You are on page 1of 163

The Pennsylvania State University

The Graduate School


College of Engineering

DEVELOPING A STABILITY ASSESSMENT METHOD FOR

POWER ELECTRONICS-BASED MICROGRIDS

A Dissertation in
Mechanical Engineering
by
Peter M. Austin

© 2015 Peter M. Austin

Submitted in Partial Fulfillment


of the Requirements
for the Degree of

Doctor of Philosophy

December 2015
The dissertation of Peter M. Austin was reviewed and approved∗ by the following:

Thomas G. Hughes
Senior Research Associate Emeritus of Applied Research Laboratory
Dissertation Co-Adviser and Co-Chair of Committee

Susan W. Stewart
Research Associate and Assistant Professor of Aerospace Engineering and
Architectural Engineering
Dissertation Co-Adviser and Co-Chair of Committee

Hosam K. Fathy
Assistant Professor of Mechanical Engineering

Jeffrey S. Mayer
Associate Professor of Electrical Engineering

Horacio Perez-Blanco
Professor of Mechanical Engineering

James A. Turso
Senior Research Associate of Nuclear Engineering
Special Member

Karen Thole
Mechanical and Nuclear Engineering Department Head


Signatures are on file in the Graduate School.

ii
Abstract

Modern microgrids with microsources and energy storage are dependent on power
electronics for control and regulation. Under certain circumstances power electron-
ics can be destabilizing to the system due to an effect called negative incremental
impedance. A careful review of the theory and literature on the subject is pre-
sented. This includes stability criteria for both AC and DC systems, as well as a
discussion on the limitations posed by the analysis. A method to integrate stability
assessment with higher-level microgrid architectural design is proposed. Crucial
to this is impedance characterization of individual components, which was accom-
plished through simulation. DC and AC impedance measurement blocks were
created in Matlab simulink to automate the process. A detailed switching-level
model of a DC microgrid was implemented in simulink, including wind turbine
microsource, battery storage, and three phase inverter. Maximum power point
tracking (MPPT) was included to maximize the efficiency of the turbine and was
implemented through three rectifier alternatives and control schemes. The sta-
bility characteristics of each was compared in the final analysis. Impedance data
was collected individually from the components and used to assess stability in the
system as a whole. The results included the assessment of stability, margin, and
unstable operating points to demonstrate the feasibility of the proposed approach.

iii
Table of Contents

List of Figures vii

List of Tables xi

Acknowledgments xii

Chapter 1
Background 1
1.1 Recent Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Microgrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Power Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Negative Incremental Input Impedance . . . . . . . . . . . . . . . . 12
1.6 Design Challenge . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.7 Proposed Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.8 Dissertation Objective . . . . . . . . . . . . . . . . . . . . . . . . . 16

Chapter 2
Stability 18
2.1 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.1 Linear Time-Invariant (LTI) Assumption . . . . . . . . . . . 19
2.1.2 Internal and External Stability . . . . . . . . . . . . . . . . 22
2.2 DC system Stability . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2.1 Electrical System Transfer Function . . . . . . . . . . . . . . 24
2.2.2 Nyquist Stability Criteria . . . . . . . . . . . . . . . . . . . 26
2.2.3 Stability Criteria . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.4 Passivity-Based Stability Criteria . . . . . . . . . . . . . . . 30
2.2.5 Generalized Admittance Space Analysis . . . . . . . . . . . . 31
2.3 AC bus Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.3.1 Generalized Nyquist . . . . . . . . . . . . . . . . . . . . . . 32

iv
2.3.2 Stability Criteria . . . . . . . . . . . . . . . . . . . . . . . . 33
2.4 Proposed Stability Assessment Method . . . . . . . . . . . . . . . . 34
2.4.1 Minimizing Conservatism . . . . . . . . . . . . . . . . . . . . 34
2.4.2 Stability Margins . . . . . . . . . . . . . . . . . . . . . . . . 37
2.4.3 Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.4.4 Operating Points . . . . . . . . . . . . . . . . . . . . . . . . 40

Chapter 3
Simulated Impedance Measurement 43
3.1 Nyquist Contour and Impedance . . . . . . . . . . . . . . . . . . . 43
3.2 DC Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2.1 Current Perturbation Injection . . . . . . . . . . . . . . . . 48
3.2.2 DC Impedance Measurement Block . . . . . . . . . . . . . . 49
3.2.3 Results and Verification . . . . . . . . . . . . . . . . . . . . 52
3.3 AC System Impedances . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.3.1 The dq0 Synchronous Reference Frame . . . . . . . . . . . . 56
3.3.2 DQ Impedance Matrix . . . . . . . . . . . . . . . . . . . . . 56
3.3.3 AC Impedance Measurement Block . . . . . . . . . . . . . . 57
3.3.4 Results and Verification . . . . . . . . . . . . . . . . . . . . 59

Chapter 4
Notional Microgrid 64
4.1 Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2 Notional Microgrid Layout . . . . . . . . . . . . . . . . . . . . . . . 68
4.3 Wind Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.4 Rectification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.4.1 Passive Diode Bridge . . . . . . . . . . . . . . . . . . . . . . 73
4.4.2 Boost Converter . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.4.3 Switching Bridge . . . . . . . . . . . . . . . . . . . . . . . . 84
4.5 Battery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.6 Inverter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.6.1 "Stiff" grid P plus resonant Inverter . . . . . . . . . . . . . . 94
4.6.2 "Stiff" grid dq0 Inverter . . . . . . . . . . . . . . . . . . . . . 99
4.6.3 "weak" grid dq0 Inverter . . . . . . . . . . . . . . . . . . . . 102

Chapter 5
Results and Conclusions 106
5.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.1.1 P Resonant Inverter and Stability Margins . . . . . . . . . . 106
5.1.2 Notional Wind Microgrid and Operating Point Stability . . . 109

v
5.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

Chapter 6
Recommendations 118

Bibliography 121

Appendix A
Grid Control 129
A.1 Grid vs. Microgrid . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A.2 Electrical Grid Control . . . . . . . . . . . . . . . . . . . . . . . . 130
A.2.1 AGC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
A.2.2 CERTS microgrid control . . . . . . . . . . . . . . . . . . . 131
A.3 High Level Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

Appendix B
Matlab Code 135
B.1 Matlab Impedance Fitting Code . . . . . . . . . . . . . . . . . . . 135
B.2 Impedance Data Interpolation . . . . . . . . . . . . . . . . . . . . 136
B.3 Matlab Nyquist Plotter . . . . . . . . . . . . . . . . . . . . . . . . 137

Appendix C
Voltage Source Converters 140
C.1 Two Level Voltage Source Three Phase Converter . . . . . . . . . 140
C.2 dq0 Inner Current Control Loop Design . . . . . . . . . . . . . . . 141
C.3 PMSG control Design . . . . . . . . . . . . . . . . . . . . . . . . . 143
C.3.1 dq0 Outer Voltage Control Loop Design . . . . . . . . . . . 145
C.3.2 Symmetrical Optimum Method . . . . . . . . . . . . . . . 147
C.4 Component Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
C.4.1 Current Ripple Design . . . . . . . . . . . . . . . . . . . . 148
C.4.2 Capacitor Sizing . . . . . . . . . . . . . . . . . . . . . . . . 148

vi
List of Figures

1.1 Structure of the Grid . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2 Some Basic DC-DC Converter Topologies . . . . . . . . . . . . . . . 9
1.3 The Half Bridge Circuit . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4 HOMER Microgrid . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5 High Level Design and Stability Analysis Relationship . . . . . . . . 16

2.1 Switching and Control Frequency Issues . . . . . . . . . . . . . . . 20


2.2 Buck Converter w & w/o Filter Capacitor . . . . . . . . . . . . . . 21
2.3 Wind System and Six Pulse Rectifier . . . . . . . . . . . . . . . . . 21
2.4 DC Electrical Diagrams . . . . . . . . . . . . . . . . . . . . . . . . 24
2.5 Thévenin Equivalents . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.6 Nyquist Plot Construction . . . . . . . . . . . . . . . . . . . . . . . 27
2.7 Visual Summary of DC Stability Criteria . . . . . . . . . . . . . . . 30
2.8 Assessing Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.9 DC LTI Circuit Diagram . . . . . . . . . . . . . . . . . . . . . . . . 36
2.10 DC LTI Example Circuits Time Domain . . . . . . . . . . . . . . . 36
2.11 DC LTI example Stabe Nyquist Plot N = +1 . . . . . . . . . . . . 37
2.12 DC LTI example Unstabe Nyquist Plot N = −1 . . . . . . . . . . . 37
2.13 Stability Margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.14 Interface Choice and Nyquist Plots . . . . . . . . . . . . . . . . . . 39
2.15 Military EMS Microgrid Concept . . . . . . . . . . . . . . . . . . . 41

3.1 Reduced Nyquist Contour . . . . . . . . . . . . . . . . . . . . . . . 45


3.2 Impedance Fitting Results . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 DC System Current Injection . . . . . . . . . . . . . . . . . . . . . 49
3.4 DC Impedance Measurement Block . . . . . . . . . . . . . . . . . . 50
3.5 Frequency Sweep Diagnostics . . . . . . . . . . . . . . . . . . . . . 51
3.6 Inside DC IM Block . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.7 Simple LTI Circuit Test Diagrams . . . . . . . . . . . . . . . . . . . 52
3.8 Simple LTI Impedance Results . . . . . . . . . . . . . . . . . . . . . 53

vii
3.9 Diode Bridge Circuit Test Diagrams . . . . . . . . . . . . . . . . . . 54
3.10 "Source" type Diode Bridge Impedance Results . . . . . . . . . . . . 54
3.11 "Load" type Diode Bridge Impedance Results . . . . . . . . . . . . . 55
3.12 dq0 Transformation of Three Phase Voltage . . . . . . . . . . . . . 57
3.13 Injector Block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.14 AC Current Perturbation . . . . . . . . . . . . . . . . . . . . . . . . 58
3.15 Simple AC LTI Circuit Diagram . . . . . . . . . . . . . . . . . . . . 60
3.16 Simple AC LTI Circuit Impedance Magnitudes . . . . . . . . . . . . 61
3.17 Simple AC LTI Circuit Impedance Angles . . . . . . . . . . . . . . 61
3.18 Six Pulse Diode Bridge Diagram . . . . . . . . . . . . . . . . . . . . 62
3.19 Six Pulse Diode Bridge Input Impedance Magnitudes . . . . . . . . 62
3.20 Six Pulse Diode Bridge Input Impedance Angles . . . . . . . . . . . 63

4.1 General Hybrid Microgrid . . . . . . . . . . . . . . . . . . . . . . . 65


4.2 Power Electronics-based Microgrid . . . . . . . . . . . . . . . . . . 66
4.3 DC "leg" Architecture Examples . . . . . . . . . . . . . . . . . . . . 67
4.4 Conventional Grid Architecture Example . . . . . . . . . . . . . . . 68
4.5 Notional Microgrid Layout . . . . . . . . . . . . . . . . . . . . . . . 69
4.6 Example Cp vs TSR Curve . . . . . . . . . . . . . . . . . . . . . . . 70
4.7 Inside Wind Turbine Simulink Block . . . . . . . . . . . . . . . . . 71
4.8 Wind Turbine - Motor Structure . . . . . . . . . . . . . . . . . . . . 72
4.9 Turbine-Motor Rectifier Architectures . . . . . . . . . . . . . . . . 73
4.10 Turbine Diode Rectifier Diagram . . . . . . . . . . . . . . . . . . . 75
4.11 Diode Bridge Impedance Measurement Diagram . . . . . . . . . . . 75
4.12 Diode Bridge Impedances . . . . . . . . . . . . . . . . . . . . . . . 76
4.13 Diode Bridge Impedances with Constant Voltage . . . . . . . . . . . 77
4.14 Diode Bridge Impedances with Constant Speed . . . . . . . . . . . 77
4.15 Diode Mechanical Power Impedance Surface . . . . . . . . . . . . . 78
4.16 Diode Electrical Power Impedance Surface . . . . . . . . . . . . . . 78
4.17 Motor Rectifier Boost Subsystem . . . . . . . . . . . . . . . . . . . 79
4.18 Boost Converter Circuit . . . . . . . . . . . . . . . . . . . . . . . . 80
4.19 Motor Rectifier Boost Control Flow . . . . . . . . . . . . . . . . . . 81
4.20 Turbine Start Up with Boost MPPT . . . . . . . . . . . . . . . . . 81
4.21 Boost MPPT and Variable Wind . . . . . . . . . . . . . . . . . . . 82
4.22 Motor-Rectifier-Boost Time Domain . . . . . . . . . . . . . . . . . 82
4.23 Boost MPPT Rectifier Impedance Magnitude . . . . . . . . . . . . 83
4.24 Boost MPPT Rectifier Impedance Angles . . . . . . . . . . . . . . . 84
4.25 Switching Bridge Rectifier . . . . . . . . . . . . . . . . . . . . . . . 85
4.26 Switching Bridge Rectifier Control Flow . . . . . . . . . . . . . . . 86
4.27 Switching Bridge Rectifier: Startup . . . . . . . . . . . . . . . . . . 87

viii
4.28 Switching Bridge Rectifier: Variable Wind . . . . . . . . . . . . . . 88
4.29 Switching Bridge Rectifier Diagram . . . . . . . . . . . . . . . . . . 88
4.30 Switching Bridge Rectifier Impedance Magnitudes . . . . . . . . . . 89
4.31 Switching Bridge Rectifier Impedance Angles . . . . . . . . . . . . . 90
4.32 Battery Equivalent Circuit . . . . . . . . . . . . . . . . . . . . . . . 92
4.33 Battery Nominal Impedance Plot . . . . . . . . . . . . . . . . . . . 92
4.34 Battery Equivalent Circuit Impedances . . . . . . . . . . . . . . . . 93
4.35 Two-Level Three Phase Voltage Sourced Converter Diagram . . . . 94
4.36 P Resonant Controller . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.37 P Resonant Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.38 P Resonant Time Domain Simulation . . . . . . . . . . . . . . . . . 97
4.39 P Resonant DC Impedance Magnitudes . . . . . . . . . . . . . . . . 98
4.40 P Resonant DC Impedance Angles . . . . . . . . . . . . . . . . . . 98
4.41 P Resonant AC Impedances: P = 2.5 MW Q = 0.5 MVAR . . . . . 99
4.42 "Stiff" dq0 Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.43 "Stiff" dq0 Time Domain Results . . . . . . . . . . . . . . . . . . . 101
4.44 "Weak" dq0 Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.45 "Weak" dq0 Time Domain Simulation . . . . . . . . . . . . . . . . . 103
4.46 "Weak" dq0 DC Impedance Magnitudes . . . . . . . . . . . . . . . . 104
4.47 "Weak" dq0 DC Impedance Angles . . . . . . . . . . . . . . . . . . 105
4.48 "Weak" dq0 AC Impedance . . . . . . . . . . . . . . . . . . . . . . . 105

5.1 P Resonant Inverter System . . . . . . . . . . . . . . . . . . . . . . 107


5.2 P Resonant Inverter Nyquist Plots . . . . . . . . . . . . . . . . . . 108
5.3 P Resonant Inverter System Time Domain . . . . . . . . . . . . . . 108
5.4 P Resonant Inverter System Margins . . . . . . . . . . . . . . . . . 109
5.5 DC Bus Stability Assessment Interfaces . . . . . . . . . . . . . . . 110
5.6 Microgrid Component Impedances . . . . . . . . . . . . . . . . . . . 111
5.7 Nyquist Plots: Notional Wind System with Diode Rectifier . . . . . 112
5.8 Nyquist Plots: Notional Wind System with Boost Rectifier . . . . 113
5.9 Nyquist Plots: Notional Wind System with Switching Bridge Rec-
tifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.10 Diode Rectifier and High Resistance Battery Line . . . . . . . . . . 114
5.11 Boost Rectifier and High Resistance Battery Line . . . . . . . . . . 115
5.12 Switching Bridge Rectifier and High Resistance Battery Line . . . . 115
5.13 Comparison of Rectifier Impedances . . . . . . . . . . . . . . . . . 117

6.1 AC Microgrid Stability Diagrams . . . . . . . . . . . . . . . . . . . 120

A.1 CERTS Microgrid Microsource Subsystem Diagram . . . . . . . . . 132

ix
A.2 Simple Hybrid Wind System . . . . . . . . . . . . . . . . . . . . . 133
A.3 Simple Hybrid Wind System Time Domain Simulation Example . . 134

C.1 Two Level VSC Diagram . . . . . . . . . . . . . . . . . . . . . . . . 140


C.2 Post VSC Filter Diagram . . . . . . . . . . . . . . . . . . . . . . . . 144
C.3 Current Loop Control . . . . . . . . . . . . . . . . . . . . . . . . . . 144
C.4 PSMG Control Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 145
C.5 Current Ripple Diagram . . . . . . . . . . . . . . . . . . . . . . . . 149
C.6 Voltage Ripple Diagram . . . . . . . . . . . . . . . . . . . . . . . . 149

x
List of Tables

1.1 Grid vs. Microgrid . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

3.1 Simple DC LTI Parameter Values . . . . . . . . . . . . . . . . . . . 52


3.2 Diode Bridge Parameter Values . . . . . . . . . . . . . . . . . . . . 55

4.1 Microgrid Component Types . . . . . . . . . . . . . . . . . . . . . . 67


4.2 Notional Microgrid Nominal Bus Values . . . . . . . . . . . . . . . 68
4.3 Cp Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.4 Diode Rectifier Operating Ranges . . . . . . . . . . . . . . . . . . . 76
4.5 Motor Rectifier Boost Parameters . . . . . . . . . . . . . . . . . . . 80
4.6 Boost Rectifier Operating Ranges . . . . . . . . . . . . . . . . . . . 83
4.7 Switching Bridge Rectifier Parameters . . . . . . . . . . . . . . . . 86
4.8 Switching Bridge Rectifier Operating Ranges . . . . . . . . . . . . . 87
4.9 Battery Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.10 P Resonant Parameters . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.11 P Resonant Operating Range . . . . . . . . . . . . . . . . . . . . . 99
4.12 "Stiff" dq0 Inverter Parameters . . . . . . . . . . . . . . . . . . . . . 101
4.13 "Weak" dq0 Inverter Parameters . . . . . . . . . . . . . . . . . . . . 103

5.1 DC System Operating Variables . . . . . . . . . . . . . . . . . . . . 112


5.2 Notional Microgrid Stability Margins . . . . . . . . . . . . . . . . . 112
5.3 Unstable Operating Point . . . . . . . . . . . . . . . . . . . . . . . 114

xi
Acknowledgments

First I would like to thank my dissertation advisor, Dr. Tom Hughes who has over
the years taken considerable time to meet with me weekly and provide guidance
on this research as it evolved. This project would simply not have been possible
without his support. Also there is co-advisor Dr. Susan Stewart, present since the
very beginning, who has consistently supported me through rough patches. She
also introduced me to some of the interesting aspects of wind energy and encour-
aged me to pursue work and opportunities associated with it.

I would like to thank all the committee members for taking the time to review
my work and provide feedback. Special thanks are in order for Dr. Jeffrey Mayer
and Dr. James Turso who have put in countless hours over the past year advising
me on the electrical engineering matters that my background lacked.

ARL also deserves acknowledgment, providing me with the opportunity to con-


duct this work through the Walker graduate assistantship.

Finally, I must thank my wife who has been more than patient over the years
as I pursued this work. Whenever a difficult task was encountered or progress
stalled, she was there to provide the necessary encouragement and confidence to
power through.

xii
Epigraph

" The wise man doesn’t give the right answers; he poses the right questions."
- Claude Levi-Strauss

xiii
Chapter 1 |
Background

One could argue that it is the largest machine ever created. Colloquially known
as "the grid", it is a large network of electrical devices that despite its staggering
complexity is so ubiquitous and familiar to the average person it is practically a ne-
cessity. Generators, motors, and other devices are all electrically coupled together
in a network that crosses borders and continents. Although the technologies and
fuels it uses change with the times, the basic structure has been largely unchanged
since its creation. Electricity is generated by large centralized power plants using
synchronous generators. It is distributed across regions using a high voltage alter-
nating current (AC) three phase transmission system. At substations the power
steps down to lower voltages and is distributed over a local area. Finally the elec-
tricity makes its way to our homes and businesses in the (relatively) low voltage
distribution system that lines our streets with poles, lines, and transformers. This
structure, shown in Figure 1.1, was developed early in the twentieth century and
with the conclusion of the current wars, in favor of Tesla’s AC power, has persisted
since [1].

1.1 Recent Trends


Recent changes in the economic and regulatory landscape have shaken the utility
industry. Shifts are occurring due to swings in the price of natural resources, dereg-
ulation, and the implementation (as well as expectation) of environmentally based
subsidies, taxes, and regulations. This confluence of events is creating a sense of
uncertainty within the industry and this naturally creates anxiety about invest-
ing in the capital intensive grid projects that have traditionally been required for

1
Figure 1.1: Structure of the Grid

growth, like long distance transmission lines and large centralized plants. A pref-
erence for more flexible solutions and technologies has appeared. The following are
some technologies and ideas that have been used to delay traditional investments
and allow the grid to supply an increasing customer base.

• Energy Efficiency: In addition to improvements in generation and trans-


mission, end-use devices have become much more efficient over the years,
with everything from televisions to lighting consuming a fraction of the power
they once did. Note however this, disturbingly, may only be counted on in
the short term as historically increased efficiency usually leads to increased
aggregate demand, referred to as the rebound effect [2].

• Demand Response (DR): To better balance and stabilize the grid some
customers voluntarily decrease consumption in exchange for financial com-
pensation. The ubiquitous use and availability of fast digital communication
allows DR to be quickly implemented when the grid is stressed. This allows
the transmission system to be more fully utilized.

