mode class-D
power amplifier at 1 GHz,” in Microwave Conference (EuMC), 2011 36st European,
Sep. 2006, pp. 176–179. → pages 24

[45] J.-Y. Kim, D.-H. Han, J.-H. Kim, and S. Stapleton, “A 50 W LDMOS current mode 1800
MHz class-D power amplifier,” in IEEE MTT/S International Microwave Symposium,
June 2005. → pages 24

[46] A. Samulak, G. Fischer, and R. Weigel, “Basic nonlinear analysis of class-S power
amplifiers based on GaN switching transistors,” in German Microwave Conf., Mar. 16–
18 2009, pp. 1–4. → pages 24

[47] R. Leberer, R. Reber, and M. Oppermann, “An AIGaN/GaN Class-S amplifier for RF-
communication signals,” in IEEE MTT-S International Microwave Symposium, Jun.
15–20 2008, pp. 85–88. → pages 24

[48] V. Tyler, “A new high efficiency high power amplifier,” Marconi Rev., vol. 21, pp.
96–109, 1958. → pages 25

[49] D. Snider, “A theoretical analysis and experimental confirmation of the optimally loaded
and overdriven RF power amplifier,” IEEE Trans. Electron Devices, vol. ED-14, pp.
851–857, 1967. → pages 25

[50] F. H. Raab, “Introduction to class-F power amplifiers,” R.F. Design, vol. 19, pp. 79–84,
May 1996. → pages 25

[51] F. Raab, “Class-F power amplifiers with maximally flat waveforms,” IEEE Trans.
Microwave Theory Tech., vol. 45, no. 11, pp. 2007 – 2012, Jun. 1997. → pages 25

[52] ——, “Maximum efficiency and output of class-F power amplifiers,” IEEE Trans.
Microwave Theory Tech., vol. 49, no. 6, pp. 1162 – 1166, May 2001. → pages 25

147
Bibliography

[53] V. Carrubba, A. Clarke, M. Akmal, J. Lees, J. Benedikt, P. Tasker, and S. Cripps, “On
the extension of the continuous class-F mode power amplifier,” IEEE Trans. Microwave
Theory Tech., vol. 59, no. 5, pp. 1294 – 1303, May 2011. → pages 25, 26

[54] H. M. Zhe, A. A’ain, and A. Kordesch, “An integrated 2.4GHz CMOS class F power
amplifier,” in IEEE International Conference on Semiconductor Electronics, Oct. 2006,
pp. 537 – 540. → pages 25, 26

[55] A. Ouyahia, C. Duperrier, C. Tolant, F. Temcamani, and P. Eudeline, “A 71.9% power-


added-efficiency inverse class-F LDMOS,” in IEEE MTT/S International Microwave
Symposium, June 2006, pp. 1542 – 1545. → pages 25, 26

[56] S. Abbasian and T. Johnson, “Effect of second and third harmonic input impedances
in a class-f amplifier,” Progress In Electromagnetics Research C, vol. 56, pp. 39 – 53,
Feb. 2015. → pages 26, 61

[57] T. Hwang, K. Azadet, R. Wilson, and J. Lin, “Characterization of class-f power am-
plifier with wide amplitude and phase bandwidth for outphasing architecture,” IEEE
Microwave and Wireless Components Letters, vol. 24, no. 3, pp. 188 – 190, Jun. 2014.
→ pages 26

[58] J. Kim and Y. Park, “Design of a compact and broadband inverse class-f−1 power
amplifier,” Progress In Electromagnetics Research C, vol. 46, pp. 75 – 81, Jan. 2014. →
pages 26

[59] K. Chen and D. Peroulis, “A 3.1-GHz class-F power amplifier with 82% power-added-
efficiency,” IEEE Microwave and wireless components letters, vol. 23, no. 8, pp. 436 –
438, Aug. 2013. → pages 26

[60] A. Garcia-Osorio, J. Loo-Yau, and J. Reynoso-Hernandez, “A gan class-F PA with


