Вы находитесь на странице: 1из 244

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/34466506

Vector control of PWM VSI based distributed resources in a microgrid

Thesis · September 2005


Source: OAI

CITATIONS READS
29 504

1 author:

Mahesh S. Illindala
The Ohio State University
96 PUBLICATIONS   1,111 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Resilient Distribution Systems With Microgrids View project

Distributed Energy Resource Planning in Community Microgrids View project

All content following this page was uploaded by Mahesh S. Illindala on 08 July 2014.

The user has requested enhancement of the downloaded file.


VECTOR CONTROL OF PWM VSI BASED
DISTRIBUTED RESOURCES IN A MICROGRID

by

Mahesh S. Illindala

A dissertation submitted in partial fulfillment


of the requirements for the degree of

Doctor of Philosophy

(Electrical Engineering)

at the

UNIVERSITY OF WISCONSIN-MADISON

2005
© Copyright by Mahesh S. Illindala 2005
All Rights Reserved
i

Abstract

Distributed Generation (DG) systems are being increasingly favored for meeting

the supplementary generation demands due to the ever-growing needs of electrical

customers. Many DG systems utilize a voltage source inverter (VSI) based on pulse-

width modulation (PWM) as the utility/load interface and therefore are able to offer

benefits like power quality conditioning in addition to providing of utility grade power.

This thesis addresses the control aspects of these inverter based DGs enabling them to

meet the wide-ranging load demands of the high-end customers by their integration with

the existing generation, transmission and distribution network.

The control of VSI is undertaken in two stages – viz., a multi-loop internal control

using stationary frame vector regulators and an external generation control. The vector

regulators are designed on three-phase complex space-vector quantities. They are used to

regulate the load terminal voltage under balanced as well as unbalanced conditions. High

quality sequence filters are incorporated into the controller for accurate compensation of

transformer impedance effects. The generation control is carried out by means of active

power-frequency and reactive power-voltage controllers. This enables decentralized

operation of the microgrid with load sharing by the DGs and seamless interchange

between stand-alone and interconnected modes of operation of the DG. Investigations on

dynamic (small-signal) stability of a microgrid consisting of n inverter based DG units


ii

are made on chain and parallel topologies. Guidelines are provided for design of the

active power-frequency and reactive power-voltage controllers to meet the IEEE P1547

performance specifications. Computer simulation and laboratory scale experimental

results are used to verify the concepts developed in the thesis.


iii

Acknowledgements

I would like to express my sincere gratitude and thanks to advisor Professor Giri

Venkataramanan for his guidance, counsel and encouragement at every stage of this

research. I thank the National Renewable Energy Laboratory (NREL) and Wisconsin

Electric Machines and Power Electronics Consortium (WEMPEC) for funding this

research. I wish to acknowledge Professor Christopher DeMarco who made helpful

suggestions that contributed in an appendix on the stability of complex transfer functions.

I have been fortunate to work with the WEMPEC professors, students and support

staff who provided an excellent working environment with all the resources that made

this research possible.

I acknowledge the cooperation and support of Paolo, Howard, Mike, Nick, Sean,

Paul, Raahul and Shashank who helped me to their fullest with the laboratory microgrid.

While working in the WEMPEC laboratory, Raymond Marion always had helpful

practical suggestions whenever I approached him.

My stay at UW-Madison was made both enjoyable and memorable by my office-

/lab-mates and a few others - Jackson, Panda, Paolo, Rajesh, Amit, Yongsug, Jun, Bin,

Yusuke, Metin, Rick, Sreesha, Shashank, Sandeep, Bhageerath, Bala, Brian and Andrew.

I enjoyed having technical discussions as well as social get-togethers with them.

I am most grateful to my wife Satya for her encouragement, understanding,


iv

patience and assistance during the significant part of this research work. In addition, I

express thanks to my parents for their efforts, encouragement and inspiration.


v

Table of Contents

Abstract ........................................................................................... i

Acknowledgements ........................................................................ iii

Table of Contents ............................................................................v

List of Figures .................................................................................x

List of Tables ................................................................................xix

Chapter 1 Introduction ................................................................ 1

1.1 ......General.........................................................................................................1

1.2 ......Distributed Generation.................................................................................2

1.3 ......DG Load Requirements and Solutions ........................................................5

1.4 ......Microgrid and Technical Challenges...........................................................7

1.5 ......State-of-Art and Literature Survey ............................................................10

1.5.1 Current/Voltage Regulators ...........................................................12

1.5.2 Harmonic/Negative Sequence Filtering.........................................14

1.5.3 Generation (Active and Reactive Power) Control .........................16

1.6 ......Outline of the Thesis..................................................................................18


vi

Chapter 2 Stationary Frame Vector Regulator ......................... 22

2.1 ......Complex Representation of Three-Phase Quantities .................................22

2.2 ......Complex-Coefficient Transfer Functions ..................................................25

2.3 ......Feedback Controllers for Tracking with Disturbance Rejection in

Balanced Three-Phase Systems .............................................................................26

2.4 ......Evaluation of System Response of Vector Regulated AC Feedback

Control Systems.....................................................................................................32

2.4.1 System Response for a First-Order Plant.......................................32

2.4.2 System Response for a Second-Order Plant ..................................34

2.5 ......Stationary Frame Vector Regulator for Unbalanced Three-Phase Systems39

2.6 ......Equivalent Vector Regulator in the Synchronous Reference Frames........42

2.6.1 Comparative Evaluation between Synchronous and Stationary

Frame Implementations .............................................................................46

2.7 ......Summary....................................................................................................49

Chapter 3 Inverter Internal Controls - Multi-Loop Vector

Regulator 52

3.1 ......Introduction................................................................................................52

3.2 ......Plant Transfer Function Modeling of the DR Filter Interface Network ....53

3.3 ......Controller Architecture ..............................................................................56

3.3.1 Filter Inductor Current Regulator ..................................................58


vii

3.3.2 Filter Capacitor Voltage Vector Regulator....................................61

3.3.3 Command Voltage Modifier..........................................................65

3.4 ......Stability of Feedback Systems Containing Complex-Coefficient Transfer

Functions................................................................................................................73

3.5 ......Simulation Results .....................................................................................79

3.6 ......Summary....................................................................................................86

Chapter 4 Generation Control and Dynamic Behavior of a

Microgrid 89

4.1 ......Introduction................................................................................................90

4.2 ......Generation Control ....................................................................................92

4.2.1 Introduction to Generation Control................................................92

4.2.2 Active Power-Frequency Control ..................................................94

4.2.3 Reactive Power – Voltage (Magnitude) Control ...........................97

4.3 ......Dynamic (Small-Signal) Behavior of a Microgrid ..................................101

4.3.1 DR Interconnected System ..........................................................102

4.3.2 Multiple DRs Connected in a Chain............................................107

4.3.2.1 Special Case: Grid-Interfaced Mode of Operation .............. 115

4.4 ......Design Considerations for Generation Controllers of a DR in a Microgrid116

4.4.1 Active Power-Frequency Controller............................................116

4.4.2 Reactive Power-Voltage Controller.............................................117

4.5 ......Simulation Modeling and Results............................................................119


viii

4.6 ......Summary..................................................................................................127

Chapter 5 Further Studies in Microgrid Dynamic Behavior... 128

5.1 ......Dynamic Behavior of Multiple DRs Connected in Parallel....................129

5.1.1 Special Case: Grid-Interfaced Mode of Operation ......................144

5.1.2 Comments on the Sufficient Conditions for Stability of Parallel

Microgrid .................................................................................................146

5.2 ......Effect of Changing Topologies on the Dynamic Behavior of Microgrid 148

5.3 ......Effect of Variation in Tie-Line R/X Ratio on the Dynamic Behavior of

Microgrid .............................................................................................................152

5.4 ......Summary..................................................................................................156

Chapter 6 Laboratory Microgrid and Experimental Results... 158

6.1 ......Introduction and Objectives of the Laboratory Microgrid.......................158

6.2 ......Microgrid Layout.....................................................................................158

6.3 ......Components of the Laboratory Microgrid ...............................................161

6.4 ......Experimental Results ...............................................................................172

6.4.1 Inverter Internal Controls.............................................................172

6.4.2 External Generation Controls ......................................................178


ix

Chapter 7 Conclusions and Future Work ................................ 185

7.1 ......Contributions ...........................................................................................188

7.1.1 Stationary Frame Vector Regulators............................................188

7.1.2 Inverter Internal Controls - Multi-Loop Regulator......................188

7.1.3 Generation Control in an Inverter Based Distributed Resource

(DR) 189

7.1.4 Dynamic Behavior of Microgrids in Various Topologies ...........189

7.2 ......Proposed Further Research ......................................................................190

7.3 ......Summary..................................................................................................192

Bibliography................................................................................193

Appendix A Battery Energy Storage for Distributed Resources

206

Appendix B Stability of Complex Transfer Functions vis-à-vis

MIMO Systems ........................................................................... 217


x

List of Figures

Figure 1-1 Block diagram of a typical inverter-embedded distributed


resource (DR) unit..............................................................................3

Figure 1-2 Circuit schematic of the three-phase PWM inverter based DR


connected to a load.............................................................................5

Figure 1-3 Single-line diagram of a microgrid consisting of two DRs ..................8

Figure 2-1 Complex vector mapping of three-phase ac sinusoidal signals


under negative sequence unbalanced conditions..............................24

Figure 2-2 Generic block diagram of a feedback controller for an LTI


system...............................................................................................27

Figure 2-3 Bode plot of loop gain TL_1(s) for (a) negative and (b) positive
sequence components of the space vector........................................29

Figure 2-4 A backward coupling form of realization of oscillator Gosc_p(s)


in the αβ reference frame.................................................................31

Figure 2-5 Bode plot of the loop gain TL1(s) (solid line) and closed-loop
gain GCL1(s) (dashed line). (Only positive sequence response
is shown.) .........................................................................................33

Figure 2-6 Complex exponential step response f1(t) of the closed-loop


transfer function GCL1(s) ..................................................................34

Figure 2-7 Bode plot of the loop gain TL2(s) (solid line) and closed-loop
gain GCL2(s) (dashed line). (Only positive sequence response
is shown.) .........................................................................................36

Figure 2-8 Complex exponential step response f2(t) of the closed-loop


transfer function GCL2(s) ..................................................................37

Figure 2-9 Magnitude and angle response of GCL2(s) to a complex


exponential step input ......................................................................37
xi

Figure 2-10 Variation of peak overshoot and settling time for vector
controlled second-order ac plants.....................................................38

Figure 2-11 Bode plot of loop gain TL_2(s)...........................................................40

Figure 2-12 Realization of stationary frame vector regulator...............................41

Figure 2-13 Block diagram illustrating realization of vector regulators (a)


Gcp(s) in anti-clockwise rotating synchronous frame and (b)
Gcn(s) in clockwise rotating synchronous frame ..............................45

Figure 2-14 Comparison of synchronous frame and stationary frame


implementations of vector regulator ................................................48

Figure 3-1 Circuit schematic of the three-phase PWM inverter based DR


system with equivalent load Z’.........................................................54

Figure 3-2 Block diagram of the power circuit of the DR filter-interface


network for a balanced load in orthogonal αβ coordinate
system...............................................................................................55

Figure 3-3 Block diagram of the power circuit of the filter-interface


network for a balanced load for space vector quantities ..................55

Figure 3-4 Block diagram of the effective plant transfer function of the
DR filter-interface network under unbalanced/interconnected
DR conditions for space vector quantities .......................................55

Figure 3-5 Controller architecture for the DR inverter system.............................58

Figure 3-6 Block diagram of the current regulator for (a) space vector
quantities, and (b) orthogonal αβ coordinate quantities ..................59

Figure 3-7 Bode plot of the loop gain for the current regulator loop ...................60

Figure 3-8 Block diagram illustrating realization of filter capacitor


voltage controller Gcv(s) for (a) space vector quantities, and
(b) orthogonal αβ coordinate quantities...........................................62

Figure 3-9 Bode plot of the loop gain for the voltage regulator loop...................63

Figure 3-10 Bode (gain) plot of the output impedance (a) without the
multi-loop controller and (b) with the multi-loop vector
controller. .........................................................................................65
xii

Figure 3-11 Block diagram of the command voltage modifier ............................66

Figure 3-12 Bode plot of the transfer function GNSF(s) for (a) negative
sequence and (b) positive sequence components of the space
vector................................................................................................68

Figure 3-13 Block diagram illustrating realization of negative sequence


filter (NSF) in the command voltage modifier.................................69

Figure 3-14 Bode plot of the loop gain for the negative sequence filter
(NSF) Gfn(s) for (a) negative sequence and (b) positive
sequence components of the space vector iLt(t) ...............................70

Figure 3-15 Block diagram illustrating realization of complex gain ‘-jω60


Lt’ for the output of the NSF in αβ coordinates...............................71

Figure 3-16 Bode plot of the negative sequence filter GNSF(s) for different
values of gain ωo (20, 200 and 2000). (a) negative sequence
and (b) positive sequence components of the space vector..............73

Figure 3-17 Block diagram illustrating realization of negative sequence


filter (NSF) as a unity feedback system ...........................................74

Figure 3-18 Nyquist path on the s-plane and the TL locus for M =
ωo/(2ω60)...........................................................................................75

Figure 3-19 Nyquist path on the s-plane and the TL locus for M =
jωo/(2ω60) .........................................................................................76

Figure 3-20 Bode plot of loop gain TL(s) for M = ωo/(2ω60) for (a)
negative and (b) positive sequence components of the space
vector................................................................................................78

Figure 3-21 Bode plot of loop gain TL(s) for M = jωo/(2ω60) for (a)
negative and (b) positive sequence components of the space
vector................................................................................................78

Figure 3-22 Effect of load imbalance (left) on the unregulated terminal


voltage (right)...................................................................................80

Figure 3-23 Operation of the voltage control loop for vCf with the
response vCf for a complex exponential step input voltage
reference vCf†....................................................................................81
xiii

Figure 3-24 Simulation waveform of line voltage vCf,AC illustrating the


response for a step change in voltage reference vCf† from
70% to 100% of rated voltage of 480 VRMS ..................................81

Figure 3-25 Load terminal line-line voltage ac & bc (top two) and load
current in a & b phases (bottom two) for a balanced 10 kW
resistive load.....................................................................................82

Figure 3-26 Simulation waveforms illustrating DR operation during


transition from no-load to a 2 kW balanced three-phase
resistive. Load terminal line-line voltage ac & bc (top two)
and load current in a & b phases (bottom two). ...............................83

Figure 3-27 Simulation waveforms illustrating DR operation during


transition from a 2 kW balanced three-phase resistive load to
no-load. Load terminal line-line voltage ac & bc (top two)
and load current in a & b phases (bottom two). ...............................83

Figure 3-28 Simulation steady-state waveforms illustrating the load


terminal voltage and load current for an unbalanced load ...............84

Figure 3-29 α- and β- components of transformer primary current, PSF


output and NSF output waveforms for an unbalanced load .............85

Figure 4-1 Voltage regulated DR’s analogy to a voltage source for power
studies...............................................................................................91

Figure 4-2 Generation of voltage command for the inverter based DR ...............92

Figure 4-3 Block diagram of the generation controller for a rotating


machine ............................................................................................93

Figure 4-4 Block diagram of active power-frequency controller for a DR ..........95

Figure 4-5 Simplified block diagram of active power-frequency controller


for a DR............................................................................................96

Figure 4-6 Steady-state voltage droop of a DR in the stand-alone mode of


operation. (a) Deviation quantities (b) actual quantities. .................97

Figure 4-7 Block diagram of the proposed reactive power controller for a
DR ....................................................................................................99

Figure 4-8 Steady-state voltage droop of a stand-alone DR. (a) Deviation


quantities (b) actual quantities. ......................................................101
xiv

Figure 4-9 Representation of single DR connected to the infinite bus...............103

Figure 4-10 Representation of two DR interconnection through an


inductive tie-line.............................................................................105

Figure 4-11 Dominant root locus of the active power-frequency control


system for a microgrid as the system order is increased from
3 through 7 .....................................................................................106

Figure 4-12 Microgrid having DRs in a chain structure.....................................108

Figure 4-13 Simulink model of the microgrid showing two DRs


connected to the grid supply ..........................................................119

Figure 4-14 Internal structure of the DG1+Rload1 subsystem...........................120

Figure 4-15 Simulation waveforms showing response to a change in load


real power demand PL when the DR is operated as a stand-
alone unit........................................................................................121

Figure 4-16 Simulation waveforms showing response to a change in load


real power ref. set-point PL-ref to match the load demand
when the DR is operated as a stand-alone unit ..............................122

Figure 4-17 Simulation waveforms showing response to a change in load


power set-point PL-ref from 0 p.u. to 0.5 p.u. when the DR is
connected to an infinite bus of frequency 59.96 Hz.......................123

Figure 4-18 Simulation waveforms illustrating the effect of load change


at the DR terminals from 0.25 p.u. to 0.5 p.u. while it is
operated in grid-connected mode. Load terminal voltage is
the top waveform and in the bottom are the load current
(solid), DR current (dashed) and tie-line current (dashdot). ..........124

Figure 4-19 Simulation waveforms showing response to a change in load


power set-point PL-ref from 0.5 p.u. to 0 p.u. when the DR is
connected to an infinite bus of frequency 59.96 Hz.......................124

Figure 4-20 Simulation waveforms showing response of DR1 to a change


in load PL1 from 0.2 p.u. to 0.4 p.u. when the two DRs are
interconnected. ...............................................................................125

Figure 4-21 Simulation waveforms showing response of DR2 to a change


in load PL1 from 0.2 p.u. to 0.4 p.u. when the two DRs are
interconnected. ...............................................................................126
xv

Figure 4-22 Simulation waveforms illustrating the two interconnected


DRs supplying their local loads with zero power flowing
along the tie-line. Load terminal voltage is the top waveform
and in the bottom are the DR current (solid) and tie-line
current (dashdot). ...........................................................................126

Figure 5-1 Single-line diagram of a microgrid parallel structure


consisting of several DRs...............................................................129

Figure 5-2 Small-signal state-variable schematic under zero input


conditions of a microgrid having DRs in parallel .........................134

Figure 5-3 Directed graph DB11 ..........................................................................141

Figure 5-4 Undirected graph GB11 ......................................................................142

Figure 5-5 Locus of the dominant eigenvalue of the real power flow in a
5-DR parallel microgrid as ωG5 is reduced below 1 rad./s. ...........148

Figure 5-6 Small-signal equivalence of 3-unit mesh to a 3-unit parallel


microgrid ........................................................................................150

Figure 5-7 Resistance and reactance values of building wires ...........................153

Figure 5-8 Single DR connected to the infinite bus through a tie-line of


R/X ratio σ .....................................................................................154

Figure 5-9 Dominant eigenvalues of the single DR connected to infinite


bus as R/X ratio is increased from 0 through 5..............................155

Figure 6-1 One line diagram of the laboratory microgrid consisting of two
DRs.................................................................................................160

Figure 6-2 Circuit schematic of the PWM inverter based DR unit with its
LCL filter .......................................................................................163

Figure 6-3 Wiring layout of the transformer, loads and the tie-line
connections to the busbar (to be modified for contactor)...............166

Figure 6-4 Distributed generator DG1 supplying power to its local load as
well as tied to the microgrid...........................................................168

Figure 6-5 Connections of the distributed generators DG2 and DG3 to the
microgrid ........................................................................................169

Figure 6-6 Intermediate tie-line load bus junction..............................................170


xvi

Figure 6-7 Static switch interface for utility/grid interconnection of the


microgrid ........................................................................................171

Figure 6-8 Operation of the voltage control loop for vCf with the response
vCf for a complex exponential step input voltage reference
vCf†. DSP obtained from PC Master™ of Metrowerks®
CodeWarrior™. ..............................................................................173

Figure 6-9 Experimental waveform of line voltage vCf,AC illustrating the


response for a step change in voltage reference vCf† from
70% to 100% of rated voltage of 480 VRMS ................................173

Figure 6-10 Load terminal line-line voltage ac & bc (top two) and load
current in a & b phases (bottom two) for a balanced 10 kW
resistive load. Current is scaled at 1 A = 1 V.................................174

Figure 6-11 Experimental waveforms illustrating DR operation during


transition from no-load to a 2 kW balanced three-phase
resistive. Load terminal line-line voltage ac & bc (top two)
and load current in a & b phases (bottom two). Current is
scaled at 1 A = 1 V.........................................................................175

Figure 6-12 Experimental waveforms illustrating DR operation during


transition from a 2 kW balanced three-phase resistive load to
no-load. Load terminal line-line voltage ac & bc (top two)
and load current in a & b phases (bottom two). Current is
scaled at 1 A = 1 V.........................................................................176

Figure 6-13 Experimental steady-state waveforms illustrating the load


terminal voltage and load current when load on phase a is
abruptly disconnected. Current is scaled at 1 A = 1 V...................177

Figure 6-14 Transformer primary current, PSF output and NSF output
waveforms in the PC Master™ for an unbalanced load.................178

Figure 6-15 Experimental waveforms showing response to a change in


load real power demand PL when the DR is operated as a
stand-alone unit ..............................................................................179

Figure 6-16 Experimental waveforms showing response to a change in


load real power ref. set-point PL-ref to match the load
demand when the DR is operated as a stand-alone unit.................180

Figure 6-17 Experimental waveforms showing response to a change in


load power set-point PL-ref from 0 p.u. to 0.5 p.u. when the
xvii

DR is connected to grid interface PCC of frequency 59.96


Hz ...................................................................................................181

Figure 6-18 Experimental waveforms illustrating the effect of load


change at the DR terminals from 0.25 p.u. to 0.5 p.u. while it
is operated in grid-connected mode. Load terminal voltage is
the top waveform and in the bottom are the load current
(solid), DR current (dashed) and tie-line current (dashdot).
Voltage is at 100 V/div. and current is at 20 A/div........................182

Figure 6-19 Experimental waveforms showing response to a change in


load power set-point PL-ref from 0.5 p.u. to 0 p.u. when the
DR is connected to the grid interface PCC of frequency
59.96 Hz. ........................................................................................182

Figure 6-20 Experimental waveforms showing response of DR1 to a


change in load PL1 from 0.2 p.u. to 0.4 p.u. when the two
DRs are interconnected. .................................................................183

Figure 6-21 Experimental waveforms showing response of DR2 to a


change in load PL1 from 0.2 p.u. to 0.4 p.u. when the two
DRs are interconnected. .................................................................183

Figure 6-22 Experimental waveforms illustrating the two interconnected


DRs supplying their local loads with zero power flowing
along the tie-line. Load terminal voltage is the top waveform
and in the bottom are the DR current (solid) and tie-line
current (dashdot). Voltage is at 100 V/div. and current is at
20 A/div..........................................................................................184

Figure A-1 Typical step response of a micro-turbine. PL - load power


demand, PS - response of the micro-turbine...................................207

Figure A-2 Block diagram of a DR system with battery storage........................209

Figure A-3 Cell discharge voltage Vs. time for a lead-acid battery [101]..........212

Figure A-4 Peak current and voltage per cell of a lead-acid battery for a
high-rate discharge [101] ...............................................................213

Figure A-5 Thevenin equivalent circuit of the battery .......................................214

Figure A-6 Instantaneous power curves of a lead-acid battery for a high-


rate discharge [101]........................................................................214
xviii

Figure B-1 Realization of complex transfer function Gcpx(s) as a two-


input two-output MIMO system.....................................................219
xix

List of Tables

Table 2-1 Comparison of linear regulators for unbalanced three-phase


systems .............................................................................................49

Table 3-1 Evaluation of voltage regulator performance under single


phasing .............................................................................................85

Table 3-2 Comparison of frequency/sequence selective filters ............................87

Table 4-1 System parameters for the generation controller................................121

Table 5-1 Eigenvalues of a 5 DR parallel microgrid for the baseline case........147

Table 6-1 Name plate details of the motor drive inverter...................................162

Table 6-2 Inverter LC filter parameters..............................................................162

Table 6-3 Transformer specifications for DR-load interface..............................164

Table 6-4 Digital signal processing platform features [72] ................................167

Table 6-5 Evaluation of voltage regulator performance under single


phasing ...........................................................................................177

Table A-1 Evaluation of voltage regulator performance under single


phasing ...........................................................................................211

Table A-2 PNGV charge sustaining HEV, battery specification [104]..............215


1

Chapter 1 Introduction

1.1 General

Traditional electrical generation systems consisting of mainly thermal, hydel or

nuclear plants with large generation capacities are located at distant locations from load

centers. The energy is transported from the generation plant to the loads through

transmission and distribution systems. These generation, transmission and distribution

equipment are becoming utilized to their maximum capacity. In the meanwhile, the

electrical power sector is continuing to witness a steady growth in the load demand. This

increased load demand requires not just an increase in generation capacity, but also an

expansion of the transmission and distribution infrastructure. Since the current

transmission and distribution network is already massive and widespread, the constraints

of space and right of way have made the system planners to look for alternative avenues

for meeting the excess electrical energy requirements for these loads. One approach that

is favored is the use of distributed generation [1]. Distributed Generation using a variety

of technologies are being considered as a supplementary source of electric power in

addition to the predominantly centralized generation in the emerging electrical utility

infrastructure [2-8]. This research is aimed at broadly addressing issues related to

particular types of distributed generation systems and their control.


2

1.2 Distributed Generation

The term Distributed Generation, or DG, involves utilization of small generators

that are distributed in a power network, to supply the electric power needs of electrical

customers. It can be used for providing energy stabilization, ride-through and

dispatchability [2]. The capacity of these generators may range between a few kW

through a few MW. These generators are located on the utility system, at the site of the

utility customer, or at an isolated site not connected to the power grid. The DG is said to

operate in the grid-interfaced mode when it is connected to the utility, and is said to

operate in stand-alone mode when it is isolated from the utility.

Today’s DG technologies include energy sources such as diesel, combustion

turbine, combined cycle turbine, low-head hydro, fuel cells and renewable sources such

as wind and solar power. While some of these technologies make use of rotating

machines, others employ a power electronic converter to derive utility grade ac power

from the primary energy source [6-8].

A block diagram of a typical power electronic inverter-embedded DG system

with a primary energy source, dc link and a dc-ac converter (or power electronic inverter)

is illustrated in Figure 1-1. As inverter based DG technologies do not involve heavy

rotating masses, they do not have the inertia that would source or absorb transients in

load demand. For this reason, the inverter’s dc link includes an energy storage device to

meet load demands while the primary source availability and/or dynamics precludes load

following [10]. The energy storage feeds to the load requirements during startup or

transient in load demand, until the primary energy source can meet the load demand. The
3

inverter based DGs when employed with suitable controls, are capable of providing a

well-regulated voltage supply against imbalances and ridethrough against short rated

power quality events at the sensitive load bus. When coordinated with proper protection

apparatus, the distributed generation improves reliability of power supplied to sensitive

loads. Thus, use of power electronic converter enables DG systems to provide premium

power [9,10], by locating them near the critical loads. In view of this, they are sometimes

also designated as distributed resources, or DRs.

3 phase
micro- dc dc-ac
source link converter
ac output

Figure 1-1 Block diagram of a typical inverter-embedded distributed resource (DR) unit

Although functionally the inverter in a DR is quite similar to that of a

Uninterruptible Power Supply (UPS) system, the extent to which they will be used to

provide general-purpose utility grade ac power requires their control and protection

features to be suitably designed [11,12]. Furthermore, a power electronic inverter has

been known to be capable of performing functions such as reactive power compensation,

active filtering, voltage sag mitigation etc [13]. When such ancillary functions are added

to the regular generation control in its operation, the DR becomes a more viable choice as

it provides a reliable and high quality power supply at a reasonable cost [14].

The typical inverter embedded DG system shown in the form of a block diagram

in Figure 1-1, is illustrated by a three-phase circuit schematic in Figure 1-2. Each such
4

unit is rated for supplying three-phase loads or single-phase loads at the power frequency,

typically 60 Hz. As seen in the figure, the dc bus is modeled as a stiff dc source (Vdc) in

the presence of adequate dc energy storage to manage load tracking [15] (refer Appendix

A on battery energy storage requirements for DR systems). The Insulated Gate Bipolar

Transistors (IGBTs) and diodes illustrated in the figure are typically operated under

Pulse-Width Modulation (PWM) strategy to synthesize sinusoidal output voltage

waveforms of appropriate frequency and amplitude. PWM switching of the power

electronic converter in each DR necessitates installation of a ripple filter, realized

through the LC filter illustrated in the figure. Furthermore, a delta-wye transformer is

utilized to essentially provide a neutral terminal for the single- and three-phase loads. Use

of the transformer also (a) assists in stepping-down the voltage to a level of the load

rating, (b) it isolates the zero sequence currents on the load side of the network, thus

protecting the inverter during line-ground faults and furthermore, (c) disallows dc

injection from the DR into the network, and (d) provides additional filtering through its

leakage impedance. The generic impedances (Za, Zb and Zc) model the load, representing

combined three-phase and single-phase loads in wye configuration. In addition to them,

the load bus at the transformer secondary may be interfaced to utility grid and/or another

DR bus, in which case the current flowing out to the rest of the system is labeled as Itie.
5

Idc SA1 SB1 SC1


Vdc
VAp VBp VCp

SA2 SB2 SC2

ILfA Lf ILfB Lf ILfC Lf


VCfA VCfB VCfC
Cf Cf Cf
Lt I LtC Lt ILtC Lt ILtC

VNf Itiea
I tieb VanL
Itiec VbnL
VcnL
ILda I Ldb I Ldc
TAC Ta TBA Tb TCB Tc
Za

Zc
Zb
VnL

Figure 1-2 Circuit schematic of the three-phase PWM inverter based DR connected to a load

1.3 DG Load Requirements and Solutions

In general, the loads in a distribution system may be linear, non-linear, single-

phase and three-phase loads as well as those that are sensitive to disturbances in power

quality, in a typical distribution system. While the passive loads, viz., resistive, inductive,

capacitive or their combination constitute linear loads, those employing power

electronics are largely non-linear in nature as they draw non-sinusoidal currents from a

sinusoidal voltage supply. Small capacity loads are typically single-phase and the large

capacity ones are predominantly three-phase loads. Sensitive loads employ hi-tech
6

machinery like computers and electronic control equipment that are susceptible to power

quality disturbances.

Among linear loads, those operating at low power factors are known to cause a

great burden to the utility for providing reactive support. Harmonic currents drawn by

non-linear loads can produce losses and over-heating of electric distribution system

apparatus. Single-phase loads when connected to a three-phase system create an

unbalance. Furthermore, faults in a system cause distortions and imbalances in the

terminal voltage. Recently, it has also been reported that the common power quality

problems such as voltage sags, voltage swells, voltage spikes and short-term outages are

estimated to cost the US economy $26 billion annually [16]. The losses are due to loss of

productivity in the downtime of sensitive loads caused by the power quality events. Thus,

it would be highly desirable for DG systems to mitigate power quality disturbances and

become an integrated solution.

The traditional approach to obtaining high quality power supply for sensitive

loads involves use of UPS systems for conditioning the input voltage. However, the

problems in battery management at high power levels, high capital costs and minimal

utilization under normal operating conditions are some of its drawbacks. These systems

also contain transfer switches whose operation is critical to providing continuous power

to loads [14]. Backup generation systems based on rotating machines can be used as an

alternative system to supply power for sensitive loads during temporary interruptions in

power supply. However, they have a slower response time and also require automatic

transfer switches for a change over between power sources in case of power quality
7

events. During the transfer between main supply and backup, there may be a loss of

continuity of power to the loads and hence these systems are generally unable to meet

stringent power quality indices. Moreover, they cause air pollution and may not be used

for more than a limited number of hours per year due to limits on exhaust emissions [17].

This generally results in suboptimal utilization of capital investment made on them.

While the inverter embedded DG systems can be made free from such concerns,

the power rating of the primary energy source they employ limits their size. This

necessitates clustering of multiple small rated DRs for serving under any load condition.

Moreover, use of multiple small-rated DRs helps in reducing the investment in meeting

the necessary n+1 redundancy requirement. Besides, as mentioned before, generators

when interconnected and dispersed such that they are located near critical loads can

improve the reliability of the power supplied to the critical/sensitive loads. Such a DG

system consisting of several interconnected DRs is called a microgrid [8,14].

1.4 Microgrid and Technical Challenges

A single-line diagram of a typical microgrid consisting of two DRs is illustrated

in Figure 1-3. Under normal operating conditions, the DRs, viz., DR1 and DR2, share the

total load in the system along with the utility grid supply. However, when the utility

supply is down, the DRs are rated to provide all the energy requirements of atleast the

critical loads.

As seen in Figure 1-3, the interconnection between any two DRs is made by

means of a tie-line interconnect that is made up of a three-phase contactor with associated


8

synchronizing and control logic circuitry. Likewise, a three-phase static switch consisting

of thyristors is utilized for the interconnection between the microgrid and the grid mains.

Before connection, the three-phase voltages on both sides of the switch (contactor or

static switch) are matched by the synchronizing logic circuitry. Not shown in the Figure

1-3 is the protection switchgear that is essential in safeguarding equipment and

personnel. Specifications and requirements for the performance, operation, testing,

commissioning, safety, and maintenance at the point of common coupling (PCC) of the

DR interconnection to the Electric Power System (EPS) are laid out in the IEEE Standard

P1547, and UL 1741 [18,19].

Load1

T1 SS 1 T2 Lt1 Lf1

Δ -Y Y-Δ Cf1
DR1
B1
Grid
Interconnect

T ie-line
Interconnect CN1

Load12

B2
B Busbar Load2
C Capacitor
CN Contactor w/ Manual Override T3
L Inductor Lt2 Lf2
R Resistor
SS Static Switch w/ Manual Override
Y-Δ Cf2
T T ransformer DR2
B3

Figure 1-3 Single-line diagram of a microgrid consisting of two DRs

Various design and operational challenges exist in the operation of a microgrid. It


9

needs to be capable of operation in both stand-alone mode as well as in grid-interfaced

mode. The individual inverters should be capable of sharing active and reactive power in

a predetermined manner and circulating power should be avoided. In order to realize this,

it is essential to have good control of the power angle and the voltage level by means of

the inverter. Control of the inverter’s frequency dynamically controls the power angle,

and the flow of the real power. On the other hand, the flow of reactive power can be

regulated by control of the voltage level.

To prevent overloading the inverter and the micro sources, it is important to

ensure that the inverter takes up load changes in a predetermined manner, without

communication. The control of inverters used to supply power to an ac system in a

distributed environment should be based on information available locally at the inverter.

In a system with many DRs, communication of information between systems is

impractical. Communication of information may be used to enhance system performance,

but must not be critical for system operation. Essentially this implies that the inverter

control should be based on terminal quantities.

Over 90% of voltage disturbances in the utility lines are single-phase voltage sags

caused by momentary line to ground faults in distribution systems. Hence the control

strategy of the inverter should be able to meet with situations when the utility grid having

residual imbalances in voltage and when the load system has imbalances in current.

It may be expected that the application centers of distributed generation systems

may consist of nonlinear loads such as rectifiers. The control strategy of the inverters

should be such that they are capable of supplying real and reactive power demands as
10

well as the harmonic currents necessary to feed such loads without exacerbating the

situation.