• Renewables: Renewables are a hedge against commodity price fluctuations.


They tend to have a large upfront costs but incur little or no marginal cost to
produce energy. Also the widespread use of small scale renewables "behind
the meter" decreases the perceived demand, although this leads to other
problems [3].

2
• Distributed Generation: Placing small scale generation and combined
heat and power (CHP) at the substation level can overcome economies of
scale by achieving high efficiency and deferring infrastructure upgrades [4]

• Gas Fired Plants: Traditionally these have been implemented to consume a


local resource or to provide peaking power only (due to expense). "Fracking"
has created a surplus of natural gas that has driven prices down to historical
lows. Unlike coal and nuclear, natural gas plants can be constructed rapidly
and inexpensively [5]. This has led to an explosion of growth despite concerns
about the sustainability of supply at such low prices.

• Rise of the Regional Transmission Organizations (RTOs): Created


by FERC order No. 2000 [6], RTOs have been adopted in many states
to operate the transmission grid and administer the (deregulated) electricity
markets. They monitor, control, and plan the transmission system to operate
stably and be more fully utilized .

• Energy Storage: Congestion and long transmission lines experience a volt-


age drop which is sometimes the limiting factor for (allowable) operation.
Energy storage and injection of reactive power at key locations can defer the
expensive need to upgrade lines or build additional power plants.

Some of these solutions can have negative effects as well as positive. This is
especially true for renewables and distributed generation, which in small amounts
have negligible adverse effects. However as penetration grows, conflicts can arise
between DG and the grid due to the existing grid’s rigid structure. Renewables
and distributed generation produce (from the utility’s perspective) highly variable
and uncontrolled power. As stated previously in Section 1, the basic structure of
the grid; one directional flow of power from centralized plants to end users through
a hierarchical AC transmission system, was not designed for two way power flow
or for accommodating scores of independently controlled generators. Although
its structure has remained largely unchanged in the face of new technologies and
fuels, there is reason to believe change is possible. Advancements in computing,
communications, and especially power electronics technology, has made decentral-
ized approaches to grid operation more feasible. Microgrids have been proposed

3
as an alternative, or at least an augmented, approach to modernizing the grid and
addressing some of its limitations.

1.2 Microgrids
What exactly is a microgrid? Although there is no agreed upon exact definition,
the general consensus states they are autonomous electrical networks with power
capacities in the kW to MW range. This definition would technically include
several related applications not usually labeled as microgrids, such as remote sites
and vehicles.

• Remote Sites: In instances where grid power is not available (cabin, cell
tower, rural community) electricity needs are generated and distributed on
site. Commercial-off-the-shelf (COTS) generators mimic properties of the
grid so that the user can connect regular loads and devices. These systems
are expensive to run as small generators are typically less efficient and require
a higher quality fuel than their larger scale counterparts [7]. Also modularity
and scalability are limited. Generator control is simple for a single unit
system. Connecting many units together, however, requires coordination to
operate smoothly and efficiently.

• Vehicles: Electrical networks play a prominent role in many vehicles; ships,


aircraft, and more recently electric cars. In contrast to the relatively sluggish
changes in the utility industry, vehicles have been quick to adopt new power
technologies and ideas suited to their goals. The Navy is experimenting with
electric propulsion drives [8] and DC distribution [9]. Modern commercial
aircraft have embraced a 400 Hz AC network and power electronics to reduce
weight [10]. Electric cars and hybrids use bidirectional converters for regen-
erative breaking [11]. Such applications require a high degree of performance
and reliability, thus inspiring research also useful to microgrid design.

In this work the modern microgrid concept of an electrical network with dis-
tributed generation will be assumed. Such systems are expected to contain renew-
ables, energy storage and be able to work independent of or in concert with the
regular grid. Incorporating microgrids into the main grid is expected to provide a
number of potential benefits to both the utility and user, such as:

4
• Reliability: A microgrid invariably must contain some generation sources.
In the case of a grid failure, the microgrid can island itself and operate
independently until power is returned. A microgrid differentiates itself from
a simple backup system in that it is expected to smoothly transition between
modes, run parallel to the grid, and sustain itself for longer periods of time [4].

• Renewables: From the viewpoint of the utility, microgrids are an attractive


solution to the problems posed by high penetrations of renewables. Coor-
dination and control of renewables and energy storage are delegated to the
microgrid. The utility is only concerned with the net power crossing the
point of common coupling (PCC), which the microgrid aims to keep within
bounds.

• Combined Heat and Power: CHP is highly efficient from an energy view-
point since "waste" heat is utilized to offset electricity demand. Heat, unfor-
tunately, is difficult to physically transport, necessitating a heat distribution
system of steam or chilled water pipes. Microgrids allow many smaller CHP
generators to be placed at load locations in lieu of a single generator with a
complex heat distribution network. [4]

• Power Quality: Microgrids can customize the power to meet the specific
needs of the load. Incoming power quality issues such as low displacement
power factor can be corrected for and loads supplied with the correct form
and quality of power that they require. Although, as will be discussed in sec-
tion 1.5, power electronics-based microgrids can introduce new power quality
problems of their own.

• Demand Response: Currently demand response is a slow process, occur-


ring over the time scale of hours. Microgrids have the potential to respond
almost immediately to a utility request as it can control local generation or
even voluntarily islanded itself if desired.

Despite being proposed to solve a variety of problems, microgrid adoption


has been slow. Most microgrids in operation are small scale or research in na-
ture [12–14]. We can attribute this, at least in part, to the custom nature of
microgrids and the difficulty in designing them to meet universally accepted stan-
dards [13]. Although mature and standard design methods have been developed

5
for the conventional grid, these techniques don’t transfer to microgrid design as
well as one would expect. The problem arises because microgrids, despite their
name, do not necessarily behave like small versions of the main grid. There are
a number of key differences between the main grid and a (modern) microgrid, as
described in Table 1.1.

Table 1.1: Grid vs. Microgrid

Property Grid Microgrid


Power Flow Unidirectional Reversible
Energy Storage Rotating Mass Low, Virtual, Electrochemical
Transmission Inductive Resistive
Power Quality Power Factor Angle Harmonic Distortion
Control Automatic Generation Control Many Methods

Note that Table 1.1 is not absolutely definitive as microgrids can blur the lines
of distinction, containing some combination but not all of the qualities mentioned.
A more detailed breakdown and explanation of the properties is presented in Ap-
pendix A.1.

1.3 Operation
Microgrids are very versatile. They are not restricted to operate in the way the
main grid does, although there are advantages to doing so. They can be DC or
AC, they can adopt a convenient fundamental frequency and voltage, they can im-
plement new control techniques. They can be freely customized to meet the needs
specific to their location and purpose. This greater versatility comes at a cost as
it has made drafting design rules and standards very difficult [13], although it has
been attempted recently with the publishing of IEEE 1547 family of standards for
"Interconnecting Distributed Resources with Electric Power Systems".

Vehicle electrical networks and microgrids actually have a lot of cross over.
In vehicles the design problem may be challenging but designers have complete
control over the components to be included in the system. Also, the final prod-

6
uct is usually to be mass produced so the design costs can be sufficiently diluted
to allow a process of iteration and improvement. This (ideally) results in highly
performing cars and aircraft, that meet expectations of safety and reliability. An
exception in this category is ships. The scale, time, and cost of construction limits
the experiment and refinement possible before completion, incentivising designers
to pre-empt problems or solve them early. Much of the work in this dissertation
is based on research originally undertaken to solve ship based electrical network
problems.

In contrast to vehicles, remote applications, and older microgrids, a modern


microgrid is expected to have three key qualities. [4, 15, 16]

• Plug and Play: System can be reconfigured or accept new hardware with-
out significant re-engineering required. Since microgrids must be adapted
to local resources, needs, loads, etc, there is no "optimal" design to be cre-
ated and mass produced. A microgrid must be capable of accepting new
generators and loads without the need to rework the whole system.

• Peer to Peer: System must be redundant in a way that there is no single


point of failure. This goes beyond simply ensuring that it contains multiple
sources and energy storage units. It effectively prohibits the necessity of a
communication based control system to function. This does not preclude
the use of communication-based control, in fact, on some levels it would be
highly recommended [14]. But the microgrid must be able to function, at
least at a basic level, without it.

• Semi-autonomous: System must seamlessly couple and decouple itself


from the grid. If unacceptable conditions are detected at the point of com-
mon coupling (PCC), the microgrid can sever itself from the grid and operate
autonomously, maintaining nominal voltage and frequency. After an event
the microgrid can automatically resynchronize and reconnect to operate in a
grid-tied mode.

A plethora of architectures and control schemes have been developed in the


literature to satisfy these requirements. Notable for AC systems are droop methods
and the CERTS microgrid concept [4]. Expanded descriptions are presented in

7
Appendix A.2.2. Although the best architecture and method of operation is open
to debate, the necessity of power electronics is not. They are the unifying common
component in modern microgrids and they present new challenges to the design
process.

1.4 Power Electronics


The microgrid concepts being proposed to solve some of todays power problems are
not especially new, however engineers in the past were limited by the technology
available. The rise of inexpensive and efficient semiconducting switches revolution-
ized the computing industry, ultimately bringing about the digital world we live in
now. Power electronics, which are essentially scaled up versions of the switches used
in microchips and electronics, allow power conversion to be done in a fundamen-
tally different way. The production, transmission, distribution, and consumption
of electrical power via alternating current quickly became standard in the United
States and around the world in a large part because of the convenience afforded
by AC generators, motors, and transformers. Transformers and synchronous gen-
erators were especially useful as they allowed for an electromechanical means to
efficiently control and transmit electricity through the grid, Appendix A.2.1. With
the advent of high power switches that can handle large amounts of current and
voltage, it becomes possible to use switching circuits to accomplish voltage steps
in place of transformers.

There exists a huge variety of power electronic converters but common to


all active devices is a switching element, often an insulated gate biploar transis-
tor (IGBT) or metal oxide semiconductor field effect transistor (MOSFET), that
rapidly changes the pathways of a circuit. The circuits have as little resistance
as possible so that electrical power at the input and output is mostly conserved.
Since the switching elements are gated with digital signals, the exchange of voltage
and current can be dynamically controlled. A general term for this kind of device
is power converter, although converter can also refer more specifically to devices
that operate between DC buses. Some simple power electronic converter circuit
topologies are presented in Figure 1.2.

8
Figure 1.2: Some Basic DC-DC Converter Topologies

(a) Boost Converter (b) Buck Converter (c) Buck-Boost Converter

The boost converter steps voltage up, the buck steps it down, and the buck-
boost can do both but inverts the polarity at the output. There are many versions
and types of power converter circuits, but the operating principles remain the
same. High power semiconductor elements create a nonlinear circuit tailored to
some electrical conversion operation. Unlike their electromechanical counterparts
in AC systems, these converters can be dynamically controlled and used to perform
many other tasks in addition to voltage transformations. For example, later in this
dissertation, see section 4.4.2, a boost converter will be designed and modeled in
Matlab to mechanically load a wind turbine, achieving maximum aerodynamic ef-
ficiency.

Critically, power electronics allow not just DC-DC voltage transformations but
also efficient power conversion between DC and AC networks. The fundamental
building block of conversion between three phase AC and DC is the half-bridge
circuit, Figure 1.3. When power is converted from AC to DC the device is called
a rectifier. When power is converted from DC to AC it is called an inverter. His-
torically most electrical loads (ex. motors) required AC power and the remainder
(ex. incandescent lighting) were indifferent to it. However today many modern
devices (virtually all electronics and computers) run principally on DC power. As
distribution of DC power is as yet not practical except in special circumstances,
the solution is to build rectifying power supplies into the loads [1, 9, 17]. Therefore
virtually all electronics that plug into an electrical outlet contain a power rectifier
of some kind.

The counterpart to the rectifier is the inverter. While not as ubiquitous, its use
is increasingly widespread. It has two primary uses, interfacing DC power sources

9
Figure 1.3: The Half Bridge Circuit

with AC networks and driving motors. AC motors, especially induction types,


tend to be inexpensive and of high efficiency and reliability [18]. However they
are tied to the AC electrical frequency (or divisor of), operating at or very close
to a set (or sets) of speeds. This inconvenience can be eliminated with the use
of variable speed drives. Although slightly less efficient than a directly connected
motor at peak condition, they can control the speed and loading of the motor to
ensure maximum efficiency over a range of conditions. This makes them ideal for
increasing the efficiency of motors that must operate over a range of speeds. Later
a drive will be designed and modeled in matlab to mechanically load a wind turbine
in order to achieve maximum aerodynamic efficiency, see section 4.4.3, in addition
to the boost converter version already mentioned. The second popular application
of inverters is to transfer power from DC sources to an AC electrical network, such
as the grid. The AC voltage waveform present on the grid is created through the
physical rotation of the electrical machines, where frequency is directly governed

10
by the speed of the synchronous generators [19]. DC sources such as photovoltaics
(PV) and batteries produce voltage through electrochemical reactions and have no
moving parts. Since most loads, including DC ones, require a standard AC voltage
input, inverters are necessary whether connected to the grid or not.

In the previous paragraphs some of the more common applications of power


electronics are highlighted, but the list is by no means exhaustive. They have
many uses and can be called a technological enabler. They allow the flexibility to
use high tech devices (solar panels, electronics, motor control, static synchronous
compensator (STATCOMs)) and achieve better performance and efficiency. Unfor-
tunately power electronics are not without their disadvantages. Their active nature
and complex design can sometimes adversely affect overall system efficiency and
reliability. They not only require power to operate, but small losses occur during
the transition of each switching cycle. Although almost negligible individually,
the sum can be significant as they repeatedly occur thousands (and sometimes
hundreds of thousands) of times per second. A common problem in conventional
(rotating machine based) power systems is displacement power factor, a loss of
efficiency due to a phase mismatch between current and voltage waveforms. Power
electronics-based systems can easily correct this through proper control but their
switching introduces irregularities in waveform shape called harmonic distortion.
Although physically different, it is measured in a similar way and called distortion
power factor. Both factors reduce the systems transmission efficiency and can neg-
atively affect sensitive loads.

Some of the advantages and disadvantages of power electronics are conflicting.


They both increase efficiency and decrease it, they improve power quality but also
diminish it. Ultimately their usefulness depends on a balance of factors that are
highly situational. In certain situations the gains outweigh the losses and vice
versa. For example, in a typical residential home it is generally most efficient to
use grid supplied AC power directly. However if a DC source is available, such as
PV or a battery, it might be more efficient to form a home DC distribution system
and interface the grid and AC loads through power inverters [1].

System designers can balance the benefits and costs to build systems adapted to

11
their environment and specific circumstances. However there is a critical downside
that transcends the trade-off process; power electronics induced instability. An
unstable system will, at least under certain conditions, seriously malfunction. This
offsets possible efficiency or performance benefits.

1.5 Negative Incremental Input Impedance


System instability due to negative incremental input impedance is a peculiar inter-
dependent problem associated with incorporating power electronics into an elec-
trical system. The inner control loops of power electronics operate so quickly that
they can behave as constant power loads. As technologies have matured the effi-
ciency of converters has risen higher, further exacerbating the problem. Following
is the mathematical derivation as presented in [20].

Pin ≈ Pout for high efficiency converter

dPin = 0 = idv + vdi

rearrange and define resistance using Ohm’s Law

dv v
rin = =−
di i

In order to maintain output power, increasing amounts of current will be drawn


from an input with decreasing voltage. This forms a positive feedback loop where
the increased current draw depresses the input bus voltage, thus increasing current
draw, and so on. This makes power electronics behave as a negative resistance for
disturbances over certain time scales. This is inherently destabilizing to an elec-
trical system and presents a serious potential problem [21]. Although nonlinear
mechanisms (such as saturation) may intervene to prevent total runaway, an un-
stable system is likely to activate protections (breakers) or become trapped in a
limit cycle where a system variable such as voltage will perpetually oscillate. This
oscillation is more than a mere inconvenience and can transfer to connected rotat-
ing machinery potentially damaging them [20,22]. Note this effect occurs in single
and multiphase AC systems.

12
1.6 Design Challenge
In high level microgrid design, site specific information such as local resource and
demand data, is used to find Pareto optimal combinations of equipment. This
type of analysis can be accomplished through statistical means, as is done in
Hybrid2 [23], or via Monte-Carlo methods in the time domain, as is done in
HOMER [24]. This was explored in previous work modeling the economics of
microgrid designs under different scenarios [25]. An important convenience of such
programs is that they are adaptable to the level of information available. With
coarse data and information, the analysis will produce equivalently uncertain es-
timates. If more detailed data and information is available, results with a higher
confidence can be obtained. Work on custom models show such techniques are
limited only by the resolution of input data and modeling of dynamics, Appendix
A.3, making these analysis tools very powerful in the preliminary phases of design,
and potentially later on given enough information. A key facet of the analysis is
the combination of component modules in different quantities and architectures,
of which power electronics play a significant role, Figure 1.4.

Negative incremental input impedance caused by power electronics is not ac-


counted for by descriptions of power electronics in such high level design programs.
This presents a potential problem as instability can be difficult to detect or mitigate
once a design has been established. However, incorporation of power electronics
stability assessment into preliminary or high level analysis presents difficulties for
several reasons.

• Level of Specification: The device properties needed to assess whether


power electronic’s negative incremental input impedance will induce insta-
bility are not specified. As will be demonstrated later, the information needed
is extensive and not easily distilled to a single specification to be promoted
or advertised.

• Interdependence: Instability occurs due to the interaction of components


and is not necessarily due to a flaw in the individual components themselves.

13
Figure 1.4: HOMER Microgrid

Each component may function fine in laboratory, factory, and test settings
but fail when connected to the actual system.

• State Specific: Instability may only occur under a very specific set of rare
operating conditions, making even identification of the problem difficult in
some circumstances.

• Uncertainty: Even in the case that stability is assured through proper


design, stability properties can change in practical microgrids. Changes in
layout, software, heating, or aging of components can affect the stability
characteristics of systems.

1.7 Proposed Approach


In this work an approach to integrate power electronics stability assessment with
the higher level aspects of microgrid design is proposed, Figure 1.5 . Three goals

14
of a lower level stability assessment subroutine are identified as necessary to ac-
complish this. A technique that achieves these goals is expected to be useful to a
wide variety of microgrid design scenarios, although not all may not be necessary
for any given scenario.

• Qualitative Stability Assessment: In the simplest stability assessment, a


set of components and conditions are supplied, and an algorithm determines
whether the system is stable or not. Such a determination will be exact to
the criteria used, which may involve an inherent margin or conservativeness.

• Quantitative Stability Assessment: Optimization algorithms usually


work by minimizing a set of costs. In the case that a systems degree of
stability is a variable of interest, a quantitative measure of robustness is
needed. The stability assessment routine must returns stability margins for
a set of components and conditions.

• Failure Conditions Assessment: If the higher level optimization is iter-


ative or the system is adaptive, it would be very useful for the lower level
stability assessment to report the conditions of instability. In the case that
a highly desirable system has been determined unstable, it is worthwhile for
the optimization to know if it occurs under preventable conditions. Ulti-
mately this allows a separation between infeasible instability and penalizable
stability.

15
Figure 1.5: High Level Design and Stability Analysis Relationship

No details about the higher level design process are assumed, requiring gener-
ality and versatility on the stability assessment end.

1.8 Dissertation Objective


The overall objective of this dissertation was to develop and demonstrate stability
assessment techniques that meet the requirements outlined in the previous section.
Work was conducted in three major areas to support of this objective, they are
summarized below.

1. Stability Assessment: A thorough review of the theory and literature


was conducted to determine promising methods and criteria for assessing
the stability of electrical networks. Techniques most suited to microgrids
were selected and adapted to produce the outputs needed by a high level

16
microgrid design program. A summary of theory and methods to assess
microgrid stability as well comments and ideas are presented in Chapter 2.

2. Stability Characteristics: Tools were developed to measure the stability


characteristics of components via simulation. Chapter 3 contains background
theory and the developed automatic Matlab simulink impedance measure-
ment blocks.

3. Verification of Methods: The techniques and methods developed to mea-


sure stability characteristics and assess stability were demonstrated via sim-
ulation on non-trivial circuit models. Chapter 4 presents a detailed small
wind power system modeled in MATLAB while Chapter 5 reports results
and conclusions.

17
Chapter 2 |
Stability

Relevant stability theory is presented to solve the negative incremental impedance


problem of interacting power electronic converters. Critical assumptions such as
averaging and linearizing of power electronic converter models are justified and
implications discussed. Due to the assumed black box nature of components,
bounded input bounded output (BIBO) stability is the mathematical definition
of stability used in assessment, with determination of bounded input bounded
state (BIBS) stability possible in special cases. Governing transfer functions and
Thévenin impedance circuits are derived and the return ratio, L = Zs /Zl , is shown
to be the determining factor for stability. The Nyquist stability criteria is employed
to assess stability in the w-plane (Nyquist plane). The various stability criteria
developed in the literature are summarized along with a similar but alternative
method called passivity-based stability criteria (PBSC). The dual problem for AC
electrical networks is presented along with its stability criteria. The methods in the
literature are all shown to be conservative. A method is proposed that implements
the Nyquist stability criteria directly to assess stability, and a separate maximum
peak criteria (MPC) to determine margin. Also the ability to "knock out" unstable
operating points is proposed and illustrated.

2.1 Stability
Mathematical techniques for assessing stability can be very powerful but involve
more than a little subtlety. Although an intuitive concept, stability can be com-
plicated when viewed from a mathematically rigorous perspective. This makes it
important to be explicit with assumptions and the type of stability analyzed.