600 MHz bandwidth and 62.5% of PAE suitable for WiMAX frequencies,” in IEEE
International Microwave Worksho, Feb. 2010, pp. 1 – 4. → pages 26

[61] P. Aflaki, R. Negra, and F. Ghannouchi, “Design and implementation of an inverse class-
F power amplifier with 79% efficiency by using a switch-based active device model,” in
Radio and Wireless Symposium, Jan. 2008, pp. 423 – 426. → pages 26

[62] L. Dong, S. He, F. You, and Q. Lei, “High-efficiency class-f1 power amplifier design
with input harmonic manipulation,” in IEEE Topical Conference on Power Amplifiers
for Wireless and Radio Applications (PAWR), Jan. 2008, pp. 1–4. → pages 26

[63] J. Kim, G. Jo, J. Hoon Oh, Y. Kim, K. Lee, and J. Jung, “Modeling and design
methodology of high-efficiency class-F and inverse class-F power amplifiers,” IEEE
Trans. Microwave Theory Tech., vol. 59, no. 1, pp. 153 – 165, Jan. 2011. → pages
26

[64] D.-T. Wu and S. Boumaiza, “10W GaN inverse class F PA with input/output har-
monic termination for high efficiency WiMAX transmitter,” in Wireless and Microwave
Technology Conference (WAMICON), Apr. 2009, pp. 1 – 4. → pages 26

148
Bibliography

[65] A. Al Tanany, A. Sayed, and G. Boeck, “Design of inverse classF power amplifier using
GaN pHEMT for industrial applications,” in German Microwave Conference, Mar. 2009,
pp. 1 – 4. → pages 26

[66] Y. Woo, Y. Yang, and B. Kim, “Analysis and experiments for high-efficiency class-F
and inverse class-F power amplifiers,” IEEE Trans. Microwave Theory Tech., vol. 54,
no. 5, pp. 1969 – 1974, May 2006. → pages 26

[67] F. You, S. He, X. Tang, and X. Deng, “High-efficiency single-ended class-E/F2 power
amplifier with finite DC feed inductor,” IEEE Trans. Microwave Theory Tech., vol. 58,
no. 1, pp. 32 – 40, Jan. 2010. → pages 26

[68] W. Brown, “The history of power transmission by radio waves,” IEEE Trans. Microwave
Theory Tech., vol. 32, pp. 1230–1242, Sep. 1984. → pages 27

[69] N. Shinohara, “Rectennas for microwave power transmission,” IEICE Electronics


Express, vol. 10, no. 21, pp. 1 – 13, Nov. 2013. → pages 27

[70] A. Noda and H. Shinoda, “Compact class-F RF-DC converter with antisymmetric dual-
diode configuration,” in IEEE MTT-S Int. Symp., Jun. 2012, pp. 1–3. → pages 27

[71] J. Guo and X. Zhu, “Class F rectifier RF-DC conversion efficiency analysis,” in IEEE
MTT-S Int. Symp., Jun. 2013, pp. 1 – 4. → pages 27

[72] E. Falkenstein, M. Roberg, and Z. Popovic, “Low-power wireless power delivery,” IEEE
Trans. Microwave Theory Tech., vol. 60, no. 7, pp. 2277– 2286, Jul. 2012. → pages 27

[73] M. N. Ruiz, “A class E synchronous rectifier based on an E-pHEMT device for wireless
powering applications,” in IEEE MTT/S International Microwave Symposium, June
2012, pp. 1–3. → pages 27, 28, 31, 101, 104, 105

[74] D. Hamill, “Time reversal duality and the synthesis of a double class-E DC-DC con-
verter,” in 21st Power Electronics Specialist Conference (PESC), 1990, pp. 512–521. →
pages 27, 99, 100

[75] S. Abbasian and T. Johnson, “High efficiency GaN HEMT class-F synchronous rectifier
for wireless applications,” IEICE Electronics Express, vol. 12, no. 1, pp. 1 – 11, Jan.
2015. → pages 28, 99, 154

[76] ——, “High efficiency and high power GaN HEMT inverse class-F synchronous rectifier
for wireless power applications,” in IEEE European Microwave Conference (EuMc),
Paris, France, Sep. 2015, pp. 1–4. → pages 28, 61, 99, 142