The broad objective of this research is to address these control issues in

interconnected DG systems through a two-pronged strategy [20] –

(i) inverter internal controls to regulate the load bus voltage even amidst

power quality events

(ii) decentralized generation (active and reactive power) control to allow

seamless as well as stable operation of several generation sources

Firstly, the control of VSI employed in the DR is to regulate the load bus voltage

even in the event of imbalances and enable interconnection to other generation sources. A

stationary frame vector regulator has been designed to regulate load terminal voltage

under unbalanced conditions. This regulator is equivalent to the complex vector regulator

proposed by [21] when implemented in synchronous reference frames rotating in

opposite directions. Secondly, research presented here extends the work to address the

mechanisms for active and reactive power sharing based on frequency droop control and

voltage droop control by providing design guidelines to meet IEEE P1547 performance

specifications [18]. Further, investigations on small-signal stability are carried out on

most common microgrid architectures to ensure their stable operation.

1.5 State-of-Art and Literature Survey

The current state-of-art in the control of commercially available inverter

embedded microturbine DG systems like Capstone and Ingersoll Rand consists of


11

programmable operation in two operation modes, viz., normal grid connect mode and

intentional island mode [22,23]. In the normal grid-interfaced mode of operation, the DR

inverter is connected to the loads in parallel with the grid, and it is controlled as a current

source with the grid providing the voltage as well as frequency references. During grid

supply interruptions, the DR controller senses the loss and immediately disconnects from

both the grid and the loads. When the utility grid returns to within its specified limits, the

DR may be programmed to restart and recommence power to the connected loads. On the

other hand, when the DR is operated in the intentional island mode, it is programmed as a

voltage source based on open-loop voltage control with fixed voltage and frequency

reference. This control strategy causes interruption of power supply to loads as the DR

power needs to be recycled for switching between the two modes of operation.

In inverter based wind energy systems, the control strategy generally adopted

involves current control to output the desired real power in the normal grid-interfaced

mode of operation [24]. These systems require additional information of the grid voltage

through a phase-locked loop (PLL) to generate the reference currents for a desired power

output. It is generally assumed that the grid voltage is stiff so that the real and reactive

power generated can be controlled independently by varying the current vector. As a

result, such systems cannot operate during grid supply interruptions.

The research work conducted here applies an improved control methodology to

overcome the limitations in the existing technology. It regulates the load terminal voltage

to within the specified limits and allows a seamless operation in both grid-interfaced and

stand-alone modes. Furthermore, power quality conditioning capability is added into the
12

control strategy to take complete advantage of the power electronic inverter used as the

utility/load interface. The literature reported relevant to the work presented in this thesis

proposal can be categorized as follows:

The advances made in the control of uninterruptible power supplies (UPS), ac

motor drives and harmonic/sequence filtering has given a tremendous boost to the design

of controllers for voltage source inverter (VSI) based distributed resources (DRs). The

utilization of VSI and control strategies that accomplish a desired closed-loop

performance is the common link among all these applications. For that reason, the

literature presented here includes various advances made in UPS, ac motor drives,

harmonic/sequence filtering that have a bearing on the design of DR controller in

addition to the well established synchronous generator controllers.

1.5.1 Current/Voltage Regulators

Inverter control has generally been inspired by control techniques of

uninterruptible power supplies (UPS) or motor drive systems, by and large feeding

balanced loads [21,25-33]. In the stationary frame UPS control, the state-of-art

controllers utilized P or PI regulators for feedback controller design. These controllers

provide poor performance except at high bandwidth as they do not give a zero steady-

state error for ac sinusoidal quantities. For this reason, the design philosophy is generally

based on regulators operating in the synchronously rotating d-q reference frame that have

been long established for controlling VSIs [25,26,21]. They require transformation of ac

quantities into a synchronously rotating dq reference frame to form dc quantities and then
13

a PI controller is typically employed to achieve a zero steady-state error. However, as a

general-purpose utility grade generator, a DR inverter needs to operate under typical

ambient industrial or residential load conditions that have a broader degree of variability

[34]. These conditions include persistent imbalance in loads, especially due to single-

phase distribution circuits, and steady-state and transient imbalances in the utility grid

voltage [34,35]. Under such conditions the system requires implementation of separate

controllers in positive and negative rotating reference frames that collectively supply the

reference voltage of the inverter, thereby complicating system design further [37].

Recently, some authors have reported regulators in the stationary reference frame that

would bring about zero steady-state error and give a response similar to the synchronous

frame regulator consisting of PI controller [27,32,38,39]. Different schemes of multi-loop

regulators for single- and three-phase UPS that decouple the LC filter dynamics of the

inverter have been reported [29,30,33].

Employing “complex” mathematical models to three-phase systems has several

analytical and computational advantages. It has been applied in the phasor dynamic

representations of three-phase power flow studies [40,41,42]. Recently, high performance

regulators that are devised on three-phase space-vector quantities and realized in the

synchronous frame have been reported [21,36,37,43]. These vector regulators utilize

complex vector modeling of the dynamics of ac electrical machines [44,45] in

synchronously rotating reference frame for the regulator design. They implement a more

desirable pole/zero cancellation as they decouple the cross-coupling between the d and q

axes of the synchronous frame model of the plant [21].


14

On the other hand, the H∞ repetitive controller was employed in [46,47,48,49] for

improved performance of inverter based distributed generators under non-linear loads.

Liang et al [46] presented a comparison of the H∞ repetitive controller with PI regulators

in stationary/synchronous reference frames. However, design of such controllers requires

detailed system design [49] and their response time is approximately 5 cycles following a

disturbance [46]. For current regulation in non-linear loads, Yuan et al [38] and

Allmeling [39] have identified the problems associated with transforming the quantities

into multiple rotating frames and proposed stationary frame transfer functions that enable

fast controller implementations.

Further, the voltage regulator design under conditions such as load/line

imbalances and non-linear loads may require extraction of harmonic/negative sequence

quantities, as discussed in the next section.

1.5.2 Harmonic/Negative Sequence Filtering

The harmonic/negative sequence filtering concepts are of interest in DR control

because typical operating conditions include persistent imbalance in loads, especially due

to single-phase distribution circuits, harmonics due to likely non-linear loads and steady-

state and transient imbalances in the utility grid voltage. Harmonic filtering concepts

have been originally proposed for active filters. The traditional approach to eliminate the

fundamental frequency component in a three-phase current waveform involved use of

notch filters at the fundamental frequency. However, they provide an output that contains

severe distorted harmonics. The instantaneous reactive power theory proposed by Akagi
15

et al [50] circumvented this problem, but it however fails to extract current harmonics

accurately under voltage distorted conditions. An improved method for harmonic

extraction was proposed by Bhattacharya et al [51] involving transformation of the non-

linear current waveform to a synchronously rotating frame such that the fundamental

frequency component becomes a dc quantity in this new reference frame. The dc quantity

is filtered by a low-pass filter and the filtered output is transformed back into the

stationary frame. For extraction of individual harmonics, the same approach can be

extended to transform the three-phase quantities to a rotating frame at the angular

velocity corresponding to the harmonic and filtering the dc quantity [52]. In a lightly

unbalanced three-phase system, such a filter for extraction of a 60 Hz quantity along

negative sequence would also output a significant amount of the positive sequence

content. Reduction of the bandwidth, while improving the steady-state filtering, has poor

dynamic response and hence is unacceptable. Cascaded sections of such complex filters

have also been employed and their complex transfer function representation presented

[53]. Similar methods have been applied in electric machine diagnostics as well [54].

However, use of cascaded filters gives rise to phase errors especially when applied at low

frequencies like 60 Hz.

Fast stationary frame current regulators for active filters have been proposed by

Yuan et al [38] and Allmeling [39]. They employed generalized stationary frame

integrators to realize equivalent integral controllers for harmonic filtering that would

otherwise have been implemented in rotating reference frame quantities.

Sequence Filtering has also been employed in detecting spatial saliencies, i.e.,
16

asymmetries in ac machines. Information on the saliency, i.e., imbalance in the leakage

inductance of the ac machine can be detected from the positive and negative sequence

carrier signal current components drawn by the machine when a balanced high frequency

carrier voltage is applied [55,56]. The three-phase currents are transformed to rotating

reference frame and low-pass/high-pass filters are employed for filtering of the required

component [57,53,58].

1.5.3 Generation (Active and Reactive Power)

Control

The generation control in VSI based DRs follows the concepts originally

proposed in parallel/distributed operation of interconnected UPS units. Parallel operation

of multiple UPS units to achieve a higher power rating as well as better reliability

through redundancy has been presented by Holtz et al [59]. In this multiple-inverter UPS

system, each UPS module has its own controller that communicates with the controllers

of other modules. Chandorkar [60] proposed a decentralized generation control strategy

for distributed UPS units with no communication between them. The inverters’

controllers include active power-frequency droop characteristics similar to those of large

ac synchronous machines [61,62], but with no equivalent inertia. Based on the power

ratings of the individual UPS units, design guidelines are formulated [60].

In power flow studies conducted on an interconnected distributed generator

system, having arbitrarily fixed voltage references for the DR can cause excessive
17

reactive power circulation between the generators [63]. This fact has been recognized

earlier in the paralleling of synchronous generators as well as SVC/Statcom [64]. In order

to minimize the reactive power flow between the two generators due to discrepancies in

the voltage set-points, it is necessary to have soft set-points that can be realized by a

reactive power-voltage controller.

Tuladhar et al [65] proposed a variation for control of parallel operation of single-

phase inverters in a distributed ac power system without control interconnections. For

reactive power sharing, a small ac voltage signal is injected into the system as a control

signal. The reactive power to be shared among the paralleled inverters determines the

frequency of this control signal. The small active power due to the control signal is made

use of in adjusting the voltage magnitude of the inverter. Similar approach is used for

sharing of distortion power in the presence of harmonic loads.

Liang et al [48] proposed a hybrid control for an island mode distributed

generation system. Inverters connected to the same node of the distribution system are

controlled in a master-slave framework with the master having a voltage controller and

the slaves consisting of current controllers with communication between all of them. H∞

repetitive controllers are employed for the voltage/current control. On the other hand,

power sharing for inverters at remote nodes, where it is difficult to have communication,

is achieved through traditional means like frequency droop.

Piagi [66] presented microgrid controls that enable inverters to track either active

power set-point or frequency based on the requirement. Furthermore, a mechanism to

vary the power set-point is proposed when an inverter based DR’s power limits are
18

reached.

On microgrids containing combustion turbines, Cardell and Tabors [67] have

reported that microgrids that were stable with only couple of combustion turbines became

unstable when more combustion turbines were added in the system.

1.6 Outline of the Thesis

This thesis begins with an analysis of the frequency response characteristics of

three-phase complex space-vector quantities and devises a stationary frame vector

regulator that provides zero steady-state error for three-phase systems even under

unbalanced conditions. It presents a multi-loop stationary frame control strategy using

vector regulators for voltage regulation at the sensitive load bus of a distributed

generation system in the presence of imbalances.

Several propositions have been made from dynamic (small-signal) behavior

studies conducted on various microgrid architectures to ensure its stability. A design-

oriented approach is presented for designing the active power-frequency droop and the

reactive power-voltage droop characteristics in a DR to enable its seamless operation in

stand-alone as well as interfaced modes. Design variables are correlated to the IEEE

P1547 performance specifications [18]. Simulation and experimental results on a

laboratory microgrid test-bed consisting of two VSI-based DRs are presented for

verification of the proposed control strategies.

The organization of this report is as follows:

Chapter 2 describes the characteristics of the complex vector quantities and their
19

transfer functions for a three-phase system. Requirements of a feedback controller for

tracking and disturbance rejection of three-phase ac space-vector quantities are discussed.

Stationary frame based vector controllers are proposed that give zero steady-state error

for space-vector quantities are proposed. Effects of second-order plants on the

performance of the vector regulated system are discussed. Realization and comparative

evaluation against equivalent regulators in synchronously rotating reference frames are

presented.

In Chapter 3, a multi-loop stationary frame control strategy for the DR inverter is

proposed, based on the stationary frame vector controller, for regulation of the voltage at

sensitive load bus under load as well as line imbalances and faulted conditions. It consists

of an inner current loop for the filter inductor current and an outer voltage loop for the

filter capacitor voltage. State and disturbance feedback decoupling is employed to make

the system immune to varying microgrid network configurations as well as to different

load conditions. Frequency domain analysis techniques based on Bode plots for both

positive and negative sequence quantities are employed for the design of the regulators in

the stationary frame. A high fidelity complex filter is proposed for extraction of

positive/negative sequence components in the three phase system. Moreover, stability

analysis of the complex filter is presented by means of Nyquist and Bode diagrams.

Finally, selected results from simulation of the vector controller in Matlab®

SIMULINK™ are presented. Various practical conditions including imbalances are

simulated to verify the steady-state and dynamic response of the vector regulator.

A generation control scheme consisting of active power-frequency and reactive


20

power-voltage controllers is presented in Chapter 4. Assuming every DR in a microgrid

is equipped with such a generation controller, dynamic (small-signal) behavior analysis

for microgrids having a chain topology are also covered in this chapter. Sufficient

conditions are developed by means of eigenvalue analysis to guaranty small-signal

stability of n DR microgrids. These studies consider both real and reactive power flows

in the microgrid. The grid-interfaced mode of operation of the n DR microgrid is also

covered. Guidelines are provided for design of the active power-frequency droop and

reactive power-voltage droop characteristics. Besides the results of dynamic behavior

analysis, the IEEE P1547 performance specifications [18] are also utilized in the design

process. Finally, selected results from simulation of DRs equipped with generation

controls in Matlab® SIMULINK™ are presented.

Chapter 5 contains further studies on the dynamic (small-signal) behavior of

microgrids. The parallel connection of DRs in the microgrid to a point of common

coupling (PCC) is dealt in this chapter. Effect of change in topology to a mesh is

analyzed for the simplest case of a 3-DR mesh network based on its analogy to the 3-DR

parallel network. Since a non-zero R/X ratio is typical of wires in medium voltage

distribution systems, effects of R/X ratio are investigated in the active and reactive power

flow.

The assembly and layout of the laboratory microgrid is given in Chapter 6. It

contains the design and specifications of various equipment used. Selected experimental

results demonstrating the performance of the DRs controlled with the internal vector

regulator and external generation controls are also presented in this chapter.
21

Finally, the conclusions of this dissertation and further research are presented in

Chapter 7.
22

Chapter 2 Stationary Frame Vector


Regulator

It is well known that three-phase quantities can be combined to form a single

‘complex’ space-vector quantity [68]. However, the application of complex transfer

functions in analysis and developing controls and regulators is relatively new

[21,36,39,43]. Such transfer functions have complex coefficients of s and operate on

complex space-vector quantities. This chapter reviews the complex signals and systems

for three-phase power systems and subsequently develops feedback controllers that offer

zero steady-state error at any arbitrary frequency under balanced and unbalanced

conditions. The focus is on establishing reference frame independence and benchmark

responses. Frequency-domain approach based on Bode plots is employed to design the

vector controllers. Furthermore, the time response of proposed controller is evaluated on

first-order as well as second-order plants. Relation between the time response

characteristics such as peak overshoot or settling time and the phase margin information

available from the Bode plots is analyzed.

2.1 Complex Representation of Three-Phase

Quantities

Generally, three-phase voltages and currents can be represented in a complex


23

space-vector form. If fx(t) [x = a, b, c] denotes the instantaneous value of either a voltage

or current, the complex space-vector is obtained from the three-phase quantities as

[44,45]

2
f(t) = 3 {fa(t) + γ fb(t) + γ2 fc(t)} (2-1)

where γ = ej2π/3 and γ2 = e-j2π/3 and j = -1.

The space-vector can also be expressed in Clarke’s αβ coordinates as

f(t) = fα(t) + j fβ(t) (2-2)

⎡ fα(t) ⎤ ⎡ fa(t) ⎤
where ⎢ ⎥ = [C] ⎢ fb(t) ⎥ (2-3)
⎣ fβ(t) ⎦
⎣ fc(t) ⎦
-1 -1
⎡1 ⎤
and
2⎢
[C] = 3
2 2
⎥ (2-4)
⎢0 3 - 3⎥
⎣ 2 2 ⎦

It is to be noted that the above transformation can be made power invariant by

replacing 2/3 with 2/3.

In the presence of negative sequence unbalanced quantities, which are common in

three-phase three-wire power distribution systems, the complex space-vector in polar

form can be represented as

f(t) = Ap ejω60t + An e-jφ e-jω60t (2-5)

or f(t) = Ap cos(ω60 t) + An cos(ω60 t + φ) + j {Ap sin(ω60 t) - An sin(ω60 t + φ)} (2-6)

Here fα(t) = Ap cos(ω60 t) + An cos(ω60 t + φ) (2-7)


24

and fβ(t) = Ap sin(ω60 t) - An sin(ω60 t + φ) (2-8)

Figure 2-1 illustrates the three-phase quantities fx(t) (x = a, b, c) and their

complex space-vector f(t). The trajectory followed by the space-vector f(t) is an ellipse

with the major and minor axes of length (Ap + An) and (Ap – An), respectively. The

Fourier spectrum of the complex space-vector is also displayed in Figure 2-1.

fb(t) β Ap - An
fa(t)

t f
φ/2
α

fc(t) Ap + An

Three-phase quantities Space-vector trajectory

Magnitude Phase

Ap
An
ω ω
−ω60 ω60 −ω60 ω60

Fourier spectrum of space-vector


Figure 2-1 Complex vector mapping of three-phase ac sinusoidal signals under negative sequence
unbalanced conditions

Thus, the three-phase quantities can be represented by a single complex space-

vector. Likewise, three-phase physical systems like ac electrical machines that are multi-

input/multi-output (MIMO) systems can be modeled as single-input/single-output (SISO)


25

systems using complex-coefficient transfer functions [44,45]. Filters with complex

transfer functions have also been reported in the fields of communications and signal

processing [69,70].

2.2 Complex-Coefficient Transfer Functions

Consider the transfer function of a linear time-invariant (LTI) system for space-

vector quantities given by

N(s)
G(s) = D(s) (2-9)

The transfer function is a frequency domain representation linear time invariant

systems that can be described by an ordinary differential equation (ODE) with constant

coefficients assuming zero initial conditions [71]. For an ODE with constant coefficients

in the real plane R, the transfer function G(s) has poles and zeros that are either real or

complex conjugate pairs. For example, a wye-connected impedance Z(s) (= R + s L) per

phase illustrates this case. The ODE in space-vector variables illustrating the response of

series RL load is

d
v(t) = R i(t) + L dt i(t) (2-10)

Instead, consider the ODEs for a three-phase induction machine in the stator

(fixed) reference frame for space-vector quantities [44,45,27]

⎛ d⎞ Md
vs(t) = ⎜Rs + σ Ls dt⎟ is(t) + L dt ψr(t) (2-11)
⎝ ⎠ r

⎛ d⎞
0 = Rr ir(t) + ⎜- j ωr + dt⎟ ψr(t) (2-12)
⎝ ⎠
26

where ωr is the rotor speed in electrical rad./s, ψr(t) is the rotor flux-linkage vector and σ

= 1 – M2/(LsLr).

The ODEs in (2-11) and (2-12) have constant coefficients in the complex plane C

if the rotor speed is maintained constant. Accordingly, they represent an LTI system that

can be expressed with complex-coefficient transfer functions. The transfer function

between the stator current and voltage vectors is given by [27]

Ιs(s) s + (1/τr – j ωr)


GIM(s) = = 2 (2-13)
Vs(s) s σ Ls + s (Rs + Ls/τr – j σ Ls ωr) + Rs (1/τr – j ωr)

where τr = Lr/Rr, is the rotor time-constant.

Transfer function GIM(s) in (2-13) contains zeros/poles that do not occur in

complex conjugate pairs. Typically, transfer functions with real coefficients are

formulated for either the real or imaginary parts of the system, rather than viewing the

complex system in its entirety. Nevertheless, the complex-coefficient transfer functions

can be used to process the three-phase complex space-vector quantities.

The next section deals with the development of complex vector regulators

reference frame independent for balanced systems.

2.3 Feedback Controllers for Tracking with

Disturbance Rejection in Balanced Three-Phase

Systems

Briz et al [21] have introduced complex vector regulators for control of balanced

three-phase ac machines. The synchronous frame complex vector regulator proposed in


27

[21] gives improved performance as compared to the synchronous frame PI regulator.

However, it requires transformation of all the feedback quantities into a synchronous

rotating reference frame by Park’s transformation [68]. This section presents an approach

to realize the same complex vector regulator in the stationary reference frame quantities.

The block diagram of a feedback controller for a three-phase system (plant) in

complex space-vector quantity variables is illustrated in Figure 2-2. Assuming the plant

with input fin(t) and output fout(t) to constitute an LTI system, it can be represented by a

transfer function Gp(s) (say). A complex space-vector regulator transfer function Gc(s)

can be designed for specifications such as steady-state error and bandwidth in regulating

the output fout(t) to a desirable value fref(t). Moreover, the controller must provide a high

quality rejection of disturbance fdx(t) at the signal excitation frequency.

fin(t)
fdx(t)
fref(t) fout(t)
Gc(s) Gp(s)

TL(s)
Figure 2-2 Generic block diagram of a feedback controller for an LTI system

Assuming a unity gain feedback, the loop gain of the feedback control system in

Figure 2-2 is given by [71]

TL(s) = Gc(s) Gp(s) (2-14)

If Laplace Transform is designated by _{.}, and if _{fref(t)} = Fref(s), _{fdx(t)} =

Fdx(s) and _{fout(t)} = Fout(s), the Laplace Transform of the output is [71]

TL(s) Gp(s)
Fout(s) = 1 + T (s) Fref(s) + 1 + T (s) Fdx(s) (2-15)
L L
28

It is known that balanced three-phase systems contain only positive sequence

quantities. If it is desired to regulate with a zero steady-state error as well as achieve

disturbance rejection of three-phase 60 Hz sinusoidal quantities, the complex space-

vector regulator transfer function Gc(s) can be determined to achieve a loop gain of

ωo
TL_1(s) = (2-16)
s – jω60

where ωo denotes the bandwidth of the system.


In this case, TL_1(s) ⎪⎪ → ∞ (2-17)
⎪s = +jω60

The closed-loop transfer function provides a unity gain at the frequency of ω60

rad./s and is determined as

1
GCL_1(s) =
⎛s - j ω60⎞ (2-18)
1+⎜ ⎟
⎝ ωo ⎠

Figure 2-3 illustrates the Bode plots of the transfer function TL_1(s) for space-

vector quantities with solid lines denoting positive sequence response and dashed lines

denoting negative sequence response. As seen in this figure, the Bode plot displays a

varied frequency response to the positive sequence and negative sequence quantities. The

frequency response for the positive sequence complex exponential quantities is indicated

by the characteristics for ω > 0, and those for negative sequence complex exponential

quantities by the characteristics for ω < 0.


29

60 60

40 40
Magnitude (dB)

Magnitude (dB)
20 20

0 0

20 20

40 40

60 60
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

135 135

90 90
Phase Angle (deg)

Phase Angle (deg)


45 45

0 0

45 45

90 90

135 135
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

(a) (b)
Figure 2-3 Bode plot of loop gain TL_1(s) for (a) negative and (b) positive sequence components of the
space vector.

Consider a plant consisting of three-phase wye-connected series RL load. The

plant transfer function between the applied voltage and the current drawn by the series

RL load is

1
Gp(s) = R + s L (2-19)

The complex space-vector regulator transfer function for the plant Gp(s) to

achieve the loop gain TL_1(s) is determined as

TL_1(s)
Gc_1(s) = ^ (2-20)
Gp(s)

^ ^
ωo(R + s L)
which gives Gc_1(s) = (2-21)
s – jω60
30

⎛ ^
ωo R⎞⎟ ⎛ 1
⎜ ^ ⎞
Gc_1(s) = ωo L + s ⎜
or ⎝ ⎠ ⎛– jω60⎞⎟ (2-22)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ s ⎠⎠

⎛ Ki⎞
i.e., Gc_1(s) = ⎜Kp + s ⎟ Gosc_p(s) (2-23)
⎝ ⎠

ω R
^
^
where Kp = ωo L, Ki = o (2-24)
s

⎛ 1 ⎞
Gosc_p(s) = ⎜
and ⎛– jω60⎞⎟ (2-25)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ s ⎠⎠

In the above equations, the estimated parameters of the series RL load used in the

controller are designated by a hat ‘^’.

The complex space-vector regulator Gc_1(s), thus derived, would be identical to

the complex regulator proposed by Briz et al [21], if transformed to a synchronously

rotating reference frame.

Thus, a stationary frame vector regulator requires an oscillator Gosc_p(s) cascaded

with PI regulator to become capable of tracking positive sequence quantities along with

disturbance rejection. Oscillator Gosc_p(s) can be realized in a backward coupling form as

illustrated in Figure 2-4. It consists of two integrators along the coupled feedback paths

for α and β quantities.


31

fα(t) wα(t)

uα(t) -ω60 ⌠

uβ(t) ω60 ⌠

fβ(t) wβ(t)

Figure 2-4 A backward coupling form of realization of oscillator Gosc_p(s) in the αβ reference frame

It can be shown that oscillator Gosc_p(s) produces forced oscillations of increasing

amplitude for only positive sequence quantities in the input even with zero initial

conditions to the integrators. Likewise, if it is desired to instead perform identical

functions for negative sequence quantities, a simple reversal of α and β terminals can

achieve it. As a result, the negative sequence oscillator would have the transfer function

⎛ 1 ⎞
Gosc_n(s) = ⎜
⎛jω60⎞⎟ (2-26)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ s ⎠⎠

An observation from the Bode plots of loop gain TL_1(s) (refer Figure 2-3) is the

occurrence of two gain crossover frequencies – one of positive slope and the other of

negative slope. The next section analyses the impact of such occurrences on the dynamic

response of the vector regulated systems.


32

2.4 Evaluation of System Response of Vector

Regulated AC Feedback Control Systems

2.4.1 System Response for a First-Order Plant

Consider a closed-loop system employing the vector regulator for control of a

first-order plant. The vector regulated system tracks an ac complex exponential input of

frequency ω60 = 2π(60) rad./s. The loop gain of such a unity feedback gain control

system is

Gosc_p(s)
TL1(s) = (2-27)
⎛s⎞
⎜ω ⎟
⎝ o⎠

⎛ 1 ⎞
Gosc_p(s) = ⎜
where ⎛– jω60⎞⎟ (2-25)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ s ⎠⎠

Closed-loop transfer function of the feedback system with a loop gain TL1(s) is

given by

1
GCL1(s) =
⎛ - j ω60⎞
s (2-28)
1+⎜ ⎟
⎝ ωo ⎠

Bode plots of the loop gain TL1(s) and closed-loop gain GCL1(s) for positive

sequence quantities alone are plotted in Figure 2-5 for ωo = 2π(40) rad./s. In this figure,

the loop gain plot is displayed by a solid line and closed-loop gain by a dashed line. As

seen in the figure, the loop gain plot has an infinite magnitude at frequency ω60 = 2π(60)

rad./s and therefore the closed-loop gain is unity with zero phase delay at this frequency.
33

Furthermore, the gain-crossover frequencies are displaced equally on either side of the

center frequency ω60 by ±ωo and the phase margin is 90o at both these gain-crossover

frequencies. The loop gain crosses with rising edge of single slope at the lower gain-

crossover frequency of (ω60 - ωo) with a phase advance of 90o and it crosses with falling

edge of single slope at the higher gain-crossover frequency of (ω60 + ωo) with a phase

delay of 90o.

20 180
135
90

Phase Angle (deg)


Magnitude (dB)

45
0 0
45
90
135 φm
20 180 3
10 100 1 .10
3 10 100 1 .10
Frequency (Hz) Frequency (Hz)

Figure 2-5 Bode plot of the loop gain TL1(s) (solid line) and closed-loop gain GCL1(s) (dashed line). (Only
positive sequence response is shown.)

To determine the complex exponential step response for positive sequence, the

input quantity to the complex filter is given by

⎧0 for t ≤ 0
cu(t) = ⎨ j ω t (2-29)
⎩e 60 for t > 0

Taking Laplace Transforms,

1
CU(s) = _ {ej ω60 t u(t)} = (2-30)
s – j ω60

The complex exponential step response f1(t) of closed-loop system with transfer

function GCL1(s) for complex exponential positive sequence ac quantities of frequency


34

ω60 = 2π(60 Hz) is therefore determined as

⎧⎪ GCL1(s) ⎫⎪ -1⎧⎛ 1 ⎞⎛ 1 ⎞⎫⎪


f1(t) = _ -1⎨ ⎬ = _ ⎪⎜ ⎟⎜ ⎟ (2-31)
⎩⎪s – j ω ⎪
60⎭ ⎨⎝ s – j ω60⎠ ⎛ s - j ω60⎞ ⎬
⎜ 1 + ⎜ ⎟ ⎟
⎩⎪ ⎝ ⎝ ωo ⎠⎠⎭⎪

i.e., f1(t) = ejω60t (1 - e-ωot) (2-32)

Figure 2-6 illustrates the complex exponential step response f1(t) for closed-loop

transfer function GCL1(s). As seen in the figure, the response is of first order with a

bandwidth of ωo rad./s and a settling time of Ts = 5/ωo seconds. Furthermore, f1(t) has no

overshoot as the phase margin is 90o at the gain-crossover frequencies.

1.5

0.5
β
f1α(t) t
1.5
0 0.012 0.024 0.036 0.048 0.06
1 0.5
f1
1
0.5
1.5
α
1.5 1 0.5 0 0.5 1 1.5 1.5
0.5
1
1 0.5
t
1.5 f1β(t)
0 0.012 0.024 0.036 0.048 0.06
0.5

1.5
Figure 2-6 Complex exponential step response f1(t) of the closed-loop transfer function GCL1(s)

2.4.2 System Response for a Second-Order Plant

While the previous section demonstrated the response of a first order plant, higher

order physical systems produce overshoot and oscillations in their time response. As
35

second-order plants are the simplest systems that exhibit such a behavior, consider a

closed-loop system employing the vector regulator for control of a second-order plant.

The vector regulated system tracks an ac complex exponential input of frequency ω60 =

2π(60) rad./s. The loop gain of such a feedback control system is

Gosc_p(s)
TL2(s) = and (2-33)
s ⎛ s⎞

ωo ⎝
1+ ⎟
ω2⎠

the closed-loop transfer is given by

1
GCL2(s) = (2-34)
⎛s – jω60⎞ s ⎛s – jω60⎞
1+⎜ ⎟+ ⎜ ⎟
⎝ ωo ⎠ ω2 ⎝ ωo ⎠

Bode plots of the loop gain TL2(s) and closed-loop gain GCL2(s) for positive

sequence quantities alone are plotted in Figure 2-7 for ωo = 2π(40) rad./s and ω2 = 2π(60)

rad./s. As seen in the figure, the loop gain plot has an infinite magnitude at frequency ω60

= 2π(60) rad./s and therefore the closed-loop gain is unity with zero phase delay at this

frequency. However, contrary to the earlier case, the gain-crossover frequencies are not

displaced equally from the center frequency ω60, and the phase margin is more than 90o

at the positive slope and less than 90o at the negative slope. As a result, the closed-loop

response shows peaking at the negative slope gain-crossover frequency alone giving rise

to asymmetry in its Bode plot about 60 Hz frequency. This is in contrast to the

conventional second-order system response in dc systems, which exhibit a symmetric

frequency response about the zero frequency and have an identical phase margin for

positive and negative frequencies.


36

20 180
135
90

Phase Angle (deg)


Magnitude (dB)

45
0 0
45
90
135
φm
180 3
20 3 10 100 1 .10
10 100 1 .10
Frequency (Hz) Frequency (Hz)

Figure 2-7 Bode plot of the loop gain TL2(s) (solid line) and closed-loop gain GCL2(s) (dashed line). (Only
positive sequence response is shown.)

The complex exponential step response f2(t) of closed-loop system with transfer

function GCL2(s) for complex exponential positive sequence ac quantities of frequency

ω60 = 2π(60 Hz) is given by

⎪⎧ GCL2(s) ⎪⎫
f2(t) = _ -1⎨ ⎬ (2-35)
⎪⎩s – jω60⎪⎭

After simplification, the time response may be expressed as

⎡ 1
f2(t) = ejω60t ⎢1 +
2γ2
{(γ1 - γ2) e(γ1 + γ2) t - (γ1 + γ2) e(γ1 − γ2) t}⎤⎥⎦ (2-36)

(ω2 + jω60)
where γ1 = - 2 (2-37)

(ω2 - jω60)2 - 4ω2(ωo - jω60)


and γ2 = 2 (2-38)

Figure 2-8 illustrates the complex exponential step response f2(t) for closed-loop

transfer function GCL2(s) at ωo = 2π(40) rad./s and ω2 = 2π(60) rad./s. The absolute value

and argument of f2(t) are plotted in Figure 2-9(a) and (b), respectively. As seen in these

figures, the complex exponential step response displays an oscillatory response in both
37

magnitude and angle derivative before settling at unity magnitude and 2π(60 Hz)

frequency.

1.5

1
β
0.5
1.5
f2α(t) t
1 0 0.012 0.024 0.036 0.048 0.06
f2 0.5
0.5 1
α 1.5
1.5 1 0.5 0 0.5 1 1.5 1.5
0.5
1
1 0.5

1.5 f2β(t) t
0 0.012 0.024 0.036 0.048 0.06
0.5

1.5
Figure 2-8 Complex exponential step response f2(t) of the closed-loop transfer function GCL2(s)

1.25 3.14

1
1.57
0.75
|f2(t)| ∠f2(t) t
0.5 0 0.012 0.024 0.036 0.048 0.06

1.57
0.25

t
3.14
0 0.012 0.024 0.036 0.048 0.06

(a) (b)

Figure 2-9 Magnitude and angle response of GCL2(s) to a complex exponential step input

A time response such as the one in Figure 2-8 and Figure 2-9 is associated with

the characteristics of peak overshoot (Pk) and settling time (Ts) that vary as a function of

frequencies ω2 and ωo. Figure 2-10 illustrates the results obtained for a 3% tolerance in
38

magnitude as well as angle using MathCAD® software. The phase margin was affected

by varying the parameter ω2 for different values of ωo.


1.45 0.35
ω o = ω 60/5
1.4 ω o = ω 60/3 ω 2/ ω o = 2
0.3 ω 2/ ω o = 4
ω o = ω 60
1.35 2nd order dc s/m ω 2/ ω o = 8
Peak overshoot (p.u.)

0.25 ω 2/ ω o = 16
5/ ω

Settling time (s)


1.3 o

1.25 0.2

1.2 0.15

1.15
0.1
1.1 ω60
0.05
1.05

1 0
0 10 20 30 40 50 60 70 80 90 50 100 150 200 250 300 350 400

Phase margin (deg) ωo (rad./s)

Figure 2-10 Variation of peak overshoot and settling time for vector controlled second-order ac plants

As seen in Figure 2-10, the peak overshoot is affected by the phase margin (φm)

that can be easily determined from the Bode plot of loop gain TL2(s) (refer Figure 2-7).