18
2.1.1 Linear Time-Invariant (LTI) Assumption
In reality true linearity is rare, yet analysis based on that assumption dominates
the field of control theory. This is primarily due to the powerful methods available
to analyze linear systems. To assess the stability of a microgrid we also use this
assumption, but there are two important limitations that must be considered.

First, any linear stability assessment is valid only in an operating region where
the system behaves approximately linearly. Components in the system are more
or less "black boxes", so no knowledge of their linearity can be assumed. Stability
assessment is therefore only valid in the neighborhood of the operating point lin-
earized around, hence the term small signal stability. This is overcome by carrying
out the stability assessment over all the operating points in a sufficient density. As
will be shown, this is a measurement and computational burden but establishing
stability for a linearized system makes the problem more tractable.

Second, the switching of the power electronics means the system is not time
invariant. This problem is solved by assuming averaged models of the power elec-
tronic converters. System states (voltage and current) are averaged over a switch-
ing cycle allowing them to take on smooth continuous trajectories. Averaging
the state space models of power electronics is commonly used to simplify analy-
sis [18, 26]. However in the frequency domain, the averaged model breaks down
when approaching the switching frequency of the components, which in turn in-
validates small signal stability analysis there. Since the the negative incremental
impedance effect of power electronics is due to the inner control loop, the danger
of instability is only present over frequencies at which active control occurs. The
results of stability analysis can therefore be trusted when the switching time is
much less than the time constant associated with control action. This is simple
to ensure for a single device but in a microgrid all components must be consid-
ered. Therefore technically for the analysis to hold, the fastest control action of
the fastest device must be slower than the slowest switching device, Figure 2.1

19
Figure 2.1: Switching and Control Frequency Issues

Practically speaking this restriction is not absolute as power electronics are


typically filtered in such a way that the switching effects result in only a small
(to specification) ripple of voltage and current, Figure 2.2. At the accessible in-
terface the system is very nearly represented physically by the averaged model.
Thus mostly removing the problem posed by components with very low switching
frequency given they are adequately filtered. This negation under appropriate cir-
cumstances is important as passive diode bridge rectifiers are ubiquitous in small

20
PMSG wind power systems [27–31], but slow turbines speeds lead to a correspond-
ingly low electrical frequency and therefore also a low switching frequency, Figure
2.3.

Figure 2.2: Buck Converter w & w/o Filter Capacitor

Figure 2.3: Wind System and Six Pulse Rectifier

In summary, to ensure the microgrid can be approximated as an LTI system


the following factors must be considered when structuring a procedure to assess
the stability of a microgrid.

• Operating Points: Adoption of linear methods requires that a series of


"plants" be assessed for stability at given operating points. In terms of mi-
crogrid components, measurement of stability characteristics is only possible

21
under quasi-steady state conditions. This restricts the validity of a small
signal stability assessment to at most the common set of operating point
equilibriums between interacting components. In practice this is further re-
stricted to the common operating points at which stability characteristics
can be measured, although some aspects of generalized admittance space
analysis seek to mitigate this, section 2.2.5 [32].

• Switching, Control, and Filters: Valid analysis requires that power elec-
tronics behave linearly, meaning an assumption of averaged operation is re-
quired. The degree to which a power electronic device obeys this is dependent
on the amount of filtering. In the case that stability needs to be assessed
at an interface with little filtering then the switching frequencies must be
greater than the bandwidth of the control (fastest) control system.

2.1.2 Internal and External Stability


Mathematically there are many types of stability, but they can be separated into
two categories; internal and external stability [33].

• Internal: Concerned with the response of the systems internal states, which
generally correspond to (or some combination of) modes of energy storage,
to a set of initial conditions. Notable is Lyapunov stability which ensures
internal states remain bounded around an equilibrium point when initially
displaced from it.

• External: Concerned with the response of the system’s outputs, which are
usually measurable variables of interest, to system inputs. Also applicable
to state space methods, but more easily understood as an assessment of a
system’s transfer functions.

Internal stability is difficult to apply to the microgrid problem being studied.


To be practical, components are considered "black boxes", Knowledge of or ac-
cess to internal states may not be possible. External methods are therefore used,
namely bounded input bounded output (BIBO) stability [20].

For the general state space system

22
ẋ(t) = A(t)x(t) + B(t)u(t)

y(t) = C(t)x(t) + D(t)u(t)

BIBO Stability [33]: For any bounded input u(t) and initial condition x0 ,
there exists a finite constant N such that ky(t)k ≤ N for all t

In the case that the system is a LTI continuous system, section 2.1.1. A proper
transfer function can be constructed for a given set of inputs and outputs.

C(sI − A)−1 B + D = H(s)

The system is BIBO stable for those inputs and outputs if the poles of the
transfer function are in the open left hand plane (LHP).

To analyze the stability of power electronics-based networks, most techniques


rely on establishing BIBO stability. An exception is the recently developed pas-
sivity based stability criteria (PBSC), section 2.2.4. BIBO stability is convenient
because it can be determined for easily measured quantities like interface bus
voltages. It does have some weakness that can be of concern when applied to
microgrids.

BIBO only guarantees that selected finite inputs do not result in selected out-
puts that go to infinity. System states or even the response of the system to no
input can theoretically be unstable for BIBO systems under certain conditions [33].
This deficiency is mostly overcome by testing stability using impedance techniques
at multiple interfaces in a system.

If transfer function H(s) is co-prime, then its state space realization is mini-
mal [34], and it completely characterizes all the internals states. BIBO stability
of the transfer function H(s) then indicates the more stringent bounded input
bounded state (BIBS) stability. This is stronger than both BIBO and Lyapunov
types and more rigorously establishes stability.

23
Since microgrid components in this work are "black boxes", explicitly establish-
ing a transfer function developed from measured data as co-prime is practically
impossible. Controllability and observability would need to be tested, requiring
the true state space realization. If such knowledge were available other simpler
techniques could be used to determine stability (ex. eigenvalue analysis). Being
electrical devices however, we can assume that the dominant energy storage modes
are functions of voltage and current. Impedance transfer functions, which account
for the relationship of these two variables, will likely capture the dynamics of these
modes, and therefore allow the presumption of actual stability from BIBO stability
(at least locally).

2.2 DC system Stability


DC stability assessment, which was pioneered by Dr. David Middlebrook in the
70’s, provides a method of determining stability of constant power loads connected
to DC buses. It also provides the basis for the more recently developed AC stability
assessment methods.

2.2.1 Electrical System Transfer Function


To begin stability assessment the electrical system must be "cut" at an interface
and a transfer function created. The resulting transfer function varies depending
on the the method used. Originally Middlebrook did so using a constant power
load and a filtered voltage source, Figure 2.4 [35]. Alternatively the electrical
system has also been conceived more generally as back to back voltage sourced
converters [32, 36].

Figure 2.4: DC Electrical Diagrams

(a) Input Filtered CPL (b) Back to Back Converters

24
The transfer function is derived from the Thévenin equivalent circuit, see Fig-
ure 2.5.

Figure 2.5: Thévenin Equivalents

(a) Input Filtered CPL (b) Back to Back Converters

Despite beginning with different structures and containing different outputs,


the stability of the derived transfer functions depend on the same factor if compo-
nents are assumed to be stable by themselves.

For filtered source and constant power load [20].

Zl Voc vload
vlink = =
Zl + Zs hl

voc = hs vs
hs hl
vl = vs
1 + ZZsl
Disturbances in the source voltage are the input, resulting load voltage is the
output. Transfer function hs is a passive filter so it will never have right-hand
plane (RHP) poles. Transfer function hl will not have RHP if the converter was
designed to be stable with an ideal voltage source, which can be expected [20].

For the back-to-back converter model [32].

Zl Zs
vlink = vs + vl
Zs + Zl Zs + Zl

Ns
Zs =
Ds
Nl
Zl =
Dl

25
Nl Ds vs + Ns Dl vl
v=
Nl Ds (1 + ZZsl )
The link voltage is the output, while both source and load voltage disturbances
act as the input. If the source is designed to be stable when connected to a constant
load current and the load is designed to be stable when connected to a constant
voltage source then Nl and Ds do not have any RHP zeros.

As established in the previous Section, 2.1.2 BIBO stability at the interface is


ensured if the transfer function has no RHP poles. In both formulations the poles
are determined solely by the ratio of the source and load impedances.

1
1 + ZZsl

Or equivalently by the RHP zeros of the denominator.

Zs
1+
Zl

2.2.2 Nyquist Stability Criteria


The Nyquist criterion was developed at Bell Telephone Laboratories by Harry
Nyquist in 1932. It is a general technique that is used to determine the stability of
LTI feedback systems. The technique is graphical and is applied to a wide array
of problems. While it displays more information than the popular bode diagram,
it can be difficult to decipher the actual cause of instability.
The Cauchy argument principle forms the theoretical basis of the Nyquist Sta-
bility Criterion. It states that a contour in a complex function’s s-plane can used
to determine the relative number of enclosed poles and zeros. In a Nyquist plot
this contour is mapped to a new plane, the w-plane, so that the determination can
be done visually.

A graphical understanding of how the Nyquist plot is constructed mathemat-


ically is presented in Figure 2.6. For a single input single output (SISO) system
the mapping is accomplished by drawing vectors from a contour point to each pole
and zero. The resulting point in the w-plane is located on vector of zero product

26
vectors, divided by pole product vectors. By sweeping through the contour, a new
contour is created in the w-plane. The encirclements of the origin indicates the
relative number of poles and zeros encircled in the s-plane. By choosing a con-
tour, called the Nyquist contour, that encircles the entire RHP a Nyquist plot is
created. The Nyquist stability criterion states that the number of zeros is equal
to the number of critical point encirclements plus the number of poles.

Z =N +P

The critical point is the origin by default, but changes as the w-plane is shifted
to account for denominator constants in the transfer function. So in the derived
electrical system transfer function, and typical feedback systems, the desire is to
find zeros of 1 + G(s). The plane is shifted and the critical point becomes −1 + 0j.

Figure 2.6: Nyquist Plot Construction

If we assume, as in section 2.2.1, that the source is stable with a constant


current load, and the load is stable with a constant voltage source we can also
assume that ZZsl has no RHP zeros in the denominator, and therefore that 1 + ZZsl

27
has no RHP poles.

Ns
v(s) = Zs (s)i(s) = i(s) → Ds 6= RHP zeros
Ds

v(s) Dl
i(s) = = v(s) → Nl 6= RHP zeros
Zl Nl
Zs Ns Dl
= 6= RHP poles
Zl Ds Nl
The plotted transfer function has no poles. This by itself does not have a direct
effect on stability but prevents the zeros from "hiding" behind poles in the RHP.
This simplifies analysis because any unstable poles with manifest in encirclements
of the critical.

2.2.3 Stability Criteria


For the purposes of design, graphical encirclements of mapped functions aren’t
particularly conducive to the formulation of constraints. Since the development
of these stability methods, a series of criteria have been created to ease analysis.
Each work by creating a forbidden region in the w-plane to restrict the Nyquist
plot of the impedance ratio ZZsl , also termed the return ratio.

• Middlebrook Criteria [35]: Developed as a means to prevent instability in


systems with converters that behave as CPLs. The return ratio is constrained
to have a magnitude less than one. Or in the case that a stability margin
is desired, a magnitude less then some fraction of one. This criteria is the
most conservative as it forbids the Nyquist contour from occupying much of
the space. The chief advantage is its simplicity, which lends itself to design
specification. In contrast to all other criteria, it requires no knowledge of
impedance phase angle, which can be more difficult to measure.

• GMPM [36]: Gain margin and phase margin criterion was developed in
1995 to be less conservative then the Middlebrook criterion by forbidding
less of the w-plane. It enforces a desired gain and phase margin but allows
the return ratio to inhabit the space outside of the region allowed by the
Middlebrook criterion, as long these margins are maintained.

28
• Opposing Argument Criterion [37]: Actually more conservative then
the preceding GMPM criterion, but more easily applied to parallel loads [38].
The forbidden region is denoted by a vertical line in the w-plane. The location
of this line is specified for individual loads and is in terms of their proportion
of power consumption and the source impedance.

• ESAC [32]: The energy source analysis consortium criterion (ESAC) is


less conservative then both the GMPM and opposing argument criterion. It
imposes a gain and phase margin, as in the GMPM criterion, but further
reduces the forbidden region so the assessment is made less susceptible to
grouping effects due to interface choices. Grouping effects will be discussed
in greater detail in section 2.4.3. The ESAC plays an important role in
admittance space analysis which is discussed in section 2.2.5. Recently the
root exponential stability criterion (RESC) has been developed by using a
continuous function to define a boundary equivalent to the ESAC criterion,
solving some computational problems [39].

• Necessary and Sufficient Stability Criterion [40]: Separates the as-


sessment into two parts; qualitative stability and quantitative margin. The
forbidden region is shrunk to inhabit only the real line for values < −1. This
determines qualitatively whether the system is stable or not but provides no
margin of error. To meet stability margin requirements, a circular forbidden
region is draw around the critical point. In contrast to the other criteria
presented in this list, margin is defined in terms of maximum peak criteria
(MPC) rather than gain and phase margins.

Over time there has been a general trend of decreasing conservatism and in-
creasing computational requirement. This reflects the rapid strides made in com-
puter power since the Middlebrook criteria was proposed in 1976. Although more
computationally intensive, the less conservative criteria allow for less bulky pas-
sive filters to ensure stability [38]. A visual summary of the stability criteria is
presented in Figure 2.7.

29
Figure 2.7: Visual Summary of DC Stability Criteria

2.2.4 Passivity-Based Stability Criteria


The analysis in Section 2.2.1 and resulting criteria are all based on preventing the
factor 1+1Zs from containing poles in the RHP. This factor, called the minor loop
Zl
gain, is dependent on the ratio of impedances. The recently proposed method for
analyzing microgrid stability, called the passivity-based stability criteria (PBSC),
depends on individual or paralleled impedances instead [41].

Passivity [42]: The system y = h(x, u) is termed passive if uT y ≥ 0 for all x


and u.

For electrical systems let the system input be voltage, output be current, and
the storage function be a measure of the energy stored in the system.

v(t)i(t) ≥ 0
Z t
v(τ )i(τ )dτ ≥ 0 for all t
0

The electrical system will be passive if it can only absorb energy. As derived

30
in [41] this reduces to a requirement in the frequency domain that

1Z∞
Re[Z(jω)]kI(jω)k2 dω > 0 for all I
π 0
This is true for a one port network if the Nyquist plot of Z(jω) lies entirely in
the closed RHP [41].

Establishing passivity of the network and therefore stability of interacting com-


ponents is done by ensuring the paralleled impedance of each device meets this
requirement.

Z(jω) = Z1 (jω)kZ2 (jω)k . . . kZn (jω) 6= LHP Nyquist Contour

The passivity-based stability criteria has several properties that make it more
general then the other techniques presented.

• Unaffected by the layout of components whether in parallel or series.

• Does not distinguish between components as "sources" or "loads"

• Bi-directional power flow is allowed

There are some issues however, it can be difficult to ensure stability margin
using PBSC, although modifications have been offered to overcome this [43].

Although theory and formulation are different, the stability characteristics


needed (component impedances) are identical to that of the minor loop gain meth-
ods. The bulk of this work will therefore also apply to PBSC type analysis and it
is mentioned here as an alternative for that reason.

2.2.5 Generalized Admittance Space Analysis


Generalized admittance space analysis is a method that systematically applies a
chosen DC stability criteria to a generalized sets of source and load impedance
data [32]. The method addresses some of the shortcomings of the minor loop gain
stability techniques.

31
• Load Impedance Specification: The source impedance is applied to a
selected minor loop gain criteria, section 2.2.3, to specify an allowable range
for the load admittance (reciprocal of impedance). Thus making it design
oriented.

• Generalized Impedances: Both the source and load impedances can be


generalized. That is, a family of impedances created by uncertainty or
variations in operating conditions are enclosed by a shape in the w-plane
impedance and admittance spaces. Stability assessment is carried out on
these shapes rather than individual impedance plots.

• Visually Interpretable Results: Determination of stability can be visu-


ally determined by noting the intersection of the generalized load admittance
with the source-stability criteria derived forbidden region.

Additionally the method has been implemented and released open source as
a Matlab toolbox. The DC Stability Toolbox is coordinated by Dr. Scott Sud-
hoff and available online [44]. This practical tool automates many aspects of the
stability assessment and graphics, requiring only that impedance data be provided.

2.3 AC bus Stability


As stated in section 1.4, power electronics play an important role in AC microgrids
but present the same potential instability problems. In [20] DC minor loop gain
stability methods are adapted to create analogous methods for AC systems.

2.3.1 Generalized Nyquist


Analysis of the DC power electronics-based system results in a scalar transfer func-
tion, section 2.2.1, that is analyzed using the Nyquist criterion. In an equivalent
three phase system, the resulting transfer function is multiple input multiple out
(MIMO) due to the transformation of variables [20]. The transformation of voltage
and current to the synchronous reference frame, which will be discussed in section
3.3.1, leads to a matrix transfer function with matrix impedances describing the
filter-CPL scenario in Figure 2.4 part a).

32
" #−1
Zdqs
vdq = I+ Hdq vdq
Zdql
The transfer function is analogous to that derived in section 2.2.1, except the
terms are now 2x2 matrices. Again the numerator transfer function Hdq is consid-
ered a priori stable if the source is stable by itself, which will always be the case for
a RLC filter no matter how complicated [43]. Stability determination is likewise
dependent on the the matrix transfer function minor loop gain.

[I + Ldq (s)]

The generalized Nyquist criterion (GNC) was developed by MacFarlane and


Postlethwaite to allow stability assessment of matrix transfer functions [20, 45]. A
full description of the mathematics involved can be found in [45].The end result is
similar to the SISO case except the Nyquist plot is traced by the minor loop gain
denominator eigenvalues. To ensure stability the Nyquist plots, there will be two
in this case, cannot encircle the origin.

Nyquist Plots = eig[I + Ldq (s)] for s = Nyquist contour

Or more simply, stability is ensured if the critical point −1 + j0 is not encircled


by the Nyquist plots of the return ratio matrix, Ldq (s).

2.3.2 Stability Criteria


As in the DC case, AC system stability can be determined directly through exam-
ination of the Nyquist plots. Such analysis is not conducive to design formulation,
therefore a number of stability criteria have been developed to remedy this. These
criteria can be categorized into two groups; those that consider only the dd values
of the impedance matrices, and those that limit some norm of the impedance ma-
trices.

D Channel Criterion [46]: In the case that the three phase system has high
power factor, stability can be determined exclusively by the dd channel impedances
values; Z(s)sdd and Z(s)ldd . The D channel can be constructed and analyzed using
the DC stability criteria, section 2.2.3. This forms a powerful duality with the well

33
developed area of DC stability criteria. Although it must be noted that translation
reduces the SISO Nyquist results to being sufficient but no longer necessary for
stability [47]. That is, there is a built in conservatism that is above and beyond
the margins already present in the DC criteria.

Impedance Norm Criteria: These methods are analogous to the Middle-


q 
brook criteria [35] which constrains the 2-norm real(L)2 + imag(L)2 of the
return ratio, thus creating the unit circle as an allowed space in the w-plane, see
Figure 2.7. In the AC system the appropriate norm to apply is not obvious [20].
This has led to opportunities for researchers to develop, apply, and compare the
effectiveness of norm based criteria for AC systems [20, 47].

2.4 Proposed Stability Assessment Method


A meld of the presented techniques and theory are adapted to accomplish the
dissertation objective, section 1.8. The procedure of the developed approached
can be summarized as follows.

1. Collect impedance data about components to build an off-line database

2. Make impedances of interacting components compatible by interpolating


data to match frequency and operating points.

3. Construct families of Nyquist plots from matched return ratios.

4. Check qualitatively for stability using Nyquist stability criterion (or GNC
for AC systems)

5. Apply maximum peak criteria (MPC) definition of stability margin to quan-


titatively measure robustness

6. Report system’s unstable operating points

2.4.1 Minimizing Conservatism


In the developed approach artificial conservatism is practically eliminated because
no stability margin is actually required. The system (or subset of operating condi-
tions) is determined stable or unstable using the Nyquist criterion as a necessary

34
and sufficient condition. This is done similarly to the method proposed in [40],
except additional conservatism is removed because conditional stability is allowed.
This is accomplished using a simple algorithm that tracks zero crossings of the
boundary (−∞ + 0j, −1 + 0j), Figure 2.8.

pn = positive to negative zero crossing of real line < −1 + 0j

np = negative to positive zero crossing of real line < −1 + 0j

X X
N= pn − np

Nyquist criterion satisfied if N ≥ 0

Figure 2.8: Assessing Stability


(a) Minimal Forbidden Region (b) Consecutive Crossing Restriction

This is applied to the simple DC LTI circuit in Figure 2.9, an example borrowed
from a Virginia Tech power electronics stability tutorial [48]. Although the system
a) is stable, as shown in time domain study, Figure 2.10, the resulting Nyquist plot
fails all the stability criteria listed in section 2.2.3. Application of the simple zero
crossing algorithm to the Nyquist plots correctly predicts the stability, figures 2.11
and 2.12

35
Figure 2.9: DC LTI Circuit Diagram

Figure 2.10: DC LTI Example Circuits Time Domain

(a) Stable: R = −0.1Ω (b) Unstable: R = −0.2Ω

36
Figure 2.11: DC LTI example Stabe Nyquist Plot N = +1

(a) (b) Close Up

Figure 2.12: DC LTI example Unstabe Nyquist Plot N = −1

(a) (b) Close Up

2.4.2 Stability Margins


Techniques presented so far, including the stability criteria in section 2.2.3 and
modified PBSC in section 2.2.4, either intrinsically contain conservatism or have
stability margin specified as a constraint. The exception being the admittance
space analysis in section 2.2.5 that allows for a visual determination of stability
margin. In the proposed method, margin for a given combination of components
and set of operating conditions is based on the minimum distance between the

37
return ratio Nyquist plot and the critical point, Figure 2.13. This is based on the
MPC and work in [40].