[77] S. Dehghani, S. Abbasian, and T. Johnson, “Tracking load to optimize power efficiency
in RF to DC rectifier circuits,” in IEEE Wireless Power Transfer Conference(WPTC),
May 2015, pp. 1 – 3. → pages 28

[78] M. Litchfield, S. Schafer, T. Reveyrand, and Z. Popovic, “High-efficiency X-band MMIC


GaN power amplifiers operating as rectifiers,” in IEEE MTT-S Int. Symp., Jun. 2014,
pp. 1 – 4. → pages 28, 104, 105

149
Bibliography

[79] M. N. Ruiz and J. A. Garcia, “An E-pHEMT self-biased and self-synchronous class E
rectifier,” in IEEE MTT-S Int. Symp., Jun. 2014, pp. 1 – 4. → pages 28, 105

[80] R. Ishikawa and K. Honjo, “Microwave power transfer evaluation at 2.45 GHz using
a high-efficiency GaAs HEMT amplifier and rectifier,” in 43rd European Microwave
Conference (EuMc), Oct. 2013, pp. 916 – 919. → pages 28, 105, 106

[81] ——, “Reversible high efficiency amplifier/rectifier circuit for wireless power transmis-
sion system,” in Asia-Pacific Microwave Conference (APMC), 2013, pp. 74 – 76. →
pages 28, 105, 106

[82] M. Roberg, T. Reveyrand, I. Ramos, E. Falkenstein, and Z. Popvoic, “High-efficiency


harmonically terminated diode and transistor rectifiers,” IEEE Trans. Microwave
Theory Tech., vol. 60, no. 12, pp. 4043–4052, Dec. 2012. → pages 28, 105, 106

[83] C. Gómez, J. A. Garcia, A. Mediavilla, and A. Tazón, “A high efficiency rectenna


element using E-pHEMT technology,” in 12th GaAs Symp., 2004, pp. 315 – 318. →
pages 28, 105

[84] S. N. Ali and T. E. Johnson, “RF switch-mode power amplifier with an integrated
diplexer for signal reconstruction and energy recovery,” in IEEE MTT-S Symp. Dig.,
Montréal, Canada, Jun. 17–22, 2012, pp. 1–3. → pages 30, 130

[85] A. Long, J. Yao, and S. I. Long, “A 13W current mode class D high efficiency 1 GHz
power amplifier,” in Midwest Symposium on Circuits and Systems, vol. 1, Aug. 4–7
2002, pp. 33–36. → pages 34

[86] H. Kobayashi, J. M. Hinrichs, and P. Asbeck, “Current-mode class-D power amplifiers


for high-efficiency RF applications,” IEEE Trans. Microwave Theory Tech., vol. 49,
no. 12, pp. 2480–2485, Dec. 2001. → pages 34

[87] F. Ghannouchi, S. Hatami, P. Aflaki, M. Helaoui, and R. Negra, “Accurate power


efficiency estimation of ghz wireless delta-sigma transmitters for different classes of
switching mode power amplifiers,” IEEE Trans. Microwave Theory Tech., vol. 58, no. 11,
pp. 2812 –2819, Nov. 2010. → pages 36

[88] G. Dambrine, A. Cappy, F. Heliodore, and E. Playez, “A new method for determining
the FET small-signal equivalent circuit,” IEEE Trans. Microwave Theory, vol. 36, no. 7,
pp. 1151–1159, 1998. → pages 62

[89] D. Lovelace, J. Costa, and N. Camilleri, “Extracting small-signal model parameters of


silicon MOSFET transistors,” in IEEE MTT-S Dig., 1994, pp. 865–868. → pages 62

[90] “CGH60015D GaN HEMT die,” Cree, Data Sheet, Rev. 3.1, 2013. [Online]. Available:
http://www.cree.com → pages 66