The peak overshoot plot also contains as baseline that of second-order dc systems. As

seen in this plot, the peak overshoot for second-order ac vector controlled systems is

lesser than that for the second-order dc systems at phase margins less than 70o. Besides

peak overshoot, the settling time for second-order ac vector control systems is also

plotted against ωo in the same Figure 2-10. In this plot, however the first-order systems

that produce ideal response with a settling time equal to 5/ωo are used as the baseline. As

seen in Figure 2-10, the settling time varies considerably only for values of ωo

significantly less than ω60 if ω2/ωo < 8.

Until now this chapter dealt with balanced three-phase systems. The next section

deals with unbalanced three-phase systems that are widespread throughout in power

distribution systems.
39

2.5 Stationary Frame Vector Regulator for

Unbalanced Three-Phase Systems

The stationary frame vector regulator dealt in an earlier section gives high

performance with zero steady-state error for balanced systems alone. Its transfer function

was

⎛ Ki⎞
Gc_1(s) = ⎜Kp + s ⎟ Gosc_p(s) (2-23)
⎝ ⎠

However, in power system applications, persistent imbalance in loads, especially

due to single-phase distribution circuits, and steady-state and transient imbalances in the

utility grid voltage is quite common [34,35]. Therefore, it is generally desired to regulate

both positive and negative sequence components of three-phase 60 Hz sinusoidal

quantities with zero steady-state error. Under such conditions, the controller is to be

designed to achieve a loop gain of the sum of two terms each giving an infinite gain for

each sequence component. Hence,

ωo/2 ωo/2 ωo s
TL_2(s) = + = 2 (2-39)
s – jω60 s + jω60 s + ω602


and TL_2(s) ⎪⎪ → ∞ (2-40)
⎪s = ± jω60

Figure 2-11 illustrates the Bode plot of the transfer function TL_2(s). Since this

loop gain is a real transfer function its characteristics are identical for positive and

negative frequencies; and hence Bode plot is displayed here for only positive frequencies.

In this plot, ωo was chosen as 400. As seen in the figure, the infinite gain at ω60 = 2π(60)
40

rad./s implies a zero steady-state error for both the sequence quantities of 60 Hz

frequency.

The vector regulator transfer function for the plant Gp(s) to achieve the loop gain

TL_2(s) is determined as

TL_2(s)
Gc_2(s) = ^ (2-41)
Gp(s)

^ ^
ωo s (R + s L)
which gives Gc_2(s) = (2-42)
s2 + ω602

⎛ ^
ωo R⎞⎟ ⎛ 1
⎜ ^ ⎞
Gc_2(s) = ωo L + s ⎜ 2⎟
or ⎝ ⎠ (2-43)
⎜1 + ⎛⎜ω60⎞⎟ ⎟
⎝ ⎝ s ⎠⎠

⎛ Ki⎞
i.e., Gc_2(s) = ⎜Kp + s ⎟ Gosc(s) (2-44)
⎝ ⎠

ω R
^
^
where Kp = ωo L, Ki = o (2-24)
s

⎛ 1 ⎞
Gosc(s) = ⎜ 2⎟
and ⎜1 + ⎛⎜ω60⎞⎟ ⎟ (2-45)
⎝ ⎝ s ⎠⎠

60 135

40 90
Phase Angle (deg)
Magnitude (dB)

20 45

0 0

20 45

40 90

60 135
3 4 3 4
0.01 0.1 1 10 100 1 .10 1 .10 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

Figure 2-11 Bode plot of loop gain TL_2(s)

The overall block diagram of the proposed stationary frame vector regulator
41

Gc_2(s) is illustrated in Figure 2-12. As seen in the figure, it is employed for the output

vector fout(t) to track the reference stationary frame vector fref(t) with zero steady-state

error for both positive and negative sequence components.

fαout(t) -ω60 ⌠ ω60 ⌠


⌡ ⌡

fαref(t) Ki
Kp + s

Gosc(s)

Ki
fβref(t) Kp + s

-ω60 ⌠ ω60 ⌠
fβout(t) ⌡ ⌡

Figure 2-12 Realization of stationary frame vector regulator

The vector regulator Gc_2(s) shown above is different from the P + resonant

stationary frame controller proposed by Zmood et al [32]. The P + resonant controller

transfer function of [32] is given by

2 Ki s
Gc_2Zm(s) = Kp + (2-46)
s2 + ω602

Bolognani and Zigliotto [36] proposed a space-vector based stationary frame

controller given by

(1 + sT)(s + α)
Gc_2Bo(s) = K (2-47)
s2 + ω602

However, details of controller realization and design of constant α were not provided in

[36]. The above transfer function is equivalent to the vector regulator Gc_2(s) in (2-44) if
42

the zero is at s = -α = 0.

Thus, in order to track both positive and negative sequence quantities along with

their disturbance rejection, vector regulator Gc_2(s) has a PI cascaded with a real

coefficient oscillator Gosc(s). In the next section, the realization of proposed vector

regulator in synchronous frame quantities is described and a comparative evaluation is

made between the stationary and synchronous frame implementations.

2.6 Equivalent Vector Regulator in the Synchronous

Reference Frames

The transformation of three-phase quantities in αβ stationary reference to dq

synchronous reference frame rotating anti-clockwise at ω60 rad./s is made by Park’s

transformation [68]
e+
⎡ fd (t) ⎤ ⎡ fα(t) ⎤
⎢ e+ ⎥ = [Pe+] ⎢ ⎥ (2-48)
⎣ fq (t) ⎦ ⎣ fβ(t) ⎦

⎡ cos(ω60 t) sin(ω60 t) ⎤
where [Pe+] = ⎢ ⎥ (2-49)
⎣ -sin(ω60 t) cos(ω60 t) ⎦

In terms of space-vector quantities, if fs(t) and fe+(t) denote them in stationary and

synchronous reference frames where the synchronous reference frame rotates in the anti-

clockwise direction, (2-48) is equivalent to

fe+(t) = e-jω60t fs(t) (2-50)

If the transformation is made instead to a synchronous reference frame rotating in

the clockwise direction at ω60 rad./s, the equations are


43
e-
⎡ fd (t) ⎤ ⎡ fα(t) ⎤
⎢ e- ⎥ = [Pe–] ⎢ ⎥ (2-51)
⎣ fq (t) ⎦ ⎣ fβ(t) ⎦

where [Pe–] = [Pe+]-1 (2-52)

In space-vector quantities, the relation between synchronous reference frame

quantities rotating in opposite directions is ‡

fe–(t) = (fe+(t))* (2-53)

Suppose the complex transfer function in the stationary reference frame between

f s (t) and fs(t) is Gs(s), the equivalent transfer functions in the synchronous reference
err c

e+ e+ e– e–
frame, between vector quantities ferr(t) and f c (t), and ferr(t) and f c (t) are determined as

respectively

Ge+
( s) =
{
_ f (t)} _ {e
e+
c
=
–jω60t s
f (t)
c } = G (s + j ω
s
60) (2-54)
_{f (t)} _{e
e+ –jω60t s
f (t)}
err err

Ge–
_
( s) =
{f (t)} = _ {e
e–
c
jω60t s
f (t)
c } = G (s – j ω
s
and 60) (2-55)
_{f (t)} _{e (t)}
e– jω60t s
f
err err

Alternatively, if a stationary reference frame complex transfer function can be

represented as a sum of partial fractions

Gs(s) = G1s (s) + G2s (s) , (2-56)

e+ e– s s
and G 1 (s) and G 2 (s) as the equivalent transfer functions of G1(s) and G2(s) respectively,

in the anti-clockwise and clockwise rotating synchronous reference frames, then


* denotes complex conjugate
44

Gs(s) = Ge+1(s – jω60) + Ge–2(s + jω60) (2-57)

s s s s
where fc1(t) and fc2(t) are the outputs of the transfer functions G1(s) and G2(s)

s
respectively, and ferr(t) is their common input quantity.

The vector regulator in the stationary reference frame can be represented in the

form of (2-56) as

⎛ Ki⎞ 1 Kps + Ki⎞ 1 ⎛Kps + Ki⎞


Gc(s) = ⎜Kp + s ⎟ Gosc(s) = ⎛⎜
⎝ ⎠ 2 ⎝ s - j ω60 ⎟⎠ + 2 ⎜⎝ s + j ω60 ⎟⎠ (2-58)

Applying (2-54) on the first term, the equivalent anti-clockwise rotating

synchronous reference frame transfer function can be realized as

1⎛ Ki Kp ω60⎞
Ge+ (s) = 2 ⎜Kp + s + j s ⎟ (2-59)
cp ⎝ ⎠

Likewise, (2-55) can be applied on the second term and the equivalent clockwise

rotating synchronous reference frame transfer function can be realized as

1⎛ Ki Kp ω60⎞
Ge– (s) = 2 ⎜Kp + s - j s ⎟ (2-60)
cn ⎝ ⎠

e+ e–
Thus, the regulators Gcp(s) and Gcn(s) derived above are identical to the complex

vector regulators proposed by Briz et al in [21] when implemented in synchronous

frames that are rotating in anti-clockwise and clockwise directions. Figure 2-13 illustrates

this realization. In the next subsection a comparative evaluation is made between the

stationary frame and synchronous frame realizations of the vector regulator.


45

e+
fdout(t)

e+
fderr (t)
e+
fdref (t) 1⎛ Ki⎞ e+
fdc1(t)

2⎝ Kp+
s ⎟⎠

1 ⎛Kpω60⎞
⎜ ⎟
2⎝ s ⎠

1 ⎛Kpω60⎞
⎜ ⎟
2⎝ s ⎠

e+ 1⎛ Ki⎞ e+
fqc1(t)
fqref (t) 2 ⎜⎝Kp + s ⎟⎠
e+
fqerr(t)
e+
fqout(t) (a)

e–
fdout(t)

e–
fderr (t)
e–
fdref (t) 1⎛ Ki⎞ e–
fdc1 (t)
2 ⎜⎝Kp + s ⎟⎠

1 ⎛Kpω60⎞
⎜ ⎟
2⎝ s ⎠

1 ⎛Kpω60⎞
⎜ ⎟
2⎝ s ⎠

e– 1⎛ Ki⎞ e–
fqc1 (t)
fqref (t) ⎜
2⎝ Kp+
s ⎟⎠
e–
fqerr(t)
e–
fqout(t) (b)

Figure 2-13 Block diagram illustrating realization of vector regulators (a) Gcp(s) in anti-clockwise rotating
synchronous frame and (b) Gcn(s) in clockwise rotating synchronous frame
46

2.6.1 Comparative Evaluation between Synchronous

and Stationary Frame Implementations

Based on the transfer function relationships between each reference frame dealt in

the earlier section, a comparison can be made between the synchronous frame and

stationary frame based implementations of the vector regulator. As digital signal

processors (DSPs) generally consider a multiply accumulate as a single cycle operation

[72], the criteria chosen as the basis for making this comparison is the number of

multiply-accumulates (MACs) and numerical integrations (ITGRs) in each of these

implementations. Such a criteria would be applicable in determining the merits and

demerits in case the controller is applied in analog circuitry (using operational amplifiers)

as well. The final impact of such a comparative study for both analog circuitry and digital

code would be on equipment cost and planning as it evaluates the memory usage as well

as computational burden involved in the two implementations of the controller.

Accordingly, each MAC or ITGR is considered as a unit and the total number of

such units is determined for both synchronous and stationary frame implementations. It is

to be noted that the comparison of the proposed stationary frame controller is made

against the synchronous frame implementation that was followed earlier in this section,

although couple of other methods also exist.

At the outset, every synchronous frame implementation requires transformation of

feedback variables that were originally in stationary reference frame into a synchronously

rotating reference frame. Every such transformation, according to (2-48) and (2-51),

comprises of two multiply-accumulates (MACs). The controller does this transformation


47

of variables immediately after sensing and scaling the feedback signals.

After the transformation to synchronous frame quantities, each of the complex

regulators (refer Figure 2-13(a) or Figure 2-13(b)) contains two error determining adders,

two PI regulators and two adders that produce the final controller output, thus totaling to

six multiply-accumulates (MACs). Further, each of the complex regulators requires four

integrators (ITGRs) to obtain the controller output along with MACs. The final controller

output in stationary reference frame is obtained after transforming back to the stationary

reference frame and adding the two regulator outputs, which takes another five MACs.

On the other hand, the stationary frame implementation contains only two PI

regulators and two oscillators (refer Figure 2-12). After determining the error between

the reference and feedback quantities using two MACs, the PI regulator plus oscillator

requires another four MACs and six ITGRs.

The comparison of synchronous frame and stationary frame implementations of

the vector regulator are tabulated in Figure 2-14. As seen in this figure, the total units

required for the stationary frame implementation is less than half that necessary for a

synchronous frame implementation.


48

Units
Type of
Description
Implementation
MACs ITGRs Total

2 MACs f e+ (t)
out
f s (t)
out
2 MACs f e– (t) 4 0 4
out

‡ Refer (2-48) and (2-51) for details

f e+ (t)
ref
6 MACs
+
fe+(t)
c1
Synchronous 4 ITGRs
Frame f e+ (t)
out

17 8 25
5 MACs fs(t)
c

f e– (t)
ref
6 MACs
+ fe–(t)
c2

f e– (t) 4 ITGRs
out

Total 21 8 29
‡ Refer Figure 2-13 for details

f s (t)
ref
6 MACs
+
fs(t)
Stationary c
6 6 12
Frame s 6 ITGRs
f (t)
out

‡ Refer Figure 2-12 for details

Figure 2-14 Comparison of synchronous frame and stationary frame implementations of vector regulator
49

2.7 Summary

In this chapter, the concepts of feedback controllers for dc quantities have been

extended for application towards three-phase ac space-vector quantities. Vector

regulators on stationary frame quantities were derived. These vector regulators have a PI

controller cascaded with an oscillator. They give a zero steady-state error when the

oscillator frequency is equal to the excitation frequency. The time response

characteristics of peak overshoot and settling time for vector regulated systems are

evaluated against the frequency-domain parameters like the phase margin. When

transformed to synchronous rotating frames, the vector regulators are equivalent to the

high performance complex vector regulators proposed by Briz et al [21]. A comparative

evaluation of the synchronous frame and stationary frame implementations of the vector

regulator is also covered. Further comparison against various linear regulators of the

features offered for three-phase unbalanced systems is given in Table 2-1.

Table 2-1 Comparison of linear regulators for unbalanced three-phase systems

Linear Regulator Features

Stationary frame PI regulator


Ki
f s (t) Kp + s fs(t)
ref c Steady-state error
in magnitude and
phase angle
f s (t)
out
50

Linear Regulator
Features
Synchronous frame PI regulator [26]

e-jω60t ejω60t

Zero steady-state
Ki error with
Kp + s
f s (t) fs(t) frequency
ref c
dependent transient
Ki response in
Kp + s magnitude and
phase angle [21]

f s (t) ejω60t e-jω60t


out

P + Resonant regulator [32]


Simplified
s
f (t) implementation of
ref 2 Ki s synchronous frame
KP + fs(t) PI regulator in
s + ω602
2
c
stationary reference
frame for
f s (t) unbalanced systems
out

Synchronous frame complex vector regulator [21]

e-jω60t ejω60t

Ki Kp ω60 Zero steady-state


Kp + s + j
s
f (t) s fs(t) error with improved
c
ref transient response
in magnitude and
Ki Kp ω60 phase angle
Kp + s - j s

f s (t) ejω60t e-jω60t


out
51

Linear Regulator Features

Stationary frame vector regulator


f s (t)
ref KI Simplified
KP + s Gosc(s) fs(t)
c implementation of
complex vector
regulator for
1 unbalanced systems
where Gosc(s) = 2
ω60
f s (t) 1 + s2
out
52

Chapter 3 Inverter Internal Controls -


Multi-Loop Vector Regulator

3.1 Introduction

A voltage source inverter (VSI) in a distributed resource (DR) is operated in

pulse-width modulation (PWM). The interface between inverter and the load/utility is

typically an LCL filter network and a transformer that are necessary for filtering and

conditioning to a utility grade voltage. However, addition of such filters increases the

order of system and may further cause resonance in the system. The inverter internal

controls described in this chapter are high speed loops aimed at achieving terminal

voltage regulation for the DR according to its generation requirements. Such controls can

also be designed to offer additional benefits of power quality conditioning and thus make

the inverter based DR a more attractive solution. In this chapter, a new approach is

presented to regulate the terminal voltage at the sensitive load bus even under load and

line imbalance conditions by applying the stationary frame vector regulator proposed in

the previous chapter.


53

3.2 Plant Transfer Function Modeling of the DR

Filter Interface Network

The power circuit schematic of an inverter based DR and its filter interface

network is illustrated in Figure 3-1. This circuit is a particular case of that shown in

Figure 1-2, when the generalized impedances are balanced in all three phases and they

are transformed to the primary side of the transformer. Besides, the transformer

secondary is not connected to any other network. A block diagram of this power circuit in

the orthogonal αβ coordinate system for a balanced load Z’(s) is illustrated in Figure 3-2.

Its equivalent SISO representation for space vector quantities is shown in Figure 3-3. As

seen in Figure 3-3, the plant of the DR filter interface network between the inverter

output vinv(t) and load terminal voltage vload(t) is of third order since the system consists

of three dynamic energy storage elements (Lf, Cf and Lt), in addition to the dynamic

elements constituting the load (Z’). Moreover, if the plant is modified to contain

unbalanced loads in the three phases and/or multiple DRs, the block diagram in Figure

3-3 is unworkable.

In order to deal with situations such as unbalanced loads or interconnected DRs,

the control output quantity is chosen to be vCf(t). The effect of load unbalance or

interconnected DR can be viewed as a disturbance input iLt(t) and the block diagram

representing the linear time-invariant (LTI) system of the plant transfer function for space

vector quantities can be represented as shown in Figure 3-4. Thus, imbalances and effects

of interconnected DRs are relegated to the complex space-vector quantities, and the
54

complex transfer functions for space-vector quantities remain linear and time-invariant.

The condition under which this assumption is valid is when the voltage controller

contains a feedforward quantity iLt(t) that makes the effective plant transfer function

unaffected by the dynamics in the energy storage element Lt, and in addition to the

dynamic elements constituting the unbalanced load [73].

SA1 DA1 SB1 DB1 SC1 DC1


Idc

Vdc
VAp VBp VCp

SA2 DA2 SB2 DB2 SC2 DC2

ILfA LfA ILfB LfB ILfC LfC

VCfA VCfB VCfC

CfA CfB CfC


ILtA ILtB ILtC

VNf
I'Lda
V'
I'Ldb anL
V'
I'Ldc bnL
V'cnL
I'Lda I'Ldb I'Ldc
Z'

Z'

Z'

V'nL

Figure 3-1 Circuit schematic of the three-phase PWM inverter based DR system with equivalent load Z’
55

iLf,α(t)
vinv,α(t) 1 1 1
Lf s Cf s Z’(s)
Lt s
vCf,α(t) vload,α(t)
iLt,α(t)

vinv(t) = vinv,α(t) + j vinv,β(t) vCf(t) = vCf,α(t) + j vCf,β(t) vload(t) = vload,α(t) + j vload,β(t)

iLf(t) = iLf,α(t) + j iLf,β(t) iLt(t) = iLt,α(t) + j iLt,β(t)

iLt,β(t)
vinv,β(t) vCf,β(t)
1 1 1
Lf s Cf s Z’(s)
Lt s
vload,β(t)
iLf,β(t)

Figure 3-2 Block diagram of the power circuit of the DR filter-interface network for a balanced load in
orthogonal αβ coordinate system

iLf(t)
vinv(t) 1 1 1
Lf s Cf s Z’(s)
Lt s
vCf(t) vload(t)
iLt(t)
Figure 3-3 Block diagram of the power circuit of the filter-interface network for a balanced load for space
vector quantities

iLf(t)
vinv(t) 1 1 1
Lf s Cf s Z’(s)
Lt s
vCf(t) vload(t)

iLt(t)
Figure 3-4 Block diagram of the effective plant transfer function of the DR filter-interface network under
unbalanced/interconnected DR conditions for space vector quantities

The effective plant transfer function of the DR filter interface network between

vinv(t) and vCf(t) that is depicted by the block diagram in Figure 3-4 is of second order,

which may cause stability problems in the design of a regulator [74]. Moreover, the plant
56

response would contain beat frequencies corresponding to the excitation frequency and

the resonant frequency of the LfCf filter, unless the non-ideality such as series resistance

of the inductor Lf or an external bleeder resistance of the capacitor Cf exists. Multi-loop

controllers have been found to give superior response under such conditions [29,30,33].

On similar lines, for the DR filter interface, a multi-loop regulator is designed with an

inner iLf(t) current loop and an outer vCf(t) voltage loop. The next section provides a

detailed controller design for the DR.

3.3 Controller Architecture

The objective of the DR internal controller is to receive voltage magnitude and

phase angle commands of vload†(t) from an external system controller and ensure an

accurate reproduction of the same at the load bus of the system, as vload(t). Furthermore,

the steady state and dynamic properties of the load and the system are different for

negative sequence and positive sequence excitations, making the design of the regulator

challenging.

The controller illustrated in Figure 3-5, uses a hierarchal approach to ensure

adequate steady state and dynamic performance with time varying three phase sinusoidal

command signals even in the presence of unbalanced operating conditions. At the outset,

the feedback voltage regulator is built to regulate the filter capacitor voltage, vCf(t). This

enables the feedback controller to be designed around a second order system with

reasonable performance attributes. The heart of the controller is a space-vector

modulator, which provides the duty ratio commands dinv(t) for various throws of the
57

IGBT switches of the inverter, which is developed from the voltage command for the

inverter vinv†(t). This controller incorporates a dc voltage feedforward component to

reject any dc voltage disturbances from affecting the dynamic performance of the

regulator. The space-vector modulator operates at a switching frequency of 4 kHz. The

realization of the space-vector modulator is typically carried out using a digital signal

processing system. At the next higher level in the controller is a current regulator that

acts as an inner feedback loop and regulates the filter inductor Lf current. In the current

loop design, capacitor voltage vCf(t) is added to the amplified error of the inductor Lf

current to decouple the voltage vCf(t) and the current iLf(t) in the plant. Since the filter

inductor current is not the ultimate quantity of interest, the steady state error performance

is not critical and hence uses a simple proportional regulator. The command current for

the inner current regulator loop itself iLf†(t) is derived from a filter capacitor Cf voltage

regulator loop. The dynamic performance and stability of the closed-loop system is

improved by adding the feedforward current of inductor Lt [75]. The design of this

regulator incorporates a controller that ensures accurate command following and

disturbance rejection with sinusoidal inputs.

The outermost segment containing a command voltage modifier is designed to

compensate for the voltage drop vcvm(t) across the impedance Lt representing the

interface reactor, which includes the effect of leakage inductance of the transformer by

modifying the voltage command to the capacitor voltage regulator. The modified

command is derived by appropriately multiplying the measured current with the series

impedance in the complex domain.


58

Command iLt,A
Voltage
Modifier vinv,AC Vdc vinv,A iLt,B
vload,AC iLf,A
PWM To
Voltage Current
vload,BC Inverter vinv,B Transformer
Regulator Regulator and Load
iLf,B
vinv,BC vinv,C iLfB
iLfA

vCf,BC

vCf,AC

Figure 3-5 Controller architecture for the DR inverter system

Figure 3-5 also gives a pictorial representation illustrating the implementation of

this control scheme for the inverter based DR. Stationary frame current and voltage

regulators are implemented with currents in phases A and B, and line voltages across

phases AC and BC, since two variables are adequate to describe a three-phase three-wire

system. For ease of implementation of sequence filters, these three-phase quantities are

transformed to Clarke’s αβ coordinates in the stationary reference frame. The design of

each of the segments in the multi-loop regulator are elaborated below starting with the

inner current regulator for Lf.

3.3.1 Filter Inductor Current Regulator

The block diagram illustrating the innermost Lf current loop in the orthogonal αβ

coordinate system is shown in Figure 3-6. The current regulator gives the voltage

reference to the space vector modulator vinv†(t). As seen in Figure 3-6, a simple controller

with a constant gain Kc is employed for the purpose. The current loop has vCf(t) as a

feedforward signal to remove the dependency of current iLf(t) on voltage vCf(t). Then, the

net transfer function as seen by the current loop controller is the integral gain transfer
59

function [1/(Lf s)]. The conventional feedback controller solution for an integral gain

plant is a proportional regulator. The same solution is used for this control loop despite

the fact that it does not give a zero steady-state error for sinusoidal ac signals. Thus,

Gci(s) = Kc (3-1)

The gain Kc is related to the bandwidth ωci of the current loop as

^ ω
Kc = L (3-2)
f ci

The fixed switching frequency of the PWM inverter is 4 kHz and so the

bandwidth of current loop is set below this frequency at ωci = 2π(600 Hz).

iLf(t) vCf(t)

iLf†(t) Kc vinv†(t)

(a)

iLf,α(t) vCf,α(t)

iLf,α†(t) Kc vinv,α†(t)

iLf†(t) = iLf,α†(t) + j iLf,β†(t) vinv†(t) = vinv,α†(t) + j vinv,β†(t)

iLf,β†(t) Kc vinv,β†(t)

iLf,β(t) vCf,β(t)
(b)

Figure 3-6 Block diagram of the current regulator for (a) space vector quantities, and (b) orthogonal αβ
coordinate quantities
60

The loop gain of the current regulator is determined as

Kc
TL_i(s) = ^ (3-3)
Lf s

Bode-plot of the transfer function TL_i(s) is illustrated in Figure 3-7 for a gain of

Kc = 3.66. As seen in figure, there exists a simple pole at zero frequency indicating a zero

steady-state error for dc quantities only in iLf†(t). The ac sinusoidal quantities at

frequency ω60 have a steady-state error that is determined by the bandwidth ωci.

Furthermore, since the phase plot of the loop gain is confined to be within –90o, it has a

large phase margin of 90o and gives a first-order response.

The closed-loop transfer function of the current regulator loop is determined as

1
Gi_cl(s) =
⎛ Lf ⎞ (3-4)
1 + s ⎜K ⎟
⎝ c⎠

The inner current loop with closed-loop transfer function Gi_cl(s) behaves as a

first-order low-pass filter with cut-off frequency equal to the bandwidth of the current

loop ωci.

100 135

80 90
Phase Angle (deg)
Magnitude (dB)

60 45

40 0

20 45

0 90

20 135
3 4 3 4
0.01 0.1 1 10 100 1 .10 1 .10 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

Figure 3-7 Bode plot of the loop gain for the current regulator loop
61

3.3.2 Filter Capacitor Voltage Vector Regulator

Subsequent to the design of the current feedback loop, design of the vCf(t) voltage

regulator is undertaken. By having the current iLt(t) as a feedforward signal in the vCf(t)

voltage loop design, the dependency of voltage vCf(t) on current iLt(t) is removed for

frequencies below the bandwidth of the current loop ωci. Assuming that the current iLt(t)

has no appreciable harmonic content for frequencies |ω| > |ωci|, if Gcv(s) denotes the

transfer function of the feedback regulator for voltage vCf(t), the loop gain of the voltage

regulator is determined as

1
TL_v(s) = Gcv(s) Gi_cl(s) C s (3-5)
f

If the dynamics of the voltage loop are limited to lower frequencies than the

bandwidth of the current loop ωci, the transfer function Gi_cl(s) can be neglected in (3-5).

1
Therefore, TL_v(s) ≈ Gcv(s) C s (3-6)
f

In the voltage vCf(t) control loop design from (3-6), the plant as seen by the

inductor current iLf(t) is the capacitance Cf that has an integral gain transfer function

[1/(Cf s)]. Unlike the current controller design, the voltage loop controller Gcv(s) is

designed to track with a zero steady-state error as well as reject disturbances of both

positive and negative sequence 60 Hz components. Therefore, a stationary frame vector

regulator for unbalanced systems that was described in the earlier chapter is used for this

purpose. As the effective plant to the voltage regulator is an integral gain transfer

function, a simple proportional gain cascaded with an oscillator would suffice. The

voltage regulator transfer function is therefore


62

Gcv(s) = Kv Gosc(s) (3-7)

where Gosc(s) is the oscillator transfer function in (2-45). A zero integral gain (Ki) was

found sufficient here due to disturbance input feedback decoupling employed here.

Figure 3-8 gives the block diagram illustrating realization of filter capacitor

voltage controller Gcv(s). Such a voltage controller is equipped to regulate the three-phase

voltage when disturbance inputs contain unbalanced quantities.

vCf(t) iLt(t)

vCf†(t) Kv Gosc(s) iLf†(t)

(a)

-ω60 ⌠ ω60 ⌠
⌡ ⌡
vCf,α(t) iLt,α(t)

vCf,α†(t) iLf,α†(t)
Kv

† † † † † †
vCf (t) = vCf,α (t) + j vCf,β (t) iLf (t) = iLf,α (t) + j iLf,β (t)

Kv

vCf,β (t) iLf,β†(t)

vCf,β(t) -ω60 ⌠ ω60 ⌠ iLt,β(t)


⌡ ⌡

(b)
Figure 3-8 Block diagram illustrating realization of filter capacitor voltage controller Gcv(s) for (a) space
vector quantities, and (b) orthogonal αβ coordinate quantities

Bode plot of the loop gain of the voltage regulator given by TL_v(s) from (3-5),

incorporating the controller Gcv(s) from (3-7) is illustrated in Figure 3-9 for Kv = 0.02.
63

This plot shows that fundamental frequency components @60Hz have an infinite gain.

The phase margin (φm) at the negative slope gain crossover frequency is observed to be

more than 50o thus limiting the peak overshoot to less than 5% (referring to Figure 2-10).

60 180

40 135

Phase Angle (deg)


90
Magnitude (dB)

20
45
0 0

20 45
90
40
135 φm
60 180
3 4 3 4
0.01 0.1 1 10 100 1 .10 1 .10 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

Figure 3-9 Bode plot of the loop gain for the voltage regulator loop

The above design procedure for the capacitor voltage regulator loop was made

assuming that there exist no components in iLt(t) above the bandwidth of the current loop

ωci. The current iLt(t), is otherwise treated as a disturbance input in the design. In order to

ascertain the disturbance iLt(t) rejection, the output impedance of the DR system both

without and with the controller is examined. The output impedance is given by the

transfer function between the capacitor voltage vCf(t) output and the reactor current iLt(t)

input. When no controller is employed for voltage regulation, the output impedance is

determined as

⎛ -1 ⎞ ⎛ 1 ⎞
Go_1(s) = ⎜C s⎟ ⎜ 1 ⎟
⎝ f ⎠ (3-8)
⎜1 + ⎟
⎝ Lf Cf s2⎠

From (3-8), it is clear that the output impedance of the LfCf filter is maximum

(infinity) at the corner frequency of the filter, viz., ω = ± 1/(LfCf) and is minimum at

the frequency ω = 0.
64

By contrast, the output impedance with the controller when vload†(t) is zero is

given by


⎛ 1 ⎞⎛ 1 ⎞
Go_2(s) ⎪⎪ =⎜ ⎟ ⎜ ⎟ (3-9)
⎝Cf (s + ωci)⎠ 2 Kv ωci s
⎪Vload†(s) = 0 ⎜1 + 2 2 ⎟
⎝ Cf (s + ωci) (s + ω60 )⎠

As evident from (3-9), the multi-loop controller improves the output impedance.

The output impedance with the controller is minimum at frequencies ω = 0 and ω =

± ω60. This is on account of infinite gain of Gcv(s) at ω = ± ω60. Furthermore, there exists

no infinite gain at ω = ± 1/(LfCf). Thus, the parallel resonance of the LfCf filter is

eliminated by the action of the multi-loop controller.

The Bode (gain) plot of output impedance without and with the controller is

illustrated in Figure 3-10(a) and Figure 3-10(b), respectively. As seen in Figure 3-10,

without the controller LfCf filter exhibits parallel resonance to iLt(t) at its corner

frequency of 933 Hz. By employing the proportional controller with capacitor voltage

vCf(t) feedforward in the inner current loop, this parallel resonance is damped.

Furthermore, use of the inductor current iLt(t) feedforward in the outer voltage loop

provides a low impedance for all frequencies well below the bandwidth of the current

loop ωci. The voltage regulator having singularities at ω = ±ω60 ensures a transmission

zero for 60 Hz components of iLt(t) complex space vector. This is indicated by a low gain

at the fundamental frequency in the response. Thus, the controller Gcv(s) provides good

disturbance rejection in addition to command tracking for fundamental quantities.


65

80 80
60 60
40 40
Magnitude (dB)

Magnitude (dB)
20 20
0 0
20 20
40 40
60 60
80 80
3 4 3 4
0.01 0.1 1 10 100 1 .10 1 .10 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

(a) (b)
Figure 3-10 Bode (gain) plot of the output impedance (a) without the multi-loop controller and (b) with the
multi-loop vector controller.

3.3.3 Command Voltage Modifier

The controller thus designed, regulates the filter capacitor voltage on the primary

side of the transformer. However, the principal objective of designing a controller for the

DR system is to regulate the load bus voltage on the secondary side of the transformer

under all conditions to remain within adequate tolerance limits. For that reason, the

voltage drop vcvm(t) across transformer leakage and reactor Lt is estimated and added to

the capacitor voltage reference as it improves the load bus voltage regulation. The

expanded block diagram of the command voltage modifier block (refer Figure 3-5) is

illustrated in Figure 3-11. As seen in Figure 3-11, in the command voltage modifier

block, positive and negative sequence components of reactor current iLt(t) are extracted

by filters PSF and NSF, respectively. The impedance of transformer leakage together

with the reactor Lt at frequency ω60 is also shown in Figure 3-11.

The extraction of positive and negative sequence components in the reactor

current is essential to determine the voltage drop in the reactor and transformer leakage.

The three-phase complex space-vector of this voltage drop is positive for the positive
66

sequence component in iLt(t) and negative for the negative sequence component in iLt(t),

when the impedance of the reactor and transformer leakage is assumed to be purely

inductive.

Positive
jω60 Lt sequence
filter (PSF)
iLt(t)
Negative
-jω60 Lt sequence
filter (NSF)

vcvm(t)
vload†(t)
Gcv(s) iLf†(t)

vCf(t)
Figure 3-11 Block diagram of the command voltage modifier

Sequence filtering has been classified into two categories, viz., averaging and

delaying techniques [76]. The sequence filter proposed in this section can be categorized

as an improved method of averaging techniques. The averaging techniques proposed

earlier include single stage band-pass/band-reject filters [57,58] and cascaded sections of

complex band-pass/band-reject filters [53]. These methods are observed to give high

performance for high frequency fundamental frequencies that are used in carrier-signal-

injection-based sensorless techniques [53]. On the other hand, the delaying techniques

involve combinations of the three phase quantities, with a few of those delayed by a

definite interval of time period resulting in unacceptable phase delays and complicating

the design of controllers [76].


67

Among the averaging techniques, the complex band-pass filters have been state-

of-the-art in industry. The complex band-pass filters, originally proposed for

Communications and Signal Processing [69,70]], have been employed for

positive/negative sequence filtering of 60 Hz components by Hochgraf [77] and Yuan et

al [38]. Allmeling [39] has extended this transfer function for a complex band-pass filter

that passes all the chosen harmonics. The method applied by Hochgraf [77] involved

heterodyning using synchronous reference frame extractors requiring transformations of

all quantities. Yuan et al [38] have described a means to perform positive sequence

filtering in stationary reference frame by filter that can be mathematically expressed on

the lines of (2-18) as

⎛ 1 ⎞
GY-PSF(s) = ⎜
⎛s – j ω60⎞⎟ (3-10)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ ωo ⎠⎠

where ω60 = 2π(60 Hz) is the power frequency in rad./s.