Figure 2.13: Stability Margins

(a) Stable System (b) Unstable System

Note: that stability margin does not indicate stability. In fact as evidenced
in Figure 2.13 part b), a highly unstable system can have a margin better than a
stable one.

2.4.3 Interfaces
Describing the stability margin and assessing stability can be complicated by the
choice of measurement interfaces. The problem lies in the sensitivity of the return
ratio to component groupings. This can be demonstrated for the stable LTI DC
circuit in Figure 2.9. The circuit is simple so the return ratio can be analytically
determined and plotted for interfaces both pre and post capacitor. Let interface 1
be post-capacitor and interface 2 be pre-capacitor, then the return ratios will be
as below.

Zs1 Rs + sLs
L1 = =
Zl1 (1 + sRs C + s2 Ls C)(Rl + sLl )

Zs2 (Rs + sLs )(1 + sRl C + s2 Ll C)


L2 = =
Zl2 Rl + sLl

As can be seen, the return ratio transfer functions differ significantly. The re-
turn ratios can be compared to each other in the w-plane, see Figure 2.14. Both

38
return ratio plots are of the same system, so despite their differences they describe
a system with the exact same stability margin.

Figure 2.14: Interface Choice and Nyquist Plots

This effect has two important implications. First, stability criteria with con-
servatism may fail at some interfaces even if the system is stable. Ostensibly this
is solved by considering a system to be stable if a stability criteria is met at any
interface. However, as described in section 2.1.2, the system states can only be de-
termined stable if the transfer function is a minimal realization of the system [20].
This is generally not the case and even if it were, it requires observability and
controllability to be confirmed, which is difficult to do using empirically obtained
impedance data. It does guarantee that interfaces with a common voltage can be
determined stable if either interfaces meets the stability criteria. Therefore inter-
faces separated only by a capacitor will be stable if either interface meets the given
stability criteria. Interfaces separated by other components, such as inductors or
resistors, will need to be checked if they are not part of a purely passive network.

The second implication is that stability margins cannot be used to describe the

39
system in an absolute way. Different interfaces result in very different values of
stability margin, making a quantitative comparison of different systems difficult.
Development of a new stability margin definitions to solve this issue remains an
open problem [40].

In this work the problem is addressed by checking interfaces as close to the


power electronics component as possible. This makes capacitors part of the source
filter, which will always be stable as a passive RLC network [43]. Also comparison
of stability margins is restricted to like interfaces and between like components.

2.4.4 Operating Points


The operating and loading conditions of power electronics will change their impedances,
and therefore their stability characteristics. This adds significantly to the compu-
tational burden of stability assessment as each set of conditions must be considered.
Generalized admittance space stability analysis [39] addresses this but requires the
combinations of all possible conditions to be stable. This may be overly stringent
as the system integrator can limit the operating points or even specific combina-
tions of operating points the system undergoes. This presents an opportunity for
systems with a rare or preventable instability to be identified and stabilized.

Motivating Example: A Military Microgrid EMS [7]


Consider the remote microgrid concept in [7], Figure 2.15. An energy man-
agement system (EMS) seeks to improve the fuel efficiency of a small kW class
microgrid. The EMS interfaces battery energy storage with the single phase AC
bus through a bidirectional buck-boost inverter. The AC bus is directly connected
to a set of autonomous generators. By forcing the battery to inject or absorb power,
the EMS can more efficiently load the generators and utilize smaller generators.
This can greatly increase fuel efficiency; a 50 % reduction of fuel consumption
predicted in [7].

40
Figure 2.15: Military EMS Microgrid Concept

In this remote microgrid system the bidirectional buck-boost inverter must


provide substantial power and therefore introduces the power electronics stability
problem. Generators are expected to be commercial-off-the-shelf (COTS) and are
regulated only by a thyristor switch at the interface (they are on-off devices). If
a system were found (or predicted) to be unstable at certain component loadings,
those conditions could simply be avoided by the EMS in the same way it avoids
inefficient conditions.

In this dissertation the concept of operating point "knock-out" is proposed as


a way for higher level optimization to consider systems that have an instability. If
a highly desirable system is identified, or an existing system needs to be adapted,
the EMS could avoid operating steady state in those unstable conditions. This
can be achieved by pinging between two stable loadings that produce the required
amount of power on average. By passing through the unstable point sufficiently

41
fast, the negative effects may be avoided. This would be an additional advantage
to adopting an intelligent EMS and to making microgrids more adaptable to future
components and situations that are not now anticipated.

42
Chapter 3 |
Simulated Impedance Measure-
ment

In chapter 2 it was shown that armed with detailed component impedance infor-
mation, the inter-component compatibility of power electronics can be determined.
In this chapter the details of impedance extraction for DC and AC systems are
covered. An impedance measurement tool built in Matlab’s SimPowerSystems
is presented. It is both a simulation of practical impedance measurement and a
practical tool for determining the operating point dependent impedances of devices
built in simulink. It determines the impedances of arbitrarily (subject to compu-
tational limitations) complex AC and DC power electronic converters. Tests and
verification of the simulated impedance measurement blocks are provided.

3.1 Nyquist Contour and Impedance


In section 2.2.2 it was shown that the Nyquist contour allows RHP poles in the
minor loop gain to be found. Construction of this contour implies that the return
ratio be measured for complex frequencies of nonzero real and imaginary parts.
That is ...

Z(s)s
L(s) = for s = σ + jω swept through Nyquist Contour
Z(s)l
Impedance is only considered in the context of frequency (jω) since practical
measurements have to be taken under steady state conditions, meaning σ = 0 [49].
This makes it impossible to construct the Nyquist contour in the s-plane as pre-

43
scribed by the Nyquist stability criterion, section 2.2.2. Fortunately we can exploit
the properties of a strictly proper transfer function.

Strictly Proper [34] : If a transfer function H(s) is rational it can be ex-


pressed in terms of a numerator and denominator H(s) = N (s)
D(s)
. If the degree of the
denominator is greater than the degree of the numerator than it is termed strictly
proper. A strictly proper transfer function approaches zero as the magnitude of s
approaches infinity.

deg(D(s)) > deg(N (s)) then H(∞) = 0

This property is very useful as it guarantees that the return ratio Nyquist
plot will close through zero if it is a strictly proper transfer function. This can be
ensured by selecting appropriate interfaces to measure. Although the impedance
transfer functions will be unknown, grouping capacitive elements and filters with
the source will increase the return ratios denominator order, section 2.4.3. With
the assumption that the return ratio is strictly proper, impedance measurements to
an arbitrarily high frequency will close the Nyquist plot through zero. Thus effect
of the positive real part sections of the Nyquist contour becomes unimportant,
Figure 3.1.

Z(jω)s
L(s) for s = Nyquist contour ≈ L(jω) = for ω = (−∞, +∞)
Z(jω)l

This section provides justification for applying impedances to the Nyquist sta-
bility criteria as well as further encouragement to choose measurement interfaces
wisely.

44
Figure 3.1: Reduced Nyquist Contour

3.2 DC Systems
Impedance can be defined using generalized Ohm’s law.

V = IZ

It is the effective resistance of a circuit to an alternating current or voltage. Un-


like resistance it contains real and imaginary parts, so it can relate the magnitude
and phase of the voltage waveform to a current waveform. Impedance describes
a linear electrical network and is a function of frequency. Power electronic con-
verters are highly non linear, section 2.1.1, so impedance is valid for the device
only for small signals and for a specific set of operating conditions. A power elec-
tronics component must therefore be tested under a multitude of frequencies and
conditions to be sufficiently characterized to assess stability.

Ẑ(ω, xop )

The choice of variables to include in xop and the ranges to measure ω and xop
are dependent on the system characteristics and to some degree the investigator’s
judgment.

Ideally impedance information would be available over the full frequency range,

45
Figure 3.2: Impedance Fitting Results
(a) Raw Data (b) Fitted Data

ω = (−∞, +∞). This can be accomplished by measuring impedances at certain


frequencies and fitting the data to a transfer function. This was successfully uti-
lized to assess the stability of a simple LTI circuit using only low fidelity simulated
measurements, Figure 3.2. Matlab code to automatically process the impedance
data is presented in Appendix B.1. Unfortunately the transfer function fitting pro-
cess requires knowledge of numerator and denominator orders. A major aspect of
the dissertation’s goals is application to "back box" components. This complicates
the use of the developed fitting technique since the transfer function structure is
assumed unknown. Although future work may allow for a program to make esti-
mates about impedance transfer function orders.

Measurement of impedance over the frequency range of ω = (−∞, +∞) is ob-


viously not practically possible. We can immediately remove the requirement for
negative frequencies as the return ratio’s Nyquist plot is mirrored across the real
axis. Also measurements only need to be conducted up to some sufficiently high
frequency. As stated previously this frequency, ωmax must be high enough to allow
the return ratio Nyquist plot to close [50]. But it must also be significantly higher
than frequencies connected power electronics controls operate at. The negative in-
cremental impedance instability problem is associated with the power electronics
control, not the power electronics components themselves. Outside control fre-
quencies, power electronics are passive and small signal stability is not a problem.

46
It may be enticing to simply select an extremely large value for ω, but there
are factors in opposition to this.

• Capability: Measuring the impedance at higher frequencies will become


difficult. Via simulation, as is done in this work, higher frequency measure-
ments become computationally intensive, requiring ever larger amounts of
time to simulate. In physical implementations, impedance measurement re-
quires a high power electrical perturbation. Higher frequency perturbations
will require higher frequency switching equipment and higher frequency sen-
sors, resulting in an escalation of costs as the measurement unit is designed
to accommodate higher frequencies.

• Switching: Both the measurement of and meaning of impedances becomes


questionable at frequencies approaching and exceeding that of the device
switching frequency. This occurs because of the breakdown in the assump-
tion of an averaged power electronics model, section 2.1.1. A discrepancy
between analytically derived impedances based on the averaged assumption
and physical measurements has been noted in the literature [51].

Recommendations can therefore be supplied for impedance measurement of


power electronic converter loads.

1
ωs > ωmax >
τ
where τ is fastest connected controller’s time constant and ωs is the device’s
switching frequency. Note this requirement is soft for "source" type power elec-
tronic components because their stability characteristics are assumed to based on
their output filter.

The selection of operating test points X̂op and ranges is even more nuanced
than selection of ω. In this work operating points are separated into two categories;
general and specific.

• General: These are uncontrollable or uncertain variables that affect impedance.


They can be totally exogenous variables or just variables that are generalized

47
for convenience. Temperature, age, and manufacturing variability are excel-
lent examples of exogenous variables to be generalized since they effect the
impedance but can’t be easily controlled. In generalized admittance space
stability analysis, section 2.2.5, component loading ranges are also general-
ized [39].

• Specific: A proposed designation for operating variables that can be re-


stricted by the design of a microgrid or controlled by the EMS. A natural
choice for this would be the loadings of dispatchable power electronics de-
vices.

The range covered by each variable, whether general or specific, should be repre-
sentative of the system’s steady operating conditions. Some engineering judgment
is required to differentiate between transient operating points and quasi-steady
ones.

3.2.1 Current Perturbation Injection


Measurement of impedance is accomplished using perturb and observe methods. A
sinusoidal voltage or current is injected into the interface of the system to disturb
it. The current and voltage waveforms at the perturbation frequency are extracted
and the impedance is calculated using generalized Ohm’s law.

V(jω)
Z(jω) =
I(jω)
Where ω is the perturbation frequency and V and I are the resulting phasors
measured at that frequency.

In this work shunt (parallel) current injection is chosen as the means to disturb
the system, Figure 3.3, series voltage injection being the alternative. In practical
implementations voltage injection can allow simpler sensors while current injec-
tion requires much less equipment to create the disturbance [50]. In this work
measurement is via simulation so those attributes are irrelevant. Current injection
is chosen because it convenient for testing loads, which most components in the
microgrid will be.

48
Figure 3.3: DC System Current Injection

3.2.2 DC Impedance Measurement Block


To achieve the dissertation’s goals, section 1.8, a Matlab tool was developed to
determine the stability characteristics of arbitrary power electronic devices. The
electrical network techniques studied, Nyquist based and passivity based, are de-
pendent on determining the small signal impedance of devices at quasi-steady
operating points. This was accomplished using a combination of Matlab script
and custom simulink blocks. The procedure is automated and is described below.

1. Set component operating conditions

2. Set injection frequency

3. Inject current disturbance at interface

4. Continuously measure voltage and current

5. Continuously extract voltage and current phasors at perturbation frequency


using Fourier block

6. Terminate simulation when voltage and current phasors are within bounds
for n cycles

7. Record impedance and save to data file with operating point information

49
Figure 3.4: DC Impedance Measurement Block

8. Repeat process for all frequencies and operating condition combinations

The block was made to be user friendly, allowing custom source and load cir-
cuits to be easily connected, Figure 3.4. The block can also run for a single set of
operating conditions and frequency to troubleshoot problems or optimize speed.
Diagnostic results are automatically generated and opening the block will yield a
number of useful plots and measurement values, figures 3.5 and 3.6.

Although alternative methods are available to determine the impedance of de-


signed components, the presented tool is by comparison adaptable and simple to
use. Matlab and other commercial simulation programs have the capability to
automatically calculate the impedance of implemented circuits, but they must be
linear and averaged. This makes them unsuitable for determining the impedance
of power electronics and other nonlinear components. This is not an issue for the
presented DC impedance measurement block, which treats the source and load as
"black boxes", only measuring voltage and current at the interfaces. The connected
device can be arbitrarily complex with switching elements and controls included,
allowing a Matlab simulink user freedom to rapidly create and test devices via sim-
ulation. This adaptability does however come at a premium computational cost
since the circuit needs to be fully simulated in the time domain, complicated or fast
switching devices can require significant computing time to complete. It should be
noted that in a physical implementation of the impedance measurement technique
this issue is alleviated since transients will settle quickly and measurement time
will depend only on perturbation frequency.

50
Figure 3.5: Frequency Sweep Diagnostics

Figure 3.6: Inside DC IM Block

51
Figure 3.7: Simple LTI Circuit Test Diagrams
(a) Circuit (b) Simulink Implementation

3.2.3 Results and Verification


The DC impedance measurement block was applied to a number of test circuits
from the literature and the results compared to verify its efficacy. The first test is
of the simple LTI circuit shown in Figure 3.7. The circuit is simple enough that
the source and load impedance transfer functions can be easily found analytically.

Zl (jω) = Rl + jωLl
1 Rs + jωLs
Zl (jω) = (Rs + jωLs )k →
jωC jωRs C + 1 − ω 2 Ls C
Plots of the source and load impedances for the analytical results and those
measured with the impedance measurement block is presented in Figure 3.8. The
parameter values are presented in Table 3.1.

Table 3.1: Simple DC LTI Parameter Values

Parameter Value
Vs 100 V
Rs 1Ω
Ls 5mH
C 200 µF
Rl −0.1Ω
Ll 2mH

For a more challenging test with a nonlinear device, the impedance measure-
ment block was applied to the output of a six pulse diode bridge. In [51] the
analytical impedance of a filtered diode bridge is presented. The diode bridge
from this work is modified to create two versions, a "source" type and "load" type

52
Figure 3.8: Simple LTI Impedance Results

rectifier, Figure 3.9.


!
1
Zs = Rc + k(R∗ + jωL∗ )
jωC

3
R∗ = Lac ωs + R + 2Rac
π

L∗ = L + 2Lac

53
Figure 3.9: Diode Bridge Circuit Test Diagrams
(a) "source" type (b) "load" type

Figure 3.10: "Source" type Diode Bridge Impedance Results

3
Zl = Lac ωs + 2Rac + j2ωLac
π
The "source" type has a capacitive output filter, making it ideally paired with
a constant current load. The "load" type is stripped of this output filter making it
ideally paired with a constant voltage source. The results can be found in figures
3.10 and 3.11, and the parameters used are presented in Table 3.2.

Good agreement is obtained between the impedance plots produced by the


impedance measurement block and the analytically derived plots for both a simple
LTI system and a nonlinear switching system. This verification is important as
later circuits will be more complex and analytical results will not be available to

54
Figure 3.11: "Load" type Diode Bridge Impedance Results

Table 3.2: Diode Bridge Parameter Values

Parameter Value
Vrms 120 V
Rac 30mΩ
Lac 0.22mH
L 0.9 mH
R 60mΩ
Rc 30mΩ
C 1.05 mF

confirm the validity of measurements.

3.3 AC System Impedances


With some modifications, AC systems can be analyzed in a similar way to DC
systems. The stability assessment methods presented so far required the system to
be linear and time invariant (LTI). Linearization of an AC system is problematic as
variables of interest, voltage and current, are not stationary in steady conditions.
The solution is to transform the system variables into a reference frame where they
are stationary. This is accomplished by transforming the per phase variables (abc)

55
to the synchronous reference frame using the dq0 transformation.

3.3.1 The dq0 Synchronous Reference Frame


The dq0 transformation is the combination of two reference frame rotations. The
first is the Clarke tranformation, sometimes referred to as αβ transformation.
 
s 1 − 12 − 21 
2 √ √
3
TClarke =  0 − 23 
 
3 1 2 
√ √1 √1
2 2 2

Under the conditions that the three phase system is balanced, this transforma-
tions results in the three time varying quantities (abc) being reduced to two (αβ)
and one stationary zero (γ = 0). The two time varying quantities can then be
made stationary by rotating the system about γ with the fundamental frequency
(typically 60Hz for terrestrial domestic systems).
 

cos(θ) sin(θ) 0 
Trotate =  − sin(θ) cos(θ) 0 
 
 
0 0 1
Z τ
θ= ωdt , where ω is the fundamental frequency
0

Now a steady AC system is described by stationary variables of voltage and


current. Note that in the case of the three phase system being balanced, the
third variable in the Clarke transformation is always zero and can be dropped.
This results in the dq0 transformation producing only two variables of interest,
the direct (d) and in-quadrature (q). The transformation process is illustrated in
Figure 3.12.

3.3.2 DQ Impedance Matrix


The generalized Ohm’s law, along with perturb and observe methods, are also
used in AC systems. However after dq0 transformations, current and voltage are
no longer scalars, they are 2x2 matrices, resulting in a 2x2 impedance matrix.

56
Figure 3.12: dq0 Transformation of Three Phase Voltage
(a) abc (b) αβ (c) direct quadrature

A single measurement is no longer sufficient to determine the impedance. Two


independent perturbations and measurements must be made [52].
    −1
Zdd Zdq   Vd1 Vd2   Id1 Id2 
Zdq = =
Zqd Zqq Vq1 Vq2 Iq1 Iq2
As in the DC case, the system is perturbed by injecting voltage or current
at an interface, but each injected disturbance must be separate and independent.
This can be accomplished by injecting a wave on the d axis and q axis separately,
although any independent linear combination will suffice. The voltage and current
measurement matrices are used to calculate the impedance matrix, each has two
columns. The first column results from the first perturbation, usually the d axis
injection for simplicity, and the second column results from the q axis perturbation.

3.3.3 AC Impedance Measurement Block


A counterpart to the DC impedance measurement block was created for AC sys-
tems. It operates on the same principles but is structured differently. Again shunt
current injection is the method chosen to disturb the system. This strategy is
considered the most practical for physical implementation, with numerous circuits
developed in the literature to do so [53,54]. Note that the developed AC impedance
measurement block is amenable to testing such techniques in simulation by simply
switching out the simulink perturbation block with a user designed perturbation
source, figures 3.13 and 3.14.

57
Figure 3.13: Injector Block

Figure 3.14: AC Current Perturbation

The AC impedance block follows the same procedure as the DC block, section
3.2.2, except with two key differences.

• Synchronization: To inject perturbations, a reference angle needs to be


measured and supplied to transform the d and q axis waves to the natural
abc reference frame.

• Simultaneous Simulations: The two independent perturbations are car-


ried out simultaneously in parallel simulations. This greatly increases the
speed by cutting the number of model initializations in half.

Running simulations in parallel affects performance but synchronization, more


critically, affects accuracy. In three phase electrical systems the reference angle is
the system’s synchronous frequency integrated with the d axis aligned 90 degrees

58
behind the A phase (by convention). In this work we determine the reference angle
θ using either a phase locked loop (PLL) or by assuming a proportional relationship
with the more easily measured rotor reference frame.

• PLL: A phase locked loop uses a low bandwidth PID controller to synchro-
nize the phase of an exogenous signal with an internal reference. This internal
reference is supplied to transform abc measurements into the dq frame and
dq injection commands into the abc frame. PLLs have been used in AC
impedance measurement systems [52], but they can negatively interact with
the disturbance, leading to errors. These errors can cause significant inaccu-
racy in the resulting impedance measurement [55]. The solution to this is to
ensure the bandwidth of the PLL is much lower than the lowest frequency
being measured. This however increases the time needed to synchronize and
reduces the PLL’s ability to follow changes.

• Rotor Reference: In this work, as in [20], the rotor reference frame and syn-
chronous reference frame of synchronous generators is assumed to be rigidly
coupled, at least in quasi-steady circumstances. This allows the rotor posi-
tion to be used as an alternative source for synchronous reference angle. In
practice this might be avoided due to the extra effort of mechanical mea-
surement, but in simulation this is not a problem. It obviates the need for a
PLL, removing the associated error and computational expense.