[91] A. Ramadan, T. Reveyrand, A. Martin, J. Nebus, P. Bouysse, and L. Lapierre, “Exper-


imental study on effect of second-harmonic injection at input of classes F and inverse F
GaN power amplifiers,” IET Electronics letters, vol. 46, no. 8, pp. 570–572, TBD 2010.
→ pages 70, 71

150
Bibliography

[92] P. White, “Effect of input harmonic terminations on high efciency class B and class F
operation of phemt devices,” in IEEE MTT-S Int. Symp., vol. 3, 1998, pp. 1611 – 1614.
→ pages 70, 71

[93] S. Goto, T. Kunii, A. Ohta, A. Inoue, Y. Hosokawa, R. Hattori, and Y. Mitsui, “Ef-
fect of bias condition and input harmonic termination on high eciency inverse class-F
amplifiers,” in 31st IEEE European Microwave Conference (EuMc), 2001, pp. 1–4. →
pages 70, 71

[94] P. Colantonio, F. Giannini, G. Leuzzi, and E. Limiti, “Theoretical facet and experimen-
tal results of harmonic tuned PAs,” Int. J. RF Microw. Comput.-Aided Eng., vol. 13,
no. 6, pp. 459–472, TBD 2003. → pages 70

[95] P. Colantonio, F. Giannini, and E. Limiti, “An approach to harmonic load- and source-
pull measurements for high-efficiency PA design,” IEEE Trans. Microwave Theory,
vol. 52, no. 1, pp. 191–198, 2004. → pages 70, 71

[96] P. Colantonio, F. Giannini, G. Leuzzi, and E. Limiti, “Multiharmonic manipulation for


highly efficient microwave power amplifiers,” Int. J. RF Microw. Comput.-Aided Eng.,
vol. 11, pp. 366–384, 2001. → pages 70, 71

[97] S. Gao, P. Butterworth, S. Ooi, and A. Sambell, “High-efficiency power amplifier design
including input harmonic termination,” IEEE Microwave And Wireless Components
Letters, vol. 16, no. 2, pp. 81–83, 2005. → pages 70, 71

[98] P. Colantonio, F. Giannini, G. Leuzzi, and E. Limiti, “Theoretical facet and experimen-
tal results of harmonic tuned PAs,” Int. J. RF Microw. Comput.-Aided Eng., vol. 13,
no. 6, pp. 459–472, TBD 2003. → pages 71

[99] V. Vadala, A. Raffo, S. D. Falco, G. Bosi, A. Nalli, and G. Vannini, “A load pull
characterization technique accounting for harmonic tuning,” IEEE Trans. Microwave
Theory, vol. 61, no. 7, pp. 2695–2704, 2013. → pages 71

[100] P. Colantonio, F. Giannini, R. Giofré, and L. Piazzon, “A design technique for con-
current dual band harmonic tuned power amplifier,” IEEE Trans. Microwave Theory,
vol. 56, no. 11, pp. 2545–2555, 2008. → pages 77

[101] J. Rollett, “Stability and power-gain invariants of linear two ports,” IEEE Trans. Circuit
Theory, vol. 9, no. 1, pp. 29–32, 1962. → pages 78

[102] H. Nemati, C. Fager, M. Thorsell, and H. Zirath, “High-efficiency LDMOS power-


amplifier design at 1 GHz using an optimized transistor model,” IEEE Trans. Microwave
Theory, vol. 57, no. 7, pp. 1647–1654, 2009. → pages 79

[103] P. Saad, C. Fager, H. Nemati, H. Cao, H. Zirath, and K. Andersson, “A highly efficient
3.5 GHz inverse class-F GaN HEMT power amplifier,” Int. J. of Microwave and Wireless,
vol. 2, no. 3-4, pp. 317–324, 2010. → pages 79

[104] P. Saad, C. Fager, H. Cao, H. Zirath, and K. Andersson, “Design of a highly efficient 2-4
GHz octave bandwidth GaN-HEMT power amplifier,” IEEE Trans. Microwave Theory,
vol. 58, no. 7, pp. 1677–1685, 2010. → pages 79, 87, 88, 89, 90, 95