However, it can be shown that in order to operate satisfactorily as a band-pass

filter for the positive sequence 60 Hz quantities under unbalanced conditions, GY-PSF(s)

requires a small gain ωo. Furthermore, if a similar transfer function is used to design a

negative sequence filter for a three-phase system containing chiefly positive sequence

components, it allows residual positive sequence components along with phase distortion.

Therefore, an alternative method using complex-coefficient filters that gives an improved

response is proposed here.

Negative Sequence Filter (NSF)

The proposed sequence filter incorporates both band-pass and band-reject


68

functions. The NSF is designed to maintain a unity gain for negative sequence

component and provide a transmission zero for the positive sequence. Thus, it passes

(with unity gain) only the 60 Hz quantity of the negative sequence and entirely eliminates

the 60 Hz quantity of positive sequence. The NSF complex transfer function is given by

⎛ 1 ⎞
GNSF(s) = ⎜
⎛-2jω60 (s + j ω60)⎞⎟ (3-11)
⎜1 + ⎜ ⎟⎟
⎝ ⎝ ωo (s – j ω60) ⎠⎠

Bode-plots of the NSF transfer function in (3-11) for positive and negative

sequence components is illustrated in Figure 3-12. As seen in the figure, the NSF gives a

unity gain for negative sequence components at frequencies near 60 Hz, while

attenuating the positive sequence components at frequencies near 60 Hz.

20 20

0 0
Magnitude (dB)

Magnitude (dB)

20 20

40 40

60 60

80 80

100 100
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

135 135

90 90
Phase Angle (deg)

Phase Angle (deg)

45 45

0 0

45 45

90 90

135 135
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

(a) (b)
Figure 3-12 Bode plot of the transfer function GNSF(s) for (a) negative sequence and (b) positive sequence
components of the space vector

Rewriting (3-11) in the form


69

Gfn(s)
GNSF(s) = 1 + G (s) , (3-12)
fn

the transfer function GNSF(s) can be realized as a unity gain feedback controller of

forward gain Gfn(s), where

⎛ ωo (s – j ω60) ⎞
Gfn(s) = ⎜ ⎟ (3-13)
⎝-2jω60 (s + j ω60)⎠

Simplifying (3-13) further,

⎛ ωo ⎞ ⎪⎧⎨ω60 ⎪⎫⎬ ⎛ 1 ⎞
Gfn(s) = ⎜ ⎟ +j ⎜
⎝2ω60⎠ ⎩⎪ s ⎭⎪ jω ⎟ (3-14)
⎜1 + 60⎟
⎝ s ⎠

The equation (3-14) consists of three factors: (i) the first factor, (ωo/(2ω60)) is a

constant gain; (ii) the second factor is a complex gain {(ω60/s) + j} of the forward

coupling form (Refer Figure 3-13); and (iii) the third factor is the oscillator Gosc_n(s) in

(2-26) that provides forced oscillations for negative sequence components while

neglecting positive sequence components. The negative sequence filter, GNSF(s), may be

realized in αβ coordinates as in Figure 3-13, with the three factors clearly outlined.

iLt,α(t) iLt,αn(t)
ωo/(2ω60) ω60 ⌠

-ω60 ⌠
iLt(t) = ⌡
(i) (ii) (iii) iLt,n(t) =
iLt,α(t) + j iLt,β(t)
iLt,αn(t) + j iLt,βn(t)
ω60 ⌠

iLt,β(t) iLt,βn(t)
ωo/(2ω60) ω60 ⌠

Figure 3-13 Block diagram illustrating realization of negative sequence filter (NSF) in the command
voltage modifier
70

The loop gain of the NSF is equal to Gfn(s), due to the unity gain feedback. The

Bode plot of the open-loop gain of the NSF for ωo = 200 is shown in Figure 3-14(a) and

Figure 3-14(b) for negative and positive sequence quantities of the complex space vector

iLt(t).

As seen in Figure 3-14, the dashed lines indicating the negative sequence

response show an infinite gain at ω = -ω60; and the solid lines indicating the positive

sequence response show a transmission zero at ω = +ω60. Therefore, the NSF allows only

the negative sequence components of iLt(t) to pass with a unity gain while blocking the

positive sequence components entirely.

80 80
60 60
40 40
Magnitude (dB)

Magnitude (dB)

20 20
0 0
20 20
40 40
60 60
80 80
100 100
120 120
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

135 135

90 90
Phase Angle (deg)

Phase Angle (deg)

45 45

0 0

45 45

90 90

135 135
4 3 3 4
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)

(a) (b)
Figure 3-14 Bode plot of the loop gain for the negative sequence filter (NSF) Gfn(s) for (a) negative
sequence and (b) positive sequence components of the space vector iLt(t)

In the NSF shown in Figure 3-13, the negative sequence component of the current

iLt(t) is iLt,n(t). The negative sequence current iLt,n(t) is multiplied by a complex gain ‘-
71

jω60 Lt’ to determine the negative sequence component of vcvm(t) as shown in Figure 3-15

in forward coupling method of realization of complex filter.

iLt,αn(t) -ω60 Lt vcvm,αn(t)


iLt,n(t) = vcvm,n(t) =
iLt,αn(t) + j iLt,βn(t) vcvm,αn(t) + j vcvm,βn(t)
iLt,βn(t) ω60 Lt vcvm,βn(t)

Figure 3-15 Block diagram illustrating realization of complex gain ‘-jω60 Lt’ for the output of the NSF in
αβ coordinates

Positive Sequence Filter (PSF)

Similar to the NSF, the positive sequence filter (PSF) is designed to maintain

unity gain for the positive sequence component while providing a transmission zero for

the negative sequence component in iLt(t). The PSF complex transfer function is given by

⎛ 1 ⎞
GPSF(s) = ⎜
⎛2jω (s – j ω60)⎞⎟ (3-15)
⎜1 + ⎜ 60 ⎟⎟
⎝ ⎝ ωo (s + j ω60) ⎠⎠

The PSF can be realized by the NSF block diagram in Figure 3-13 by swapping

the α and β terminals at both the input and output. Thereafter, the positive sequence

component of the current iLt,p(t) thus extracted is multiplied by a complex gain ‘jω60 Lt’

to determine the positive sequence component of vcvm(t). As a result, the overall

command voltage modifier output vcvm(t) is the sum of the estimated positive and

negative sequence voltage drops across the leakage inductance of the transformer plus the

reactor Lt. Accordingly, it is determined as

vcvm(t) = vcvm,p(t) + vcvm,n(t) (3-16)


72

Sequence Filter Response:

The transient response of the sequence filter proposed in the earlier section is

related to the gain ωo. Figure 3-16 gives the Bode plot of the negative sequence filter for

values of gain ωo = 20, 200 and 2000. As seen in figure, when ωo is increased the width

of the pass-band is also increased but that of the stop-band is decreased. The transient

response of the negative sequence quantities is influenced by the pass-band width while

that of the positive sequence by the stop-band width of the frequency response. Thus, the

design of gain ωo is made according to the transient response requirements of either

positive or negative sequence or both components. Nevertheless, for all values of ωo a

unity gain and zero phase lag for the negative sequence and a total elimination of positive

sequence 60 Hz quantities is guaranteed at steady-state. In addition, if this filter is used in

a close-loop system, the phase lag/lead introduced by it in the loop gain transfer function

can be reduced by increasing the gain ωo for stabilizing the closed-loop system against a

degradation of only the transient performance of the positive sequence stop-band width.
73
20 20

Magnitude (dB) 0 0

Magnitude (dB)
20 20

40 40

60 ωo ωo 60 ωo ωo
80 80

100 100
3 4
4 3
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10 100 1 .10 1 .10
Frequency (Hz) Frequency (Hz)
135 135
90 90
Phase Angle (deg)

Phase Angle (deg)


45 ωo 45
ωo
0 0
45 ωo 45
ωo
90 90
135 135
4 3
1 .10 1 .10 100 10 1 0.1 0.01 0.01 0.1 1 10
3
100 1 .10 1 .10
4

Frequency (Hz) Frequency (Hz)


(a) (b)
Figure 3-16 Bode plot of the negative sequence filter GNSF(s) for different values of gain ωo (20, 200 and
2000). (a) negative sequence and (b) positive sequence components of the space vector

3.4 Stability of Feedback Systems Containing

Complex-Coefficient Transfer Functions

The realization of proposed sequence filters, which contain complex-coefficient

transfer functions in the previous subsection, involved a feedback loop. Therefore, it is

necessary to establish the stability of such a filter. In this section, stability of one of the

two proposed filters, viz., the negative sequence filter (NSF) is investigated. While the

stability of feedback systems with real-coefficient transfer functions using Nyquist

criterion [78] has been exhaustively dealt in several textbooks on control systems [71,79],

stability of complex-coefficient systems is relatively new [80,81]. Gataric et al [81] have


74

have applied the Nyquist diagram on complex-coefficient transfer functions for three-

phase power systems. Extending this approach, guidelines to ensure the stability of such

transfer functions based on important observations in the Nyquist diagram of complex

transfer functions are presented here. For an interpretation of the stability of complex

transfer functions with regard to the multi-input multi-output (MIMO) systems theory,

refer to the Appendix B.

The vector block diagram of NSF realization explicitly showing the feedback

loop is illustrated in Figure 3-17. As seen in the figure, the loop gain TL(s) is also the

forward gain Gfn(s) of the unity feedback NSF system and it is of the form

⎛M (s – j ω60)⎞
TL(s) = ⎜ ⎟ (3-17)
⎝ (s + j ω60) ⎠

f(t) fn(t)
Gfn(s)

Figure 3-17 Block diagram illustrating realization of negative sequence filter (NSF) as a unity feedback
system

where M is a constant in the complex plane C. Two cases of complex constant M are

investigated for stability, viz., purely real and purely imaginary.

ωo ωo ⎛s – j ω60 ⎞
Case (i): M = : TL(s) = ⎜ ⎟
2ω60 2ω60 ⎝s + j ω60⎠

The locus of the Nyquist path in the s-plane and the TL plane corresponding to

this case are displayed in Figure 3-18. The transfer function TL(s) has a pole at s = -jω60

and a zero at s = jω60. Since there exists a singularity of TL(s) on the jω axis at s = -jω60,
75

the Nyquist path is modified to pass along a semicircle of infinitesimal radius around s =

-jω60. The Nyquist path is divided into 5 segments and the corresponding segments in the

TL locus are marked with the same numbers. It is observed that unlike real-coefficient

transfer functions, the TL locus is not symmetric for positive and negative frequencies for

complex-coefficient transfer functions.

jω Im{TL}
s-plane TL-plane
j∞
5
jω60 s = -j∞
1 1
-jω60 ∞ σ (-1,0) Re{TL}
2 4 3
3 5
4 2
-j∞
s = -jω60
P=0 N≠0

Figure 3-18 Nyquist path on the s-plane and the TL locus for M = ωo/(2ω60)

As seen in Figure 3-18, while the number of poles of TL(s) in the right half of the

s-plane is P = 0, the number of encirclements of the TL locus around (-1,0) is denoted by

N ≠ 0. This is because the TL locus passes through the point (-1,0). Therefore, the number

of zeros of 1 + TL(s) in the right half of the s-plane Z = N + P ≠ 0. Hence, according to

the Nyquist stability criterion the feedback system with this loop gain is not stable.

ωo ⎛ s – jω60 ⎞
Verification: 1 + TL(s) = 1 + ⎜ ⎟ , which has a zero on the imaginary axis and
2ω60 ⎝s + jω60⎠

hence the closed-loop system is not stable.

jωo jωo ⎛ s – j ω60 ⎞


Case (ii): M = : TL(s) = ⎜ ⎟:
2ω60 2ω60 ⎝s + j ω60⎠
76

In this case, the loop gain transfer function in Case (i) is multiplied by the

imaginary operator ‘j’. The Nyquist path in the s-plane and the TL locus corresponding to

this case are displayed in Figure 3-19. As seen in the figure, multiplying the loop gain by

‘j’ has resulted in the TL locus to be rotated anti-clockwise by 90o as compared to that in

Case (i). While the number of poles of TL(s) in the right half of the s-plane P = 0, the

number of encirclements of the TL locus about (-1,0) in this case is N = 0. Therefore, the

number of zeros of 1 + TL(s) in the right half of the s-plane Z = N + P = 0. Hence, the

feedback system with this loop gain is stable.

jω Im{TL}
s-plane TL-plane
j∞ s = −j∞ 3
5 s = -jω60
jω60 4
1 s = j∞
5

-jω60 2 ∞ σ (-1,0) 1 Re{TL}


3
4
-j∞ 2

P=0 N=0

Figure 3-19 Nyquist path on the s-plane and the TL locus for M = jωo/(2ω60)

jωo ⎛s – jω60 ⎞
Verification: 1 + TL(s) = 1 + ⎜ ⎟, which has a zero in the left half of
2ω60 ⎝s + jω60⎠

the s-plane, and hence the closed-loop system is stable.

A key observation made from stability analysis is that it is possible to stabilize a

feedback system containing complex-coefficient transfer functions by simply modifying

the phase of loop gain TL(s) with the complex coefficient M. However, it is not always

necessary to use Nyquist stability criterion for determination of stability of complex-

coefficient transfer functions. If the number of poles of TL(s) in the right half of the s-
77

plane P = 0, it is also possible to evaluate stability on the basis of Bode plots [71,82]. The

Bode plots of loop gain of TL(s) for Cases (i) and (ii) are displayed in Figure 3-20 and

Figure 3-21, respectively. These plots have log sweep of frequency as it would provide

the magnitude and phase variation over a wide frequency range that is beneficial for

visualizing broad frequency phenomena in power systems. The left section of each Bode

plot represents response for negative frequencies, and right section that for positive

frequencies of space vector. As noted earlier, the frequency response for positive

sequence quantities is indicated by the characteristics for positive frequencies, and those

for negative sequence quantities by the characteristics for negative frequencies.

As seen in Figure 3-18, the Nyquist diagram for Case (i) M = ωo/(2ω60) showed

instability as the TL locus passed through (-1,0) on the negative real axis. The

corresponding Bode plot of loop gain TL(s) is displayed in Figure 3-20. In the Bode

magnitude plot, the phase angle at the gain crossover frequency is +180o. Hence, the

feedback system with loop gain of Case (i) is not stable.

On the other hand, the Nyquist diagram for Case (ii) M = jωo/(2ω60) in Figure

3-19 showed that the TL locus does not encircle the point (-1,0) and therefore the

feedback system is stable. This is indicated in the Bode plot of Figure 3-21 by a 90o

phase margin (from +180o) at the gain crossover frequency.

From the stability analysis presented in this subsection, it has been proved that the

proposed negative sequence filter is stable as its forward gain Gfn(s) has a loop gain TL(s)

identical to that of Case (ii) M = jωo/(2ω60).


78

80 80
60 60

Magnitude (dB)
Magnitude (dB)
40 40
20 20
0 0
-20 -20
-40 -40
-60 -60
-80 -80
3 2 1 0 0 1 2 3
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

180 180
135 135
Phase Angle (deg)

Phase Angle (deg)


90 90
45 45
0 0
-45 -45
-90 -90
-135 -135
-180 -180
3 2 1 0 0 1 2 3
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)
(a) (b)
Figure 3-20 Bode plot of loop gain TL(s) for M = ωo/(2ω60) for (a) negative and (b) positive sequence
components of the space vector

80 80
60 60
Magnitude (dB)

Magnitude (dB)

40 40
20 20
0 0
-20 -20
-40 -40
-60 -60
-80 -80
3 2 1 0
10 10 10 10 0 1 2 3
Frequency (Hz) 10 10 10 10
Frequency (Hz)
180 180
135 135
Phase Angle (deg)
Phase Angle (deg)

90 90
45 45
0 0
-45 -45
-90 -90
-135 -135
-180 -180
3 2 1 0 0 1 2 3
10 10 10 10 10 10 10 10
Frequency (Hz) Frequency (Hz)

(a) (b)
Figure 3-21 Bode plot of loop gain TL(s) for M = jωo/(2ω60) for (a) negative and (b) positive sequence
components of the space vector
79

3.5 Simulation Results

This section presents selected results from simulation studies conducted on the

performance of the controllers designed for voltage regulation under imbalance are

presented. The various controllers tested include filter inductor current regulator, filter

capacitor voltage regulator, and the command voltage modifier containing the positive

and negative sequence filters. Digital simulation has been carried out in Matlab®

SIMULINK™ software package [83]. The SimPowerSystems™ package of the software

has been utilized for creation of various power quality events in the simulation.

The simulation model of the DR system assumes an averaged model of the pulse-

width modulator and the inverter together with dc bus voltage feedforward (refer Figure

3-5). The physical plant consisting of the DR filter interface network is modeled by linear

ODEs describing the three-phase circuits. The stationary frame current and voltage

regulators are implemented with currents in phases A and C, and line voltages across

phases AB and CB, since two variables are adequate to describe a three-phase three-wire

system. Voltage imbalances at the load terminal are caused by load imbalances in the

three phases due to unequal rms currents drawn from each phase of the voltage source.

The per-unit phasor plots in Figure 3-22 portray a typical case of unequal single-phase

loads connected to the secondary terminals of the delta-wye transformer, when

unregulated.
80

1.5 1
0.95∠-31
1
0.92∠-153 0.5
0.5 0.95∠88
1.5 1 0.5 0 0.5 1 1.5 1 0.5 0 0.5 1

0.67∠-76 0.5
0.5
1.3∠162 1

0.83∠43 1.5 1

ILa VanL
ILb VbnL
ILc VcnL

Figure 3-22 Effect of load imbalance (left) on the unregulated terminal voltage (right)

The effect of voltage imbalances or sags is a variation in the reactor current iLt.

The performance of the controller is tested by simulating a disturbance that causes a

variation in the transformer current iLt.

(i): Response of the multi-loop vector regulator:

Figure 3-23 illustrates the performance of the voltage regulator for an incremental

step change in voltage command from 70% to 100% of the rated voltage of 480 VRMS.

Waveforms displayed in Figure 3-23 are the αβ components of reference voltage, actual

voltage that tracks the reference and the error between the reference and actual values. As

seen in the figure, the capacitor voltage response reaches close to the steady state in about

one and half 60 Hz cycles. During this transient, the actual voltage of filter capacitor

vCf(t) across lines A and C is displayed in Figure 3-24.


81

800

α variables
400

-400

-800
0.3 0.32 0.34 0.36 0.38 0.4 0.42 0.44 0.46

800
β variables

400

-400

-800
0.3 0.32 0.34 0.36 0.38 0.4 0.42 0.44 0.46
Time (s)
Figure 3-23 Operation of the voltage control loop for vCf with the response vCf for a complex exponential
step input voltage reference vCf†.

2000

1500

1000
vCf,AC (V)

500

-500

-1000

-1500

-2000
0.1 0.2 0.3 0.4 0.5

Time (s)
Figure 3-24 Simulation waveform of line voltage vCf,AC illustrating the response for a step change in
voltage reference vCf† from 70% to 100% of rated voltage of 480 VRMS

(ii): Balanced three-phase load test:

The first test in this case involves a three-phase 10 kW resistive load. Figure 3-25

illustrates the steady-state voltage waveforms at the load terminals across phases ac and

bc as well as line currents through phases a and b.


82
400
200

vac(V)
0
-200
-400
0.5 0.52 0.54 0.56 0.58 0.6 0.62 0.64 0.66 0.68 0.7
400
200

vbc(V)
0
-200
-400
0.5 0.52 0.54 0.56 0.58 0.6 0.62 0.64 0.66 0.68 0.7
80
40
ia(A)
0
-40
-80
0.5 0.52 0.54 0.56 0.58 0.6 0.62 0.64 0.66 0.68 0.7
80
40
0
ic(A)

-40
-80
0.5 0.52 0.54 0.56 0.58 0.6 0.62 0.64 0.66 0.68 0.7

Time (s)
Figure 3-25 Load terminal line-line voltage ac & bc (top two) and load current in a & b phases (bottom
two) for a balanced 10 kW resistive load.

Another test carried out to observe the dynamic response of the controller under

balanced conditions involved abrupt load change. Figure 3-26 demonstrates the response

when the DR is initially operating on no-load and a 2 kW three-phase balanced resistive

load was abruptly connected using an three-pole single-throw switch. As seen in Figure

3-26, load terminal voltages on the secondary side of 480 V-Δ : 208 (120 V-Y)

transformer respond fairly well with minor glitches at the instant of no-load to load

switching. If it is assumed that the load is switched ON at t = tON, as the current is zero at

t = tON– the voltage is also zero; and as the ac current begins to flow through the load at t

= tON+ the voltage at load terminals regains its sinusoidal form. This voltage recovery

takes place within a small fraction of a 60 Hz cycle.


83
400
200

vac(V)
0
-200
-400
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
400
200

vbc(V)
0
-200
-400
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
20
10
ia(A)
0
-10
-20
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
20
10
0
ic(A)

-10
-20
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14

Time (s)
Figure 3-26 Simulation waveforms illustrating DR operation during transition from no-load to a 2 kW
balanced three-phase resistive. Load terminal line-line voltage ac & bc (top two) and load current in a & b
phases (bottom two).

The dynamic response of DR against abrupt load-change from loaded condition to

no-load is illustrated Figure 3-27. As seen in this figure, the DR responds satisfactorily

without any voltage glitch during this transient.

400
200
vac(V)

0
-200
-400
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
400
200
vbc(V)

0
-200
-400
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
20
10
ia(A)

0
-10
-20
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
20
10
ic(A)

0
-10
-20
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14

Time (s)
Figure 3-27 Simulation waveforms illustrating DR operation during transition from a 2 kW balanced three-
phase resistive load to no-load. Load terminal line-line voltage ac & bc (top two) and load current in a & b
phases (bottom two).
84

Thus, an inverter-based DR equipped with the multi-loop vector controller is

shown to satisfactorily regulate the terminal voltage under balanced load conditions. The

next case involves tests under an unbalanced three-phase loading of the DR.

(iii): Unbalanced three-phase load tests:

Inorder to test the performance of the controllers under unbalanced conditions, the

phase a connection of a 10 kW load is abruptly removed. Figure 3-28 illustrates the

steady-state voltage waveforms at the load terminals across phases ac and bc as well as

line currents through phases a and b. Table 3-1 provides a comparison of the regulated

load terminal voltage with that of unregulated case. As seen in Table 3-1, the line-line

voltages are regulated fairly close to 208 V in all three phases. Small inconsistencies

between different line voltages can be attributed to discrepancies in the scaling of

voltage/current sensing circuits as well as in estimation of the leakage inductance of the

second reactor and transformer.

400
200
vac(V)

0
-200
-400
0.9 0.92 0.94 0.96 0.98 1 1.02 1.04 1.06 1.08 1.1
400
200
vbc(V)

0
-200
-400
0.9 0.92 0.94 0.96 0.98 1 1.02 1.04 1.06 1.08 1.1
80
40
ia(A)

0
-40
-80
0.9 0.92 0.94 0.96 0.98 1 1.02 1.04 1.06 1.08 1.1
80
40
ic(A)

0
-40
-80
0.9 0.92 0.94 0.96 0.98 1 1.02 1.04 1.06 1.08 1.1

Time (s)
Figure 3-28 Simulation steady-state waveforms illustrating the load terminal voltage and load current for
an unbalanced load
85

Table 3-1 Evaluation of voltage regulator performance under single phasing

Description Vab (V) Vbc (V) Vca (V) Imbalance


(%)
w/o vector controller 214 200 200 4.6
w/ vector controller 208 208 208 0

During unbalanced conditions, the operation of voltage command modifier

consisting of sequence filters PSF and NSF has a major impact on the effectiveness of

regulation of load terminal voltage. The functioning of PSF and NSF are demonstrated in

Figure 3-29. It illustrates the filter input which is the transformer primary side current,

PSF output and NSF output waveforms.

40
20
Input

0
-20
-40
0.96 0.98 1 1.02 1.04 1.06 1.08 1.1 1.12
40
PSF output

20
0
-20
-40
0.96 0.98 1 1.02 1.04 1.06 1.08 1.1 1.12
40
NSF output

20
0
-20
-40
0.96 0.98 1 1.02 1.04 1.06 1.08 1.1 1.12
Time (s)
Figure 3-29 α- and β- components of transformer primary current, PSF output and NSF output waveforms
for an unbalanced load

As seen in Figure 6-14, the PSF and NSF satisfactorily extract the positive and

negative sequence components thus aiding in accurate compensation of the reactor and

transformer leakage impedance voltage drop.


86

3.6 Summary

In this chapter, a multi-loop regulator employing the stationary frame vector

regulator is applied to inverter based distributed resources to regulate the load terminal

voltage with zero steady-state error even under unbalanced conditions. The multi-loop

regulator consists of three segments – inner current regulator for the filter inductor

current, outer voltage vector regulator for the filter capacitor voltage, and the command

voltage modifier to compensate for the transformer leakage impedance voltage drop.

Frequency domain analysis techniques in the form of Bode plots for individual sequence

components are utilized in the design procedure.

Further, by making use of complex transfer functions, novel negative and positive

frequency/sequence selective filters that contain complex band-pass as well as complex

band-reject sections are proposed. The stability of these filters has been explained using

Nyquist diagrams and Bode plots. Table 3-2 gives a comparison of their frequency

response characteristics against the state-of-the-art synchronous frame low-pass filter and

cascaded filters. As seen in Table 3-2, the synchronous frame negative sequence filter for

60 Hz continues to pass significant amounts of positive sequence 60 Hz, and the

cascaded filter results in significant phase distortion of its output. The proposed

frequency/sequence selective complex filter gives high quality output without any phase

distortion. On the downside, its response to passband frequency components and stop-

band frequency components is complementary in nature. For a very high gain ωo in the

negative sequence filter, while the time response for passing negative sequence 60 Hz

components is greatly improved, the time response for rejecting positive sequence 60 Hz
87

components gets worsened. Nevertheless, the positive sequence 60 Hz components in its

output get nullified in steady state.

Simulation results are presented from digital simulation in Matlab®

SIMULINK™ software that establish the application of DR systems to mitigate the

effects of power quality phenomena at the sensitive load bus. In this chapter, the vector

regulators in stationary frame were introduced as a solution to the control of three-phase

ac systems.

Table 3-2 Comparison of frequency/sequence selective filters


Frequency/Sequence Selective Filter Response
20
Synchronous frame based negative sequence 60 Hz filter
Magnitude (dB) 0

20

e-jω60t
jω60t
e
40
180 120 60 0 60 120 180
1 Frequency (Hz)

f s (t) s f s (t) 135


1+ out
ωo
in
90
Phase Angle (deg)

45
0

45

90

135
180 120 60 0 60 120 180
Frequency (Hz)
88

Frequency/Sequence Selective Filter Response

Cascaded negative sequence 60 Hz filter 20

Magnitude (dB)
20
ejω60t e-2jω60t 40

60

80
s 1
f (t) 100
in 1 + s/ωo 180 120 60 0 60 120 180
Frequency (Hz)
135
90

Phase Angle (deg)


s/ωo 45
f s (t)
1 + s/ωo out 0
45
90
jω60t
e 135
180 120 60 0 60 120 180
Frequency (Hz)
20
Proposed negative sequence 60 Hz filter 0
Magnitude (dB)

20

40

60

80

1 100
⎛ ⎞ 180 120 60 0 60 120 180
s
f (t) ⎜ ⎛ -2j ω60 (s + j ω60 ⎞
) ⎟ f (t)s
Frequency (Hz)
in ⎜1 + ⎜ ⎟⎟ out
⎝ ⎝ ωo (s – j ω60) ⎠⎠ 135

90
Phase Angle (deg)

45

0
45

90

135
180 120 60 0 60 120 180
Frequency (Hz)
89

Chapter 4 Generation Control and


Dynamic Behavior of a Microgrid

In the previous chapter, the design of a multi-loop regulator for control of the load

terminal voltage in a DR network to a desired voltage reference has been demonstrated.

Such a DR controller requires the ac voltage vector as input. The reference voltage space-

vector can be set according to the nominal system specifications under stand-alone

operation. However, when several such DRs are interconnected in a microgrid

configuration, it is necessary to device a methodology to determine the magnitude and

phase of voltage vector such that DRs can be dispatched to share the overall load in a

predetermined manner.

This chapter deals with the generation control of the DR which provides the

magnitude and phase information to the multi-loop regulator. The generation control

involves active and reactive power controllers that allow parallel operation of DRs in the

microgrid. The dynamic (small-signal) behavior analysis for microgrids having a chain

topology is also covered in this chapter. Both real power and reactive power flows in the

microgrid are considered. Guidelines are provided for design of the active power-

frequency and reactive power-voltage controls on the basis of IEEE P1547 performance

specifications [18]. Finally, selected results from simulation of DRs equipped with these

generation controls in Matlab® SIMULINK™ are presented.


90

4.1 Introduction

A simplified schematic of a DR containing the IGBT based power electronic

inverter with LCL filter and a transformer is illustrated in Figure 4-1. Several such DR

networks can be interconnected in a microgrid configuration as in Figure 1-3. Since the

DR is operated to regulate the load bus voltage to a value within the tolerance limits. For

that reason, the DR along with the LC filter and transformer is approximated by a voltage

source as illustrated in the figure while conducting theoretical analysis on power studies.

This assumption is especially valid since the external power generation controls of the

inverter are designed to have a bandwidth lower by atleast one order as compared to the

internal multi-loop regulator dealt in the previous chapter. Furthermore, investigations

conducted by [60,64] have shown that the generation control would interact with the

internal controls if their bandwidths are not far apart. The power studies in the microgrid

are conducted with all variables in p.u.. It is to be noted that the regulated voltage on both

the high- and low-voltage sides of the transformer would be the same in p.u. after

compensating for the leakage inductance of the transformer.

As seen in Figure 4-1, the voltage space-vector vt specification at each DR bus

comprises the magnitude (V) and frequency (ω60) information. Such a specification

constitutes the voltage command that needs to be provided to each inverter based DR in

the microgrid. A block diagram illustrating the generation of voltage command for each

inverter based DR is shown in Figure 4-2. As seen in Figure 4-2, the generation

controller determines the incremental voltage magnitude and frequency information

based on the real and reactive powers drawn from the DR. These incremental variables
91

are added to their nominal values to determine the voltage command vector to the

inverter. The instantaneous real and reactive powers [50] at the DR terminal, denoted as

PL and QL, respectively, can be easily computed from the measured quantities vt and iLtp,

which is the positive sequence component of iLt. By the use of the positive sequence filter

output iLtp instead of iLt, it is ensured that the power measurement does not contain ac

ripple under unbalanced conditions. The following section in this chapter provides further

details on the generation controller present in each inverter based DR.

Inverter LC Filter Transformer


vt
iLf iLt
PL,
QL

vCf

DSP
Controls

vt = V ejω60t
PL,
QL

Figure 4-1 Voltage regulated DR’s analogy to a voltage source for power studies
92

Vo

ΔV
PL V

Generation

Controller
Δω †
QL ω60

ωo

Figure 4-2 Generation of voltage command for the inverter based DR

4.2 Generation Control

4.2.1 Introduction to Generation Control

Control of generation in interconnected power systems comprising conventional

large rotating machines has been well established and reported in various literature

[61,62,64]. The frequency droop features naturally offered by the components of the

system against an increase in load support paralleling of generators and sharing of the

active power. Figure 4-3 demonstrates in deviation values, indicated by a “Δ”, how a

change in load demand (ΔPL) causes a deviation in speed/frequency (Δω) for a large

rotating machine based synchronous generator [61]. The generator modeling is made by

means of the relationships between deviations in mechanical power ΔPM and speed Δω.

The input labeled ‘ΔPL-ref’ is the load ref. set-point that is the control input to shift the

generator’s governor characteristic in order to give the reference frequency at any desired

power output. Frequency droop follows after a change in load demand had been initially
93

provided by the inertia of the rotating mass and there is a delay in the response of the

prime-mover to accommodate the change in load. Governors are employed to allow the

controller to detect the change in the machine speed and thereby vary the valve input of

the prime-mover. There are two forms of governing in a power system, viz., generator

governing and load governing [62]. When a load increase causes a decrease in machine

speed, there is a decrease in the supply frequency, which eventually results in a net

decrease in the load. This self-governing attribute of a rotating machine generator is the

principal basis that makes possible paralleling of generators.While the load governing

provides transient droop in frequency, the generator/speed governing permits to some

extent restoration of frequency.

Gen. inertia and load


governing
ΔPL-ref ΔPM Δω
1 1 1
1 + sTG 1 + sTCH Ms

Prime-
mover ΔPL D
Speed governing

Figure 4-3 Block diagram of the generation controller for a rotating machine ‡

As seen in Figure 4-3, the speed governor for generation control is modeled by a

gain ρ for the speed sensor and a first-order lag with time-constant TG. The prime-mover

driving the generator unit may be a steam- or a hydro-turbine. It is modeled as a first-

order lag transfer function with time-constant TCH (known as charging time-constant) that


Source: [61]
94

signifies the delay in the production of the mechanical power output after change in the

valve position. The effect of generator inertia is modeled by the angular momentum of

the machine M and the influence of load governing by the frequency dependency factor

D. It is to be noted that manual coordination or secondary frequency control [64], also

known as automatic generation control (AGC) [61,84,85] that is generally employed in

large power systems, is not dealt here.

4.2.2 Active Power-Frequency Control

Thus, generation control in a conventional rotating machine is dependent largely

on the physical device characteristics such as the governor, prime-mover, machine inertia

and the load governing characteristics. Alternatively, generators consisting of power

electronic inverters offer a good deal of flexibility in the design of control parameters. A

key difference between a microgrid consisting of inverter based DRs and a large power

system with rotating machines is that the DRs do not possess inertia as they contain

micro-sources such as fuel cells or micro-turbines and power electronic converters. The

inertia in a rotating machine aids in supplying transient load support and further offers a

speed droop that is used by the speed governor to change the power input to the prime-

mover in response to the change in load demand. In a DR, the energy storage on the dc

link of the inverter can provide support in meeting the transient load demands until the

micro-source behind it (fuel cell or micro-turbine) responds by a change in its power

output.

The active power-frequency controller for a power electronic converter based


95

distributed UPS was proposed by Chandorkar [60]. This controller is illustrated in the

block diagram shown in Figure 4-4 and applied for the active power frequency control of

the DR. It has a frequency droop with a proportional gain transfer function (b > 0), which

provides the necessary load governing functionality that is beneficial for paralleling DR

units. The active power controller includes a frequency restoration loop whose function is

analogous to the speed governor in a rotating machine generator (refer Figure 4-3). A

notable difference from the controller for rotating machine is that the DR controller does

not have a first-order lag transfer function corresponding to that of the prime-mover with

a charging time-constant.