Matlab’s SimPowerSystems toolbox provides a PLL optimized for three phase


60 Hz systems [56]. It is used to supply the reference angle for interfaces with a
"stiff" grid connection. For scenarios where the frequency varies dynamically, such
as a motor or wind turbine generator, the rotor reference frame is used instead.

3.3.4 Results and Verification


The AC impedance measurement block was also verified against analytical results
of a linear and nonlinear circuit. The source and load impedance of a simple LTI
circuit, Figure 3.15, were found using the block and compared against the ana-
lytical result. Since this is an AC system, the impedances are described by a 2x2
matrix, resulting in a set of four plots for impedance magnitude and set of four

59
Figure 3.15: Simple AC LTI Circuit Diagram

plots for phase angle, see figures 3.16 and 3.17. There is excellent agreement be-
tween the plots.

The input impedance of the six pulse diode bridge in [51] was also found via
simulation, Figure 3.18. An analytical solution for the impedance using an aver-
aged model, although proven to be incorrect a higher frequencies, was available
for reference in [51]. The results of the AC impedance measurement block can be
found in figures 3.19 and 3.20.

60
Figure 3.16: Simple AC LTI Circuit Impedance Magnitudes

Figure 3.17: Simple AC LTI Circuit Impedance Angles

61
Figure 3.18: Six Pulse Diode Bridge Diagram

Figure 3.19: Six Pulse Diode Bridge Input Impedance Magnitudes

62
Figure 3.20: Six Pulse Diode Bridge Input Impedance Angles

63
Chapter 4 |
Notional Microgrid

A small scale notional microgrid containing a wind turbine, battery, and inverter
is developed and presented as a subject for stability investigation. It was modeled
in detail down to the switching level elements and inner current control loops
in Matlab and simulink. Alternate rectifiers and inverters are implemented for
comparison. The performance of the system is simulated along with the impedance
measurement blocks from chapter 3. The impedance characteristics of each device
are systematically determined for a range of conditions and the data stored in files
for easy assessment by stability routines later.

4.1 Architecture
In chapters 2 and 3 tools and methods to analyze the stability of a microgrid were
developed. In this chapter a DC power electronics-based microgrid is developed
for the demonstration in Chapter 5.

In this work microgrid architecture refers not just to physical layout but also
the selection of nominal interface variables and dispatch strategy. The architec-
ture of a microgrid can take many forms, and design is based on a number of
factors. The literature is rich with design strategies [57, 58] and control methods
to automate and regulate DC microgrids [11, 59–63] and AC microgrids [64, 65]. A
general hybrid microgrid architecture, Figure 4.1, is often assumed in higher level
optimization [24]. Adoption of true hybrid microgrids like this for non-vehicle ap-
plications has been slow [12, 13], due at least in some part to the ubiquity of AC
systems and lack of mature commercial DC protection devices [17, 66]. Microgrids

64
Figure 4.1: General Hybrid Microgrid

with high penetrations of renewables and power electronics will naturally contain
many DC buses, Figure 4.2. With proper control, the voltages of the DC buses
vary proportionally with each other, effectively operating like Figure 4.1 from a
high level perspective. The distinction is however important for assessing power
electronics stability. To ease stability analysis the grid can be broken into smaller
parts, in this chapter the analysis of a single DC leg will be presented.

There is great flexibility in how a microgrid can be laid out, however to be


viable all connecting buses must have some degree of regulation. From this fact
some general definitions and rules were developed to ease the creation of microgrid
layouts.

For architecture design purposes, microgrid components can be classified into


one of four categories depending on how they interact with a connected bus, Ta-
ble 4.1. These categories are applied to some simple networks for clarity. For a
DC "leg" of a microgrid, Figure 4.3, two ways of incorporating a renewable energy
source (such as solar or wind) are shown. In both cases the renewable energy
source is intermittent and injects power as it becomes available, making it a "hard
grid taker". To "make" the bus there is either a battery or capacitor. For short
time scales and nominal conditions, the battery regulates the bus to a constant
voltage, thus it is classified as a "hard grid maker". Whether added intention-

65
Figure 4.2: Power Electronics-based Microgrid

ally or not, any DC bus will have at least some capacitance associated with it.
This naturally changes the voltage, a variable locally available to all connected
devices, in response to a net power imbalance, therefore making it a "soft grid
maker". The inverter can either inject power according to an exogenous signal,
"hard grid taker", or inject power in response to the bus voltage, "soft grid taker".
These designations are not absolute and depend on perspective, this is illustrated
in Figure 4.4 by applying to the conventional grid and droop control. The des-
ignation even applies to mechanical components such as the governor and prime
mover. The governor acts as a "soft grid taker" changing power in response to
local line variables (speed/system frequency). The inertia of the prime mover and
synchronous generator act as "soft grid maker" because it translates a mismatch of
power into a locally measurable bus variable, system frequency. When many units
are in parallel the energy stored in the rotating inertia becomes very large and the
frequency of the AC bus is regulated very tightly, making the collection of sources
a "hard grid maker". While these cases are simple they illustrate how the categoriz-
ing of microgrid components can be useful in creating a more complicated network.

66
Figure 4.3: DC "leg" Architecture Examples

Table 4.1: Microgrid Component Types

Property "Maker" "Taker"


Changes bus variables Changes power flow
"Soft" in response to bus energy in response to bus variables

Regulates bus variables Exogenous signals


"Hard" by changing power flow controls power flow

In developing a microgrid architecture of arbitrary complexity two rules are


proposed to simplify the process.

• Every bus (common node) must connect at least one "maker" and "taker"
component.

• Every bus (common node) can only connect to one "hard grid maker" which
will negate the effects of any "soft grid makers".

67
Figure 4.4: Conventional Grid Architecture Example

4.2 Notional Microgrid Layout


The first scenario in Figure 4.3 is selected to be modeled and simulated in detail.
The renewable power source is a small wind turbine. This is a good compromise
between the possible alternatives; combustion generator and PV, as it incorporates
some of the qualities of each. A general schematic of the notional microgrid is illus-
trated in Figure 4.5. The system is a low voltage microgrid supplying a larger AC
microgrid either as a grid forming agent or as a dispatchable renewable generator
unit. Table 4.2 contains the nominal variables for the DC and (post inverter) AC
buses.

Table 4.2: Notional Microgrid Nominal Bus Values

Parameter Value
DC bus voltage 300 V
AC bus voltage 120 Vrms L-N
AC bus frequency 60 Hz

To ensure a convincing demonstration, detailed switching models of each com-

68
Figure 4.5: Notional Microgrid Layout

ponent (along with variants and alternatives) were designed and implemented in
Matlab simulink.

4.3 Wind Turbine


The wind turbine is the prime mover of the microgrid "leg". An aero-mechanical
model of the wind turbine was developed and implemented in simulink. Dynamics
of the system are dominated by the turbine’s large inertia and the variable aerody-
namic efficiency. In quasi steady conditions the wind power extracted is described
by the power coefficient Cp , which is dependent on the tip speed ratio λ.

1
PW = Cp (λ)ρAV 3
2
ωR
λ=
V
For modeling purposes it is convenient to fit the power coefficient and tip speed
ratio data to a third order polynomial. An example of this is presented in Figure
4.6. Each wind turbine will have unique aerodynamic characteristics and therefore
have a unique set of coefficients to describe the power coefficient function, Cp (λ).
The turbine parameters are presented in Table 4.3.

69
Figure 4.6: Example Cp vs TSR Curve

Table 4.3: Cp Coefficients

Parameter Value
Radius 1m
kg
Air Density ρ 1.2 m 3

Moment of Inertia J 0.5 kgm2


Viscous Friction D 0.001 Nrad ms

a3 -0.0022
a2 0.0207
a1 0.0098
a0 0.0124

Cp (λ) = a3 λ3 + a2 λ2 + a1 λ + a0

Applying conservation of angular momentum, an equation can be balanced


for the summed torques acting on the turbine. Rearranging this equation and
substituting it into the derivative of the square of angular speed leads to a robust
mechanical model of the turbine.

Te : Electromagnetic torque from motor

70
Figure 4.7: Inside Wind Turbine Simulink Block

TW : Torque from wind turbine

TD = Dω : viscous torque (friction)


X
T = J ω̇ = TW − Te − TD
1
ω̇ = [TW − Te − Dω]
J
dω 2
= 2ω ω̇
dt
dω 2 2ω 2
= [TW − Te − Dω] → [TW ω − Te ω − Dω 2 ]
dt J J
PW
PW = TW ω → TW =
ω
dω 2 2
= [PW (V, ω) − Te ω − Dω 2 ]
dt J
1
2
Z
2
ω= [PW (V, ω) − Te ω − Dω 2 ]
J
Using the above equation the wind turbine is modeled in Matlab simulink to
create a turbine block; with inputs of wind speed V and electromagnetic torque
Te , Figure 4.7.

An important element of the turbine is the electric motor or generator. Matlab

71
SimPowerSystems includes a library of motor models, each capable of receiving
rotor speed as an input and having electromagnetic torque as an output Te . A
wide range of machine preset models are available as well as the capability of
defining parameters, negating the need to create a custom model.

• PMSM: 16 preset permanent magnet synchronous machine models ranging


from nominal power ratings of 250 W to 40 kW

• Synchronous Machine: 24 preset fundamental machines ranging from


nominal power ratings of 8 kW to 2.5 MW

The two models are combined in an algebraic loop to form a combined wind
turbine and motor model that connects to the variable voltage and frequency three
phase bus, Figure 4.8.

Figure 4.8: Wind Turbine - Motor Structure


(a) Diagram (b) Simulink

4.4 Rectification
The rectifier fulfills two important functions for the wind turbine and motor.

• Rectification: Transforms the variable voltage, variable frequency three


phase AC power from the motor into a DC power.

• MPPT: Controls the amount of current to properly load the turbine for
given wind conditions

72
As shown previously in Figure 4.6, the aerodynamics of the blades result in
the turbine being very sensitive to the ratio of wind speed and rotation speed.
To produce the maximum power, the wind turbine would ideally be spinning at
the optimum speed for every given set of wind conditions. This process is called
maximum power point tracking (MPPT). It is accomplished by varying the motor
currents drawn to speed or slow the turbine to the optimal speed. A wealth of
MPPT algorithms are available in the literature [67–70]. For this work the torque
control MPPT from [70] is implemented.

T (ω) = Kopt ω 2
1 R5
Kopt = Cpmax ρπ 3
2 λopt
The speed of the MPPT algorithm is limited by the turbine inertia, which
is large. This is beneficial from a stability viewpoint as it won’t interact with
the relatively high speed power electronics controls that cause instability. Thus
allowing the MPPT control to be neglected in the small signal analysis. However
the means of implementing the MPPT torque control does effect the stability
characteristics. Three rectifiers are designed, each with a different power electronic
converter structure, but equally valid components for the microgrid, Figure 4.9.

Figure 4.9: Turbine-Motor Rectifier Architectures


(a) Diode Bridge (b) Boost Rectifier (c) Switching Bridge

4.4.1 Passive Diode Bridge


Small wind turbines are frequently passive systems using a six pulse diode bridge
to rectify the power leaving the motor, Figure 4.10. In these systems there is no
active MPPT, but careful design can mitigate this [71]. The lack of sensors and
switching components required leads to low cost and high reliability but comes at

73
the price of converter efficiency [71]. Other more sophisticated hardware exists,
but even then the diode bridge is still present with the additional active compo-
nents, Sections 4.4.2 and 4.4.3.

Since the passive diode bridge rectifier has no active controls, proper design is
critical to achieving acceptable overall efficiency of the system. This is primarily
accomplished through the selection of the DC bus voltage. Until the motor line-to-
line voltage exceeds that of the DC bus, the diodes will prevent any current (and
therefore electrical power) flowing from the motor. The produced motor voltage
amplitudes depend on the rotational speeds. A convenient expression describing
this uses a constant Cv (L-N volts per rad/s) that is specific to a given motor.


VLL = 3Cv ω

Until the turbine reaches the speed necessary to overcome the DC bus voltage,
the motor is essentially open circuited, the turbine free spinning. Once sufficient
speed is obtained, current begins to flow and the resulting electrical counter torque
resists further increases in speed. Under quasi steady conditions the wind speed
and turbine speed are therefore linked. TSR for a given speed and overall power
production of the system is therefore dependent on selection of the DC bus voltage.

ω = f (Vdc , Vwind )
ωR
λ= = f (Vdc , Vwind )
Vwind
1
PW = Cp (λ)ρAV 3 = f (Vdc , Vwind )
2
The selection of DC bus voltage will be subject to the availability of DC storage
options (battery string voltages) and the requirements of the post rectifier inverter.
A problem of competing specifications can be overcome with the addition of gear-
ing between the motor and turbine; making PW = f (GR , Vwind ).

The DC impedance measurement block, section 3.2.2, is connected to the mo-


tor and diode bridge in simulink. A controllable DC dummy voltage source is
also connected to allow setting operating points, Figure 4.11. Only two operating
variables can be independently specified, motor speed ω and DC bus voltage V .

74
Figure 4.10: Turbine Diode Rectifier Diagram

These operating points are varied over their expected operating ranges, Table 4.4.
The mechanical and electrical power are measured in simulation along with other
variables so that reduced order impedance functions can be made.

Figure 4.11: Diode Bridge Impedance Measurement Diagram

Once the impedance measurement runs are complete, a data file containing
impedances and operating points was analyzed for trends. Figure 4.12 displays
impedance plots for different operating point combinations and a pattern is ev-
ident. A parametric analysis, Figures 4.13 and 4.14, shows the dependence of
impedance on each of the input variables, ω and V . In Figure 4.13, each pathway
represents a set of impedances with constant voltage V and varying motor speed
ω. Red indicates high values, blue low. Figure 4.14 is the same with the roles of

75
Table 4.4: Diode Rectifier Operating Ranges

Parameter Min Max


Output Voltage 250 V 350 V
Motor Speed 240 rpm 3800 rpm
Power Mech 0W 4900 W
Power Elec 0W 4500 W

voltage and speed reversed. It is evident that impedance is not easily predicted
from speed or voltage set points alone.

Figure 4.12: Diode Bridge Impedances

Three dimensional interpolation between power, frequency, and impedance was


used to create impedance surfaces, Figures 4.15 and 4.16. Examination of the sur-
faces reveals them to be relatively smooth with power points very close. This
indicates power is a very good predictor of the diode bridge’s impedance. The

76
Figure 4.13: Diode Bridge Impedances with Constant Voltage

Figure 4.14: Diode Bridge Impedances with Constant Speed

high impedance for low power is due to the diode blocking, essentially creating an
open circuit when the bus voltage is higher than that produced by the motor at a
given speed.

The results presented demonstrate a method for determining factors important


to impedance. In more complex components with control, the variables dictating

77
Figure 4.15: Diode Mechanical Power Impedance Surface

Figure 4.16: Diode Electrical Power Impedance Surface

impedance are be more complicated.

4.4.2 Boost Converter


Diode rectification with a current controlled converter is the preferred method of
small wind MPPT in several sources [67, 68, 70]. It is a middle ground between
the passive diode rectifier and the full three phase switching bridge converter.

78
It allows MPPT control and increased efficiency but without as many switching
components, sensors, or computational requirements as a full switching bridge rec-
tifier [72]. Generally speaking, the converter does not necessarily need to be of
boost type, but in this case the battery voltage is higher than that produced by
the motor under typical operation. Voltage boosting alleviates this issue without
the need for mechanical gearing, a preferred choice in small scale systems.

A current controlled boost converter was implemented in Matlab simulink. It


combines motor, diode rectifier, filter, and boost converter into a single subsystem,
Figure 4.17. Note that impedance measurement is taken at the output of the
subsystem but other interfaces marked by the dotted lines could also be tested
to ensure stability within the subsystem. The analysis would be identical to the
procedures presented but at a different interface. Details of the boost converter
and circuit elements are presented in Figure 4.18 and Table 4.5.

Figure 4.17: Motor Rectifier Boost Subsystem

The control structure is summarized in Figure 4.19. Torque commands are


periodically issued by the MPPT control and translated into a current set point
using PI control. Hysteresis control was applied to the boost converter switch to
develop an average current that approximately matches the set point.

79
Figure 4.18: Boost Converter Circuit

Table 4.5: Motor Rectifier Boost Parameters

Parameter Value
Filter C 200 µF
Boost L 100 µ H
Relay Tol ± 0.25 A
MPPT kopt 0.004757
Current Control kp 10

Relay operation (hysteresis control):

I < I ∗ − Itol : close switch

I > I ∗ + Itol : open switch

Where I is the measured current at filter output, Figure 4.18.

Simulation of the switching model over long time scales is impractical, requiring
about 5 minutes of calculation per second of simulated time. To test the MPPT,
the boost converter is replaced by a controllable ideal current source. Time do-
main simulations were then carried out, Figure 4.20 illustrates the control system

80
Figure 4.19: Motor Rectifier Boost Control Flow

effectively administering MPPT for a wind turbine start up scenario with constant
wind speed (7 m/s). Figure 4.21, illustrates the system following changing wind
speeds. The switching model was simulated over a short time period (about 1
second) to confirm the current control was working as intended, Figure 4.22.

Figure 4.20: Turbine Start Up with Boost MPPT

81
Figure 4.21: Boost MPPT and Variable Wind

Figure 4.22: Motor-Rectifier-Boost Time Domain

82
Impedance measurement at the boost converter output is conducted as in Fig-
ure 4.17. The wind turbine model is excluded because wind is a poor choice for
an input variable given the decoupling between turbine speed and wind speed.
Motor speed ω and current command I ∗ are selected as operating point variables.
Sweeps are conducted over nominal operating point ranges, Table 4.6, using the
DC impedance measurement block. As in section 4.4.1, impedance is characterized
as a function of electrical power by aggregating results together and creating an
impedance surface using multidimensional interpolation. Results are presented in
Figure 4.23 and 4.24. Note high (infinite) impedance values are rounded down to
make the plots easier to read.

Table 4.6: Boost Rectifier Operating Ranges

Parameter Min Max


Output Current 0A 15 A
Motor Speed 1000 rpm 3800 rpm
Power Elec 0W 5000 W

Figure 4.23: Boost MPPT Rectifier Impedance Magnitude

83
Figure 4.24: Boost MPPT Rectifier Impedance Angles

4.4.3 Switching Bridge


This type of wind turbine rectifier places power electronic switches in anti-parallel
with the diodes of the six pulse bridge from section 4.4.1. Many variants exists [30],
but for this work a general architecture based on ideal switches is assumed, Figure
4.25.

The popular technique of motor torque loading through q-axis current com-
mands is adopted to apply the MPPT [73–75].

3 3
Te = (Ld − Lq )id iq + λm iq
2 2
λm = λP where P = motor pole paris and λ = flux linkage

Ld = Lq for a non-salient motor


2Te
i∗q =
3λm
An advantage of the switching bridge is tighter control over the motor loading.
The direct axis current does not provide additional torque in an ideal non-salient

84
Figure 4.25: Switching Bridge Rectifier

motor. By eliminating the d axis current, the motor can produce a maximum
torque for a given total current magnitude purely through the q axis current.
To control the switches a pulse width modulated (PWM) PI current controller
with decoupling terms is designed and implemented in Matlab simulink. A flow
diagram of the control is presented in Figure 4.26. Rotor angle is used to establish
a synchronous reference frame to transform the abc variables into the dq frame. In
the dq frame a permanent magnet synchronous generator (PMSG) can be modeled
using the following equations [73, 75].

did
Ld = −Ris + Lq ωr iq + Vd
dt
diq
Lq = −Riq − Ld ωr id − λm ωr + Vq
dt
A full derivation of the PMSG equations along with PI controller design equa-
tions are presented in Appendix C.3. The parameter values and resulting controller
gains are summarized in Table 4.7.

Time domain simulations with the turbine connected are carried out to confirm
proper MPPT operation. Figure 4.27 shows system response to start up with con-
stant wind speed of 7 m/s and Figure 4.28 shows MPPT operation during variable
wind speeds of 5 m/s to 12 m/s.

85
Figure 4.26: Switching Bridge Rectifier Control Flow

Table 4.7: Switching Bridge Rectifier Parameters

Parameter Value
Motor R 0.9585 Ω
Motor L 0.00525 H
Motor Poles 6
Motor λ 0.1218 V s
PWM fs 10 kHz
Control gain, kp 9.25E1
Control gain, ki 958E1

Impedance sweeps are conducted using the DC impedance measurement block


as shown in Figure 4.29. The operating point ranges tested are presented in Table
4.8. Using electrical power as a predictor, impedance surfaces are generated via
multidimensional interpolation, Figures 4.30 and 4.31.

86
Figure 4.27: Switching Bridge Rectifier: Startup

Table 4.8: Switching Bridge Rectifier Operating Ranges

Parameter Min Max


Output Current 0A 15 A
Motor Speed 240 rpm 3800 rpm
Power Mech 0 kW 7.4 kW
Power Elec 0 kW 6.5 kW

87
Figure 4.28: Switching Bridge Rectifier: Variable Wind

Figure 4.29: Switching Bridge Rectifier Diagram

88
Figure 4.30: Switching Bridge Rectifier Impedance Magnitudes

89
Figure 4.31: Switching Bridge Rectifier Impedance Angles

4.5 Battery
As discussed, section 4.1, every bus needs a grid "making" component. For the
DC bus, this component will regulate voltage. For this notional system a battery
is chosen as the regulator. Individual cells are assumed to be placed in series
and parallel to be nominally 300 V across terminals and capable of handling the
expected power requirements without an unacceptable rise or drop in voltage.
Batteries themselves are dynamic, changing with age, temperature, and charge [76].

90
Battery modeling is a sophisticated area of study, being dependent on a number
of complicated electrical, chemical, and mechanical factors [77].