151
Bibliography

[105] S. A. Bassam, M. Helaoui, and F. M. Ghannouchi, “2-D Digital predistortion (2-D-DPD)


architecture for concurrent dual-band transmitters,” IEEE Trans. Microwave Theory,
vol. 59, no. 10, pp. 2547–2553, 2011. → pages 81

[106] H. Yu, X.and Jiang, “Digital predistortion using adaptive basis functions,” IEEE Trans.
Circuits Syst. I, vol. 60, no. 12, pp. 3317–3327, TBD 2013. → pages 81

[107] L. Xiao, S. Abbasian, and T. Johnson, “All-digital encoders for RF switch-mode power
amplifier applications,” in IEEE Wireless and Microwave Tech. Conf. (WAMICON),
Jun. 2014, pp. 1 – 6. → pages 87

[108] M. Moazzam and C. Aitchison, “A high gain dual-fed single stage distributed amplifier,”
in IEEE MTT-S Int. Microw. Symp. Dig., 1994, pp. 1409–1412. → pages 87

[109] F. Lin, Q.-X. Chu, and Z. Lin, “A novel tri-band branch-line coupler with three con-
trollable operating frequencies,” Microw. Wireless Compon. Lett., vol. 20, no. 12, pp.
666–668, Dec. 2010. → pages 87

[110] B. S. Virdee, A. S. Virdee, and B. Y. Banyamin, Broadband Microwave Amplifiers.


Norwood, MA: Artech House, 2004. → pages 87

[111] E. Saphiro, J. Xu, A. Naga, F. Williams, U. Mishra, and R. York, “A high efficiency
traveling-wave power amplifier topology using improved power-combining technique,”
IEEE Microw. Guided Wave Lett., vol. 8, no. 3, pp. 133–135, Mar. 1998. → pages 87

[112] J. Gassmann, P. Watson, L. Kehias, and G. Henry, “Wideband, high efficiency GaN
power amplifiers utilizing a non-uniform distributed topology,” in IEEE MTT-S Int.
Microw. Symp. Dig., Jun. 2007, pp. 615–618. → pages 87

[113] V. Carrubba, A. Clarke, M. Akmal, J. Lees, J. Benedikt, P. Tasker, and S. Cripps, “On
the extension of the continuous class-F mode power amplifier,” IEEE Trans. Microwave
Theory, vol. 59, no. 5, pp. 1294–1303, May 2011. → pages 87

[114] J. Chen, S. He, F. You, R. Tong, and R. Peng, “Design of broadband high-efficiency
power amplifiers based on a series of continuous modes,” IEEE Microw. Wireless
Compon. Lett., vol. 24, no. 9, pp. 631–633, Sep. 2014. → pages 87

[115] P. Colantonio, F. Giannini, R. Giofre, and L. Piazzon, “High-efficiency ultra-wideband


power amplifier in GaN technology,” Electron. Lett., vol. 44, no. 2, pp. 130–131, Jan.
2008. → pages 87

[116] N. Tuffy, L. Guan, A. Zhu, and T. J. Brazil, “A simplified broadband design method-
ology for linearized high-efficiency continuous class-F power amplifiers,” IEEE Trans.
Microwave Theory, vol. 60, no. 6, pp. 1952–1963, Jun. 2012. → pages 87

[117] R. M. Fano, “Theoretical limitations on the broadband matching of arbitrary


impedances,” J. Franklin Inst., vol. 249, no. TBD, p. TBD, TBD 1950. → pages 88

[118] D. E. Dawson, “Closed-form solutions for the design of optimum matching networks,”
IEEE Trans. Microwave Theory, vol. 57, no. 1, pp. 121–129, Jan. 2009. → pages 88, 89

152
Bibliography

[119] A. Vatankhahghadim, “Design of high efficiency broadband adjusted class AB power


amplifier,” PhD Dissertation, University of Waterloo, vol. TBD, no. TBD, p. TBD,
TBD 2010. → pages 90

[120] R. W. Rhea, HF Filter Design and Computer Simulation. New York: Noble, 1994. →
pages 93

[121] D. M. Pozar, Microwave Engineering-4th ed. Reading, Massachusetts: John Wiley &
Sons, Inc., 2012. → pages 93