1/β

ΔPL-ref 1 ΔPR Δω
1 + s/ωG b

ΔPL
Frequency droop
Frequency restoration

Figure 4-4 Block diagram of active power-frequency controller ‡ for a DR

A more simplified block diagram of the same active power-frequency controller

for DR in Figure 4-4 is illustrated in Figure 4-5. The parameters in Figure 4-5 are

⎡ b/β ⎤
Kbβ = ⎢ ⎥ (4-1)
⎣1 + b/β⎦

and ωb = ωG (1 + b/β) (4-2)


adapted from the distributed UPS controller proposed by Chandorkar [60]
96

It is to be noted that parameter Kbβ represents the extent of frequency restoration

in a DR upon load transient, and 0 < Kbβ < 1 so that a very low value Kbβ → 0 denotes

almost no frequency restoration whereas Kbβ → 1 denotes total frequency restoration.

From the small-signal model block diagram in Figure 4-5, the transfer function

between the deviation in frequency Δω and deviation in load demand ΔPL alone is a lead-

lag filter

-Δω(s) ⎡1 + s/ωG⎤
= b (1 - Kbβ) ⎢ ⎥ (4-3)
ΔPL(s) ⎣ 1 + s/ωb ⎦

Likewise, the transfer function between the deviation in frequency Δω and

deviation in load ref. set-point ΔPL-ref alone is a first-order lag function

Δω(s) ⎡ 1 ⎤
= b (1 - Kbβ) ⎢ ⎥ (4-4)
ΔPL-ref(s) ⎣1 + s/ωb⎦

1 Δω
PL-ref 1 + s/ωG b

1
PL Kbβ Δωh
1 + s/ωb

(a) Large-signal implementation

1 Δω
ΔPL-ref 1 + s/ωG b

1
ΔPL Kbβ Δωh
1 + s/ωb

(b) Small-signal model

Figure 4-5 Simplified block diagram of active power-frequency controller for a DR

The two transfer functions show a first-order time response of time constant 1/ωb
97

and a final steady-state deviation in frequency of b(1 - Kbβ). The steady-state frequency

droop can be graphically represented as shown in Figure 4-6(a) and Figure 4-6(b) for the

incremental variables and actual variables, respectively. As seen in these figures, the

straight line can be shifted along the ΔPL (PL) axis by varying ΔPL-ref (PL-ref). The rated

load is denoted by PLR in Figure 4-6(b).

Δω ω
PL_ref < PLR
ΔPL-ref = 0
Slope = - b Slope = - b

ΔPL ωo

Slope = Slope =
- b(1-Kbβ) - b(1-Kbβ)
PL
0 1.0 p.u.

(a) (b)

Figure 4-6 Steady-state voltage droop of a DR in the stand-alone mode of operation. (a) Deviation
quantities (b) actual quantities.

As the DRs in a microgrid employ semiconductor devices in the power electronic

converters, their current capacity is limited by the device ratings. It is therefore necessary

to control the apparent power drawn from a DR in a microgrid. This can be achieved by

having a controller for regulating reactive power drawn from the DR along with the

active power regulator.

4.2.3 Reactive Power – Voltage (Magnitude) Control

It has been proved beyond doubt through power flow studies conducted on an

interconnected distributed generator system that having arbitrarily fixed voltage

references for the DR can cause excessive reactive power circulation between the
98

generators [63]. This fact has been recognized earlier in the paralleling of synchronous

generators as well as SVC/Statcom [64]. In order to minimize the reactive power

exchange between the two generators due to discrepancies in the voltage set-points, it is

essential to have soft set-points in voltage magnitude. The proposed reactive power-

voltage controller achieves this task – a reactive load increase causes a decrease in

voltage magnitude set-point that eventually results in a net decrease in the reactive load.

Reactive power-voltage controller exploits the dependency of the reactive power

supplied by the DR on the voltage magnitude at the load bus. Figure 4-7 shows a block

diagram of the proposed reactive power controller. As seen in this figure, a simple

feedback controller containing first-order lag is employed for the deviation in the reactive

power. Unlike the active power-frequency controller, the reactive power-voltage

controller proposed here does not contain a voltage restoring loop. Moreover, it does not

achieve a zero steady-state error but provides a voltage droop upon an increase in the

reactive load. As seen in Figure 4-7, the input labeled ‘ΔQL-ref’ is the load ref. set-point

that is the control input to shift the DR’s voltage regulator characteristic in order to give

the reference voltage (magnitude) at any desired reactive power output.


99

1 1 ΔV
QL-ref Dq 1 + s/ωq

QL
Voltage droop

(a) Large-signal implementation

1 1 ΔV
ΔQL-ref
Dq 1 + s/ωq

ΔQL
Voltage droop

(b) Small-signal model

Figure 4-7 Block diagram of the proposed reactive power controller for a DR

Voltage droop is a first-order lag function and not instantaneous. The active

modulation control, which varies the modulation index according to the dc bus voltage

variations in the inverter of the DR, assists in providing a fixed voltage magnitude

against load variation for some duration before other controls take precedence.

Similar to the frequency droop curves, it is possible to represent the steady-state

voltage droop at a particular bus by voltage droop curves. In order to generate the voltage

droop characteristics of the DR with the proposed controller, the relation between the

deviation in voltage and the change in reactive load is re-examined. This relation is the

sum of the response of ΔV to ΔQL and ΔQL-ref when acting alone, and is determined as

⎛ 1 ⎞⎛ 1 ⎞
ΔV(s) = ⎜D ⎟⎜ ⎟ (ΔQL-ref(s) – ΔQL(s)) (4-5)
⎝ q⎠⎝1 + s/ωq⎠

The steady-state deviation in voltage response is obtained by setting s = 0 in the


100

above equation. The steady-state change in voltage is obtained as

ΔQL-ref – ΔQL
ΔV = Dq (4-6)

This equation represents a straight line and can be graphically represented as

shown in Figure 4-8(a). As seen in the figure, the straight line can be shifted along the

ΔQL axis by varying the load ref. set-point. The same curves are displayed in Figure

4-8(b) for the actual voltage V and power demand QL variables. Representing the

nominal values of the quantities by a subscript “o”, the actual quantities are determined as

QL = QLo + ΔQL (4-7)

and V = Vo + ΔV (4-8)

Thus, Figure 4-8(b) characterizes the steady-state voltage droop feature of the

DR. As seen in Figure 4-8(b), the actual quantity corresponding to the load ref. set-point

is QL_ref. The value of QL_ref corresponding to a load ref. set-point of zero is QL_ref = 0. A

positive value of load ref. set-point would refer to QL_ref > 0 and a negative value to QL_ref

< 0. By changing QL_ref the controller can be set to give nominal voltage at any desired

reactive load condition. According to the convention employed, the load is inductive

when it draws a positive reactive power, and capacitive when it draws a negative reactive

power.
101

V
ΔV QL_ref = 0
Slope =
-Dq-1 QL_ref > 0
ΔQL-ref > 0 Vo

ΔQL Slope =
-Dq-1
QL_ref < 0
ΔQL-ref < 0
ΔQL-ref = 0 QL
0 1.0 p.u.
-1.0 p.u.
(a) (b)

Figure 4-8 Steady-state voltage droop of a stand-alone DR. (a) Deviation quantities (b) actual quantities.

Each inverter based DR is equipped with the reactive power-voltage (magnitude)

controller presented in this section along with the active power-frequency controller. The

dynamic behavior of a microgrid consisting of several such DRs is investigated in the

following section.

4.3 Dynamic (Small-Signal) Behavior of a Microgrid

The previous sections discussed the operation of the active power-frequency

controller and reactive power-voltage (magnitude) controller in an inverter based DR. As

these controllers are programmed in a digital signal processor, their parameters need to

be designed to meet conditions that guaranty stable decentralized operation when DRs

are connected in wide varying microgrid configurations. In this section, a set of design

constraints are developed for the DR controller gains that would guaranty stable

operation of the microgrid. The following assumptions are made in the analysis of

dynamic behavior of the microgrid —

(i) The analysis is on linearized small-signal models of the power system


102

(ii) Inverter based DRs operate within their maximum capacity limits

(iii) Dynamics of the inverter internal controls are fast as compared to the

external power controls so that the internal controls can be neglected in

power flow analysis

(iv) Tie-lines between any two sources in the microgrid are purely inductive

in nature

4.3.1 DR Interconnected System

Consider the operation of a single DR when connected to the grid supply through

an inductive tie-line as shown in Figure 4-9(a). Grid supply generally has enormous

generation capacity as compared to the DR. Accordingly, it can be approximated to an

infinite bus that is characterized by a stiff frequency. Therefore, any load variation at the

grid supply end does not cause a droop in its frequency in either transient- or steady-state.

The small-signal state-variable schematic of this system under zero input conditions is

given in Figure 4-9(b). While the real power loop is depicted in the right half of this

figure, the reactive power loop is in its left half.


103

PL1, QL1
X1,g
Infinite Bus

V1∠δ1 E∠0o

DR1
(a) Single-line diagram

ΔV1(s) Δω 1(s)
ωq1
1/Dq1 b1
s

ω b1
Kbβ1
s
Dqtie1,g
Δωh1(s)

Δδ1(s) 1
Po1,g
s
(b) Small-signal state-variable schematic under zero input conditions

Figure 4-9 Representation of single DR connected to the infinite bus

The system differential equations of Figure 4-9(b) are

d
dt Δωh1 = - ωb1 Δωh1 - b1 b1/β1 ωG1 (Po1g Δδ1g) , (4-9)

d
dt Δδ1g = - Δωh1 - b1 (Po1g Δδ1g) (4-10)

d ⎛⎛ Dqtie1g⎞ ⎞
and dt ΔV1 = - ωq1 ⎜⎝⎜⎝1 + Dq1 ⎟⎠ ΔV1⎟⎠ (4-11)

V1o E 2V1o - E
where Po1g = X and Dqtie1g = X1g (4-12)
1g

Representing the system differential equations in the form

d
dt x = A x + B u (4-13)

The state vector x is given by


104

xT = [Δωh1 Δδ1g ΔV1] (4-14)

Of primary interest is the system matrix, A, from which the system stability can

be determined. System matrix in this case is

⎡ -ω-1
b1 -b1(b1/β1)ωG1Po1g 0

A=
⎢ -b1Po1g 0 ⎥ (4-15)
⎢ 0 Dqtie1g
-ωq1 ⎛1 + D ⎞⎥
⎣ 0
⎝ q1 ⎠⎦

From eigenvalue studies on matrix A, it has been observed that the single grid-

interfaced DR characterized by the above system matrix has real distinct eigenvalues and

is stable if the constants b1, β1, ωG1, ωq1, Dq1 are all positive and the maximum droop in

voltage is less than the nominal voltage.

Then, consider the operation of a two DRs interconnected through an inductive

tie-line as shown in Figure 4-10(a). The small-signal state-variable schematic of this

system under zero input conditions is given in Figure 4-10(b).

For the state vector xT = [Δωh1 Δδ12 Δωh2 ΔV1 ΔV2], the system matrix in the case of

the two DR system in Figure 4-10(b) can be derived as

⎡ -1 ⎤
-ωb1 -b1(b1/β1)ωG1Po12 0 0 0

⎢0 ⎥
-(b1+b2)Po12 1 0 0
b2(b2/β2)ωG2Po12 -ωb2 0 0
A=
⎢0 0 0
Dqtie12
-ωq1 ⎛1 + D ⎞
ωq1V1o ⎥ (4-16)

⎢ ⎝ q1 ⎠

Dq1X12

⎣0 0 0
ωq2V2o
Dq2X12 ⎝
Dqtie21
-ωq2 ⎛1 + D
q2 ⎠⎦

V1o V2o 2V1o - V2o 2V2o - V1o


where Po12 = X12 , Dqtie12 = X12 and Dqtie21 = X12 (4-17)
105

PL1, QL1 PL2, QL2


Xtie1,g

V1∠δ1 V2∠δ2

DR1 DR2
(a) Single-line diagram

ΔV1(s) ωq1 Δω 1(s)


1/Dq1 b1
s

Dqtie1,2 ω b1
Kbβ1
s
V1o
X1,2 Δω h1(s)

V2o
Δδ 1,2(s) 1
X1,2 Po1,2
s
Dqtie2,1
ΔV2(s) ω q2
1/Dq2 b2
s Δω 2(s)
ωb2
Kbβ2
s
Δω h2(s)

(b) Small-signal state-variable schematic under zero input conditions

Figure 4-10 Representation of two DR interconnection through an inductive tie-line

In the case of a two-DR system also, it has been observed through eigenvalue

studies that the system matrix has real distinct eigenvalues and is stable if the constants

(k = 1, 2) bk, βk, ωGk, ωqk, Dqk are all positive and the maximum droop in voltage is less

than the nominal voltage. On similar lines, system matrices can be constructed for

different microgrid configurations including the higher-order systems.

Cardell and Tabors [67] have shown that microgrids that were stable with only

few combustion turbine DRs can become unstable when more combustion turbines were
106

added in the system. They further suggested that frequency stability problems may arise

only as the number of DRs increases.

To investigate the frequency stability of inverter based DRs, the eigenvalues of

the active power-frequency control flow section have been calculated for typical control

gains in systems of order 3 through 7 etc. Figure 4-11 illustrates the dominant

eigenvalues as the order of system matrix is increased. As seen in this figure, the

eigenvalues appear to be advancing towards the imaginary axis in the left half of complex

s-plane as the system order is increased. However, it is not certain whether the

eigenvalues might cross into the right half of s-plane if the system order increases

greatly.
90
120 60
0.8

0.6
150 30
( )
max r3
0.4

( )
max r4
0.2

( )
max r5 180 0 0

( )
max r6

( )
max r7
210 330

240 300

270

Figure 4-11 Dominant root locus of the active power-frequency control system for a microgrid as the
system order is increased from 3 through 7

In the following subsections, the dynamic behavior of microgrid chain

architectures containing multiple inverter based DRs are studied based on eigenvalue

analysis and sufficient conditions for their stability are established.


107

4.3.2 Multiple DRs Connected in a Chain

In this subsection, investigation of dynamic behavior is carried out on a microgrid

consisting of several interconnected DRs in a chain formation. Each of them is equipped

with the active power-frequency and reactive power-voltage controllers for decentralized

operation. Chandorkar [60] solved the stability problem for a generalized n distributed

UPS system connected in a chain formation, where each UPS inverter was equipped with

an active power-frequency controller alone. However, the system model considered in

[60] was a simplified one neglecting the frequency restoration loop of the active power-

frequency controller; and furthermore the reactive power-voltage controller was absent in

that model.

The dynamic behavior is investigated in this section for the microgrid architecture

shown in Figure 4-12(a), which is designated as a chain formation. It is a radial network

of n DRs with the infinite bus at one end and loads connected at each DR bus. It is

assumed that the tie-line between any two buses i and j has an inductive reactance Xi.j.

The operation of this chain microgrid is analyzed below with grid-interfaced mode as a

special case.

The state variable schematic of the chain microgrid illustrating the DRs

connected by an inductive tie-line is given in Figure 4-12(b). These state variables are

directly depicted for two of the units, viz., kth and k+1th in Figure 4-12(b), with the dashed

lines indicating interconnections to the rest of the microgrid.


108

DR2 DRk+1

V2∠δ2 Vk∠δk Vk+1∠δk+1


V1∠δ1 Vn∠δn

DR1 DRk DRn


(a) Single-line diagram

Δω k(s)
ΔVk(s) ω qk
1/Dqk bk
s

Dqtiek,k+1 ωbk
Kbβk
s
Vko
Xk,k+1 Δω hk(s)

Vk+1o
Δδk,k+1(s) 1
Xk,k+1 Pok,k+1
s
Dqtiek+1,k
Δω k+1(s)
ΔVk+1(s) ω
qk+1
1/Dqk+1 bk+1
s
ω bk+1
Kbβk+1
s
Δω hk+1(s)

(b) Small-signal state-variable schematic under zero input conditions

Figure 4-12 Microgrid having DRs in a chain structure

For the state vector xT = [Δωh1 Δδ12 Δωh2 … Δδn-1.n Δωhn ΔV1 ΔV2 … ΔVn], the system

matrix for the microgrid in Figure 4-12 is given by

⎡ [A11]2n-1x2n-1 [0]2n-1xn ⎤
A = ⎢⎢ ⎥
⎥ (4-18)
⎣ [0]nx2n-1 [A22]nxn ⎦

where the partitioned submatrices of A refer to the active and reactive power flows,

respectively, and are given by


109

⎡ ⎤
-ωb1 -b1(b1/β1)ωG1Po12 0 0 … 0

⎢ -1 -(b1+b2)Po12 1 b2Po23 … 0

=⎢ ⎥
0 b2(b2/β2)ωG2Po12 -ωb2 -b2(b2/β2)ωG2Po23 … 0
A11 (4-19)
0 b2Po12 -1 -(b2+b3)Po23 … 0

⎢ : : : ⎥
⎣ 0 0 0 … bn(bn/βn)ωGnPo.n-1.n -ωbn ⎦
ωq1V1o
⎡ ⎤
Dqtie12
-ωq1⎛1+ D ⎞ 0 0 … 0
⎝ q1 ⎠ Dq1X12
⎢ ωq2V2o
-ωq2⎛1+
Dqtie21+Dqtie23⎞ ωq2V2o ⎥
⎢ ⎥
0 … 0
Dq2X12 ⎝ D q2 ⎠ Dq2X23
A22 =
⎢ ⎥
ωq3V3o Dqtie32+Dqtie34⎞ ωq3V3o
0 -ωq3⎛1+ D X … 0
Dq3X23 ⎝ D q3 ⎠ q3 34

⎢ ⎥
: : :

⎣ 0 0 0 0

Dqtie.n.n-1
… -ωqn⎛1+ D
qn ⎠⎦

(4-20)

Matrix A can be represented as a direct sum that is represented mathematically as

A = [A11]2n-1x2n-1 ⊕ [A22]nxn , (4-21)

and the eigenvalues λ11 (of A11) and λ22 (of A22) are given by the solution of

det(λ11 I - A11) det(λ22 I - A22) = 0 (4-22)

The properties of the eigenvalues of A11 and A22 are determined separately below,

beginning with A22 as it is easily solvable with known properties.

Proposition 1: The sufficient conditions for all eigenvalues of Α22 of a microgrid in chain

configuration to be negative and real are ΔVk_max < Vko and ωqk, Dqk > 0 (k = 1, 2, …, n),

where ΔVk_max is the maximum droop in voltage of kth DR unit when operated in stand-

alone mode and Vko is its nominal voltage.


110

Proof: Let a22.i,j represent the element in the ith row and jth column of matrix Α22, which is

a Jacobi matrix [86].

For the controller constraints ωqk, Dqk > 0 (k = 1, 2, …, n), consider a diagonal

matrix D22 = diag(d22.1, d22.2, …, d22.n) with elements

d22.i a22.i.i+1
(for i = 1, 2, …, n) (4-23)
d22.i+1 = a22.i+1.i

Then there exists a similar matrix B22 = D22-1 A22 D22 that is real symmetric given

by

⎡ ( ) ⎤
Dqtie12 V1oV2oωq1ωq2
-ωq1 1+ 0 … 0
Dq1 Dq1Dq2X122

⎢ V1oV2oωq1ωq2
Dq1Dq2X122 (
-ωq2 1+
Dqtie21+Dqtie23
Dq2 ) V2oV3oωq2ωq3
Dq2Dq3X232
… 0 ⎥
B22 =
⎢ 0
V2oV3oωq2ωq3
(
-ωq3 1+
Dqtie32+Dqtie34
) … 0

⎢ ⎥
Dq2Dq3X232 Dq3
: : :

⎣ 0 0 …
Vn-1oVnoωqn-1ωqn
Dqn-1DqnXn-1.n2 (
-ωqn 1+
Dqtie.n.n-1
Dqn )⎦
(4-24)

It is well known that similar matrices have the same eigenvalues, and that a real

symmetric matrix has only real eigenvalues [86]. Since B22 is a real symmetric matrix

that is similar to A22, all the eigenvalues λ22 (of A22) are real.

Further, it is observed in (4-20) that A22 has only negative eigenvalues according

to Gerschgorin theorem [86] as it is diagonally dominant with all diagonal elements

strictly negative for k = 1, 2, …, n,

Vko Vko
if (Dqk + Dqtie.k.k-1 + Dqtie.k.k+1) > X +X (4-25)
k-1.k k.k+1

⎛ 2Vko - Vk-1o 2Vko - Vk+1o⎞ Vko Vko


i.e. if ⎜Dqk + + ⎟ > + (4-26)
⎝ Xk.k-1 Xk.k+1 ⎠ Xk-1.k Xk.k+1
111

⎛Vko - Vk-1o Vko - Vk+1o⎞


i.e. if Vko ⎜ X + X ⎟ > -Dqk Vko (4-27)
⎝ k.k-1 k.k+1 ⎠

The expression in the left hand side of above inequality is a measure of total

reactive power generated by the kth DR under nominal conditions. If Qko denotes this

nominal value of generated reactive power, matrix A22 becomes diagonally dominant

Qko
if - D < Vko (4-28)
qk

or if ΔVk_max < Vko (4-29)

since the ratio of Qko and Dqk gives the steady-state voltage droop in the stand-alone

mode of operation of the DR (refer Figure 4-7).

The above inequality signifies that the condition for diagonal dominance of A22 is

when the (maximum) steady-state droop in voltage at every DR terminal under stand-

alone mode is strictly less than its nominal value. This is generally true as the controller

gains are seldom designed to result in a voltage droop larger than the nominal value.

Consequently, under these conditions A22 has only negative real eigenvalues λ22.

Proposition 2: The sufficient conditions for all the eigenvalues of Α11 of a microgrid in

chain configuration to be negative and real are bk, βk, ωGk > 0 (k = 1, 2, …, n).

Proof: At first it is proved that all the eigenvalues of A11 are real and later it is proved

that they are negative.

For bj, βj, ωGj > 0 (j = 1, 2, …, n), consider a diagonal matrix D11 given by
112

⎡ ⎤
b1(b1/β1)ωG1 0 0 0 … 0

⎢ 0
1
Po12
0 0 … 0 ⎥
D11 =
⎢ 0 0 b2(b2/β2)ωG2 0 … 0 ⎥ (4-30)
⎢ 0 0 0
1
… 0 ⎥
⎢ ⎥
Po23
: : :
⎣ 0 0 0 0 … bn(bn/βn)ωGn ⎦
Then there exists a similar matrix B11 = D11-1 A11 D11 that is real symmetric given

by

⎡ ⎤
-ωb1 - b1(b1/β1)ωG1Po12 0 0 … 0

⎢ - b1(b1/β1)ωG1Po12 -(b1+b2)Po12 b2(b2/β2)ωG2Po12 b2 Po12Po23 … 0



⎢ ⎥
0 b2(b2/β2)ωG2Po12 -ωb2 - b2(b2/β2)ωG2Po23 … 0
B11 =
0 b2 Po12Po23 - b2(b2/β2)ωG2Po23 -(b2+b3)Po23 … 0
⎢ : : : ⎥
⎣ 0 0 0 … bn(bn/βn)ωGnPo.n-1.n -ωbn ⎦
(4-31)

As it is well known, similar matrices have same eigenvalues and a real symmetric

matrix has real eigenvalues; since B11 is a real symmetric matrix that is similar to A11, all

the eigenvalues λ11 (of A11) are real.

After having proved that all the eigenvalues of Α11 are real, it is proved that all of

them are negative as follows:

Matrix A11 can be simplified to a tridiagonal form by elementary column

transformations as
113

-ωb1 b1Po12ωG1
⎡ -1 ⎤
0 0 … 0

⎢0 0 1
-b2Po12ωG2 -ωb2 b2Po23ωG2
0 … 0

⎢0 ⎥
… 0
A11 T= (4-32)
0 -1 0 … 0

⎢ : : : ⎥
⎣ 0 0 0 … -bnPo.n-1.nωGn -ωbn ⎦
1 -b1Po12 0 0 … 0
⎡0 1 0 0 … 0 ⎤
where T=
⎢0 b2Po12 1 -b2Po23 … 0 ⎥
(4-33)
⎢ 0: 0 0 1 … 0

⎣0 1 ⎦
: :
0 0 … bnPo.n-1.n

The determinant of matrix T can be easily evaluated as det(T) = 1. The

characteristic polynomial of A11 is determined by solving the equation

det(λ11 I - A11) = 0 (4-34)

which is the same as det(λ11 I - A11) det(Τ) = 0 (4-35)

or det(λ11 T - A11 Τ) = 0 (4-36)

Let ϕ2n-1(λ11) = det(λ11 T - A11 Τ) represent the characteristic polynomial for the

2n-1 x 2n-1 matrix A11. The characteristic polynomial is evaluated as

λ11+ωb1 -b1Po12(λ11+ωG1)
⎪ ⎪
0 0 … 0

⎪ 1 λ11 -1
b2Po12(λ11+ωG2) λ11+ωb2 -b2Po23(λ11+ωG2)
0 … 0

⎪ ⎪(4-37)
0 … 0
ϕ2n-1 (λ ) =
11
0 0 1 λ11 … 0
⎪ : : : ⎪
⎪ 0 0 0 … bnPo.n-1.n(λ11+ωGn) λ11+ωbn ⎪
For the above tridiagonal matrix, the determinant satisfies a recursive formula

[86] that is determined as


114

ϕ2k-1(λ11) = (λ11 + ωbk) ϕ2k(λ11) + bk Po.k-1.k (λ11 + ωGk) ϕ2k-3(λ11)

ϕ2k-2(λ11) = λ11 ϕ2k-3(λ11) + bk Po.k-1.k (λ11 + ωGk-1) ϕ2k-4(λ11)


⎫⎪
ϕ1(λ11) = (λ11 + ωb1)
⎬ (for k = 1, 2, …, n)

ϕ0(λ11) = 0
⎪⎭
(4-38)

The eigenvalues λ11 (of A11) are given by the roots of the characteristic equation

ϕ2n-1(λ11) = 0. As seen from the recursive formula in (4-38), the coefficients of all powers

of λ11 are positive. By the Descartes’ rule of signs [79], it is determined that the number

of positive real roots of ϕ2n-1(λ11) = 0 is zero. Therefore, all the eigenvalues λ11 (of A11)

are negative real.

Proposition 3: The sufficient conditions for ensuring stability of a microgrid in chain

configuration are bk, βk, ωGk, ωqk, Dqk > 0 and ΔVk_max < Vko (for k = 1, 2, …, n).

Proof: It was determined in the previous two propositions that the sufficient condition for

stability of real power control in the chain microgrid is

bk, βk, ωGk > 0 (k = 1, 2, …, n) ,

and that for stability of reactive power control in the chain microgrid is

ωqk, Dqk > 0 and ΔVk_max < Vko (k = 1, 2, …, n)

Combining the above two inequalities, the sufficient conditions for stability of the

microgrid chain are

Bk, βk, ωGk, ωqk, Dqk > 0 (k = 1, 2, …, n) , (4-39)

and ΔVk_max < Vko (k = 1, 2, …, n) (4-40)


115

4.3.2.1 Special Case: Grid-Interfaced Mode of Operation

The grid utility is considered as an infinite bus with no incremental change in

voltage, angle or frequency, i.e. ΔV1 = 0, Δδg = 0 and Δωg = 0. Accordingly, if the chain

microgrid is connected to the grid utility at the nth DR, an additional state variable –

voltage angle between the nth DR and grid supply (Δδng) is considered and the state vector

now becomes xT = [Δωh1 Δδ12 Δωh2 … Δδn-1.n Δωhn Δδng ΔV1 ΔV2 … ΔVn]. The modified system

matrix is then determined as

A = [A11]2nx2n ⊕ [A22]nxn (4-41)

⎡ ⎤
-ωb1 -b1(b1/β1)ωG1Po12 0 0 … 0

⎢ -1
0
-(b1+b2)Po12 1 b2Po23
b2(b2/β2)ωG2Po12 -ωb2 -b2(b2/β2)ωG2Po23 …
… 0
0

where A11 =⎢ ⎥ , (4-42)
0 b2Po12 -1 -(b2+b3)Po23 … 0
⎢ : : : ⎥
⎣ 0 0 0 … -1 -bnPong ⎦
ωq1V1o
⎡ ⎤
Dqtie12
-ωq1⎛1+ D ⎞ 0 0 … 0
⎝ q1 ⎠ Dq1X12
⎢ ωq2V2o
-ωq2⎛1+
Dqtie21+Dqtie23⎞ ωq2V2o ⎥
⎢ ⎥
0 … 0
Dq2X12 ⎝ D q2 ⎠ Dq2X23
A22 =
⎢ ⎥
ωq3V3o Dqtie32+Dqtie34⎞ ωq3V3o
0 -ωq3⎛1+ D X … 0
Dq3X23 ⎝ D q3 ⎠ q3 34

⎢ ⎥
: : :

⎣ 0 0 0 0 … -ωqn⎛1+

Dqtie.n.n-1 + Dqtie.n.g
D qn ⎠⎦

(4-43)

As seen in the above equations, parameters Pong and Dqtie.n.g in the above matrices

correspond to the grid-interface tie line. However, it can be proved that the Proposition 1

through Proposition 3 are still valid after modifications are made in matrices A11 and A22.
116

Accordingly, the various propositions presented in this section establish sufficient

conditions to ensure stability of a chain microgrid. Based on these conditions, the

parameters of active power-frequency and reactive power-voltage (magnitude)

controllers of its constituent DRs are designed in the following section.

4.4 Design Considerations for Generation

Controllers of a DR in a Microgrid

A design approach for active power-frequency and reactive power-voltage

controllers of the DRs in the microgrid is developed in this section. It is the outcome of

the investigation on dynamic behavior of microgrid conducted in the previous section. In

addition to that, dynamics of inverter internal controls that were covered in the previous

chapter as well as the IEEE P1547 standards [18] are taken into consideration for the

development of these guidelines. The design guidelines are presented below for the

active power-frequency and reactive power-voltage controllers on an individual basis.

4.4.1 Active Power-Frequency Controller

The various controller parameters to be designed for the active power-frequency

controller are – b, β and ωG. Among these parameters, b can be designed to satisfy the

IEEE P1547 standards [18], according to which the frequency of generation is to be

controlled to remain within with window of 59.3 Hz to 60.5 Hz. Therefore, the droop

gain bk for kth DR of rating PRk (W) is chosen as

bk = 0.5/PRk (Hz/W) (4-44)


117

The steady-state sharing of real power in a microgrid by the constituent DRs is in

accordance with their steady-state droops, (bk || βk). As it is generally desired to have the

steady-state power sharing by DRs in proportion to their ratings, for k = 1, 2, …, n,

(bk || βk) PRk = constant (4-45)

i.e. bk (1 - Kbβk) PRk = constant (4-46)

Since, bk has been already chosen according to (4-44), Kbβk is designed to be a

constant equal to ½. This value would cause frequency restoration of every DR in the

microgrid to half their initial (transient) droop upon occurrence of a load change. Thus,

Kbβk = ½ (4-47)

This restoration gain of Kbβk = ½ implies that the value of βk is identical to gain

bk.

The only parameter that remains to be designed for active power-frequency

control is the restoration loop time constant 1/ωGk. This time-constant is chosen taking

into consideration the inverter internal controller dynamics of the multi-loop regulator. It

was observed in Chapter 3 that the inner loop bandwidth is about one 60 Hz cycle.

Therefore, the active power-frequency controller restoration loop filter corner frequency

ωGk is chosen as 1 rad./s that is well below the inner controller’s bandwidth.

4.4.2 Reactive Power-Voltage Controller

The various controller parameters that are to be designed for reactive power-

voltage controller are Dq and ωq. Since the voltage droop of the DR is maximum in the

stand-alone mode of operation, this maximum droop should meet the IEEE P1547
118

standards governing DRs [18]. According to the IEEE P1547 standards, the permitted

range for voltages at the DR terminal is 88% to 110% (as a percentage of nominal

voltages). In per unit, the permitted range is given by 0.88 p.u. to 1.10 p.u. or in terms of

incremental variables maximum droop allowed in ΔV is -0.12 p.u. to 0.10 p.u..

Therefore, Dqk is chosen as 10 p.u. that would give a maximum droop in ΔV is -0.10 p.u.

to 0.10 p.u..

With the control parameter Dq fixed at 10, the other parameter ωq is designed so

that the steady-state is reached in a certain duration. The time-constant of the response of

the deviation in voltage to load change is given by the ratio 1/ωq. This time-constant is

chosen taking into consideration the inverter internal controller dynamics of the multi-

loop regulator. It was observed in Chapter 3 that the inner loop bandwidth is about one

60 Hz cycle. Therefore, the reactive power-voltage controller filter corner frequency ωqk

is chosen as 1 rad./s that is well below 2π(60 Hz).

The load reference set-point for an active power-frequency controller PL_ref is set

according to the desired generation from the DR under grid-interfaced mode of operation.

It is generally assumed that the DR power rating is well above this set-point so that the

DR is not driven into generating beyond its capacity. On the other hand, for the reactive

power-voltage controller, the load reference set-point QL_ref is ideally initialized to zero

as this would cause the voltage (magnitude) reference to decrease below nominal voltage

for inductive loads and to increase above the nominal voltage for capacitive loads.
119

4.5 Simulation Modeling and Results

Based on the guidelines presented in the previous section, simulation of a

microgrid is conducted in Matlab® SIMULINK™ software [83]. The model of the

microgrid illustrating two DRs along with the grid supply is displayed in Figure 4-13.

The subsystems DG1+Rload1 and DG2+Rload2 are the two DRs each with its local load.

Grid Network subsystem contains the model of a stiff three-phase voltage source in series

with an inductive reactance signifying the grid supply and the Thevenin impedance of the

grid network. The subsystem Tie line12 contains the modeling of the 100 yard tie-line as

a series RL line. An intermediate load is included in between the line. Synchronizing

logic for the connection between the two DRs is generated at the block Tie12_En, and the

same for the connection between the DR1 and the Grid Network is generated at the block

Grd_En.

Figure 4-13 Simulink model of the microgrid showing two DRs connected to the grid supply
120

Figure 4-14 gives a detailed structure of the simulation model for a DR with its

local load. The DR is an averaged model of an inverter that feeds the resistive load. The

power controller consists of the active power-frequency and reactive power-voltage

controllers described in Figure 4-4 and Figure 4-7, respectively. The system parameters

for the three-phase 16 kW/208 V/60 Hz system are tabulated in Table 4-1. This

simulation model assumes that the bandwidth of the internal inverter control is high

enough to approximate it as a unity gain.

Figure 4-14 Internal structure of the DG1+Rload1 subsystem

The simulation of microgrid is carried out in three modes to demonstrate the

operation of the generation controls. These are (i) single DR in stand-alone mode, (ii)

single DR in grid-interfaced mode and (iii) two DR interconnected mode. Selected results

are presented below during each mode of operation.


121

Table 4-1 System parameters for the generation controller


V1o 1.0 p.u.
V2o 1.0 p.u.
Vgrid 1.0 p.u.
b π rad./p.u.
Kbβ 1/2
Dq 10 p.u.
ωG 1 rad./s
ωq 1 rad./s
(i) Single DR in Stand-Alone Mode of Operation:

The simulation results illustrating the frequency response against change in PL

and PL-ref are plotted in Figure 4-15 and Figure 4-16, respectively. The initial value of PL-

ref is set at zero and therefore at a load of 0.25 p.u. the frequency is 59.9375 Hz. As seen

in Figure 4-15, a step change in PL from 0.25 p.u. to 0.5 p.u. gives a maximum transient

deviation in frequency of 0.125 Hz (determined by controller gain b) from the initial

frequency of 59.9375 Hz and a steady-state deviation in frequency of 0.0675 Hz

[determined by b(1 - Kbβ)*PL]. The time-constant is determined as (1 - Kbβ)/ωG = 500 ms.