• aging effects

• memory and stratification

• cycling and SOC

• mass transport

• electric double layer

• electric and magnetic effects

Over the timescales that power electronic converter instability occurs (fractions
of a second) many of those battery effects are irrelevant. A common modeling tech-
nique is to develop an equivalent circuit to represent the battery [78]. Given the
premise that impedance measurements will be in the 1 Hz or higher range, an
equivalent circuit model including electric double layer and high frequency electric
and magnetic effects is created, Figure 4.32. Battery parameters are chosen to
produce reasonable results, Table 4.9, they will later serve as centers for ranges
of battery values. Figure 4.33 shows the impedance plot of the battery nominal
values with real and reverse imaginary axes, as is convention in the literature,
for comparison. Note, the semicircular curve in the upper quadrant is due to the
charge resistance and double layer capacitance while the bottom is leg is due to
the series inductance (a high frequency effect).

Table 4.9: Battery Parameters

Parameter Value
R 100 mΩ
L 0.1 µH
RCT 0.5 Ω
CDL 1 mF

Impedance measurement can be accomplished in the same way as the other


components using the DC impedance measurement block. Results are presented in

91
Figure 4.32: Battery Equivalent Circuit

Figure 4.33: Battery Nominal Impedance Plot

Figure 4.34, showing the impedance curves for different sets of battery conditions.
In this case the model is a linear circuit, making loading irrelevant. Instead the
charge transfer resistance RCT and double layer capacitance CDL are varied to
mimic the uncertainty associated with the exogenous factors affecting batteries;
age, cycle, etc.

92
Figure 4.34: Battery Equivalent Circuit Impedances

4.6 Inverter
The inverter transfers power from the DC bus to the three phase 120/208 V AC bus.
In this notional microgrid all the inverters act as a power port from the perspective
of the DC bus. Injecting real and reactive power according to an exogenous signal,
therefore classifying them as "hard grid takers" for the DC bus. Two scenarios will
be considered for the AC side of the system, a "stiff" grid connection and a "weak"
grid connection.

• "Stiff" Grid Scenario: Represents the case where the connected AC grid is
tightly controlled and regulated. It could be a direct connection to the macro-
grid or an autonomous microgrid with another "hard grid maker" present.
In this scenario the inverter injects real and reactive power according to a
communicated set point, accepting whatever voltage conditions are at the
interface. This is a common mode for small inverters to run in, especially
if they are connected to a renewable source, as they export the maximum
power available at all times. The downside to this type of inverter is that it
only functions in the presence of an AC bus regulated by other entities.

• "Weak" Grid Scenario: Represents the case where the connected AC grid

93
is not regulated. An example would be an isolated microgrid that has is-
landed from the main grid. In this scenario the inverter seeks to maintain
nominal voltage and frequency by injecting an appropriate amount of power.
Therefore the inverter is a "hard grid maker" from the AC bus’ perspective.
This allows the renewable battery microgrid "leg" to operate an autonomous
AC microgrid either alone or in conjunction with the grid and other genera-
tors (with the addition of droop controls, Appendix A.2.2).

Three inverter alternatives for the notional microgrid are designed and imple-
mented in Matlab. Each is a two-level three phase voltage sourced converter but
with different control strategy. All enact their switching using pulse width mod-
ulation (PWM). Figure 4.35 shows the basic hardware structure of the inverters.
It contains six power electronic switches, typically IGBTs or MOSFETS. The DC
bus is divided with a neutral point for analysis purposes, Appendix C.1 for more
details.
Figure 4.35: Two-Level Three Phase Voltage Sourced Converter Diagram

4.6.1 "Stiff" grid P plus resonant Inverter


The first inverter implemented is unique because it’s control is carried out in the
αβ reference frame, with non constant variables in steady state. The inverter con-

94
sists of the two level three phase converter plus an output filter, Figure 4.37. The
inductive output filter is critical to the control, since a current control scheme is
implemented. Transformed into the αβ frame, the inverter power output is de-
scribed by the following equations [75].

αβ AC power:
3
P = [Vα iα + Vβ iβ ]
2
3
Q = [−Vα iβ + Vβ iα ]
2
Rearranged, current commands can be derived from real and reactive voltage and
power set points:
2 Vα 2 Vβ
i∗α = 2
P∗ + Q∗
2
3 Vα + Vβ 3 Vα2 + Vβ2
2 Vβ 2 Vα
i∗β = 2 2
P∗ − 2 2
Q∗
3 Vα + Vβ 3 Vα + Vβ
di
The inductor equation, L dt = ∆V , in the αβ reference frame [75]

diα VDC
L = −Riα + mα − Vα
dt 2

diβ VDC
L = −Riβ + mβ − Vβ
dt 2
A filter and current controller is designed in [75] using the αβ frame equations,
a diagram of the control is presented in Figure 4.36. Most of the components in this
work are self developed but this design is taken as is and implemented in Matlab
simulink, allowing the simulated impedance measurement blocks to be run on a
more realistic, externally developed design. The parameter values are provided in
Table 4.10 for reference, and the control transfer function follows.

1258(s + 16.34)(s + 966)(s + 2)


kα = kβ =
(s2 + 3772 )(s + 5633)(s + 0.05)

95
Table 4.10: P Resonant Parameters

Parameter Value
R 1.63 mΩ
L 100 µH
VDC 1450 V
fs 25 kHz
VAC 480 Vrms LL

Figure 4.36: P Resonant Controller

The inverter is was successfully implemented and simulated with different power
commands. Figure 4.38 shows setting of P=1MW and Q = 0.5 MVAR. Impedance
sweeps on the DC an AC side are conducted while varying the real and reactive
power commands, Figures 4.39 ,4.38, and 4.41. Table 4.11 contains the input
ranges swept.

96
Figure 4.37: P Resonant Diagram

Figure 4.38: P Resonant Time Domain Simulation

97
Figure 4.39: P Resonant DC Impedance Magnitudes

Figure 4.40: P Resonant DC Impedance Angles

98
Table 4.11: P Resonant Operating Range

Parameter Min Max


AC Real Power 0 MW 3 MW
AC Reactive Power 0 MVAR 3 MVAR

Figure 4.41: P Resonant AC Impedances: P = 2.5 MW Q = 0.5 MVAR

(a) Magnitudes (b) Angles

4.6.2 "Stiff" grid dq0 Inverter


Control in the dq0 reference frame involves conversion of signals from the abc
frame, thus requiring a phase locked loop (PLL) and additional computational
power. DQ Voltage and current are constant in steady state, allowing control to
be achieved using proportional integral (PI) methods. This greatly simplifies the
controller design process (compared to the P resonant control in section 4.6.1) [75].
This allows for the quick creation of inverter circuits and controls for experimen-
tation in this work.

In this section, a kW-class inverter is designed in Matlab simulink and will later
serve as the base for the "weak" dq0 inverter in section 4.6.3. It is a "stiff" inverter
delivering power into a frequency and voltage regulated AC bus, Figure 4.42 shows
basic structure. The current commands are obtained from the power commands
in the dq0 frame [75].

99
3
P = [vd id + vq iq ]
2
3
Q = [−vd iq + vq id ]
2
Since the PLL aligns the d axis with the source voltage, vq = 0 in steady state,
simplify and rearrange

2P
i∗d =
3vd
2Q
i∗q = −
3vd
The inner control loop is designed to be fast by pole-zero canceling a stable (but
slow) transfer function pole. Details and derivation of the inner current loop are
in Appendix C.2. The PI gains are defined in terms of the component parameters
and an arbitrary time constant τ . Theoretically with the PI control, the actual
currents will have a first order response to current commands, described by time
constant τ . Selection of smaller τ ensures tighter control but power electronics
stability considerations from section 3.2 will be affected.

L
kp =
τ
R
ki =
τ
Circuit elements are designed to meet current ripple requirements given a se-
lected switching frequency and power loading. Procedure for this is presented in
Appendix C.4.1. This along with the preceding PI control gain equations allow
inverter customization for a given application. A set of parameters is designed for
3kVA rated inverter, allowing 5% ripple, Table 4.12.

The inverter is implemented in Matlab simulink and simulated in the time


domain for different power set points. Figure 4.43 shows the AC side power injected
with set point of P = 2.5kW and Q = 0.5kV AR.

100
Figure 4.42: "Stiff" dq0 Diagram

Table 4.12: "Stiff" dq0 Inverter Parameters

Parameter Value
Power 3 kVA
Allowed Current Ripple 5%
VDC 300 V
fs 200 kHz
R loss 1.5 mΩ
τ 1 × 10−5
Outputs
L 0.6 mH
kp 600
ki 150

Figure 4.43: "Stiff" dq0 Time Domain Results

101
4.6.3 "weak" grid dq0 Inverter
In this section an grid regulating kW class inverter is designed, implemented, and
impedance tested in Matlab. In contrast to the P resonant and "stiff" dq inverters,
sections 4.6.1 and 4.6.2, this inverter regulates the voltage and frequency of the
AC bus under variable loading. It can be classified as a "hard grid maker" and
can support an islanded AC microgrid. This is accomplished by adding an output
capacitor and feeding forward the load current, Figure 4.44.

Figure 4.44: "Weak" dq0 Diagram

In addition to the inner control loop, an outer voltage control loop is added.
Details of the derivation are in Appendix C.3.1. PI control is again used, the
method of symmetrical optimum determining the gains, Appendix C.3.2. The ca-
pacitor is sized using voltage ripple requirements. The details are presented in
Appendix C.4.2. The resulting element and design parameters are presented in
Table 4.13.

The designed inverter is implemented in Matlab simulink, Figure 4.45 shows


the time domain results for power commands P = 5kw and Q = 2.5kV AR. The

102
Table 4.13: "Weak" dq0 Inverter Parameters

Parameter Value
Power 3 kVA
Allowed Current Ripple 5%
VDC 300 V
fs 200 kHz
R loss 1.5 mΩ
τ 1 × 10−5
δm 53◦
Ouputs
L 0.6 mH
C 6.9 µF
kp 0.0462
ki 2239

simulated impedance measurement blocks are applied for varying power outputs,
the results are presented in Figures 4.46, 4.47 and 4.48.

Figure 4.45: "Weak" dq0 Time Domain Simulation

103
Figure 4.46: "Weak" dq0 DC Impedance Magnitudes

104
Figure 4.47: "Weak" dq0 DC Impedance Angles

Figure 4.48: "Weak" dq0 AC Impedance


(a) Magnitude (b) Angle

105
Chapter 5 |
Results and Conclusions

In this chapter the techniques and data developed thus far were used to demon-
strate stability assessment, stability margin, and operating point stability for non-
trivial circuit models. First a circuit containing the P resonant inverter and an
ideal source with variable line resistance was analyzed with emphasis placed on
measurement of stability margin. Second, a more complex notional wind power
microgrid was analyzed. This system contains component choices and a wider
range of operating conditions. The usefulness of the unstable operating point as-
sessment is illustrated. Finally conclusions are presented about both the notional
wind microgrid under study and the developed methods in general.

5.1 Results

5.1.1 P Resonant Inverter and Stability Margins


To demonstrate the viability of the MPC for stability margins, the P resonant
controlled inverter from section 4.6.1 was analyzed. An ideal voltage source, this
could be a DC generator or bus, supplies the inverter through a resistive transmis-
sion line, Figure 5.1. Recall that this MW class inverter was obtained from [75],
and was included so that the methods presented could be used on an externally
created design. The impedance plot of the inverter differs from the dq0 versions
in that it has a relatively low impedance, Figure 4.39, making it especially prone
to instability.

The line resistance is varied, and Nyquist plots generated. Note separate com-

106
Figure 5.1: P Resonant Inverter System

ponent testing, as practiced in this work, is required as an unstable system can’t


be measured. The collection of plots show instability occurs for some conditions,
Figure 5.2, which is confirmed in the time domain, Figure 5.3.

The stability margin is tracked as a function of the line resistance for different
loadings, Figure 5.4. Stability margins rapidly deteriorate as the line resistance
increases beyond 0.1 Ω, although higher loadings are shown to be affected more.
Note that this resistance corresponds to less than 50 ft of 14 gauge power wire.
Time domain simulations confirm instability approximately at the indicated resis-
tances and loadings.

107
Figure 5.2: P Resonant Inverter Nyquist Plots

Figure 5.3: P Resonant Inverter System Time Domain

108
Figure 5.4: P Resonant Inverter System Margins

5.1.2 Notional Wind Microgrid and Operating Point Stability


Assessment techniques were also applied to the notional wind microgrid developed
in Chapter 4. Design variables included component choices as well as operating
points. For this more complicated system, assessment must be carried out at
multiple interfaces to ensure stability. The three necessary locations are shown
in Figure 5.5. While components can be further separated to allow for more in-
terfaces for testing, this is not necessary if they are considered stable by themselves.

Throughout this work the term "source" and "load" have been used in quota-
tions deliberately. The definitions are given in section 2.2.1 and can lead to counter
intuitive descriptions, as is the case for this notional microgrid. The definitions
lead the wind turbine and inverter subsystems to be classified as loads while the
battery is considered a source. Although the wind turbine is obviously a power
source, with the connected power electronics it behaves like a constant current
source. It is ill designed to be connected with a constant current load, section 4.4,

109
Figure 5.5: DC Bus Stability Assessment Interfaces

as would be required to be defined as a "source" in the power electronics sense.


Although the battery can absorb or produce power, it is ideally suited to be consid-
ered the system "source" because it is modeled as a filtered voltage source, section
4.5. Since any linear RLC network is passive [43], the source is guaranteed to
be stable by itself when connected to a constant current load thus satisfying the
definition of source.

Note that designations are not exclusive and there is some flexibility in how
a system is divided up. The battery is also stable when connected to a constant
voltage source, making it equally applicable as a "load". However batteries gener-
ally have very low impedance and therefore would make a poor choice for a "load"
subsystem. Designations and interface choices obviously don’t affect the actual
stability of the system, but they do change the shape of the Nyquist plot, causing
conservative criteria to fail for some selections. In general dual components with
high impedance should be considered "loads" and low impedance ones considered
"sources" to minimize conservatism.

A set of operating conditions including component loadings, element variabil-


ity, and component choices were selected to conduct the stability analysis. To

110
Figure 5.6: Microgrid Component Impedances

construct the return ratio, component impedance data was loaded from data files
and processed. Collected impedance data will vary in resolution and range, no
attempt to standardize collection was desired, Figure 5.6. Data is interpolated for
the set of operating conditions considered, resized and re-sampled in the frequency
domain to make compatible, Appendix B.2

The selected operating parameters and ranges are presented in Table 5.1. Since
the system has energy storage the inverter’s rated power can safely be less than
the wind turbine subsystem’s rated power. The battery, being a modeled linear
circuit, does not vary with loading so instead circuit element parameters are var-
ied an order of magnitude around the nominal values developed in section 4.5 to
simulate aging and uncertainty.

A Matlab script was developed, Appendix B.3, to generate and stack Nyquist
plots for an interface. The results for each component combination and interface
are presented in Figures 5.7, 5.8, and 5.9. The number of variables leads to an
almost incomprehensible number of plots, although for this microgrid, the com-
ponents are very well behaved, making visual determination of stability possible.
Minimum stability margin for each system and interface are presented in Table 5.2.

111
Table 5.1: DC System Operating Variables

Parameter Min Max


Wind Power 0 kW 4.5 kW
Inverter Power 0 kW 3 kW
Battery RCT 10 mΩ 1Ω
Battery R 10 mΩ 1Ω
Battery CDL 0.1 mF 10 mF

Note that interface return ratios can be calculated off-line from the individual
component impedance data for the interfaces by the following equations:

~ op )
Zbattery (ω, X
~ op ) = Zs =
L1 (ω, X
Zl ~ op )kZinverter (ω, X
Zturbine (ω, X ~ op )

~ ~ ~
~ op ) = Zs = Zbattery (ω, Xop )kZturbine (ω, Xop )(ω, Xop )
L2 (ω, X
Zl ~ op )
Zinverter (ω, X

Zs ~ op )kZinverter (ω, X
Zbattery (ω, X ~ op )(ω, X
~ op )
~ op ) =
L3 (ω, X =
Zl Zturbine (ω, X~ op )

Table 5.2: Notional Microgrid Stability Margins

System Rectifier Type Interface 1 Interface 2 Interface 3


Diode 0.7594 0.9162 0.8032
Boost 0.6647 0.9194 0.6658
Switching 0.9192 0.9165 0.9665

Figure 5.7: Nyquist Plots: Notional Wind System with Diode Rectifier
(a) Interface 1 (b) Interface 2 (c) Interface 3

112
Figure 5.8: Nyquist Plots: Notional Wind System with Boost Rectifier
(a) Interface 1 (b) Interface 2 (c) Interface 3

Figure 5.9: Nyquist Plots: Notional Wind System with Switching Bridge Rectifier
(a) Interface 1 (b) Interface 2 (c) Interface 3

Comparison of the rectifier choices using the resulting stability margins and
Nyquist plots is difficult because they are all assessed to be very stable for the given
conditions. In the previous section it was shown that increasing the line resistance
destabilized the P resonant controlled inverter. A similar tactic was applied to
stress the wind system. A 5Ω line resistance added to the battery, corresponding
to about 200 ft of 24 gauge std hookup wire. The stability analysis is repeated,
producing more interesting results, Figures 5.10, 5.11, and 5.12. The switching
and boost converter are largely unaffected but the system with the passive diode
rectifier becomes unstable. Furthermore, analysis of the operating points show
that the diode rectifier system experiences this instability for only one of the 405
combinations of conditions tested, Table 5.3. Note that this instability was only
detected for inverter power output of 2.5 kW. All other variables held constant,
the system was assessed stable when the inverter was producing less that 1.25
kW or more than 3.75 kW. If the battery were used as part of an EMS, section
2.4.4, steady inverter operation in this range could be avoided, thus avoiding the
instability.

113
Table 5.3: Unstable Operating Point

Parameter Value
Wind Power 2.25 kW
Inverter Power 2.50 kW
Battery RCT 1Ω
Battery R 1Ω
Battery CDL 0.1 mF

Figure 5.10: Diode Rectifier and High Resistance Battery Line


(a) Interface 1 (b) Interface 2 (c) Interface 3

5.2 Conclusions
The overall dissertation objective was largely achieved and unique contributions
delivered in the form of a Matlab impedance measurement block and high level op-
timization framework that includes power electronics stability assessments. Small
signal stability assessment techniques were adapted for microgrids and in this chap-
ter successfully applied to non-trivial models. The three requirements for a lower
level stability routine in the proposed approach; qualitative stability assessment,
quantitative stability assessment, and failure conditions assessment, were demon-
strated via simulation.

The Nyquist stability criterion (and GNC for AC systems) was selected and
shown to be effective in predicting the stability of both linear an nonlinear compo-
nents in AC and DC electrical systems. Although alternatives exist, this method
was particularly useful because it requires only information about individual com-
ponent impedances; data that can be captured empirically from any arbitrary
device.

114
Figure 5.11: Boost Rectifier and High Resistance Battery Line
(a) Interface 1 (b) Interface 2 (c) Interface 3

Figure 5.12: Switching Bridge Rectifier and High Resistance Battery Line
(a) Interface 1 (b) Interface 2 (c) Interface 3

Study of underlying theory revealed some limitations. The assumption of sys-


tem linear time-invariance (LTI) can be problematic for nonlinear switching de-
vices. To ensure the validity of analysis, the switching frequency of connected
devices must be greater than the control bandwidths, however filtering (with ca-
pacitors for example) can soften this requirement. Mathematically the stability
assessed is bounded input bounded output (BIBO), one of the weaker types. As
applied in this work, it only guarantees small disturbances do not result in an infi-
nite valued voltage response. In special cases this is sufficient to ensure stability of
the whole system, but information to determine this would not be available in the
proposed microgrid application. Therefore assessment of all accessible interfaces
is recommended.

Time domain simulations showed that while the Nyquist criterion alone could
be used for stability assessment, there was some error associated with the inex-
act impedance data measured in simulation. This necessitates the use of stability
margins. The criteria presented in the literature all intrinsically contain some
margin and conservatism. Maximum peak criteria was chosen to assess stability

115
margin, and was done separately from the stability analysis. Results in section
5.1.1, showed how it could be used to quantitatively compare the the stability of
different conditions. This verified the intuitive conclusion that higher source resis-
tance and higher inverter loading lead to lower stability.

Much of the work encountered in the literature was aimed at electrical engi-
neers working on component circuit design and specification. The concerns of a
microgrid integrator are not so focused, needing to consider a range of factors
and trade-offs simultaneously. According to the proposed methods, fast off-line
determination of stability would use a database of component stability character-
istics. This was accomplished by measuring impedances as a function of operating
point and frequency via repeated simulations. The data was saved in files and
readily used by the developed Matlab assessment scripts, taking only seconds to
generate the results presented in this dissertation. This makes relatively complex
networks, such as microgrids, tractable problems once the stability characteristics
are obtained. Note however, that acquiring the necessary impedance data for some
of the developed components in this work required hours or even days of simulation.

The wind turbine subsystem studied was implemented with MPPT and three
alternative rectifier architectures. The advantages and disadvantages of each are
well documented in the literature for small wind applications. In this work a
comparison of their stability characteristics was sought. By themselves impedance
information was not particularly useful, as each had a similarly low impedance,
Figure 5.13. When incorporated into the notional system and assessed, the diode
bridge was shown to be particularly susceptible to instability if the battery was
connected by a long (200 ft) line. The other rectifiers, boost rectifier and switching
bridge, were unaffected according to the resulting Nyquist plots.