[122] D. C. Hamill, “Time reversal duality in dc-dc converters,” in Power Electronics


Specialists Conference (PESC), June 1997, pp. 789–795. → pages 100

[123] G. Callet, J. Faraj, O. Jardel, C. Charbonniaud, J. C. Jacquet, T. Reveyrand, E. Mor-


van, S. Piotrowicz, J. P. Teyssier, and R. Quéré, “A new nonlinear HEMT model for
AlGaN/GaN switch applications,” Int. J. Microw. Wirel. Tech., vol. 2, no. 3-4, pp.
283–291, Jul. 2010. → pages 104

[124] M. Coffey, S. Schafer, and Z. Popovic, “Two-stage high-efficiency X-band GaN MMIC
PA/ rectifier,” in IEEE MTT-S Int. Symp., May 2015, pp. 1 – 3. → pages 105

[125] R. Langridge, T. Thornton, P. Asbeck, and R. Larson, “A power reuse technique for
improved efficiency of outphasing microwave power amplifiers,” IEEE Trans. Microwave
Theory Tech., vol. 47, no. 8, pp. 1467–1470, Aug. 1999. → pages 129

[126] Z. Zhang, L. Larson, P. Asbeck, and R. Langridge, “Analysis of power recycling tech-
niques for RF and microwave outphasing power amplifiers,” IEEE Trans. Circuits Syst.
II, vol. 49, no. 5, pp. 312–320, May 2002. → pages 129

[127] P. A. Godoy, D. J. Perreault, and J. L. Dawson, “Outphasing energy recovery amplifier


with resistance compression for improved efficiency,” IEEE Trans. Microwave Theory
Tech., vol. 57, no. 12, pp. 2895–2906, Dec. 2009. → pages 129

[128] M. Litchfield and Z. Popovi, “X-band outphasing GaN MMIC PA with power recycling,”
in IEEE MTT-S Int. Symp., May 2015, pp. 1–4. → pages 129

[129] S. N. Ali and T. E. Johnson, “A new high efficiency RF switch-mode power amplifier
architecture for pulse encoded signals,” in IEEE Wireless and Microwave Tech. Conf.
(WAMICON), Cocoa Beach FL, USA, Apr. 15–17, 2012, pp. 1–6. → pages 130

[130] D. Calvillo-Cortes, M. van der Heijden, M. Acar, M. Langen, R. Wesson, F. van Rijs,
and L. de Vreede, “A package-integrated Chireix outphasing RF switch-mode high-
power amplifier,” IEEE Trans. Microwave Theory Tech., vol. 61, no. 10, pp. 3721–3732,
Oct. 2013. → pages 143

153
Appendix A

Measurement Results for Another


Class-F PA
In this appendix, the measurement results for another class-F power amplifier which was
published in [75] are presented.

80

75
Efficiency (%)

70

65

60
20 21 22 23 24 25
Input Power (dBm)

Figure A.1: Measured drain efficiency as a function of input power for a CW test signal.

154
Appendix A. Measurement Results for Another Class-F PA

90

80

Efficiency (%) 70

60

50

40

30

20
940 960 980 1000 1020 1040
Frequency (MHz)

Figure A.2: Measured drain efficiency as a function of frequency for a CW test signal.

−10
Power spectrum density (dB/RBW)

−20

−30

−40
a
b
c
−50

−60

−70
975 980 985 990 995 1000 1005
Frequency (MHz)

Figure A.3: Measured output spectrums for a WCDMA signal at three different output power
levels: (a) 34.2 dBm (b) 32.4 dBm and (c) 30.5 dBm

155
Appendix A. Measurement Results for Another Class-F PA

60 −34

(a)
50 −35

40 (b) −36
Efficiency (%)

ACLR (dBc)
(a)
30 (c) −37

20 −38

(b)
10 (c) −39

0 −40
30 31 32 33 34 35
Output Power (dBm)

Figure A.4: Measured drain efficiency and ACLR as a function of output power for a WCDMA
signal.

156

You might also like