1

0.75
PL (p.u.)

0.5

0.25

0
5 5.25 5.5 5.75 6 6.25

60.25

60
f (Hz)

59.75

59.5
5 5.25 5.5 5.75 6 6.25
Time (s)
Figure 4-15 Simulation waveforms showing response to a change in load real power demand PL when the
DR is operated as a stand-alone unit
122

Figure 4-16 illustrates the response for a change in PL-ref from 0 p.u. to 0.5 p.u.

that reflects the current load condition. As against the response to change in PL shown in

Figure 4-15, the step change in PL-ref gives a first-order lag response with a time-constant

of (1 - Kbβ)/ωG = 500 ms and a steady-state deviation in frequency of 0.0675 Hz

[determined by b(1 - Kbβ)*PL-ref].

0.75
PL-ref (p.u.)

0.5

0.25

0
5 5.25 5.5 5.75 6 6.25

60.25

60
f (Hz)

59.75

59.5
5 5.25 5.5 5.75 6 6.25
Time (s)
Figure 4-16 Simulation waveforms showing response to a change in load real power ref. set-point PL-ref to
match the load demand when the DR is operated as a stand-alone unit

(ii) Single DR in Grid-Interfaced Mode of Operation:

The grid is simulated as an infinite bus using differential equations of three-phase

voltage sources of fixed voltage amplitude and frequency in series with a tie-line of

reactance Xtie = 0.1 p.u.. As it was observed that the grid frequency in the laboratory

experiment test bed was 59.96 Hz, the same was employed in simulation also. However,

the nominal frequency set-point of the power controller is maintained at 60 Hz. The

results illustrating the response for a change in PL-ref from 0 p.u. to 0.5 p.u. at t = 5 s are

plotted in Figure 4-17. The load under this condition was 0.25 p.u.. Ideally, if the grid
123

frequency were at 60 Hz, the generated power would have followed the PL-ref set-point.

However, as the frequency is 59.96 Hz, the difference between PL-ref and generated power

is determined from the steady-state droop curves as (60-59.96)/[b(1-Kbβ)], which gives a

value of 0.16 p.u. for PL-ref = 0 p.u. and to 0.66 p.u. for PL-ref = 0.5 p.u.. The DR

frequency increases for a short duration to enable higher generation but eventually

returns to the grid frequency of 59.96 Hz.

1
PDR (p.u.)

0.5
0
-0.5
-1
5 5.25 5.5 5.75 6 6.25

60
f (Hz)

59.9

59.8
5 5.25 5.5 5.75 6 6.25
2
1.5
V (V)

1
0.5
0
5 5.25 5.5 5.75 6 6.25

Time (s)
Figure 4-17 Simulation waveforms showing response to a change in load power set-point PL-ref from 0 p.u.
to 0.5 p.u. when the DR is connected to an infinite bus of frequency 59.96 Hz

Figure 4-18 illustrates the simulation waveforms when the load PL is changed

from 0.25 p.u.to 0.5 p.u.. As seen in this figure, the DR supplies only the transient change

in load but in steady state its generation returns to a value determined by PL-ref. Figure

4-19 illustrates the response for a change in PL-ref from 0.5 p.u. to 0 p.u. at t = 10 s. As

seen in this figure, the DR generated power returns to the earlier values when PL-ref is

changed back to zero.


124

500
400
300
200
100
0
-100
-200
-300
-400
-500
9.75 9.8 9.85 9.9 9.95 10 10.05 10.1 10.15 10.2 10.25

80
60
40
20
0
-20
-40
-60
-80

9.75 9.8 9.85 9.9 9.95 10 10.05 10.1 10.15 10.2 10.25

Figure 4-18 Simulation waveforms illustrating the effect of load change at the DR terminals from 0.25 p.u.
to 0.5 p.u. while it is operated in grid-connected mode. Load terminal voltage is the top waveform and in
the bottom are the load current (solid), DR current (dashed) and tie-line current (dashdot).

1
0.5
PDR (p.u.)

0
-0.5
-1
10 10.25 10.5 10.75 11 11.25

60
f (Hz)

59.9

59.8
10 10.25 10.5 10.75 11 11.25
2
1.5
V (V)

1
0.5
0
10 10.25 10.5 10.75 11 11.25
Time (s)
Figure 4-19 Simulation waveforms showing response to a change in load power set-point PL-ref from 0.5
p.u. to 0 p.u. when the DR is connected to an infinite bus of frequency 59.96 Hz.
125

(iii) Two DR Interconnected Mode of Operation:

Figure 4-17 illustrates the results of DR1 response to a change in PL1 at its

terminals from 0.2 p.u. to 0.4 p.u.. The corresponding waveforms for DR2 that has a local

load of 0.4 p.u. are displayed in Figure 4-21. The load power set-points in both units are

the same equal to 0.4 p.u.. Hence, as seen in these two figures the overall load is shared

equally by the two DRs. The final steady-state frequency is 60 Hz as the load burden of

0.4 p.u. for each DR is equal to its PL-ref set-point.

However, the load change PL1 transient is felt by the local DR1 severely that its

generation initially increases to more than the final steady-state value. In contrast, the

farther DR2 increases gradually its generation to meet the increased load demand.
1
0.5
0
PDR1 (p.u.)

-0.5
-1
10 10.25 10.5 10.75 11 11.25

60

59.9
f1 (Hz)

59.8
10 10.25 10.5 10.75 11
2
1.5
1
V1 (V)

0.5
0
10 10.25 10.5 10.75 11 11.25
Time (s)
Figure 4-20 Simulation waveforms showing response of DR1 to a change in load PL1 from 0.2 p.u. to 0.4
p.u. when the two DRs are interconnected.

Figure 4-22 illustrates the DR1 load terminal voltage, DR1 current and the tie-line

current between the two DRs. As seen in the figure, since the two interconnected DRs are

supplying their local loads, the current through the intertie is zero.
126

PDR2 (p.u.)
0.5
0
-0.5
-1
10 10.25 10.5 10.75 11 11.25

f2 (Hz) 60

59.9

59.8
10 10.25 10.5 10.75 11 11.25
2
V2 (V)

0
10 10.25 10.5 10.75 11 11.25
Time (s)
Figure 4-21 Simulation waveforms showing response of DR2 to a change in load PL1 from 0.2 p.u. to 0.4
p.u. when the two DRs are interconnected.

500
400
300
200
100
0
-100
-200
-300
-400
-500
10.5 10.55 10.6 10.65 10.7 10.75 10.8 10.85 10.9 10.95 11

80
60
40
20
0
-20
-40
-60
-80

10.5 10.55 10.6 10.65 10.7 10.75 10.8 10.85 10.9 10.95 11

Figure 4-22 Simulation waveforms illustrating the two interconnected DRs supplying their local loads with
zero power flowing along the tie-line. Load terminal voltage is the top waveform and in the bottom are the
DR current (solid) and tie-line current (dashdot).
127

4.6 Summary

This chapter presented a generation control scheme consisting of active power-

frequency and reactive power-voltage controllers for the inverter based DRs. These are

droop-based controllers that allow decentralized operation of the microgrid without

communication between the DRs. Small-signal models were developed for microgrids

consisting of DRs connected in a chain topology. Investigations on dynamic (small-

signal) behavior of an n DR chain microgrid is carried out using eigenvalue analysis and

sufficient conditions were developed to guaranty their small-signal stability. These

eigenvalue investigations considered both the real and reactive power flows in the

microgrid. As a special case, the grid-interfaced mode of operation of the n DR microgrid

was also covered. It was proved by means of mathematical propositions that the n DR

chain microgrid is stable for all positive values of generation controller gains with the

maximum droop limited to be less than the nominal value. Guidelines were provided for

design of the active power-frequency and reactive power-voltage controllers. Besides the

results of dynamic behavior analysis, the IEEE P1547 performance specifications [18]

were also utilized in the design process. Finally, selected results are presented from

digital simulation in Matlab® SIMULINK™ software that establish the application of

DR systems to share power in a benign manner when interconnected.


128

Chapter 5 Further Studies in


Microgrid Dynamic Behavior

The dynamic behavior of a chain microgrid had been investigated in the previous

chapter. While this is one of the possible configurations of placement of DRs in a

microgrid, distribution networks of other architectures are also widespread. This chapter

is a continuation of the studies on dynamic behavior of microgrids. Practical issues are

investigated on the operation of the microgrid and preliminary results are presented. As

in the previous chapter, the following assumptions are made in the analysis of dynamic

behavior —

(i) The analysis is on linearized small-signal models of the power system

(ii) Inverter based DRs operate within their maximum capacity limits

(iii) Dynamics of the inverter internal controls are fast as compared to the

external power controls so that the internal controls can be neglected in

power flow analysis

(iv) Tie-lines between any two sources in the microgrid are purely inductive

in nature

The effect of relaxing assumption (iv) is included in this chapter.


129

5.1 Dynamic Behavior of Multiple DRs Connected in

Parallel

The previous chapter covered the sufficient conditions for a microgrid having a

chain structure have been derived. This section covers the dynamic behavior for the

microgrid architecture shown in Figure 5-1, designated a parallel configuration. As seen

in the figure, the kth DR is connected to the PCC by an inductive tie-line of reactance Xk.

PCC (Vp,δp)

X1 X2 Xn

(V1,δ1) (V2,δ2) (Vn,δn)

DR1 DR2 DRn

Figure 5-1 Single-line diagram of a microgrid parallel structure consisting of several DRs

Proposition 4: (i) The incremental change in voltage magnitude ΔVp at the PCC in a

parallel microgrid is a weighted average of the incremental changes in voltage

magnitudes ΔVk (k = 1, 2, …, n) of n DR units connected to the PCC. Or mathematically,

ΔVp = c1 ΔV1 + c2 ΔV2 + … + cn ΔVn

where ck (k = 1, 2, …, n) is a constant such that ck ≥ 0 and ∑ck = 1.

(ii) The incremental change in phase angle Δδp at the PCC in a parallel microgrid is a

weighted average of the incremental changes in phase angles Δδk (k = 1, 2, …, n) of n

DR units connected to the PCC. Or mathematically,


130

Δδp = d1 Δδ1 + d2 Δδ2 + … + dn Δδn

where dk (k = 1, 2, …, n) is a constant such that dk ≥ 0 and ∑dk = 1.

(iii) Also, the incremental change in frequency Δωp at the PCC in a parallel microgrid is

a weighted average of the incremental changes in frequencies Δωk (k = 1, 2, …, n) of n

DR units connected to the PCC. Or mathematically,

Δωp = d1 Δω1 + d2 Δω2 + … + dn Δωn

Proof: (i) As seen in Figure 5-1, n DRs are connected to the infinite bus through a star

point, known as point of common coupling (PCC). The voltage at the PCC is dependent

on the voltages at all the sources connected to it by a weighted average. This PCC

voltage can be determined by the phasor relationship

~
⎛Xprll⎞ ~ ⎛Xprll⎞ ~ ⎛Xprll⎞ ~
Vp = ⎜ X ⎟ V1 + ⎜ X ⎟ V2 + … + ⎜ X ⎟ Vn , (5-1)
⎝ 1⎠ ⎝ 2⎠ ⎝ n⎠

which is the ac phasor form of Millman’s theorem.

In the above equation, the expression for equivalent parallel inductance Xprll is

obtained from

1 1 1 1
Xprll = X1 + X2 + … + Xn (5-2)

Equation (5-1) for phasor voltage PCC can be rewritten in terms of complex

exponential quantities as

⎛Xprll⎞ ⎛Xprll⎞ ⎛Xprll⎞


Vp ejδp = ⎜ X ⎟ V1 ejδ1 + ⎜ X ⎟ V2 ejδ2 + … + ⎜ X ⎟ Vn ejδn (5-3)
⎝ 1⎠ ⎝ 2⎠ ⎝ n⎠

Expansion of complex exponential functions into real and imaginary parts gives
131

⎛Xprll⎞ ⎛Xprll⎞
Vp (cos(δp) + j sin(δp)) = ⎜ X ⎟ V1 (cos(δ1) + j sin(δ1)) + ⎜ X ⎟ V2 (cos(δ2) + j sin(δ2))
⎝ 1⎠ ⎝ 2⎠

⎛Xprll⎞
+ … + ⎜ X ⎟ Vn (cos(δn) + j sin(δn)) (5-4)
⎝ n⎠

Practically the difference between the phase angles of any two voltage sources

connected by an inductive tie-line is a very small value; and a phase angle reference can

be chosen such that at every voltage node

cos(δk) = 1 ⎫
⎬ ∀ k ∈ 1, 2, …, n (5-5)
sin(δk) = δk ⎭

As a particular case, if the infinite bus is one of the sources connected to the PCC,

it can be chosen as the phase angle reference and therefore its phase angle is zero.

The expression for voltage phasor at the PCC would then become

⎛Xprll⎞ ⎛Xprll⎞ ⎛Xprll⎞


Vp (1 + j δp) = ⎜ X ⎟ V1 (1 + j δ1) + ⎜ X ⎟ V2 (1 + j δ2) + … + ⎜ X ⎟ Vn (1 + j δn)
⎝ 1⎠ ⎝ 2⎠ ⎝ n⎠

(5-6)

Equating the real parts on left- and right-hand sides of (5-6), the magnitude

condition is determined as

Vp = c1 V1 + c2 V2 + … + cn Vn (5-7)

⎛Xprll⎞
where ck = ⎜ X ⎟ ≥ 0 and ∑ck = 1.
⎝ k⎠

If incremental changes are allowed in magnitudes of voltages, the relationship

between these incremental changes in voltages is given by


132

n
ΔVp =
∑ ⎛∂Vp⎞
⎜ ⎟ ΔVk
⎝∂Vk⎠
k=1

or ΔVp = c1 ΔV1 + c2 ΔV2 + … + cn ΔVn (5-8)

⎛Xprll⎞
where ck = ⎜ X ⎟ ≥ 0 and ∑ck = 1. This is valid under the assumption that the incremental
⎝ k⎠

change in reactance ΔXk ≈ 0 (for k = 1, 2, …, n) so that Δck ≈ 0 (for k = 1, 2, …, n).

(ii) Equating the imaginary parts on left- and right-hand sides of (5-6), the phase angle

condition is determined as

δp = d1' δ1 + d2' δ2 + … + dn' δn (5-9)

⎛Vk⎞
where dk' = ⎜V ⎟ ck ≥ 0 and ∑dk' = 1.
⎝ p⎠

If incremental changes are allowed in phase angles of voltages, since δk << 1 (for

k = 1, 2, …, n), the relationship between the incremental changes in phase angles is given

by

n
Δδp =
∑ ⎛⎛∂δp⎞ ⎛ ∂δp ⎞
⎜⎜ ⎟ Δδk + ⎜
⎝⎝∂δk⎠ ⎝∂dk'⎠

⎟ Δdk'⎟

k=1

or Δδp = d1 Δδ1 + d2 Δδ2 + … + dn Δδn (5-10)

⎛Vko⎞
where dk = ⎜V ⎟ ck ≥ 0 and ∑dk = 1 . This is valid under the assumption
⎝ po⎠

⎛ ∂δp ⎞ ⎛Vpo ΔVk - Vko ΔVp⎞


⎜ ⎟ Δdk' = ck δk ⎜ Vpo2 ⎟ ≈ 0 since δk << 1.
⎝∂dk'⎠ ⎝ ⎠
133

(iii) Since the coefficients ck and dk (for k = 1, 2, …, n) are independent of time, the

expression for incremental change in frequency Δωp at the PCC, after differentiating with

respect to time on both sides of the phase angle relationship given above, is determined

as

d
Δωp = dt (Δδp) = d1 Δω1 + d2 Δω2 + … + dn Δωn (5-11)

Therefore, the incremental change in frequency at the PCC is a weighted average

of the incremental changes in frequencies of the sources connected to it. This weighting

factor is determined from the tie-line inductance and voltage magnitudes on either side of

the tie-line.

With the help of the Proposition 4 stated above, it is now possible to study the

system matrix of the parallel microgrid. The state variable schematic of the parallel

microgrid illustrating the kth DR connected to the PCC by an inductive tie-line of

reactance Xk is given in Figure 5-2. The state vector x can be chosen such that its

transpose is given by xT = [Δωh1 Δδ1p Δωh2 Δδ2p Δωh3 … Δδn-1.p Δωhn ΔV1 ΔV2 … ΔVn].

The system matrix, A, from which the stability of the interconnected parallel

microgrid can be determined. The matrix A for the microgrid in Figure 5-2 is given by

⎡ [A11]2n-1x2n-1 [0]2n-1xn ⎤
A = ⎢⎢ ⎥
⎥ (5-12)
⎣ [0]nx2n-1 [A22]nxn ⎦

Matrix A can be represented as a direct sum of its principle submatrices as

A = [A11]2n-1x2n-1 ⊕ [A22]nxn , (5-13)


134

where A11 and A22 represent the active and reactive power flows, respectively. Their

eigenvalues, λ11 (of A11) and λ22 (of A22), can be determined independently.

ΔVk(s) ω qk Δω k(s)
1/Dqk bk
s

Dqtiek,p ω bk
Kbβk
s

Δω hk(s)
ck
Δδ k,p(s) 1
Pok,p
s
dk
ΔVp(s)
Vko
Δω p(s)
Xk,p

Figure 5-2 Small-signal state-variable schematic under zero input conditions of a microgrid having DRs in
parallel

Moreover, the sum of real powers flowing from all DRs to the PCC is equal to

zero.

i.e., ΔP1 + ΔP2 + … + ΔPn = 0 (5-14)

or Po1p Δδ1p + Po2p Δδ2p + … + Po.n.p Δδnp = 0 (5-15)

The partitioned submatrices of A refer to the active and reactive power flows,

respectively, and are determined from Figure 5-2 as

⎡ -(1-d ) -[b (1-d )+b d ]P ⎤


-ωb1 -b1(b1/β1)ωG1Po1p 0 … 0 0

⎢ 0 1 1 1 n n o1p d2 … (bn-1dn-1-bndn)Po.n-1.p dn

⎢ d (b d -b d )P ⎥
0 -ωb2 … 0 0
A11 = (5-16)
1 1 1 n n o1p -(1-d2) … (bn-1dn-1-bndn)Po.n-1.p dn

⎢ . . . … . . ⎥
⎣ 0 b (b /β )ω P n n n Gn o1p 0 … bn(bn/βn)ωGnPo.n-1.p -ωbn ⎦
135

Dqtie.1.p c1ωq1V1o c2ωq1V1o cnωq1V1o


⎡ -ωq1⎛1+ D ⎞+ D X
⎝ q1 ⎠ Dq1X1 … Dq1X1 ⎤
⎢ ⎥
q1 1

c1ωq2V2o Dqtie.2.p c2ωq2V2o cnωq2V2o


-ωq2⎛1+ D ⎞+ D X …
and A22 =
⎢ Dq2X2 ⎝ q2 ⎠ q2 2 Dq2X2

⎢ ⎥
. . … .

⎣ c1ωqnVno
DqnXn
c2ωqnVno
DqnXn
Dqtie.n.p cnωqnVno
… -ωqn⎛1+ D ⎞+ D X
⎝ qn ⎠ qn n

(5-17)

The properties of the eigenvalues of A11 and A22 are determined separately below,

beginning with A22 as it is easily solvable with known properties.

Proposition 5: The sufficient conditions for all eigenvalues of Α22 of a parallel microgrid

to have negative real parts are ΔVk_max < Vko and ωqk, Dqk > 0 (k = 1, 2, …, n), where

ΔVk_max is the maximum incremental change in voltage of kth DR unit when operated in

stand-alone mode and Vko is its nominal voltage.

Proof: By applying Gerschgorin theorem [86] on Α22 = (a22.i,j), it is well known that the

n
eigenvalues of A22 lie in disks with centers a22.i.i and radii ρ22.i = ∑ |a22.i.j| (where i = 1, 2,
j=1

…, n, and j ≠ i). With all of its diagonal elements strictly negative, A22 has only

eigenvalues with negative real parts if it is diagonally dominant under the following

condition (for i = 1, 2, …, n):

|a22.i.i| > ρ22.i (5-18)

ωqiVio n ω V
⎛ Dqtie.i.p⎞ qi io
i.e., ωqi⎜1+ D ⎟ > D X
⎝ qi ⎠ qi i
∑ cj = DqiXi (5-19)
j=1
136

Vio
i.e., (Dqi + Dqtie.i.p) > X (5-20)
i

2Vio - Vpo Vio


i.e., Dqi + Xi > X (5-21)
i

⎛Vio - Vpo⎞
i.e., Vio ⎜ X ⎟ > -Dqi Vio (5-22)
⎝ i ⎠

The expression in the left hand side of above inequality is a measure of total

reactive power generated by the ith DR under nominal conditions. If Qio denotes this

nominal value of generated reactive power, matrix A22 becomes diagonally dominant

Qio
i.e., - D < Vio (5-23)
qi

or ΔVi_max < Vio (5-24)

The above inequality signifies that the condition that diagonal dominance of A22

is when the (maximum) stand-alone droop in voltage at every DR terminal is strictly less

than its nominal value. This is generally true as practically the controller gains are never

designed to result in a voltage droop larger than the nominal value. Consequently, under

these conditions A22 has only negative real eigenvalues λ22.

Proposition 6: The sufficient conditions for all eigenvalues of A11 of a parallel microgrid

to have negative real parts are

ωbk + bkPokp = C1 ⎪⎫
⎬ (for k = 1, 2, …, n)
bkωGkPokp = C2 ⎪⎭

where C1 and C2 are some positive constants.

Proof: The dense matrix A11 is characteristic of the parallel microgrid where every DR
137

unit has some form of coupling to all remaining units. This coupling obscures and makes

it difficult to evaluate the properties of matrix A11. However, the matrix A11 can be made

sparse by the similar transformation, B11 = P11 A11 P11-1 where

-(ωb1+b1Po1p) -b1ωG1Po1p
⎡ ⎤
0 0 … b1Po1p

⎢ 1
0
0
0
0
-(ωb2+b2Po2p) -b2ωG2Po2p …
0 … -1
b2Po2p

B11 =⎢ ⎥ (5-25)
0 0 1 0 … -1
⎢ . . . . … . ⎥
⎣ b11.2n-1.1 b11.2n-1.2 b11.2n-1.3 b11.2n-1.4 … b11.2n-1.2n-1 ⎦
and the transformation matrix,

⎡0 ⎤
-1 -b1Po1p 0 0 … 0

⎢0 1
0
0
-1
0
-b2Po2p
… 0
… 0

=⎢
… 0 ⎥
P11 (5-26)
0 0 0 1
⎢. . . . … . ⎥
⎣ -d 1 (dnbn-d1b1)Po1p -d2 (dnbn-d2b2)Po2p … -d ⎦ n

In the matrix B11 = (b11.i,j), the elements of the 2n-1th row are determined as

b11.2n-1.2k-1 = dk(ωbn - ωbk + bnPonp - bkPokp) ⎪⎫


⎬ (for k = 1, 2, …, n-1) (5-27)
b11.2n-1.2k = dk(bnωGnPonp - bkωGkPokp) ⎪⎭

⎛ n ⎞
and b11.2n-1.2n-1 = - ⎜ωbn + bnPonp - ∑ dhbhPohp⎟ (5-28)
⎜ ⎟
⎝ h=1 ⎠

As seen in (5-25), all rows/columns except one in B11 = (b11.i,j) already have a

distinctive sign pattern determined as


138

- - 0 0 … +
⎡ ⎤
⎢ +
0
0
0
0
-
0
-


-
+

sign(B ) = ⎢
11
⎥ (5-29)
0 0 + 0 … -
⎢ . . . . … . ⎥
⎣ x x x x … x ⎦
The entries labeled ‘x’ in sign(B11) are not known a priori. Nevertheless, once

these entries are known the sign pattern of matrix B11 is adequate to demonstrate

qualitative stability known as sign stability. A matrix B11 is called sign stable if each

matrix C11 of the same qualitative (or sign) pattern as B11 (sign b11.i,j = sign c11.i,j for all

i,j) is stable regardless of the magnitudes of b11.i,j [87]. The stability of A11 can be judged

from that of B11 as the eigenvalues of similar matrices A11 and B11 are identical.

According to Jeffries et al [87, Theorem 2], the 2n-1 x 2n-1 real matrix B11 =

(b11.i,j) is sign stable if and only if it satisfies the following five conditions:

(i) b11.ii ≤ 0 for all i.

All the diagonal entries of B11 are non-positive.

(ii) b11.ij b11.ji ≤ 0, for all i ≠ j.

The off-diagonal entries b11.ij and b11.ji must not be of same sign.

(iii) The directed graph DB11 has no k-cycle for k ≥ 3.

In the directed graph DB11 — the vertex set consists of 2n-1

elements V = {1, 2, 3, …, 2n-1} and the edge set comprises the non-zero

off-diagonal ordered pairs of B11, ED = {(i,j): i ≠ j and b11.i,j ≠ 0} with

arrows starting at i and pointing towards j. Thereafter, the directed graph


139

DB11 must not contain cycles with more than 3 edges.

(iv) In every RB11-coloring of the undirected graph GB11, all vertices are

black.

In the undirected graph GB11 — the vertex set consists of 2n-1

elements V = {1, 2, 3, …, 2n-1}, the edge set comprises the ordered pairs

with non-zero product among off-diagonal entries of B11, EG = {(i,j): i ≠ j

and b11.i,j ≠ 0 ≠ b11.j,i} and rows having non-zero diagonal entries form the

set RB11 = {i: b11.i,i ≠ 0}. According to the definition [87], elements of RB11

are painted black and no black vertex is allowed to have precisely one

white neighbor, and each white vertex must have atleast one white

neighbor. Thereafter, the undirected graph GB11 must have RB11-coloring

with all vertices in black.

(v) The undirected graph GB11 admits a (V ∼ RB11)-complete matching.

According to [87], if condition (iii) is satisfied then condition (v) is

equivalent to some term in the expansion of det(B11) being different from

zero.

It is observed that the unique solution that satisfies all the above five conditions,

(i) through (v), is the last row off-diagonal entries of B11 being equal to zero; i.e. b11.2n-1,j

= 0 (for j = 1, 2, …, 2n-2).

In this case,
140

dk(ωbn - ωbk + bnPonp - bkPokp) = 0 ⎪⎫


⎬ (for k = 1, 2, …, n-1) (5-30)
dk(bnωGnPonp - bkωGkPokp) = 0 ⎪⎭

and therefore, condition (i) is satisfied for the last row element also as shown by

⎛ n ⎞ n
⎜ ⎟
b11.2n-1.2n-1 = - ωbn + bnPonp - ∑ dhbhPohp = - ∑ ωbk (5-31)
⎜ ⎟
⎝ h=1 ⎠ h=1

i.e., b11.2n-1.2n-1 < 0 (5-32)

Condition (ii) is verified by inspection. For condition (iii), the directed graph DB11

is illustrated in Figure 5-3. It has V = {1, 2, 3, …, 2n-1} as its vertex set and ED = {(1,2),

(1,2n-1), (2,1), (2,2n-1), (3,4), (3,2n-1), (4,3), (4,2n-1), (5,6), (5,2n-1), (6,5), (6,2n-1), …,

(2n-3,2n-2), (2n-3,2n-1), (2n-2,2n-3), (2n-2,2n-1)} as its edge set. As seen in the figure,

dark lines represent the elements of edge set. Light grey lines are also displayed in Figure

5-3 that represent ordered pairs involving the last (i.e. 2n-1th) row of B11. These light grey

lines are not part of the edge set ED as presence of even a single ordered pair involving

the last (i.e. 2n-1th) row of B11 in the edge set would bring about a 3-cycle. Thus, the only

possible solution to avoid a 3-cycle in directed graph DB11 is b11.2n-1,j = 0 (for j = 1, 2, …,

2n-2).
141

2n-1

b11.1,2n-1 b11.2n-2,2n-1

b11.2n-1,1 b11.2,2n-1 b11.3,2n-1


b11.4,2n-1 b11.2n-3,2n-1

b11.2n-1,4 b11.2n-1,2n-2
b11.2n-1,3
b11.2n-1,2 b11.2n-1,2n-3

b11.1,2 b11.3,4 b11.2n-3,2n-2

1 b11.2,1 2 3 b11.4,3 4 2n-3 b 2n-2


11.2n-3,2n-2

Figure 5-3 Directed graph DB11

For condition (iv), Figure 5-4 shows the undirected graph GB11 that has V = {1, 2,

3, …, 2n-1} as its vertex set, EG = {(1,2), (3,4), (5,6), …, (2n-3,2n-2)} as its edge set and

RB11 = {1, 3, 5, …, 2n-1}. As seen in Figure 5-4, the edges in GB11 correspond to the 2-

cycles in GB11. Besides, the elements of RB11 are painted black. However, as no black

vertex is allowed to have precisely one white neighbor, and each white vertex must have

atleast one white neighbor according to the definition [87], vertices {2, 4, 6, …, 2n-2} are

also painted black to prevent the elements of RB11 from having a single white neighbor.

As a consequence, in every RB11-coloring of the undirected graph GB11, all vertices are

black. Therefore, condition (iv) is also satisfied for b11.2n-1,j = 0 (j = 1, 2, …, 2n-2).

Likewise, condition (v) can be proved for non-negative control parameters as the

undirected graph GB11 admits a (V ∼ RB11)-complete matching. This is true according to

[87] since the determinant of B11 has some non-zero term in its expansion and condition
142

(iii) is satisfied.

2n-1

1 2 3 4 2n-3 2n-2

Figure 5-4 Undirected graph GB11

Thus, it is proved that the sufficient conditions for B11 (or A11) to be a stable

matrix are b11.2n-1,j = 0 (for j = 1, 2, …, 2n-2), which are the same as

dk(ωbn - ωbk + bnPonp - bkPokp) = 0 ⎪⎫


⎬ (for k = 1, 2, …, n-1) (5-30)
dk(bnωGnPonp - bkωGkPokp) = 0 ⎪⎭

ωbk + bkPokp = C1 ⎫⎪
or ⎬ (for k = 1, 2, …, n) (5-33)
bkωGkPokp = C2 ⎭⎪

where C1 and C2 are some positive constants.

Physical Significance of Proposition 6:

The sufficient conditions developed in Proposition 6 for stability of real power

flow in the parallel microgrid are more rigid as compared to those of a chain. These are

very conservative on stability and are important as they have a deeper physical meaning.

In Proposition 4, it was determined that the excursion in the frequency/angle/voltage at

the PCC of a parallel microgrid is a weighted average of the excursions of the same
143

quantities in all constituent DRs. Hence, the nearest DR to the PCC is the dominant DR

as it has a higher weightage in determining the PCC characteristics. In such a scenario,

the sufficient conditions in (5-33) recommend the following criteria in the design of a

parallel microgrid —

(i) As the steady-state sharing of real power by the DRs is according to

their droop characteristic (that has a slope of b || β), the DR that has a

smaller droop generates more power. It is recommended to locate this

smaller droop DR of higher generation capacity electrically closer to

the PCC making it the dominant DR in the parallel microgrid.

(ii) Frequency restoration time constant (i.e. 1/ωG) of the dominant DR

has to be large enough for easier tracking of its frequency by the

(smaller rated) DRs farther from the PCC.

Proposition 7: The sufficient conditions for ensuring stability of a microgrid in parallel

configuration are ωqk, Dqk > 0, ΔVk_max < Vko (for k = 1, 2, …, n) and

ωbk + bkPokp = C1 ⎪⎫
⎬ (for k = 1, 2, …, n)
bkωGkPokp = C2 ⎭⎪

where C1 and C2 are some positive constants.

Proof: It was determined in the previous two propositions that the sufficient conditions

for stability of reactive power control in the parallel microgrid is

ωqk, Dqk > 0 and ΔVk_max < Vko (k = 1, 2, …, n)

and that for real power control in the parallel microgrid is


144

ωbk + bkPokp = C1 ⎪⎫
⎬ (for k = 1, 2, …, n)
bkωGkPokp = C2 ⎪⎭

where C1 and C2 are some positive constants.

Therefore, the sufficient conditions for stability of the overall parallel microgrid

are the combination of the above two conditions.

5.1.1 Special Case: Grid-Interfaced Mode of

Operation

The grid utility is considered as an infinite bus with no incremental change in

voltage, angle or frequency, i.e. ΔV1 = 0, Δδg = 0 and Δωg = 0. If the parallel microgrid

is connected from the PCC to the infinite bus of voltage E through a tie-line of

inductance Xg, the following relations are true

⎫⎪
n
∑ck = 1 - c g

k=1
(5-34)

⎪⎭
n
∑dk = 1 - dg
k=1

where ck and dk are computed after taking into consideration parameters (Vg and Xg) of

the grid interface tie-line.

As a result, in the grid-interfaced mode of operation voltage angle between nth DR

and PCC (Δδnp) is considered as an additional state variable and thus the state vector now

becomes xT = [Δωh1 Δδ1p Δωh2 Δδ2p Δωh3 … Δδn-1.p Δωhn Δδn.p ΔV1 ΔV2 … ΔVn]. The system matrix

is then determined as

A = [A11]2nx2n ⊕ [A22]nxn (5-35)


145

⎡ -(1-d ) ⎤
-ωb1 -b1(b1/β1)ωG1Po1p 0 … 0 0

⎢ 0 1 -b1(1-d1)Po1p d2 … dn bndnPo.n.p

⎢ d ⎥
0 -ωb2 … 0 0
where A11 = , (5-36)
1 b1d1Po1p -(1-d2) … dn bndnPo.n.p
⎢ . ⎥
⎣ d ⎦
. . … . .
1 b1d1Po1p d2 … -(1-dn) -bn(1-dn)Po.n.p

Dqtie.1.p c1ωq1V1o c2ωq1V1o cnωq1V1o


⎡ -ωq1⎛1+ D ⎞+ D X
⎝ q1 ⎠ Dq1X1 … Dq1X1 ⎤
⎢ ⎥
q1 1

c1ωq2V2o Dqtie.2.p c2ωq2V2o cnωq2V2o


-ωq2⎛1+ D ⎞+ D X …
and A22 =
⎢ Dq2X2 ⎝ q2 ⎠ q2 2 Dq2X2

⎢ ⎥
. . … .