Both stability assessment scenarios; P resonant inverter system and notional


wind microgrid, highlighted the importance of stability analysis to the higher level
microgrid design and optimization problem. In each system, stability was shown to
be negatively effected by line resistance, a property primarily a function of length
and gauge. Generator placement and line layouts will potentially affect power
electronic converter instability and likewise component choices will therefore also

116
Figure 5.13: Comparison of Rectifier Impedances
(a) Diode Bridge (b) Boost Rectifier (c) Switch Bridge

affect allowable (stable) layouts. This is further evidence that the relationship
between microgrid design and power electronics induced instability is significant.

When a long resistive line was placed between the battery and the DC bus
of the notional microgrid, the system variant with the passive diode rectifier was
assessed unstable. Although this instability was only detected for a single set of
operating conditions. Conventionally this would disqualify the architecture entirely
but in the proposed method the instability could be mitigated. The conditions
of the battery and rectifier loading are considered exogenous and uncontrollable,
but the allowable loadings of the inverter could be restricted. If a diode rectifier
architecture was strongly preferred due to other (non-stability related) factors,
the instability could be overcome given the operating point stability assessment
information generated. The microgrid scenario presented demonstrates the benefit
of such analysis.

117
Chapter 6 |
Recommendations

During the course of this research, several topics were noted as being worthy of
additional investigation.

Impedance estimation techniques were based on introducing a perturbation


and measuring the resulting waves at each frequency of interest. This method was
shown to be effective but time consuming as many tests were required to sweep
the range of frequencies. Impedance fitting was briefly investigated and alternative
techniques for impedance measurement that excite a range of frequencies simulta-
neously appear in the literature. Adoption of these has the potential to reduce the
number of test required to characterize the impedance of a device, thus making
the proposed methods more feasible.

Even if the impedance could be estimated for a range of frequencies with a


few tests, the problem of the many operating point combinations remains. The
nonlinear nature of components require many impedance characterizations to suf-
ficiently describe the device. This was especially problematic for the impedance
measurement blocks created in Matlab. Time domain simulations to character-
ize the impedance are computationally expensive requiring long calculation times.
This problem would likely be solved by empirically characterizing components
using equipment as reality occurs in real time. However practical measurement
introduces other problems associated with the limitations of sensors, noise, and
injection of perturbations that have not been fully addressed in the literature.

Some potential concerns about the interaction of devices when averaged models

118
and time invariance are no longer valid assumptions were discussed. In theory a
slowly switching device such as a passive diode rectifier violates these assumptions
and would therefore be difficult to incorporate into a stability analysis. How-
ever intuitively such a device with a large filter will behave mostly like the filter
in those frequencies, making stability assessment possible. A clear procedure for
determining when switching frequencies matter and how much would be of benefit.

Stability margins were suggested as a way to incorporate stability into the


higher level design optimization. Where systems with higher and lower margin
can be appropriately rewarded or penalized in the optimization process. The max-
imum peak criteria (MPC) was selected as a quantitative measure for this purpose.
Stability margins were assessed for a number of cases and examined, but a strong
relationship with robustness was not apparent. Other metrics may prove better at
quantitatively describing the margin, but need to be assessed for this application.

A proposed solution to instability at certain operating points is to restrict the


system from dwelling in them. However such operating points are likely unavoid-
able from a transient point of view. This was not considered a problem since it
is not uncommon in reality for systems to have an unstable point that can be un-
eventfully traversed. A more formal treatment of how fast and when such unstable
points can be safely encountered under transient conditions would further justify
the proposed methods.

The methods and techniques developed were for both AC and DC systems, the
primary difference being the non-scalar impedance characteristics of AC systems.
The assessment methods developed are modular and allow stability to be assessed
for a larger microgrid. The notional microgrid developed was conceived as part of
a large AC microgrid, Figure 6.1. AC impedance characteristics of the notional
wind subsystem could be combined with other the microgrid "leg" impedances.
Stability could than be analyzed using the same procedure as demonstrated in this
dissertation.

119
Figure 6.1: AC Microgrid Stability Diagrams

120
Bibliography
[1] Hammerstrom, D. J. (2007) “AC versus DC Distribution Systems Did We
Get it Right?” in Power Engineering Society General Meeting, Tampa, FL,
pp. 1–5.

[2] Sorrell, S. (2007) The Rebound Effect: an assessment of the eveidence for
economy-wide energy savings from improved energy efficiency, Tech. rep., UK
Energy Research Centre, iSBN 1-903144-0-35.

[3] Troester, E. (2009) “New German Grid Codes for Connecting PV sys-
tems to the Medium Voltage Power Grid,” in 2nd International Workshop on
Concentrating Photovoltaic Power Plants: Optical Design, Production, Grid
Connection, pp. 9–10.

[4] Lasseter, R. H. (2004) “Microgrid: A Conceptual Solution,” in Power


Electronics Specialists Conference, pp. 4285–4290.

[5] NREL (2010), “Supporting Data for Energy Technology Costs,” .

[6] (1999) Regional Transmission Organizations, Tech. rep., Federal Energy Reg-
ulatory Commission, docket No. RM99-2-000; Order No. 2000.

[7] Kelly, R., G. Orltl, and A. L. Jullianl “Reducing Fuel Consumption at


a Remote Military Base: Introducing an energy management system,” IEEE
Electrification Magazine, 1(2).

[8] Turso, J., W. Ainsworth, L. Dusang, D. Miller, and L. Smith (2007),


“U.S.S. Makin Island: Simulation-Based Analysis and its Role in Electric-
Plant Control System Design,” IEEE.

[9] Kanellos, F. D., G. J. Tsekouras, and J. Prousalidis “Onboard DC


grid employing smart grid technology: challenges, state of the art and future
prospects,” IET Electrical Systems in Transportation, 5(1).

[10] Louganski, K. P. (1999) Modeling and Analysis of a DC Power Distribution


System in 21st Century Airlifters, Electrical engineering, Virginia Polytechnic
Institute and State University, Blacksburg, VA.

121
[11] Zhang, L., T. Wu, Y. Xing, K. Sun, and J. M. Gurrero (2011) “Power
control of DC Microgrid using DC bus signaling,” in Applied Power Electronics
Conference and Exposition, Fort Worth, TX, pp. 1926–1932.

[12] Lidula, N. and A. Rajapakse (2010) “Microgrids research: A review of


experimental microgrids and test systems,” Renewable and Sustainable Energy
Reviews, 15, pp. 186–202, doi: 10.1016/j.rser.2010.09.041.

[13] Ustun, T., C. Ozansoy, and A. Zayegh (2011) “Recend developments in


microgrids and example cases around the world A review,” Renewable and Sus-
tainable Energy Reviews, 15, pp. 4030–4041, doi: 10.1016/j.rser.2011.07.033.

[14] Planas, E., A. G. de Muro, J. Andreu, I. Kortabarria, and I. M.


de Alegria (2013) “General aspects, hierarchical controls and droop meth-
ods in microgrids: A review,” Renewable and Sustainable Energy Review, 17,
pp. 147–159.

[15] Moyet, J. and D. Menicucci (2011) Advanced Concepts for Controlling


Energy Surety Microgrids, Tech. rep., Sandia National Laboratories, Albu-
querque, NM, sAND2011-5048.

[16] Lasseter, R. H. (2006), “Control and Design of Microgrid Components,”


PSERC, Ithaca.

[17] Cuzner, R. M. and G. Venkataramanan (2008) “The Status of DC


Micro-Grid Protection,” in Industry Applications Society Annual Meeting, Ed-
monton, Alta, pp. 1–8.

[18] Mohan, N. (2012) Electric Machines and Drives, Don Fowley.

[19] Glover, J. D., M. S. Sarma, and T. J. Overbye (2012) Power System


Analysis and Design, Christopher M. Shortt.

[20] Belkhayat, M. (1997) Stability Criteria for AC Power Systems with Regu-
lated Loads, Ph.D. thesis, Purdue University, West Lafayette, IN.

[21] Emadi, A. (2001) “Modeling and Analysis of Multi-Converter DC Power


Electronic Systems using the Generalized State Space Averaging Method,” in
Industrial Electronics Society, Denver, CO, pp. 1001–1007, the 27th Annual
Conference of the IEEE.

[22] Corporation, W. E. (1967), “Computer Analysis of Shipboard Electrical


Power Distribution Systems with Static Line Voltage Regulators,” for U.S.
Navy Marine Engineering Laboratory, report No 67-1000.

122
[23] Baring-Gould, E. I. (1996) Hybrid2 The Hybrid System Simulation Model,
National Renewable Energy Laboratory, 1617 Cole Boulevard, Golden, Col-
orado 80401-3393, 1.0 ed.
[24] National Renewable Energy Laboratory, 1617 Cole Boulevard, Golden, Col-
orado 80401-3393 (2005) Getting Started Guide for HOMER, 2.1 ed.
[25] Austin, P. (2011) Practical Use of HOMER In Preliminary Microgrid De-
sign, M.s. paper, The Pennsylvania State University.
[26] Rosado, S., R. Burgos, S. Ahmed, F. Wang, and D. Boroyevich
(2008) “Modeling of Power Electronics for Simulation Based Analysis of Power
Systems,” in Proceedings of the 2007 Summer Computer Simulation Confer-
ence, Society for Computer Simulation International, San Diego, CA, USA,
pp. 19–26.
[27] Dehghan, S. M., M. Mohamadian, and A. Y. Varjani (2009) “A New
Variable SPeed Wind Energy Conversion System Using Permanent Magnet
Synchronous Generator and Z Source Inverter,” in IEEE Transactions on
Energy Conversions, vol. 24, pp. 714–724.
[28] Haque, M. E., M. Negnevitsky, and K. M. Muttaqi (2010) “A Novel
Control Strategy for a Variable Speed Wind turbine With a Permanent Mag-
net Synchronous Generator,” in IEEE Transactions on Insustry Applications,
vol. 46, pp. 331–339.
[29] Baroudi, J. A., V. Dinavahi, and A. M. Knight (2007) “A review of
power converter topologies for wind generators,” Renewable Energy, 32, pp.
2369–2385, doi: 10.1016/j.renene.2006.12.002.
[30] de Freitas, T. R. S., P. J. M. Menegaz, and D. S. L. Si-
monetti (2011) “Converter Topologies for Permanent Magnetic Syn-
chronous Generator on Wind Energy Conversion System,” in 2011 Brazil-
ian Power Electronics Conference, IEEE, Praiamar, Brazil, pp. 936–942, doi:
10.1109/COBEP.2011.6085292.
[31] Tran, D.-H., B. Sareni, X. Roboam, and C. Espanet (2010) “Inte-
grated Optimal Design of a Passive Wind Turbine System: An Experimental
Validation,” in IEEE Transactions on Sustainable Energy, vol. 1, pp. 48–56.
[32] Sudhoff, S. D. (2000) “Admittance Space Stability Analysis of Power Elec-
tronic Systems,” in IEEE trans. Aerosp. Electron. Sys., vol. 36, IEEE, pp.
965–973.
[33] Bay, J. S. (1999) Fundamentals of Linear State Space Systems, Thomas
Casson.

123
[34] Chen, C.-T. (2013) Linear System Theory and Design, 4 ed., Oxford Uni-
versity Press.

[35] Middlebrook, R. D. (1976) “Input Filter Considerations in Design and


Application of Switching Regulators,” in IEEE Industry Applications Society,
pp. 91–107.

[36] Wildrick, C. M., F. C. Lee, B. H. Cho, and B. Choi (1995) “A Method


of Defining the Load Impedance Specification for A Stable Distributed Power
System,” in IEEE Trans. Power Electron., vol. 10, IEEE, pp. 280–285.

[37] Feng, X., S. Ye, K. Xing, F. C. Lee, and D. Borojevic (1999) “Individ-
ual load impedance specification for a stable DC distributed power system,”
in 1999 Applied Power Electronics Conference and Exposition, IEEE, Dallas,
TX, pp. 923–929, doi: 10.1109/APEC.1999.750480.

[38] Riccobono, A. and E. Santi (2012) “Comprehensive Review of Stability


Criteria for DC Distribution Systems,” IEEE, pp. 3917–3925.

[39] Sudhoff, S. D. and J. M. Crider (2011) “Advancements in Generalized


Immittance Based Stability Analysis of DC Power Electronics Based Distri-
bution Systems,” in Electric Ship Technologies Symposium, IEEE, Alexan-
dria,VA, pp. 207–212.

[40] Li, A. and D. Zhang (2014) “Necessary and Sufficient Stability Criterion and
New Forbidden Region for Load Impedance Specification,” Chinese Journal
of Electronics, 23(3).

[41] Riccobono, A. and E. Santi (2012) “A Novel Passivity-Based Stability


Criterion (PBSC) for Switching Converter DC Distribution Systems,” in Ap-
plied Power Electronics Conference and Exposition, IEEE, pp. 2560–2567.

[42] Khalil, H. K. (2002) Nonlinear Systems, 3 ed., Prentice-Hall Inc.

[43] Gu, Y. and X. H. Wuhua Li (2014) “Passivity-Based Control of DC Micro-


grid for Self-Disciplined Stabilization,” IEEE Transactions on Power Systems,
30(5), pp. 2623–2632, doi: 10.1109/TPWRS.2014.2360300.

[44] Sudhoff, S., “DC Stability Toolbox 3.0,” Purdue University School of Elec-
trical and Computer Engineering website.

[45] MacFarlane, A. G. J. (1980) Complex Variable Methods for Linear Mul-


tivariable Feedback Systems, Taylor and Francis Ltd.

124
[46] Burgos, R., D. Boroyevich, F. Wang, K. Karimi, and G. Francis
(1998) “On the AC stability of high power factor three-phse rectifiers,” in
Energy Conversion Congress and Exposition, IEEE, Atlanta, GA, pp. 511–
521.

[47] Liu, Z., J. Liu, W. Bao, and Y. Zhao “Infinity-Norm of Impedance-Based


Stability Criterion for Three-Phase AC Distribution Power Systems With
Constant Power Loads,” IEEE Transactions on Power Electronics, 30(6).

[48] Boroyevich, D. and P. Mattavelli (2012), “Small-Signal Stability and


Subsystem Interactions in Three Phase Nano-Grids,” Center for Power Elec-
tronics Systems.

[49] Amy, J. V. (1991), “A Brief Discussion of the Use of Complex Impedances


in the Dynamic Analysis of Electric Power Systems,” .

[50] Belkhayat, M. and M. L. Williams (2000) “Impedance Extraction Tech-


niques for DC and AC Systems,” in Proceedings of the Naval Symposium on
Electric Machines.

[51] Jaksic, M., Z. Shen, I. Cvetkovic, D. Boroyevich, P. Mattavelli,


M. Belkhayat, and J. Verhulst (2013) “Nonlinear Sideband Effects in
Small-Signal Input dq Admittance of Six-Pulse Diode Rectifiers,” in 2013
Applied Power Electronics Conference and Exposition, IEEE, Long Beach,
CA, pp. 2761–2768, doi: 10.1109/APEC.2013.6520687.

[52] Shen, Z., M. Jaksic, P. Mattavelli, D. Boroyevich, J. Verhulst,


and M. Belkhayat (2013) “Design and Implementation of Three-phase
AC Impedance Measurement Unit (IMU) with Series and Shunt Injection,”
in 2013 Applied Power Electronics Conference and Exposition, IEEE, Long
Beach, CA, pp. 2674–2681, doi: 10.1109/APEC.2013.6520674.

[53] Familiant, Y. A., J. Huang, K. A. Corzine, and M. Belkhayatl


“New Techniques for Measuring Impedance Characteristics of Three-Phase
AC Power Systems,” IEEE Transactions on Power Electronicsl, PE-I(2).

[54] Huang, J., K. A. Corzine, and M. Belkhayat “Small-Signal Impedance


Measurement of Power-Electronics-Based AC Power Systems Using Line-to-
Line Current Injection,” IEEE Transactions on Power Electronicsl, 24(2).

[55] Shen, Z., M. Jaksic, B. Zhou, P. Mattavelli, D. Boroyevich,


J. Verhulst, and M. Belkhayat (2013) “Analysis of Phase Locked Loop
(PLL) Influence on DQ Impedance Measurement in Three-Phase AC Sys-
tems,” in 2013 Applied Power Electronics Conference and Exposition, IEEE,
Long Beach, CA, pp. 939–945, doi: 10.1109/APEC.2013.6520326.

125
[56] MathWorks, “Matlab,” .

[57] Atwa, Y. M. (2010) Distribution System Planning and Reliability Assess-


ment under High DG Penetration, Ph.D. thesis, University of Waterloo, Wa-
terloo, Ontario, Canada.

[58] Fu, Q., L. F. Montoya, A. Solanki, A. Nasiri, V. Bhavarju, T. Ab-


dallah, and D. C. Yu “Microgrid Generation Capacity Design With Re-
newables and Energy Storage Addressing Power Quality and Surety,” IEEE
Trans. Smart Grid, 3(4).

[59] Vandoorn, T. L., B. Meersman, L. Degroote, B. Renders, and


L. Vandevelde “A Control Strategy for Islanded Microgrids with DC-Link
Voltage Control,” IEEE Trans. Power Del., 26(2).

[60] Vandoorn, T. L., B. Meersman, J. D. M. D. Kooning, B. Renders,


and L. Vandevelde “Analogy Between Conventional Grid Control and Is-
landed Microgrid Control Based on a Global DC-Link Voltage Droop,” IEEE
Transactions on Power Delivery, 27(3).

[61] Ito, Y., Y. Zhongqing, and H. Akagi (2004) “DC Microgrid Based Distri-
bution Power Generation System,” in Power Electronics and Motion Control
Conference, Xi’an, pp. 1740–1745.

[62] Anand, S., B. G. Fernandes, and J. M. Guerrero “Distributed Control


to Ensure Proportional Load Sharing and Improve Voltage Regulation in Low-
Voltage DC Microgrids,” IEEE Transactions on Power Electronics, 28(4).

[63] Kakigano, H., Y. Miura, A. Nishino, and T. Ise (2010) “Distribution


voltage control for DC Microgrid by converters of energy storages considering
the stored energy,” in Energy Conversion Congress and Exposition, Atlanta,
GA, pp. 2851–2856.

[64] Lasseter, R. H. (2010) “CERTS microgrid Laboratory Test Bed,” ,âĂİ


IEEE Transactions on Power Delivery Journal.

[65] ——— (2006), “Control of Small Distributed Energy Resources,” Patent 7


116 010 B2.

[66] Saeedifard, M., M. Graovac, R. F. Dias, and R. Iravani (2010) “DC


power systems: Challenges and opportunities,” in Power and Energy Society
General Meeting, Minneapolis, MN, pp. 1–7.

[67] Koutroulis, E. and K. Kalaitzakis (2006) “Design of a Maximum Power


Tracking System for Wind-Energy Conversion Applications,” IEEE Transac-
tions on Industrial Electronics, 53(2), pp. 486–494.

126
[68] Broe, A. M. D., S. Drouilhet, and V. Gevorgian (1999) “A Peak
Power Tracker for Small Wind Turbine in Battery Charging Applications,”
IEEE Transactions on Energy Conversion, 14(4), pp. 1630–1635.

[69] Pan, C.-T. and Y.-L. Juan (2010) “A Novel Sensorless MPPT Controller for
a High-Efficiency Microscale Wind Power Generation System,” IEEE Trans-
actions on Energy Conversion, 25(1), pp. 207–216.

[70] Mirecki, A., X. Roboam, and F. Rechardeau (2007) “Architecture Com-


plexity and Energy Efficiency of Small Wind Turbines,” IEEE Transactions
on Industrial Electronics, 54(1), pp. 660–670.

[71] Belouda, M., J. Belhadj, B. Sareni, and X. Roboam (2011) “Battery


sizing for a stand alone passive wind system using statistical techniques,”
in International Multi-Conference on Systems, Signals and Devices, IEEE,
Sousse, pp. 1–7, doi: 10.1109/SSD.2011.5767373.

[72] Ohyama, K., S. Arinaga, and Y. Yamashita (2007) “Modeling and


Simulation of Variable Speed Wind Generator System Using Boost Con-
verter of Permanent Magnet Synchronous Generator,” in European Con-
ference on Power Electronics and Applications, IEEE, pp. 1–9, doi:
10.1109/EPE.2007.4417541.

[73] Schiemenz, I. and M. Stiebler (2001) “Control of a Permanent Magnet


Synchronous Generator Used in a Variable Speed Wind Energy System,” in
IEEE international Electric Machines and Drives Conference, IEEE, Cam-
bridge, MA, pp. 872–877, doi: 10.1109/IEMDC.2001.939422.

[74] Chinchilla, M., S. Arnaltes, and J. C. Burgos (2006) “Control of


Permanent-Magnet Generators Applied to Variable-Speed Wind Energy Sys-
tems Connected to the Grid,” IEEE Transactions on Energy Conversion,
21(1), pp. 130 – 135.

[75] Yazdani, A. and R. Iravani (2010) Voltage-Sourced Converters in Power


Systems, John Wiley and Sons.

[76] Waag, W., S. Kabitz, and D. U. Sauer (2013) “Experimental inverstiga-


tion of the lithium-ion battery impedance characteristic at various conditions
and aging states and its influence on application,” Applied Energy, 102, pp.
885–897.

[77] Jossen, A. (2006) “Fundamentals of battery dynamics,” Journal of Power


Sources, 154, pp. 530–538.

127
[78] Buller, S., M. Thele, R. W. A. A. D. Doncker, and E. Karden
(2005) “Impedance-Based Simulation Models of Supercapacitors and Li-Ion
Batteries for Power Electronic Applications,” IEEE Transactions on Industry
Applications, 41(3), pp. 742–747.

[79] Lasseter, R. H. “CERTS Microgrid Laboratory Test Bed,” IEEE Trans.


Power Del., 26(1).

128
Appendix A|
Grid Control

A.1 Grid vs. Microgrid


There are signficant difference between the properties of the centralized grid and
microgrids.

• Power Flow: The centralized grid was designed in a hierarchical structure


where electricity was always flowing from the large power stations, through
the transmission system, trickling down to the end user. In microgrids the
situation is more dynamic as distributed generation and energy storage can
cause the power demands to behave atypically and even reverse direction
occasionally. This requires greater flexibility on the part of control systems.

• Energy Storage: The grid is dominated by large synchronous rotating ma-


chines on the supply side. The combined inertia of all these machines acts
like a flywheel, providing a kinetic energy storage to absorb power fluctua-
tions. In microgrids without significant amounts of directly coupled rotating
machines, energy storage must be supplied from elsewhere or be vulnerable
to disturbances. In a DC bus, capacitance can be used, but AC buses are
more complicated as the "inertia" needs to be synthesized virtually. That is,
the dampening provided virtually via control algorithms.

• Transmission Network: Line inductance dominates the connections in the


grid due to long transmission lines and liberal use of transformers. Microgrids
cover a small geographic area and therefore have short resistive connections
that don’t require transformers.

129
• Power Quality: In the conventional grid the chief concern is power factor
angle, a mismatch of voltage and current waveforms resulting from inductive
effects. This can be easily corrected for in microgrids using power electronics
but in turn suffer from harmonic distortions from the switching components.

• Control: Due to the factors mentioned in this list, control methods for the
grid and microgrids are different. The grid uses the robust and unversal
Automatic Generator Control (AGC). This uses system frequency to bring
the electrical supply and demand on the grid into balance on a moment to
moment basis. Microgrids are not standardized and a plethora of control
schemes have been suggested in the literature.

A.2 Electrical Grid Control

A.2.1 AGC
Automatic generation control is the combination of several control schemes that
work together to maintain the balance, stability, and economics of the grid. These
controls are centered around the widespread use of synchronous generators and an
inductive distribution system of long lines and transformers [19].

• Generator Voltage Control: Synchronous generators require the rotor to


produce a magnetic field. For large generators this is accomplished by an ex-
citer that delivers DC voltage to sets of rotor windings. The supplied voltage
dictates the output voltage, and therefore the reactive power, produced by
the stator. This allows, either by an outside command or by control loop, to
maintain local voltage by injecting or absorbing reactive power.

• Turbine Governor Control: The large synchronous generators of the grid


have a significant amount of rotational kinetic energy stored among them.
When demand exceeds supply, this mass supplies the power by decelerating
and when supply exceeds demand it accelerates. Since the rotor speeds are
directly coupled with the frequency of the grid, it acts as a global variable in-
dicating power mismatch. Droop control ∆pm = ∆pref −k∆f is implemented
to change the governor power setting depending on the grid frequency.

130
• Load Frequency Control: While turbine governor control can be relied
upon to balance the system, it will inevitably result in steady state frequency
error if ∆pref is not also adjusted as load changes. LFC accomplishes this by
driving the steady state frequency back to the nominal value. It also ensures
that local generators pick up most of the local load changes by keeping the
net power flow between areas to the scheduled levels.

• Economic Dispatch: This level of control updates the least frequently and
enacts solutions to the economic problem of dispatch by alterning individual
generator setpoints. The economic problem considers the cost of generation
as well as losses in the transmission lines due to the geographic location of
generators.

A.2.2 CERTS microgrid control


The Consortium for Electric Reliability Technology Solutions (CERTS) sponsors
initiatives relating to distributed energy resource microgrids. The CERTS micro-
grid concept consists of two main parts [4, 65, 79].

• Smooth and automatic connection and disconnection of microgrids from the


main grid

• plug and play, peer to peer, automatic controls for (inertia-less) inverter
based microsources

Control is conducted independently on each generator by measuring voltage


and frequency at the interface, not needing interunit communication. This is ac-
complished using P/f and V/Q droop control. The P/f droop differs from turbine
governor control in two ways. First, even if the unit is a generator, the DC bus
is tapped for energy rather than the machines rotating mass. Secondly power is
measured and output angle (frequency) is adjusted, the reverse of how turbine
governor control is implemented. The Q/V droop ensures that minor differences
in voltage setpoint do not lead to large circulating currently by lowering the volt-
age setpoint with increasing reactive power and vice versus of decreasing reactive
power. The microsource and associated power electronics and controls are illus-
trated in Figure A.1. Note an inductive connection to the AC bus is required for

131
the inverter controls to function properly. Although alternative control schemes
for resistive AC microgrids have been proposed [60].

Figure A.1: CERTS Microgrid Microsource Subsystem Diagram

A.3 High Level Design


High level microgrid design programs use time domain studies to assess viabil-
ity of designs but are limited by their time step. The hybrid microgrid design
tool HOMER conducts Monte-Carlo-like analysis over a year using one hour time
steps [24]. For microgrids with highly intermittant renewables the intra-hourly
power profiles will significantly differ from the hour average. A Matlab subrou-
tine was developed to conduct a trade-off analysis for a hybrid wind system for
a one hour period, Figure A.2. The simulation operates using simple power and
energy balance of the components while tracking key variables. Wiebull randomly
distributed wind speeds were used to create the variability along a period square
wave demand routine. Some example time domain results are presented in Figure
A.3. The goal of the analysis was to vary the generator control parameters to cre-
ate Pareto optimal options of design in terms of each components required power

132
Figure A.2: Simple Hybrid Wind System

rating, energy storage requirement (battery only), and response rate (mostly for
generator).

X
P (t) = 0 = Pbatt + Pdump + Pwind
Z τ
Ebatt = Pbatt (t)dt
0
" #
dPk (t)
P˙k = max
dt

133
Figure A.3: Simple Hybrid Wind System Time Domain Simulation Example

134
Appendix B|
Matlab Code

B.1 Matlab Impedance Fitting Code

% Store Impedance Data


ZsV = Z(:,1); % source impedance vector
ZlV = Z(:,2); % load impedance vector

% Transfer function order estimates


% Source
ONS = 1; % order of the source transfer function’s numerator
ODS = 2; % order of the source transfer function’s denominator
% Load
ONL = 1; % order of the load transfer function’s numerator
ODL = 0; % order of the load transfer function’s denominator

% Frequency Range
wmin = 1E-2; % begin sweep [Hz]
wmax = 1E5; % end sweep [Hz]

% Convert Collected Data to TF Coefficients


[B,A] = invfreqs(ZsV,freqs*2*pi,ONS,ODS);

C1 = [zeros(1,(6-numel(B))), B];

135
C2 = [zeros(1,(6-numel(A))), A];

[B,A] = invfreqs(ZlV,freqs*2*pi,ONL,ODL);

C3 = [zeros(1,(6-numel(B))), B];
C4 = [zeros(1,(6-numel(A))), A];

C = vertcat(C1,C2,C3,C4);

% Empirical Calculation
s = tf(’s’);
SN = C(1,1)*s^5+C(1,2)*s^4+C(1,3)*s^3+C(1,4)*s^2+C(1,5)*s+C(1,6);
SD = C(2,1)*s^5+C(2,2)*s^4+C(2,3)*s^3+C(2,4)*s^2+C(2,5)*s+C(2,6);
LN = C(3,1)*s^5+C(3,2)*s^4+C(3,3)*s^3+C(3,4)*s^2+C(3,5)*s+C(3,6);
LD = C(4,1)*s^5+C(4,2)*s^4+C(4,3)*s^3+C(4,4)*s^2+C(4,5)*s+C(4,6);

Zs = SN/SD;
Zl = LN/LD;

B.2 Impedance Data Interpolation

% number of frequency points to resample to


n = 50;

% resample data to allow nyquist plot of minor loop gain


% largest minimum frequency sets lower bound
lower_bound = log10(max([min(freqs_S),min(freqs_L1), ...
...min(freqs_L2),min(freqs_L3),min(freqs_L4)]));
% smallest max frequency sets upper bound
upper_bound = log10(min([max(freqs_S),max(freqs_L1), ...
...max(freqs_L2),max(freqs_L3),max(freqs_L4)]));
% new set of frequencies

136
freqs_r = logspace(lower_bound,upper_bound,n);

% interpolate for new sychronized impedance vectors


% source (the battery)
Zr_S = interp1(freqs_S,Z_S,freqs_r);
% load 1 : diode rectifier, interpolate frequency and power
d2 = linspace(min(Zp_L1(:,3)),max(Zp_L1(:,3)),3);
[xq,yq] = meshgrid(freqs_r,d2);
vq = griddata(Zp_L1(:,2),Zp_L1(:,3),Zp_L1(:,1),xq,yq);
Z_L1_int = vq’;
% load 2 : boost rectifier, interpolate frequency and power
%Z_L2_int = interp1(freqs_L2,Z_L2,freqs_r);
d2 = linspace(min(Zp_L2(:,3)),max(Zp_L2(:,3)),5);
[xq,yq] = meshgrid(freqs_r,d2);
vq = griddata(Zp_L2(:,2),Zp_L2(:,3),Zp_L2(:,1),xq,yq);
Z_L2_int = vq’;
% load 3 : switching bridge, interpolate frequency and power
%Z_L3_int = interp1(freqs_L3,Z_L3,freqs_r);
d2 = linspace(min(Zp_L3(:,3)),max(Zp_L3(:,3)),5);
[xq,yq] = meshgrid(freqs_r,d2);
vq = griddata(Zp_L3(:,2),Zp_L3(:,3),Zp_L3(:,1),xq,yq);
Z_L3_int = vq’;
% load 4 : inverter, interpolate frequency and power
d2 = linspace(min(Zp_L4(:,3)),max(Zp_L4(:,3)),5);
[xq,yq] = meshgrid(freqs_r,d2);
vq = griddata(Zp_L4(:,2),Zp_L4(:,3),Zp_L4(:,1),xq,yq);
Z_L4_int = vq’;

B.3 Matlab Nyquist Plotter


% reorganize
L1 = zeros(length(freqs_r),size(Zr_S,2)*size(Z_L1_int,2)*size(Z_L4_int,2));
Nyq1 = zeros(length(freqs_r),size(Zr_S,2)*size(Z_L1_int,2)*size(Z_L4_int,2));

137
m=1;
for i=1:size(Zr_S,2)
for j=1:size(Z_L1_int,2)
for k=1:size(Z_L4_int,2)
L1(:,m) = Zr_S(:,i).*(Z_L1_int(:,j)+Z_L4_int(:,k))./ ...
... (Z_L1_int(:,j).*Z_L4_int(:,k));
sys = frd(L1(:,m),freqs_r);
[re,im] = nyquist(sys,freqs_r);
Nyq1(:,m) = complex(re,im);
m = m + 1;
end;
end;
end;

% process nyquist plots into trajectories


% (*so that they can be "drawn" without lifting the hand)
Nyq1_m = [((Nyq1)’).’;flipud(Nyq1)];

% Calculate stability margin for each trajectory


% let col be var combo & row1 = stability row2 = stability margin
Stab1 = zeros(2,size(Nyq1_m,2));
% initialized closed nyquist plot loot (through inf <-> -inf)
temp_L1 = zeros(size(Nyq1_m,1)+1,1);
for i=1:size(Nyq1_m,2)
temp_dist1 = zeros(size(Nyq1_m,1),1);
temp_L1 = Nyq1_m(:,i);
temp_L1(size(Nyq1_m,1)+1,1)= Nyq1_m(1,i); % tack 1st value on end
for j=1:size(Nyq1_m,1)
temp_dist1(j,1) = abs(Nyq1_m(j,i)+1); % distance to critical point
if (imag(temp_L1(j,1))*imag(temp_L1(j+1,1)))<0
A3 = real(temp_L1(j,1))+((real(temp_L1(j+1,1))- ...
... real(temp_L1(j,1)))/ (imag(temp_L1(j+1,1))- ...
... imag(temp_L1(j,1))))*(-1*imag(temp_L1(j,1)));

138
if A3<-1
Stab1(1,i) = 1;
end;
end;
end;
Stab1(2,i) = min(temp_dist1);
end;

% stacked nyquist plots


scrsz = get(0,’ScreenSize’);
figure(’OuterPosition’,[0.2*scrsz(3) 0.2*scrsz(4) 0.35*scrsz(3) 0.5*scrsz(4) ])
plot(real(Nyq1_m),imag(Nyq1_m));
hold on;
plot(-1,0,’+r’,’MarkerSize’,20,’LineWidth’,4);
hold off;
title(’ Nyquist Plot Interface 1 ’);
xlabel(’ Real Axis ’);
ylabel(’ Imaginary Axis ’);
grid on;

139
Appendix C|
Voltage Source Converters

C.1 Two Level Voltage Source Three Phase Con-


verter
There are several ways to implement a three phase converter, in this work the two
level configuration is adopted for it’s simplicity. It is created by combing three
halfbridge converters in parallel, Figure 1.3 shows a half bridge converter and Fig-
ure C.1 shows them together.

Figure C.1: Two Level VSC Diagram

140
The derivations and equations provided as reference in this section are from
and were used to design the microgrid components used in this work [75].

C.2 dq0 Inner Current Control Loop Design


Design is accomplished by transforming the phasor equivalent of the inductor equa-
tion. The constitutive equation for an inductor relates voltage difference to change
in current by Faraday’s law of induction. The derivation of the dq frame equation
follows, Figure C.2 for variable definitions.

d~i
L = −R~i + V~t − V~s
dt
d~i
= −R~i + V~t − V̂s ej(ω0 t+θ0 )
L
dt
From reference dq reference frame theory

f~(t) = [fd + jfq ]ejθ(t)

Substitute and differentiate

d[id + jiq ] dθ(t) jθ(t)


Lejθ(t) +L[id +jiq ] e = R[id +jiq ]ejθ(t) +[Vtd +jVtq ]ejθ(t) −V̂s ej(ω0 t+θ0 )
dt dt

Divide by eθ(t)

d[id + jiq ] dθ(t)


L + L[id + jiq ] = R[id + jiq ] + [Vtd + jVtq ] − V̂s ej(ω0 t+θ0 −θ(t))
dt dt

Let dθ(t)
dt
= ω(t), separate real and imaginary parts, and assume the fundamental
ω(t) is approximately constant

did
L − Lω0 iq = −Rid + Vtd − V̂s cos(ωt + θ0 − θ(t))
dt
" #
diq
j L + Lω0 id = −Riq + Vtq − V̂s sin(ωt + θ0 − θ(t))
dt

141
Control inputs are Vtd and Vtq which in turn are functions of the ideal VSC dq
duty cycles md and mq
Vdc
Vtd = md
2
Vdc
Vtq = mq
2
Rearrange and solve for duty cycles
" #
2 did
md = L − Lω0 iq + Rid + Vd
Vdc dt
" #
2 diq
mq = L + Lω0 id + Riq + Vq
Vdc dt
Decouple the d and q axis by creating new control variables ud and uq

did
ud = L + Rid
dt

diq
uq = L + Riq
dt
These allow d and q currents to be independently regulated by simple PI control
k(s) = kp s+k
s
i
where the plant P (s) relates control input u and current i.
!
u(s) 1 1 1
i(s) = (Ls + R)u(s) → P (s) = = = R
i(s) Ls + R L s+ L

The system naturally contains a stable but slow pole at s = − R


L
, add PI a controller
to cancel it with a zero
!
kp s + ki 1 1
k(s)P (s) = R
s L s+ L

Draw out kp term and rearrange

ki
 
kp  s + kp  kp ki R
k(s)P (s) = R → k(s)P (s) = when =
Ls s + L Ls kp L

142
i(s)
Closed loop with of the transfer function i∗ (s)
= G(s) is therefore

kp
Ls 1
G(s) = kp → G(s) = L
1+ Ls kp
s+1

Closed loop is first order transfer fuction with time constant

L
τ=
kp

After some algebra, proportional and integral constants can be defined from circuit
elements and a selected time constant

L
kp =
τ

R
ki =
τ
The d and q axis equations are identical so these PI constants can be used for both
the i and d axis current controllers. The control output of these controllers ud
and uq can then be substituted in the previously developed equations to calculate
duty cycles md and mq in feedfoward control. See Figure C.3 for the control flow
diagram.
2
md = [ud − Lω0 iq + Vd ]
Vdc
2
mq = [uq + Lω0 id + Vq ]
Vdc

C.3 PMSG control Design


The equations describing the permanent magnet synchronous generator in the dq
reference frame are as follows:

did
Ld = −Rid + Lq ωr iq + V d
dt

diq
Lq = −Riq − Ld ωr id − λm ωr + Vq
dt

143
Figure C.2: Post VSC Filter Diagram

Figure C.3: Current Loop Control

144
Figure C.4: PSMG Control Flow

They are very similar to the equations derived in section C.2 for a VSC with RL
filter. Using the same methods and PI controlled current control the PMSG only
requires modification of the feedforward loop to account for the magnetic flux,
variable rotor speed, and lack of voltage measurement, Figure C.4.

2
md = [ud − Lωr iq + Vd ]
Vdc

2
mq = [uq + Lωr id − λm ωr + Vq ]
Vdc

C.3.1 dq0 Outer Voltage Control Loop Design


In "soft" systems the AC bus voltage and frequency will need to be regulated by the
VSC. This can be accomplished by adding an additional control loop over a VSC
with current control. The AC bus is assumed to contain some capacitance, the
layout of the "soft" system is illustrated in Figure 4.44. The phasor representation

145
of the capacitor on the AC bus is as follows.

dV~s ~ ~
C = i − iL
dt

From reference frame theory

f~(t) = [fd + jfq ]ejθ(t)

Apply the dq transformation to the capacitor equation

d[Vd + jVq ]ejθ(t)


C = [id + jiq ]ejθ(t) − [iL d + jiL q]ejθ(t)
dt

Differentiate and set dt

d[vd + jvq ]
Cejθ(t) + C[vd + jvq ]ejθ(t) jω = [id + jiq ]ejθ(t) − [iL d + jiL q]ejθ(t)
dt

Divide by ejθ(t) and separate real and imaginary components

Vd
C − CωVq = id − iLd
dt

Vq
+ CωVd + iq − iLq
C
dt
Set i = i∗ and solve for inner control loop reference

Vd
i∗d = C − CωVq + iLd
dt

Vq
i∗q = C
+ CωVd + iLq
dt
Let new control variables ud and uq be functions of voltages Vd and Vq

i∗d = ud − CωVq + iLd

i∗q = uq + CωVd + iLq

If the current control loop is fast, τ is small, then the overall voltage transfer
function including the inner current control loop can be described approximately

146
as
V (s) 1
= P (s)
u(s) Cs
where then inner control loop if designed according to section C.2 will simplify to
a first order filter, thus producing

V (s) 1 1
 
=
u(s) τ s + 1 Cs

A PI feedback, k(s), can be used, providing ud and uq from the error of Vd and Vd q
respectively. This results in the closed loop transfer function
1
!
k(s)P (s) Cs
V (s) = 1 V ∗ (s)
1 + k(s)P (s) Cs

The loop gain is


1
k(s)P (s)
Cs
substitue PI control k(s) and closed current control transfer function P (s)

s+z
k(s) = k
s
1
P (s) =
τs + 1
s+z 1 1 k s+z 1
 
l(s) = k → l(s) = −1
s τ s + 1 Cs τC s + τ s2
A method to select constants k and z to allow adequate performance is detailed in
section C.3.2.

C.3.2 Symmetrical Optimum Method


The method of symmetrical optimum allows for design of the PI constants in loop
gain with a double pole at zero [75]. A phase margin δm and time constant τ must
be selected.
1 − τz
sin(δm ) =
1 + τz

147
z can be solved for analytically. Once found, calculation of the constant k is
straight forward

k = C zτ −1

Typical choices for δm are 45◦ and 53◦ .

C.4 Component Sizing

C.4.1 Current Ripple Design


Selection of post VSC inductors were based on limiting the current ripple under
normal operation. The was derived for the half bridge in Figure 1.3. Beginning
with the inductor equation

di
L = vt − vs = ∆V
dt

Integrate and rearrange Z i2


∆V Z τ
di = dt
i1 L 0
Consider the worst case scenario for voltage when source is at crossover point and
switch is one, Figure C.5.
Vdc
∆I = τ
2L
1
Let τ = 2fs
, representing a 50% duty cycle

Vdc
Lmin =
4∆Ifs

Note: fs is the switching frequency of the device

C.4.2 Capacitor Sizing


Sizing of the AC bus capacitor is accomplished using a limitation on the maximum
voltage ripple under nominal conditions. Figure C.6 for a diagram. Beginning with
the fundamental equation of a capacitor.

dv
C = iC
dt

148
Figure C.5: Current Ripple Diagram

Figure C.6: Voltage Ripple Diagram

Z v2 Z τ
C dv = iC dt
v1 0

ic τ
V2 − V1 =
C
Assume a steady maximum current from time 0 to τ

iCmax τ
∆Vmax =
Cmin

iLmax
Cmin =
∆Vmax fs

149
Vita
Peter M. Austin

Education

Ph.D Mechanical Engineering Expected December 2015


The Pennsylvania State University

M.S. Mechanical Engineering December 2011


The Pennsylvania State Univeristy

B.S. Mechanical Engineering May 2009


Rowan University

Publications
"Study of Diminutive and Subsurface Cracks Using SOnic IR Inspection". 34th
Annual Review of Progress in Quantitative Nondestructive Evaluation. American
Institute of Physics. Vol. 975 (2008). pp. 504-511

Projects and Competitions

• NJ Clearn Energy Program: Field survey and analysis of site renewable


power potential.

• ASCE/AISC Steel Bridge Building Competition: Student Member of


steel bridge design and construction team.

• DOE National Collegiate Wind Competition: Mentor for undergrad-


uate teams in wind design competition.

You might also like