⎣ c1ωqnVno
DqnXn
c2ωqnVno
DqnXn
Dqtie.n.p cnωqnVno
… -ωqn⎛1+ D ⎞+ D X
⎝ qn ⎠ qn n

(5-37)

Proposition 5 is true for this case also as

ωqiVio n ω V
⎛ Dqtie.i.p⎞ qi io
ωqi⎜1+ D ⎟ > D X
⎝ qi ⎠ qi i
∑ cj = DqiXi (1 - cg) (for i = 1, 2, …, n) (5-38)
j=1

Likewise, Proposition 6 can be proved for this case after making the similar

transformation, B11 = P11 A11 P11-1 where

⎡0 ⎤
-1 -b1Po1p 0 0 … 0

⎢0 1
0
0
-1
0
-b2Po2p …
… 0
0

P11 =⎢ ⎥ (5-39)
. . . . … .
⎢ -d -d1b1Po1p -d2 -d2b2Po2p … -dnbnPo.n.p ⎥
⎣0 ⎦
1

d1 0 d2 … dn
146

5.1.2 Comments on the Sufficient Conditions for

Stability of Parallel Microgrid

In the case of reactive power flow in microgrid, the sufficient conditions for

stability of parallel structure are identical to those of the chain architecture dealt in the

previous chapter. The sufficient conditions derived for stability of real power flow in the

parallel microgrid (5-33), on the other hand, are rigid as compared to those of a chain. It

is to be noted that for a parallel microgrid, it is not necessary to satisfy the conditions

ωbk + bkPokp = C1 ⎫⎪
⎬ (for k = 1, 2, …, n) , (5-33)
bkωGkPokp = C2 ⎭⎪

to prevent instability. This is despite the fact that under such an event, one or more of

conditions (i) through (v) among the necessary and sufficient conditions for sign stability

[87, Theorem 2] are not met.

A sensitivity analysis is conducted in this section for the eigenvalues of a 5-DR

parallel microgrid. Consider 5 DRs to be operating in the baseline case with identical

parameters as in Table 4-1 that are designed in accordance with the guidelines provided

in the previous chapter. The tie-lines between each unit and the PCC are also assumed to

be identical for all DRs (equal to 0.1 p.u. reactance) so as to guaranty that conditions in

(5-33) are met by this baseline case. The eigenvalues of the system under this situation

are determined from MathCAD® software package as in Table 5-1. As seen in this table,

the baseline case is stable as it has all eigenvalues in the left half of the s-plane. The most

dominant eigenvalue is at -0.968. However, it is also observed from the eigenvalues that

there are repeated values among them. The eigenvalues of a chain microgrid, in contrast,
147

are real and distinct in nature.

Table 5-1 Eigenvalues of a 5 DR parallel microgrid for the baseline case


-32.448
-0.968
-2
-32.448
-0.968
-32.448
-0.968
-32.448
-0.968
For sensitivity analysis, values of one of the parameters are varied to violate one

or more of conditions (i) through (v) among the necessary and sufficient conditions for

sign stability [87, Theorem 2]. Jeffries et al have also proposed Lemmas [87, Lemma 1 –

Lemma 3] that suggest possibilities of migration of eigenvalues into the right half of s-

plane when the conditions are violated. It is observed that conditions (ii) and (iii) can get

easily violated.

The parameter chosen as the variable for this analysis is ωG5. By increasing or

decreasing ωG5 from 1 rad./s, these conditions (ii) and (iii) get violated. Using

MathCAD® software, the migration of eigenvalues is investigated as the parameter ωG5

is varied. It is observed that the dominant eigenvalue of -0.968 stays the same regardless

of the increase in ωG5. On the other hand, if ωG5 is decreased below 1 rad./s, the

migration of the dominant eigenvalue that was initially at -0.968 is towards the imaginary

axis of the complex s-plane as illustrated in Figure 5-5. However, it is verified that this

eigenvalue does not reach the imaginary axis regardless of the decrease in the value of

parameter ωG5.
148

90

120 60

0.08

0.06
150 30

0.04

0.02

eigenvals⎛ AωG5 ⎞ j 180 0 0


⎝ i⎠

210 330

240 300

270

Figure 5-5 Locus of the dominant eigenvalue of the real power flow in a 5-DR parallel microgrid as ωG5 is
reduced below 1 rad./s.

This section has thus proved for a parallel microgrid of 5 DRs that the sufficient

conditions in (5-33) of Proposition 6 even if violated do not cause any instability in the

microgrid. However, it remains to be proven that the violation does not cause any

instability for a general n DR parallel microgrid.

5.2 Effect of Changing Topologies on the Dynamic

Behavior of Microgrid

In the previous chapter as well as previous section, detailed eigenvalue studies

were carried out on the dynamic behavior of the microgrid for the chain and parallel

topologies. In this section, the dynamic behavior is investigated for the simplest of the

mesh networks, viz., a 3-DR mesh. It consists of 3-DRs with interconnection through an

inductive tie-line Xij between units i and j. Consider a 3-unit mesh network as shown in
149

Figure 5-6(a). Based on Proposition 4, the following Corollary 4.1 is proved on its small-

signal equivalence to a 3-unit parallel microgrid shown in Figure 5-6(b).

Corollary 4.1: A 3-unit mesh microgrid in Figure 5-6(a) is small-signal equivalent to a 3-

unit parallel microgrid in Figure 5-6(b) if

⎛X1 X2 + X2 X3 + X3 X1⎞
X12 = ⎜ ⎟ ,
⎝ X3 ⎠

⎛X1 X2 + X2 X3 + X3 X1⎞
X23 = ⎜ ⎟
⎝ X1 ⎠

⎛X1 X2 + X2 X3 + X3 X1⎞
and X31 = ⎜ ⎟
⎝ X2 ⎠

Proof: The small-signal real power generated by the kth DR unit of the 3-unit mesh

microgrid in Figure 5-6(a) is determined as

3
ΔPk = ∑( Po.k.i (Δδk - Δδi)) for k = 1, 2, 3 (5-40)
i=1
150

(ΔP1, ΔQ1)
(ΔV1, Δδ1)

X12 X31

(ΔV2, Δδ2) (ΔV3, Δδ3)


(ΔP2, ΔQ2) X23 (ΔP3, ΔQ3)
(a)

PCC (ΔVp, Δδp)

X1 X2 X3
(ΔV1, Δδ1) (ΔV2, Δδ2) (ΔV3, Δδ3)

(ΔP1, ΔQ1) (ΔP2, ΔQ2) (ΔP3, ΔQ3)

(b)

Figure 5-6 Small-signal equivalence of 3-unit mesh to a 3-unit parallel microgrid

On the other hand, the small-signal real power generated by the kth DR unit of the

3-unit parallel microgrid in Figure 5-6(b) is determined as

ΔPk = Po.k.p (Δδk - Δδp) for k = 1, 2, 3 (5-41)

Substituting the expression for Δδp from Proposition 4 with n = 3, the above

equation is simplified to

3
ΔPk = Po.k.p ∑(di (Δδk - Δδi)) for k = 1, 2, 3 (5-42)
i=1

The equivalence of the two small-signal models is established by equating the

right-hand sides of the above mesh and parallel equations as


151

Po.k.i = Po.k.p di for k = 1, 2, 3; i ≠ k (5-43)

Substituting the values of base powers Po.k.i, Po.k.p and coefficients di in the above

equation,

Vko Vio ⎛Vko Vpo⎞ ⎛ Vio Xprll⎞


Xki = ⎜⎝ Xk ⎟⎠ ⎜⎝Vpo . Xi ⎟⎠ for k = 1, 2, 3; i ≠ k (5-44)

where the parallel combination of reactances X1, X2 and X3 is given by

⎛ X1 X2 X3 ⎞
Xprll = ⎜X X + X X + X X ⎟ (5-45)
⎝ 1 2 2 3 3 1⎠

Consequently,

Xk Xi
Xki = X for k = 1, 2, 3; i ≠ k (5-46)
prll

⎛X1 X2 + X2 X3 + X3 X1⎞
i.e. X12 = ⎜ ⎟ , (5-47)
⎝ X3 ⎠

⎛X1 X2 + X2 X3 + X3 X1⎞
X23 = ⎜ ⎟ (5-48)
⎝ X1 ⎠

⎛X1 X2 + X2 X3 + X3 X1⎞
and X31 = ⎜ ⎟ (5-49)
⎝ X2 ⎠

Thus, the mesh-parallel analogy of a 3-unit microgrid has been proved in the case

of real power flow. Likewise, the same can also be derived for reactive power flow, but is

not covered here for purpose of brevity. The above relation between reactances in

Corollary 4.1 is identical to the well known star-delta transformation of network theory

[88].
152

5.3 Effect of Variation in Tie-Line R/X Ratio on the

Dynamic Behavior of Microgrid

The dynamic behavior of microgrid was analyzed assuming purely inductive tie-

lines. This is in accordance with the traditional power flow studies that focus on high

voltage transmission lines with R/X ratio small enough that it can be neglected [89].

However, line resistance is significant in medium voltage distribution systems and in

buildings. In such installations, the R/X ratio cannot be neglected as indicated by the

tables in Figure 5-7 [90]. In this section, the effect of variation in R/X ratio on

eigenvalues of the microgrid is analyzed for the simple case of a single DR connected to

the infinite bus.

Figure 5-8(a) displays the schematic of a single DR connected to the infinite bus

through a tie-line of inductance X1,g and R/X ratio that is denoted by σ. The small-signal

state-variable schematic of this system under zero input conditions is given in Figure

5-8(b). As seen in the figure, a finite value of σ creates coupling between the real and

reactive power loops.


153

Figure 5-7 Resistance and reactance values of building wires ‡


Source: [90]
154

PL1, QL1
σX1,g X1,g
Infinite Bus

V1∠δ1 E∠0o

DR1
(a) Single-line diagram

ΔV1(s) Δω 1(s)
ωq1
1/Dq1 b1
s
Dqtie1,g ω b1
1 + σ2 Kbβ1
s

σDqtie1,g Δωh1(s)
1+ σ 2 Po1,g
1 + σ2 Δδ1(s) 1
σPo1,g s

1 + σ2
(b) Small-signal state-variable schematic under zero input conditions

Figure 5-8 Single DR connected to the infinite bus through a tie-line of R/X ratio σ

The system differential equations of Figure 5-8(b) are

d b1 b1/β1 ωG1
dt Δωh1 = - ωb1 Δωh1 - 1 + σ2 (Po1g Δδ1g + σDqtie1g ΔV1) , (5-50)

d b1
dt Δδ1g = - Δωh1 - 1 + σ2 (Po1g Δδ1g + σDqtie1g ΔV1) (5-51)

d ⎛⎛ 1 ⎛Dqtie1g⎞⎞ σ ⎛Po1g⎞ ⎞
and ΔV1 = - ωq1 ⎜⎜1 + 2⎜ ⎟ ⎟ ΔV1- 2⎜ ⎟ Δδ1g⎟ (5-52)
dt ⎝⎝ 1 + σ ⎝ Dq1 ⎠⎠ 1 + σ ⎝ Dq1 ⎠ ⎠

The system matrix in this case is constructed as

-b1(b1/β1)ωG1σDqtie.1.g
⎡ ⎤
-b1(b1/β1)ωG1Po1g
-ωb1
1+σ2 1+σ2

A=
⎢ -1 -b1Po1g -b1σDqtie.1.g ⎥ (5-53)
⎢ 1+σ2 1+σ2

⎣0 ωq1σPo1g
Dq1(1+σ2)
-ωq1 ⎛1 +

1 ⎛Dqtie1g
1+σ2⎝ Dq1 ⎠⎠ ⎦
⎞⎞
155

As observed in Figure 5-8(b), the coupling between real and reactive power loops

(for σ ≠ 0) makes it no longer possible to decouple matrix A into A11 and A22

corresponding to the real and reactive power controls. Therefore, the eigenvalues of the

system matrix A as a whole need to be analyzed to determine the system dynamic

behavior for σ ≠ 0.

To investigate the effect of variation in R/X ratio on the dynamic behavior of the

single DR connected to infinite bus system, the dominant eigenvalues of A are plotted in

Figure 5-9 for typical system parameters as σ varies from 0 through 5 using MathCAD®

software package. As seen in Figure 5-9, the three eigenvalues advance towards the

imaginary axis in the left half of complex s-plane as the system order is increased. Thus,

the dynamic behavior of the single DR connected to infinite bus is greatly affected by

variation of the R/X ratio.

90
120 10 60
8

150 6 30
4
EIGi , 0 2

EIGi , 1 180 0 0

EIGi , 2

210 330

240 300
270

Figure 5-9 Dominant eigenvalues of the single DR connected to infinite bus as R/X ratio is increased from
0 through 5.
156

5.4 Summary

In this chapter, further eigenvalue analysis was conducted on microgrids. Small-

signal models were developed for the parallel connection of DRs in a microgrid to a

point of common coupling (PCC). It was proved that the excursion in voltage/phase

angle/frequency at the PCC are a weighted average of the excursions of the same

quantities at the DRs that are connected to it. Further, preliminary results for stability of

the parallel microgrid were reported in the form of sufficient conditions. The qualitative

result of this preliminary analysis is as follows:

The DR nearest to the PCC is the dominant DR as it has a higher weightage in

determining the excursions in voltage, phase angle and frequency at the PCC. In such a

scenario, the sufficient conditions for stability of a parallel microgrid recommend the

following criteria in the design of a parallel microgrid —

(iii)As the steady-state sharing of real power by the DRs is according to

their droop characteristic (that has a slope of b || β), the DR that has a

smaller droop generates more power. It is recommended to locate this

smaller droop DR of higher generation capacity electrically closer to

the PCC making it the dominant DR in the parallel microgrid.

(iv) Frequency restoration time constant (i.e. 1/ωG) of the dominant DR

has to be large enough for easier tracking of its frequency by the

(smaller rated) DRs farther from the PCC.

Sensitivity analysis was conducted on a 5 DR parallel microgrid to show that the

violation of sufficient conditions does not endanger the microgrid. However, it remains to
157

be proven that n DR parallel microgrids are stable over a large range of controller gains.

The mesh topology of microgrids was analyzed for the simplest case of a 3-DR mesh

network based on its small-signal analogy to the 3-DR parallel network. Finally, since a

non-zero R/X ratio is typical of wires in medium voltage distribution systems, effects of

R/X ratio were investigated in the active and reactive power flow. The microgrid system

models developed in this chapter can also be applied elsewhere to investigate the impact

of high penetration of distributed generation on the stability of interconnected power

systems.
158

Chapter 6 Laboratory Microgrid and


Experimental Results

6.1 Introduction and Objectives of the Laboratory

Microgrid

The preceding chapters dealt with the control strategy of the DR. These included

regulators for the load bus voltage as well as for the generated power of the DR. While

the voltage controller regulates the bus voltage against load variation as well as

imbalances due to unequal loads and voltage sags, the external generation controller

regulates the power flowing from the DR into the microgrid. The design of these

controllers and proof of the concept has been presented in the form of large signal

simulation done on a microgrid. In order to experimentally verify the concepts presented

in the earlier chapters of this thesis, and serve as a testbed platform to verify other control

strategies on hardware [66,91], a laboratory microgrid has been constructed [10,73]. This

chapter deals with the experimental results obtained on the laboratory microgrid. The

chapter begins with a description of the layout of this laboratory microgrid.

6.2 Microgrid Layout

A one-line diagram of the laboratory microgrid, constructed to experimentally


159

verify the controller described in this proposal, is displayed in Figure 6-1. This is same as

the schematic shown in Figure 1-3, but with protection apparatus necessary for safety of

equipment and personnel included. As seen in Figure 6-1, the microgrid consists of two

three-phase pulse-width modulated (PWM) inverter based DRs (viz., DG1 and DG2)

each having a generation capacity of 22 kVA and rated for three-phase/480 V/60 Hz.

Each inverter is equipped with LCL filters to attenuate the PWM switching ripples. The

interconnections between various DRs in the microgrid is made at 208 V level, and this is

facilitated by means of Δ-Y transformers that step down 480 V generated by each of the

DRs to the 208 V level. The two DRs are interconnected by a 100 yard tie-line with load

centers, Load1 and Load2, at each of their terminals as well as at an intermediate

location, Load12. The grid is connected to this network at the busbar B1, which is also

the load terminal of DG1 (Load1). Further, circuit breakers (CB) are provided in each

line for protection purposes, and contactors (CN) are included along those lines that

would need reconnection when both ends are active. A static switch SS1 consisting of

back-to-back thyristors is employed for connecting the microgrid to the grid mains

network as it would enable seamless transfer of the operation of the microgrid between

grid-interconnect and stand-alone modes operation.


CB3*
BS1 Load1

TX CN1 CB1 T1 SS1 CB2 T2 Lt1 Lf1

Δ-Y Y-Δ Cf1


FD1 DG1
B1
Grid
R1
Interconnect

75 yard
cable

CN2

CB4
CB5*
Load12
B Busbar
BS Bypass Switch B2 CB6*
C Capacitor Tie-line
CB Circuit Breaker w/ Shunt Trip Load2
Interconnect
CB* Main Circuit Breaker and
Breakers for Individual Loads 25 yard
T3
Lt2 Lf2
CN Contactor cable
FD Fused Disconnect
L Inductor Y-Δ Cf2 DG2
SS Static Switch B3
T Transformer
R Resistor

Figure 6-1 One line diagram of the laboratory microgrid consisting of two DRs
160
161

6.3 Components of the Laboratory Microgrid

The circuit schematic of a three-phase PWM inverter based stand-alone DR unit

(DG1/DG2 in Figure 6-1) along with its LC filter network is illustrated in Figure 6-2. The

following equipment has been employed in the construction of the lab microgrid.

Equipment ratings are given below

Voltage Source Inverter (VSI):

A low-cost off-the-shelf inverter drive designed for driving three-phase motors is

modified to operate as a distributed resource (DR. The maximum generation capacity of

each DR is 15 kW of real power and an equal amount to reactive power. For this reason,

the motor-drive inverter is overrated for 37 kW (i.e., 50 HP) to withstand possible over-

voltages and over-currents that occur in the event of fault conditions. A common dc-bus

configuration model of Allen-Bradley drive is chosen for this purpose. The name-plate

details of the motor drive inverter [92] are tabulated in Table 6-1. Dc input voltage to the

inverter is obtained from two programmable dc power supplies connected in series, each

capable of providing a maximum of 500V at 10 kW [93].


162

Table 6-1 Name plate details of the motor drive inverter ‡


Allen Bradley (Div. of Rockwell
Manufacturer
Automation)
DC input Common bus configuration

Voltage 675 - 800 V

Current 63 A

AC output 3-phase, 37 kVA

Voltage 0 – 575 V

Current 57.2 A

Power rating 50 HP/37 kW

Filter Selection:

Each of the inverter based DRs has an LC filter that is designed to attenuate the

switching ripples of the PWM inverter. The values of Lf and Cf are chosen such that the

overall filter size is optimized for a particular corner frequency. The designed values of

filter parameters for the DRs of the laboratory microgrid are tabulated in Table 6-2.

Table 6-2 Inverter LC filter parameters


Lf 0.97 mH

Rf 0.21 Ω

Cf 30 μF

fc 933 Hz


Source: [92]
INV1/2
VDC+
5 mH/ 0.97 mH/
phase phase
DC1_1/DC2_1
Lt1/2 Lf1/2
A
B
C

E1/2 IT1/2 VC1/2 I V1/2


L1/2 DC1_2/DC2_2

VDC-
Cf1/2

30 mF/phase
VC1/2 , E1/2
CTRL1/2
IL1/2 , IT1/2
DSP1/2

C Capacitor
CTRL Control and Interface Board
DC PQ500-20 DCPower Supply
DSP Motorola56F805 DSPEVM
INV AB Drive Inverter
L Inductor

Figure 6-2 Circuit schematic of the PWM inverter based DR unit with its LCL filter
163
164

Transformer Selection:

A Δ-Y transformer with each DR inverter would allow connection of both single

and three-phase loads to the DR. The specifications of this transformer in the setup are

given below [94]:

Table 6-3 Transformer specifications for DR-load interface


KVA 45
Primary 480 V (Δ)
Secondary 208V (120 V-Y)
%R 3.8
%X 3.8
%Z 5.4

The transformer secondary is connected to a three-phase ac load center (SquareD

make - Cat. #QO124L100G) that has 24 poles for various three-wire and four-wire loads.

Furthermore, a main circuit breaker with shunt trip (SquareD make - Cat.

#KAL3G1001021) is provided before the load center for disconnecting the power supply

in the event of large unbalance (>10%) in the terminal voltage, phase-loss, over-voltage

or under-voltage.

Busbar Wiring and Switchgear Selection:

Figure 6-3 illustrates the three-phase schematic of the wiring layout at the load

terminal busbar. As seen in the figure, the 480 V (Δ) primary side of the transformer

consists of the DR and its LCL filter network and its 208 V (120 V - Y) secondary side of

the transformer is connected to the four-conductor busbar. The Y-side of the three-phase

transformer allows access to the neutral thereby permitting connection of single-phase as


165

well as three-phase loads to the busbar.

Furthermore, Figure 6-3 displays the schematic of the busbar to the load and tie-

line. The various single-phase and three-phase loads are connected through a load center,

which consists of a main circuit breaker (MCB) and individual cicuit breakers (LCBs) for

each load. It is to be noted that a conductor of larger size is employed for the neutral

wiring as compared to the three phases. This is due to the possible over-loading of the

neutral phase in the event of several single-phase or harmonic loads are present in the

microgrid system [90].

DR/DR and Microgrid/Utility Interface:

A three-phase static switch employed for interconnection of the microgrid

network to the grid mains (refer Figure 6-1 for a one-line diagram of the microgrid). The

static switch consists of back-to-back thyristors in each phase [91] with associated

snubber circuitry. A synchronization logic and thyristor firing circuitry provides the gate

signals to the static switch on the basis of the three-phase voltage waveform information

on both sides of the static switch.


208 V(120V-Y) :480 V

X1
B3 H1
A”

AWG #3 H2
B” E2
Tieline
C”
X2 X0 X3
AWG #2 H3

n c b a 45 kVA / Δ -Y
MCB6 Transformer #3
LCB62

B Busbar
MCB Main CircuitBreaker
LCB61
LC Load Center
LCB Breakers for Individual Loads

LC2

Figure 6-3 Wiring layout of the transformer, loads and the tie-line connections to the busbar (to be modified for contactor)
166
167

DSP Controller Interface:

The PWM switching logic of each VSI based DR is generated using a digital

signal processor (DSP) [72]. Some important features of the DSP are tabulated in Table

6-4.

Table 6-4 Digital signal processing platform features [72]


Motorola DSP 56F805 Evaluation Board
16-bit Fixed Point DSP 512 Words Program
RAM
Upto 40 MIPS 4K Words Data Flash
Two 4-Channel, 12-bit 2K Words Data RAM
ADCs
Two 6-Channel PWM 2K Words Boot Flash
Modules
Four General Purpose Upto 64K Words Each of
Quad Timers External Program and
31.5K Words Program Data Memory
Flash

Inverter internal multi-loop regulator as well as the external generation controllers

are programmed in the DSP. Motorola Codewarrior™ and PCMaster™ are used for

debugging/editing the code.

Some photographs of the laboratory microgrid are illustrated below:


168

Figure 6-4 Distributed generator DG1 supplying power to its local load as well as tied to the microgrid
169

Figure 6-5 Connections of the distributed generators DG2 and DG3 ‡ to the microgrid


DG3 is not operational and hence is not shown in the circuit schematic of Figure 6-1
170

Figure 6-6 Intermediate tie-line load bus junction


171

Figure 6-7 Static switch interface for utility/grid interconnection of the microgrid ‡


H. Zhang [91]
172

6.4 Experimental Results

Several experiments were carried out in the laboratory microgrid for testing the

inverter internal controls and the external generation controls that were developed in the

previous chapters. Few selected waveforms are presented in this section demonstrating

the performance of the controllers designed earlier.

6.4.1 Inverter Internal Controls

In this subsection, the results from experiments conducted on the performance of

the multi-loop regulator are presented. The controllers were implemented in the

Motorola® 56F805 digital signal processor (DSP) evaluation board [72] whose features

are described in Table 6-4.

Case (i): Response of the multi-loop vector regulator:

Figure 6-8 illustrates the performance of the voltage regulator for an incremental

step change in voltage command from 70% to 100% of the rated voltage of 480 VRMS.

Waveforms displayed in Figure 6-8 are the αβ components of reference voltage, actual

voltage that tracks the reference and the error between the reference and actual values.

These are internal variables in 56F805 DSP and have been obtained from PC Master™ of

Metrowerks® CodeWarrior™. As seen in the figure, the capacitor voltage response

reaches close to the steady state in about three 60 Hz cycles. During this transient, the

actual voltage of filter capacitor vCf(t) across phases A and C was captured using a scope

and is displayed in Figure 6-9. This was followed by tests under balanced three-phase

loads at various power factors.


173

in p.u.
in p.u.

Time (ms)
Figure 6-8 Operation of the voltage control loop for vCf with the response vCf for a complex exponential
step input voltage reference vCf†. DSP obtained from PC Master™ of Metrowerks® CodeWarrior™.

Figure 6-9 Experimental waveform of line voltage vCf,AC illustrating the response for a step change in
voltage reference vCf† from 70% to 100% of rated voltage of 480 VRMS

Case (ii): Balanced three-phase load test:

The test in this case involves a three-phase 10 kW resistive load. Figure 6-10

illustrates the steady-state voltage waveforms at the load terminals across phases ac and
174

bc as well as line currents through phases a and b.

Figure 6-10 Load terminal line-line voltage ac & bc (top two) and load current in a & b phases (bottom
two) for a balanced 10 kW resistive load. Current is scaled at 1 A = 1 V.

Another test carried out to observe the dynamic response of the controller under

balanced conditions involved abrupt load change. Figure 6-11 demonstrates the response

when the DR is initially operating on no-load and a 2 kW three-phase balanced resistive

load was abruptly connected using an three-pole single-throw switch. As seen in Figure

6-11, load terminal voltages on the secondary side of 480V-Δ : 208 (120V-Y)

transformer respond fairly well with minor glitches at the instant of no-load to load

switching. If it is assumed that the load is switched ON at t = tON, as the current is zero at

t = tON– the voltage is also zero; and as the ac current begins to flow through the load at t

= tON+ the voltage at load terminals regains its sinusoidal form. This voltage recovery

took place within a small fraction of a 60 Hz cycle.


175

Figure 6-11 Experimental waveforms illustrating DR operation during transition from no-load to a 2 kW
balanced three-phase resistive. Load terminal line-line voltage ac & bc (top two) and load current in a & b
phases (bottom two). Current is scaled at 1 A = 1 V.

The dynamic response of DR against abrupt load-change from loaded condition to

no-load is illustrated Figure 6-12. As seen in this figure, the DR responds satisfactorily

without any voltage glitch during this transient.


176

Figure 6-12 Experimental waveforms illustrating DR operation during transition from a 2 kW balanced
three-phase resistive load to no-load. Load terminal line-line voltage ac & bc (top two) and load current in
a & b phases (bottom two). Current is scaled at 1 A = 1 V.

Thus, an inverter-based DR equipped with the multi-loop voltage controller is

proved to satisfactorily regulate the terminal voltage under balanced load conditions. The

next case involved unbalanced three-phase loading of the DR.

Case (iii): Unbalanced three-phase load test:

Inorder to test the performance of the controllers under unbalanced conditions, the

phase a connection of a 10 kW load is abruptly removed. Figure 6-13 illustrates the

voltage waveforms at the load terminals across phases ac and bc as well as line currents

through phases a and b. Table 6-5 provides a comparison of the regulated load terminal

voltage with that of unregulated case. As seen in Table 6-5, the line-line voltages are

regulated fairly close to 208 V in all three phases. Small inconsistencies between

different line voltages can be attributed to discrepancies in the scaling of voltage/current


177

sensing circuits as well as in estimation of the leakage inductance of the second reactor

and transformer.

Figure 6-13 Experimental steady-state waveforms illustrating the load terminal voltage and load current
when load on phase a is abruptly disconnected. Current is scaled at 1 A = 1 V.

Table 6-5 Evaluation of voltage regulator performance under single phasing

Description Vab (V) Vbc (V) Vca (V) Imbalance


(%)
w/o vector controller 217 200 200 5.8

w/ vector controller 211 203 203 2.1

During unbalanced conditions, the operation of voltage command modifier

consisting of sequence filters PSF and NSF has a major impact on the effectiveness of

regulation of load terminal voltage. The functioning of PSF is demonstrated by the DSP

internal variables displayed in Figure 6-14 that are plotted from PC Master™ of

Metrowerks® CodeWarrior™. It is to be noted that these internal variables are in p.u..


178

Figure 6-14 illustrates the filter input that is the transformer primary current, PSF output

and the NSF output waveforms.

in p.u.
in p.u.
in p.u.

Time (ms)
Figure 6-14 Transformer primary current, PSF output and NSF output waveforms in the PC Master™ for
an unbalanced load

As seen in Figure 6-14, the transformer primary current also contains harmonics

due to its magnetizing current. The sequence filters, PSF and NSF are employed to

extract the positive and negative sequence components of this distorted current, The PSF

and NSF have been designed to pass one sequence and reject the other at 60 Hz in an

efficient manner, but they do pass with some attenuation the higher-order harmonic

components. Nevertheless, Figure 6-14 demonstrates that the PSF and NSF satisfactorily

extract the positive and negative sequence components.

The next subsection deals with tests on external generation controllers in the

experimental laboratory microgrid.

6.4.2 External Generation Controls

The design of controller parameters is based on the guidelines presented in


179

Chapter 4. To demonstrate the operation of the generation controls, tests on microgrid are

carried out in three modes, viz., (i) single DR in stand-alone mode, (ii) single DR in grid-

interfaced mode and (iii) two DR interconnected mode. Selected results are presented

below during each mode of operation.

(i) Single DR in Stand-Alone Mode of Operation:

The experimental results illustrating the response of frequency against change in

PL and PL-ref are plotted in Figure 6-15 and Figure 6-16, respectively. These waveforms

were obtained from the DSP internal variables in PC Master™ of Metrowerks®

CodeWarrior™. The initial value of PL-ref is set at zero and therefore at a load of 0.25 p.u.

the frequency is 59.9375 Hz. As seen in Figure 6-15, a step change in PL from 0.25 p.u.

to 0.5 p.u. gives a maximum transient deviation in frequency of 0.125 Hz (determined by

controller gain b) from the initial frequency of 59.9375 Hz and a steady-state deviation in

frequency of 0.0675 Hz [determined by b(1 - Kbβ)*PL]. The time-constant is determined

as (1 - Kbβ)/ωG = 500 ms.


in p.u.
in Hz

Time (ms)
Figure 6-15 Experimental waveforms showing response to a change in load real power demand PL when
the DR is operated as a stand-alone unit
180

Figure 6-16 illustrates the response for a change in PL-ref from 0 p.u. to 0.5 p.u.

that reflects the current load condition. As against the response to change in PL shown in

Figure 6-15, the step change in PL-ref gives a first-order lag response with a time-constant

of (1 - Kbβ)/ωG = 500 ms and a steady-state deviation in frequency of 0.0675 Hz

[determined by b(1 - Kbβ)*PL-ref].


in p.u.
in Hz

Time (ms)
Figure 6-16 Experimental waveforms showing response to a change in load real power ref. set-point PL-ref
to match the load demand when the DR is operated as a stand-alone unit

(ii) Single DR in Grid-Interfaced Mode of Operation:

The grid frequency as observed from the point of common coupling (PCC) in the

laboratory experiment test bed was 59.96 Hz. However, the nominal frequency set-point

of the power controllers programmed in the DSP is maintained at 60 Hz. The results

illustrating the response for a change in PL-ref from 0 p.u. to 0.5 p.u. are plotted in Figure

6-17. The load under this condition was 0.25 p.u.. Ideally, if the grid frequency were at

60 Hz, the generated power would have followed the PL-ref set-point. However, as the

frequency is 59.96 Hz, the difference between PL-ref and generated power is determined
181

from the steady-state droop curves as (60-59.96)/[b(1-Kbβ)], which gives a value of 0.16

p.u. for PL-ref = 0 p.u. and to 0.66 p.u. for PL-ref = 0.5 p.u.. The DR frequency increases for

a short duration to enable higher generation but eventually returns to the grid frequency

of 59.96 Hz.
in p.u.
in Hz
in V

Time (ms)
Figure 6-17 Experimental waveforms showing response to a change in load power set-point PL-ref from 0
p.u. to 0.5 p.u. when the DR is connected to grid interface PCC of frequency 59.96 Hz

Figure 6-18 illustrates the experimental waveforms when the load PL is changed

from 0.25 p.u.to 0.5 p.u.. As seen in this figure, the DR supplies only the transient change

in load but in steady state its generation returns to a value determined by PL-ref. Figure

6-19 illustrates the response for a change in PL-ref from 0.5 p.u. to 0 p.u.. As seen in this

figure, the DR generated power returns close to its earlier values when PL-ref is changed

back to zero.
182

Figure 6-18 Experimental waveforms illustrating the effect of load change at the DR terminals from 0.25
p.u. to 0.5 p.u. while it is operated in grid-connected mode. Load terminal voltage is the top waveform and
in the bottom are the load current (solid), DR current (dashed) and tie-line current (dashdot). Voltage is at
100 V/div. and current is at 20 A/div.
in p.u..
in Hz
in V

Time (ms)
Figure 6-19 Experimental waveforms showing response to a change in load power set-point PL-ref from 0.5
p.u. to 0 p.u. when the DR is connected to the grid interface PCC of frequency 59.96 Hz.
183

(iii) Two DR Interconnected Mode of Operation:

Figure 6-20 illustrates the results of DR1 response to a change in PL1 at its

terminals from 0.2 p.u. to 0.4 p.u.. The corresponding waveforms for DR2 that has a local

load of 0.4 p.u. are displayed in Figure 6-21. The load power set-points in both units are

the same equal to 0.4 p.u.. Hence, as seen in these two figures the overall load is shared

equally by the two DRs. The final steady-state frequency is 60 Hz as the load burden of

0.4 p.u. for each DR is equal to its PL-ref set-point.


in p.u..
in Hz
in V

Time (ms)
Figure 6-20 Experimental waveforms showing response of DR1 to a change in load PL1 from 0.2 p.u. to
0.4 p.u. when the two DRs are interconnected.
in p.u..
in Hz
in V

Time (ms)
Figure 6-21 Experimental waveforms showing response of DR2 to a change in load PL1 from 0.2 p.u. to
0.4 p.u. when the two DRs are interconnected.
184

Figure 6-22 illustrates the DR1 load terminal voltage, DR1 current and the tie-line

current between the two DRs. As seen in the figure, since the two interconnected DRs are

supplying their local loads, the current through the intertie contains chiefly the

transformer magnetizing current.

Figure 6-22 Experimental waveforms illustrating the two interconnected DRs supplying their local loads
with zero power flowing along the tie-line. Load terminal voltage is the top waveform and in the bottom
are the DR current (solid) and tie-line current (dashdot). Voltage is at 100 V/div. and current is at 20 A/div.
185

Chapter 7 Conclusions and Future


Work

This thesis has presented control strategies for VSI-based distributed resources in

a microgrid. The various conclusions made in this report are as follows:

The concepts of feedback controller design for dc quantities using Bode plots of

loop gain have been extended for application towards three-phase ac space-vector

quantities. These vector regulators have a PI controller cascaded with an oscillator. They

give a zero steady-state error when the oscillator frequency is equal to the excitation

frequency. The time response characteristics of peak overshoot and settling time for

vector regulated systems are evaluated against the frequency-domain parameters like the

phase margin and gain cross over frequency. When transformed to synchronous rotating

frames, these vector regulators are equivalent to the high performance complex vector

regulators proposed by Briz et al [21]. It was shown through a comparative evaluation of

the processing involved that the synchronous frame based implementation of the vector

regulator requires more computation power as compared to stationary frame

implementation of the same regulator for the application under study.

A multi-loop regulator employing the proposed stationary frame vector regulator

was applied to inverter based distributed resources to regulate the load terminal voltage

with zero steady-state error under unbalanced conditions. Frequency domain analysis
186

techniques in the form of Bode plots for individual sequence components were utilized in

the design procedure. Further, by making use of complex transfer functions, novel

negative and positive frequency/sequence selective filters that contain complex band-pass

as well as complex band-reject sections were proposed. The stability of complex filters

has been explained using Nyquist diagrams and Bode plots. The proposed

frequency/sequence selective complex filters give a high quality output without any

phase distortion. Simulation results were presented from digital simulation in Matlab®

SIMULINK™ software that establish the application of DR systems to mitigate the

effects of power quality phenomena at the sensitive load bus.

An external generation control scheme consisting of active power-frequency and

reactive power-voltage controllers for the inverter based DRs was presented. These are

droop-based controllers that allow decentralized operation of the microgrid without

communication between the DRs. Small-signal models were developed for microgrids

consisting of DRs connected in chain and parallel topologies. Investigations on dynamic

(small-signal) behavior of an n DR microgrid structures were carried out using

eigenvalue analysis and sufficient conditions were developed to ensure their small-signal

stability. These eigenvalue investigations considered both the real and reactive power

flows in the microgrid. It was proved by means of mathematical propositions that the n

DR chain microgrid is stable for all positive values of generation controller gains with

the maximum droop limited to be less than the nominal value. Guidelines were provided

for design of the active power-frequency and reactive power-voltage controllers. Besides

the results of dynamic behavior analysis, the IEEE P1547 performance specifications
187

[18] were also utilized in the design process. Simulation results are presented from digital

simulation in Matlab® SIMULINK™ software that establish the application of DR

systems to share power in a benign manner when interconnected.

In parallel microgrids, it was proved that the excursion in voltage/phase

angle/frequency at the point of common coupling (PCC) is a weighted average of the

excursions of same quantity at the DRs that are connected to it. As a consequence, the

DR nearest to the PCC is the dominant DR as it has a higher weightage in determining

these excursions at the PCC. Preliminary results for stability of the parallel microgrid

were presented in the form of sufficient conditions. The preliminary results recommend

locating the smaller droop DR of higher generation capacity electrically closer to the

PCC making it the dominant DR in the parallel microgrid. Further, the frequency

restoration time constant (i.e. 1/ωG) of this dominant DR has to be large enough for easier

tracking of its frequency by the (smaller rated) DRs farther from the PCC. While the

violation of these sufficient conditions in a parallel microgrid has been shown to not

endanger a 5 DR microgrid, it remains to be proven when n DRs are present. The mesh

topology of microgrids was analyzed for the simplest case of a 3-DR mesh network based

on its small-signal analogy to the 3-DR parallel network. As a non-zero R/X ratio is

typical of wires in medium voltage distribution systems, effects of R/X ratio were

investigated in the active and reactive power flow.


188

7.1 Contributions

The various contributions made in this thesis proposal are as follows:

7.1.1 Stationary Frame Vector Regulators

• Systematic study of feedback regulators on space-vector quantities in

stationary reference frame and characterization of their operation under

balanced and unbalanced conditions.

• Systematic use of complex coefficient oscillators in the loop of stationary

frame feedback regulators for tracking and disturbance rejection at

selected frequencies.

• Categorization of transient response for vector controlled second-order

systems with complex coefficients to ac step commands using benchmark

indices like peak overshoot, settling time and phase margin.

• Proposal of a PI regulator cascaded with an oscillator as an alternative to

implementing two synchronous frame complex regulators in oppositely

rotating reference frames.

7.1.2 Inverter Internal Controls - Multi-Loop

Regulator

• A multi-loop stationary frame vector regulator consisting of an inner loop

for the filter inductor current and an outer loop for the filter capacitor

voltage is proposed.
189

• Provision made in the regulator for compensation of voltage drop across

transformer leakage impedance.

• High performance positive and negative sequence filters that contain

complex band-pass as well as complex band-reject sections at the chosen

frequency/sequence were proposed.

• Evaluation of stability criteria of filters with complex transfer functions

using Nyquist criteria.

7.1.3 Generation Control in an Inverter Based

Distributed Resource (DR)

• Systematic design oriented analysis of generation control structures

presented in the literature.

• Design guidelines are presented on the basis of IEEE P1547 for DR

interconnection to the Electric Power System (EPS).

7.1.4 Dynamic Behavior of Microgrids in Various

Topologies

• Developed small-signal models for general n DR microgrids consisting of

DRs connected in chain and parallel topologies. Considered both the real

and reactive power flows in the microgrid.

• Developed sufficient conditions to guaranty the small-signal stability of

chain and parallel microgrids.


190

• In parallel microgrids, proved that the excursion in voltage/phase angle/

frequency at the point of common coupling (PCC) is a weighted average

of the excursions of the same quantities at the DRs that are connected to it.

• Analyzed the mesh topology of microgrids for the simplest case of a 3-DR

mesh network based on its small-signal analogy to the 3-DR parallel

network.

• Investigated the effects of a tie-line with non-zero R/X ratio in the

dynamic behavior of a single DR connected to infinite bus system.

7.2 Proposed Further Research

The research directions developed here may be extended further along the

following lines of investigations:

1. Stationary Frame Vector Regulators

™ As the single-phase systems can be considered as a particular case of

unbalanced three-phase systems, the PI regulator cascaded with an

oscillator can be applied to control them.

™ Application of the PI regulator cascaded with oscillator to three-phase

four-wire systems where each of the three phases are controlled

individually.

™ Application of the PI regulator cascaded with a sum of oscillators for

active filtering. Each oscillator is tuned for the frequency and

sequence of the particular harmonic.


191

™ Development of frequency/sequence selective filters for harmonic

selective filtering in active filtering applications.

2. Generation (active power-frequency and reactive power-voltage) control

and dynamic behavior of microgrid

™ Integration of anti-islanding algorithms into the generation controller.

™ Evaluating the stability of microgrids with general mesh structures and

by considering tie-lines with non-zero R/X ratios.

™ The impact of integrating inverter based and synchronous machine

based distributed generators can be investigated through a dynamic

behavior analysis of the integrated system.

™ Design of a non-interacting control strategy in the generation

controller for DR for overcoming the effects of larger R/X ratio

™ Apply the technique of Lyapunov functions to establish the stability of

a general n DR parallel microgrid over a wider range of controller

gains. In this regard, the work earlier carried out on power systems

[97,98] can be extended to deal with microgrids. It is felt that the

techniques of matrix analysis using small signal models have reached

their limits to prove useful in addressing the complexities of a true

microgrid with diverse dynamic properties.

3. Field test the overall control scheme under realistic industrial grid

conditions including

™ Motor starting
192

™ Non-linear (rectifier loads)

™ Capacitor switching

™ Line-line and line-ground faults

™ Grid connect and disconnect operations

™ Loss of load

™ Loss of line

7.3 Summary

Control of voltage source inverters (VSIs) has seen tremendous advancements in

the past several decades. However, the major focus was on motor drives and

uninterruptible power supplies (UPS). While these improvements have been substantial

in advancing the state of the art, they prove to be cumbersome for employing in power

system applications. The first part of this thesis addresses this issue by taking a systems

approach in dealing with power system issues such as imbalances and building vector

regulators and sequence filters that operate on unbalanced sinusoidal systems.

In the past, several schemes for active and reactive power controls for distributed

generation in a microgrid have been developed using a heuristic approach and were

verified through simulations and lab demonstrations. Alternatively, the approach

followed in this thesis is in determining the dynamic behavior based on small-signal

models of an n DR microgrid. It is hoped that this work will provide motivation for

further detailed studies on microgrids of more general and diverse nature.


193

Bibliography

[1] J. Iannucci, “Coin of the Realm – Guest Editorial,” Distributed Energy –

The Journal for Onsite Power Solutions, March-April 2004, pp. 8-9.

[2] H. L. Willis, W. G. Scott, Distributed Power Generation Planning and

Evaluation, Marcel Dekker, Inc., New York, 2000.

[3] H. B. Puttgen, P. R. MacGregor, F. C. Lambert, “Distributed Generation:

Semantic Hype or the Dawn of a New Era?,” IEEE Power and Energy Magazine, Vol. 1,

No. 1, Jan.-Feb. 2003, pp. 22-29.

[4] R. C. Dugan, T. E. McDermott, G. J. Ball, “Planning for Distributed

Generation,” IEEE Industry Applications Magazine, Vol. 7, No. 2, March-April 2001, pp.

80-88.

[5] C. J. Hatziadaniu, A. A. Lobo, F. Pourboghrat, M. Daneshdoost, “A

Simplified Dynamic Model of Grid-Connected Fuel Cell Generators,” IEEE Trans. on

Power Delivery, Vol. 7, No. 2, April 2002, pp. 467-473.

[6] A. L. Weisbrich, S. L. Ostrow, J. P. Padalino, “WARP: A Modular Wind

Power Distributed Electric Utility Application,” IEEE Trans. on Industry Applications,

Vol. 32, No. 4, July-Aug. 1996, pp. 778-787.

[7] J. J. Bzura, “Photovoltaic Research and Demonstration Activities at New

England Electric,” IEEE Trans. on Energy Conversion, Vol. 10, No. 1, March 1995, pp.
194

169-174.

[8] R. Lasseter, “MicroGrids,” Proc. of IEEE PES Winter Meeting 2002, Vol.

1, pp. 305-308, Jan. 2002.

[9] R. Lasseter, P. Piagi, “Providing Premium Power Through Distributed

Resources,” Advanced Technology, HICSS – 33, Jan. 2000.

[10] G. Venkataramanan, M. S. Illindala, C. Houle, R. H. Lasseter, “Hardware

Development of a Laboratory-Scale Microgrid Phase 1 – Single Inverter in Island Mode

Operation,” NREL/SR-560-32527 Report, National Renewable Energy Laboratory, Nov.

2002.

[11] R. N. Friedman, “Distributed Energy Resources Interconnection Systems:

Technology Review and Research Needs,” NREL/SR-560-34259 Report, National

Renewable Energy Laboratory, Sept. 2002.

[12] L. A. Kojovic, R. D. Willoughby, “Integration of Distributed Generation

in a typical USA Distribution System,” Electricity Distribution, 2001. Part I. CIRED

2001, Vol. 4, June 2001.

[13] G. Joos, B. T. Ooi, D. McGillis, F. D. Galiana, R. Marceau, “The Potential

of Distributed Generation to Provide Ancillary Services,” Proc. of IEEE PES Summer

Meeting 2000, Vol. 3, pp. 1762-1767, July 2000.

[14] G. Venkataramanan, M. Illindala, “Microgrids and Sensitive Loads,”

Proc. of IEEE PES Winter Meeting 2002, Vol. 1, pp. 315-322, Jan. 2002.

[15] M. Illindala, G. Venkataramanan, “Battery Energy Storage for Micro-

Source Distributed Generation Systems,” Proc. of IASTED Power and Energy Systems
195

Conf. 2002, Marina del Rey, CA, 2002.

[16] Electric Light and Power, March, 1993.

[17] W. Sweet, “Will Distributed Generation Pan Out as a Panacea?

Networking Assets,” IEEE Spectrum Magazine, Vol. 38, No. 1, Jan 2001, pp. 84-86, 88.

[18] P1547 – IEEE Draft Standard for Interconnecting Distributed Resources

and Electric Power Systems. http://grouper.ieee.org/groups/scc21/1547/1547_index.html

[19] UL Standard for Inverters, Converters, and Controllers for Use in

Independent Power Systems.

http://www.eere.energy.gov/distributedpower/research/ul_1741.html

[20] M. Illindala, G. Venkataramanan, “Control of Distributed Generation

Systems to Mitigate Load and Line Imbalances,” Proc. of IEEE PESC 2002, Vol. 4, pp.

2013-2018, June 2002.

[21] F. Briz, M. W. Degner, R. D. Lorenz, “Analysis and Design of Current

Regulators Using Complex Vectors,” IEEE Trans. on Industry Applications, Vol. 36, No.

3, May/June 2000, pp. 817-825.

[22] Capstone MicroTurbine™ Technical Brochure, Capstone Turbine

Corporation, Chatsworth, CA.

[23] IR PowerWorks™ 70S Series Microturbine Technical Brochure, IR

Energy Systems, Davidson, NC.

[24] S. Song. S. Kang; N. Hahm, “Implementation and Control of Grid

Connected AC-DC-AC Power Converter for Variable Speed Wind Energy Conversion

System,” IEEE APEC Conf. Record 2003, Vol. 1, pp. 154-158, 2003.
196

[25] C. D. Schauder, R. Caddy, “Current Control of Voltage-Source Inverters

for Fast Four-Quadrant Drive Performance,” IEEE Trans. on Industry Applications, Vol.

IA-18, March/April 1982, pp. 163-171.

[26] T. M. Rowan, R. J. Kerkman, “A New Synchronous Current Regulator

and an Analysis of Current Regulated PWM Inverters,” IEEE Trans. on Industry

Applications, Vol. IA-22, July/Aug. 1986, pp. 678-690.

[27] P. M. Dalton, V. J. Gosbell, “A Study of Induction Motor Current Control

Using the Complex Number Representation,” IEEE IAS Annual Meeting 1989 Record,

Vol. 1, pp. 355-361, 1989.

[28] M. C. Chandorkar, D. M. Divan, R. Adapa, “Control of Parallel

Connected Inverters in Standalone AC Supply Systems,” IEEE Trans. on Industry

Applications, Vol. 29, No. 1, Jan./Feb. 1993, pp. 136-143.

[29] M. J. Ryan, R. D. Lorenz, “A High Performance Sine Wave Inverter

Controller with Capacitor Current Feedback and “Back-EMF” Decoupling”, IEEE Power

Electronics Specialists Conference, Vol. 1, pp. 507 – 513, 1995.

[30] M. J. Ryan, W. E. Brumsickle, R. D. Lorenz, “Control Topology Options

for Single-Phase UPS Inverters,” IEEE Trans. on Industry Applications, Vol. 33, No. 2,

March/April 1997, pp. 493-501.

[31] P. Mattavelli, G. Escobar, A. M. Stankovic, “Dissipativity Based Adaptive

and Robust Control of UPS,” IEEE Trans. on Industrial Electronics, Vol. 48, No. 2,

April 2001, pp. 334-343.

[32] D. N. Zmood, D. G. Holmes, “Stationary Frame Current Regulation of


197

PWM Inverters with Zero Steady State Error,” IEEE Trans. on Power Electronics, Vol.

18, No. 3, May 2003, pp. 814-822.

[33] P. C. Loh, M. J. Newman, D. N. Zmood, D. G. Holmes, “A Comparative

Analysis of Multi-Loop Voltage Regulation Strategies for Single and Three-Phase UPS

Systems,” IEEE Trans. on Power Electronics, Vol. 18, No. 5, Sept. 2003, pp. 1176-1185.

[34] A. von Jouanne, B. B. Banerjee, “Assessment of Voltage Unbalance,”

IEEE Trans. On Power Delivery, Vol. 16, No. 4, Oct. 2001, pp. 782-790.

[35] M. H. J. Bollen, “Fast Assessment Methods for Voltage Sags in

Distribution Systems,” IEEE Trans. on Industry Applications, Vol. 32, No. 6, Nov./Dec.

1996, pp. 1414-1423.

[36] S. Bolognani, M. Zigliotto, "A Space-Vector Approach to the Analysis

and Design of Three-Phase Current Controllers," Proc. International Symposium on

Industrial Electronics 2002, Vol. 2, pp. 645-650, 2002.

[37] C. B. Jacobina, M. B. de Rossiter Correa, T. M. Oliveira, A. M. Lima, E.

R. C. da Silva, “Current Control of Unbalanced Electrical Systems,” IEEE Trans. on

Industrial Electronics, Vol. 48, No. 3, June 2001, pp. 517-524.

[38] X. Yuan, W. Merk, H. Stemmler, J. Allmeling, “Stationary-Frame

Generalized Integrators for Current Control of Active Power Filters With Zero Steady-

State Error for Current Harmonics of Concern Under Unbalanced and Distorted

Operating Conditions,” IEEE Trans. on Industry Applications, Vol. 38, No. 2,

March/April 2002, pp. 523-532.

[39] J. Allmeling, “A Control Structure for Fast Harmonics Compensation in


198

Active Filters,” IEEE Trans. on Power Electronics, Vol. 19, No. 2, March 2004, pp. 508-

514.

[40] C. L. DeMarco, G. C. Verghese, “Bringing Phasor Dynamics into Power

System Load Flow,” Proc. of North American Power Symposium, 1993.

[41] V. Venkatasubramanian, H. Schattler, J. Zaborszky, “Fast Time Varying

Phasor Analysis in the Balanced Large Electric Power System,” IEEE Trans. on

Automatic Control, Vol. 40, No. 12, Nov. 1995, pp. 1975-1982.

[42] M. Ilić, J. Zaborszky, Dynamics and Control of Large Electric Power

Systems, John Wiley & Sons, Inc., New York, 2000.

[43] J. Holtz, J. Quan, G. Schmitt, J. Pontt O., J. Rodriguez P., P. Newman, H.

Miranda, “Design of Fast and Robust Current Regulators for High Power Drives Based

on Complex State Variables,” IEEE IAS Annual Meeting 2003 Record, Salt Lake City,

Utah, Oct. 2003.

[44] P. K. Kovacs, Transient Phenomena in Electrical Machines, Elsevier,

Amsterdam, 1984.

[45] D. W. Novotny, T. A. Lipo, Vector Control and Dynamics of AC Drives,

Clarendon Press, Oxford, 1996.

[46] J. Liang, T.C. Green, G. Weiss, Q.-C. Zhong, “Evaluation of Repetitive

Control for Power Quality Improvement of Distributed Generation,” IEEE 33rd Annual

Power Electronics Specialists Conference 2002, Vol. 4, pp. 1803 – 1808, June 2002.

[47] J. Liang, T.C. Green, G. Weiss, Q.-C. Zhong, “Repetitive Control of Power

Conversion System from a Distributed Generator to the Utility Grid,” Proc. of the 2002
199

International Conference on Control Applications, 2002, Vol. 1, pp. 13 – 18, Sept. 2002.

[48] J. Liang, T.C. Green, G. Weiss, Q.-C. Zhong, “Hybrid Control of Multiple

Inverters in an Island-Mode Distribution System,” IEEE 34th Annual Power Electronics

Specialists Conference, 2003, Vol. 1, pp. 61 – 66, June 2003.

[49] G. Weiss, Q.-C. Zhong, T.C. Green, J. Liang, “H∞ Repetitive Control of DC-

AC Converters in Microgrids,” IEEE Trans. on Power Electronics, Vol. 19, No. 1, Jan.

2004, pp. 219 – 230.

[50] H. Akagi, Y. Kanazawa, A. Nabae, "Instantaneous reactive power

compensators comprising switching devices without energy storage components," IEEE

Trans. on Industry Applications, Vol. IA-20, 1984, pp. 625-630.

[51] S. Bhattacharya, D. M. Divan, B. Banerjee, "Synchronous Frame

Harmonic Isolator Using Active Series Filter," Proc. 4th Euro. Conf. on Power

Electronics and Appln. 1991, Florence, Italy, Vol. 3, pp. 030-035, 1991.

[52] P. T. Cheng, S. Bhattacharya, D. M. Divan, "Line Harmonics Reduction in

High-Power Systems Using Square Wave Inverter-Based Dominant Harmonic Active

Filter," IEEE Trans. on Power Electronics, Vol. 14, No. 2, March 1999, pp. 265-272.

[53] F. Briz, A. Diez, M. W. Degner, "Dynamic Operation of Carrier-Signal-

Injection-Based Sensorless Direct Field-Oriented AC Drives," IEEE Trans. on Industry

Applications, Vol. 36, No. 5, Sept.-Oct. 2000, pp. 1360-1368.

[54] F. Briz, M. W. Degner, A. Zamarron, J. Guerrero, "Online Stator Winding

Fault Diagnosis in Inverter-Fed AC Machines Using High Frequency Signal Injection,"

IEEE Trans. on Industry Applications, Vol. 39, No. 4, Jul.-Aug. 2003, pp. 1109-1117.
200

[55] P. L. Jansen, The Integration of State Estimation, Control and Design for

Induction Machines, Ph. D. Thesis, Univerity of Wisconsin-Madison, 1994.

[56] M. W. Degner, Flux, Position and Velocity Estimation in AC Machines

Using Carrier Signal Injection, Ph. D. Thesis, Univerity of Wisconsin-Madison, 1998.

[57] M. W. Degner, R. D. Lorenz, "Using Multiple Saliencies for the

Estimation of Flux, Position and Velocity in AC Machines," IEEE Trans. on Industry

Applications, Vol. 34, No. 5, Sept.-Oct. 1998, pp. 1097-1104.

[58] M. Harke, Private Communication.

[59] J. Holtz, K. Werner, “Multi-Inverter UPS System with Redundant Load

Sharing Control,” IEEE Trans. on Industrial Electronics, Dec. 1990, pp. 506-513.

[60] M. C. Chandorkar, Distributed Uninterruptible Power Supply Systems, Ph.

D. Thesis, Univerity of Wisconsin-Madison, 1995.

[61] A. J. Wood, B. F. Wollenberg, Power Generation Operation & Control,

John Wiley & Sons, Inc., New York, 1984.

[62] N. Cohn, Control of Generation and Power Flow on Interconnected

Power Systems, John Wiley & Sons, Inc., New York, 1984.

[63] M. S. Illindala, Distributed Generation System – Design and Power Flow

Studies, ECE 714 Project Report, University of Wisconsin-Madison, Spring 2001.

[64] F. Saccomanno, Electric Power Systems Analysis and Control, IEEE

Press, Piscataway, NJ, 2003.

[65] A. Tuladhar, H. Jin, T. Unger, K. Mauch, "Control of Parallel Inverters in

Distributed AC Power Systems with Consideration of Line Impedance Effect," IEEE


201

Trans. on Industry Applications, Vol. 36, No. 1, Jan.-Feb. 2000, pp. 131-138.

[66] P. Piagi, Microgrid Control, Ph. D. Thesis, Univerity of Wisconsin-Madison,

2005.

[67] J. Cardell, R. Tabors, “Operation and Control in a Competitive Market:

Distributed Generation in a Restructured Industry,” Energy Policy Special Issue on

Distributed Resources, pp. 1-21, Jan 1998.

[68] R. H. Park, "Two-Reaction Theory of Synchronous Machines Generalized

Method of Analysis – Part 1," AIEE Transactions, July 1929, pp. 716-730.

[69] T. H. Crystal, L. Ehrman, “The Design and Applications of Digital Filters

with Complex Coefficients,” IEEE Trans. on Audio and ElectroAcoustics, Vol. AU-16,

No. 3, September 1968, pp. 315-320.

[70] G. R. Lang, P. O. Bracket, “Complex Analogue Filters,” Proc. Of Euro.

Conf. on Circuit Theory and Design 1981, Hague, pp. 412-419, 1981.

[71] B. C. Kuo, Automatic Control Systems, Third Edition, Prentice Hall

International, Inc., Englewood Cliffs, NJ, 1975.

[72] 56F805 Evaluation Module Hardware User’s Manual, Motorola, Inc.,

http://e-www.motorola.com/files/dsp/doc/user_guide/DSP56F805EVMUM.pdf

[73] M. S. Illindala, P. Piagi, H. Zhang, G. Venkataramanan, R. H. Lasseter,

“Hardware Development of a Laboratory-Scale Microgrid Phase 2: Operation and

Control of a Two Inverter Microgrid,” NREL/SR-560-35059 Report, National Renewable

Energy Laboratory, Oct. 2003.

[74] R. D. Middlebrook, “Input Filter Considerations in Design and


202

Applications of Switching Regulators,” IEEE IAS Conf. Record, 1976.

[75] G. Venkataramanan, D. M. Divan, T. M. Jahns, “Discrete Pulse

Modulation Strategies for High-Frequency Inverter Systems,” IEEE Trans. on Power

Electronics, Vol. 8, No. 3, July 1993, pp. 279-287.

[76] Y. Suh, Analysis and Control of Three Phase AC/DC PWM Converter

Under Unbalanced Operating Conditions, Ph. D. Thesis, Univerity of Wisconsin-

Madison, 2004.

[77] C. G. Hochgraf, Investigation of Multi-level Inverter Concepts Applied to

Regulation of Power System Voltages Including Imbalance, Ph. D. Thesis, Univerity of

Wisconsin-Madison, 1997.

[78] H. Nyquist, “Regeneration Theory,” Bell System Tech. Journal, Vol. 11,

Jan. 1932, pp. 126-147.

[79] D. Cheng, Analysis of Linear Systems, Third Edition, Addison-Wesley

Publishing Company, Inc., 1963.

[80] G. R. Hovhannisyan, “Asymptotic Stability for Second-Order Differential

Equations with Complex Coefficients,” Electronic Journal of Diff. Eqns., Vol.

2004(2004), No. 85, pp. 1-20.

[81] S. Gataric, N. R. Garrigan, “Modeling and Design of Three-Phase

Systems Using Complex Transfer Functions,” IEEE PESC Conf. Record 1999, pp. 691-

697, 1999.

[82] H. W. Bode, Network Analysis and Feedback Amplifier Design, D. Van

Nostrand Reinhold Company, New York, 1945.


203

[83] Matlab User’s Guide, The Mathworks Inc., 1997.

[84] J. Carpentier, “‘To be or not to be Modern’ that is the Question for

Automatic Generation Control (Point of View of a Utility Engineer),” Electric Power

and Energy Systems, April 1985, pp. 81-91.

[85] N. Jaleeli, L. S. VanSlyck, D. N. Ewart, L. H. Fink, A. G. Hoffmann,

“Understanding Automatic Generation Control,” IEEE Trans. on Power Systems, Vol. 7,

No. 3, August 1992, pp. 1106-1122.

[86] S. Barnett, C. Storey, Matrix Methods in Stability Theory, Thomas Nelson

and Sons Ltd, London, 1970.

[87] C. Jeffries, V. Klee, P. Van den Driessche, “When is a Matrix Sign

Stable?,” Can. J. Math, Vol. XXIX, No. 2, 1977, pp. 315-326.

[88] C. A. Desoer, E. S. Kuh, Basic Circuit Theory, International Edition,

McGraw-Hill Book Company, Singapore, 1969.

[89] A. R. Bergen, Power System Analysis, Prentice Hall, Inc., Englewood

Cliffs, NJ, 1986.

[90] National Electrical Code® Wire tables.

[91] H. Zhang, G. Venkataramanan, M. Chandorkar, “Development of Static

Switchgear for Utility Interconnection in a Microgrid,” Proc. of IASTED Power and

Energy Systems Conf. 2003, Palm Springs, CA, 2003.

[92] 1336 Plus V/Hz AC Drive User Manual, Rockwell Automation – Allen

Bradley, Milwaukee, WI, 2000.

[93] PQ Series DC Power Supplies – Operating and Service Manual, Magna-


204

Power Electronics, Inc., Fulton, NJ, 1998.

[94] 9T23B3873 Datasheets, GE Transformer Catalog, GE Industrial Systems,

Fort Wayne, IN.

[95] R. Kieferndorf, Adjustable Voltage DC Link PWM Induction Motor

Drives, Ph. D. Thesis, Univerity of Wisconsin-Madison, 2003.

[96] E. Carter, High Performance Drives Test Bed Development, M. S. Thesis,

Univerity of Wisconsin-Madison, 2000.

[97] R. J. Davy, I. A. Hiskens, “Lyapunov functions for multimachine power

systems with dynamic loads,” IEEE Trans. on Circuits and Systems I: Fundamental

Theory and Applications, Vol. 44, No. 9, Sept. 1997, pp. 796 - 812

[98] C. L. DeMarco, “A new method of constructing Lyapunov functions for

power systems,” Proc. 1988 IEEE Int. Symp. Circuits Syst., pp. 905-908, 1988.

[99] Draft Report for CERTS, Year 2000 Testing of Capstone and Honeywell

MTGs During Load Changes, Southern California Edison, Irvine, Feb 1, 2001.

[100] James Larmine, Andrew Dicks, Fuel Cell Systems Explained, John Wiley

and Sons, 2000.

[101] Technical Marketing Staff, Rechargeable Batteries Applications Handbook,

Gates Energy Products, Inc., EDN Series for Design Engineers, Butterworth-Heinemann,

MA, 1992.

[102] ABSOLYTE IIP Series Specifications, GNB Technologies, Lombard, IL.

[103] NP Series Sealed Rechargeable Lead-Acid Battery Application Manual,

Yuasa, Inc., Reading, PA.


205

[104] PNGV Battery Test Manual, U.S. Department of Energy, Idaho National

Engineering Laboratory, DOE/ID-10597, Jan. 1997.

[105] H. L. N. Wiegman, Battery State Estimation and Control for Power

Buffering Applications, Ph.D. Dissertation, University of Wisconsin-Madison, December,

1999.

[106] F. Callier, C. A. Desoer, Linear System Theory, Springer Verlag, 1991.


206

Appendix A Battery Energy


Storage for Distributed Resources

Distributed resources (DRs) can improve the reliability of power distribution

when operated in the grid-interfaced mode. This is made possible by locating them in

proximity with the critical/sensitive loads. During interruption in the grid power supply,

they need to seamlessly switch to the stand-alone mode and supply the total load demand.

Conventional high-rated generation systems like synchronous and induction generators

contain large rotating masses in which energy is stored in the form of inertia. Under load

and supply transients, the inertia of such systems delivers or absorbs any power mismatch

between generation of prime mover and load demand. In contrast, inverter based DRs are

characterized by a low inertia. As a result, they do not have sufficient energy storage

capacity to meet the demands of loads in the event of transients. Thus, in order to provide

continuous power to sensitive during abrupt changes in load power demand, it is

necessary to install auxiliary energy storage in them. If not, performance of the DRs may

deteriorate in the stand-alone mode of operation.

Transient changes in load power demand may result from faults in transmission

line and/or load switching. For instance, a 75 kW Honeywell micro-turbine takes about

35 s to respond for a 50 % change in power demand under grid connected mode of

operation [99]. On the other hand, some fuel cells require about 10 s for a 15% change in
207

power output [100]. Moreover, a fuel cell also has a recovery period of a few minutes to

establish equilibrium before it can meet another step change in power output. The typical

response that can be expected of a micro-turbine for a step change in load demand is

illustrated in Figure A-1. In the Figure, PL denotes the load power demand; PS is the

response of the micro-turbine and (PL-PS) is the shortage in power that needs to be

supplied through some means. In grid-connected mode of operation, the grid supplies the

shortage in power until the micro-source responds to a step change in power demand.

However, in stand-alone mode of operation this sudden demand can be met only if

additional storage is included in the DR system.

Figure A-1 Typical step response of a micro-turbine. PL - load power demand, PS - response of the micro-
turbine

Furthermore, addition of storage in a DR system can be used for energy

stabilization, ride-through and dispatchability [2]. Energy stabilization and ride-through

are achieved by the use of storage to provide a stable supply of power in the event of
208

failure of the micro-source, or transient changes in load power. And dispatchability is

possible when the DR with storage is controlled such that it supplies a particular

percentage of the load demand as compared to the main supply.

Energy storage can be provided in a DR system through several means such as

batteries, super-conducting magnetic energy storage (SMES), flywheel and ultra-

capacitors. Among these, the technology of batteries is the most developed and is well

established for a variety of applications. The other forms of energy storage are either still

in the prototype stage of development or are not suitable for mass production. Since a DR

system is basically comprised of a micro-source and a three-phase dc/ac inverter, its dc

link is the most appropriate location for installing storage (refer Figure 1-1). Thus,

battery storage in the dc link of the DR may be used to meet the power requirements

during abrupt changes in power demand from the load. The design issues in battery

energy storage for a DR system are presented below.

The block schematic of a general battery based DR system is given in Figure A-2,

in two possible configurations, (a) and (b). Figure A-2(a) represents an unregulated dc

bus configuration and Figure A-2(b) represents a regulated dc bus configuration. The

battery is represented in the figures as an ideal voltage source in series with an internal

resistance. Vbat is the terminal voltage of the battery, and PS, PL and Pbat are the micro-

source power, load power demand and battery power, respectively.

In the unregulated dc bus configuration, battery is connected directly to the dc

bus, and state of the battery during the charging and discharging cycles dictate the dc bus

voltage. Since the power flow in and out of the battery is for a very short duration, it
209

requires a battery with the ability to discharge to final voltage and recharge to 100% state

of charge within a short interval of time.

(a) Unregulated dc bus configuration (b) Regulated dc bus configuration


Figure A-2 Block diagram of a DR system with battery storage

In the regulated dc bus configuration, the dc bus contains a capacitor whose

voltage is regulated and is independent of the battery charging and discharging cycles. A

dc/dc converter is used as the interface between the battery and the dc bus capacitor. The

dc bus voltage can be regulated by controlling the switching scheme of either the rectifier

(at the micro-source end), the inverter or the dc/dc converter.

The dc/dc converter in the regulated dc bus configuration may be used to step-

down (buck) or step-up (boost) the voltage at the terminal of the battery. A battery

consists of several cells stacked in series, where a cell is its basic unit. It is generally

beneficial to have less current drawn from the battery, as this will maintain the

instantaneous power drawn from each cell to a minimum. So, it may be desirable to have

battery on the higher voltage side of the dc/dc converter. But the choice of too large a

battery voltage would require a large number of series connected cells resulting in charge

equalization problems.

DR system is a utility application of batteries. The currently available


210

rechargeable batteries include – lead-acid, nickel cadmium (NiCd), nickel metal hydride

(NiMH) and lithium (Li)-ion. NiCd and NiMH are not preferred in utility applications

since they have very high self discharge rate. Although, Li-ion has the highest energy

density, it is also the most expensive. The cost/kWh of a lead-acid battery is the least

among these and it can meet the power density requirements of DG applications. So, a

lead-acid battery is considered the most suitable for DG applications for further study in

this chapter.

Lead-acid batteries can be designed for deep-cycle or starting applications. Deep-

cycle batteries are used in back-up power applications where power is needed for longer

durations. These batteries are typically applied in electric vehicles and UPS systems.

Starting batteries are commonly used to start internal combustion engines. No power is

drawn from these batteries during the normal operation of the engines. They are capable

of providing large currents for a very short interval of time. Hence, starting batteries of

higher capacity are suitable for DG applications.

Battery sizing is dependent on the number of cells it contains, and the capacity of

each cell. The cell capacity is normally measured in ampere-hours denoted here as Ah. A

method of specifying charge and discharge rates of a cell is the C rate, which is defined

as the current flow rate that is numerically equal to the cell rated capacity [101]. For

example, C rate of a 1 Ah capacity cell is 1 ampere. Both charge and discharge rates are

normally represented as multiples of the C rate. A current that would nominally discharge

the battery in 1 hour is represented as 1C. A current that would nominally discharge the

battery in 10 hours is specified as 0.1C. Similarly, a current that would nominally


211

discharge the battery in a short interval of 6 minutes would be represented as 10C. The

short duration discharges typically at 3C or higher are termed as high-rate discharges.

Lead-acid batteries are usually rated at 8, 10 or 20 hour rates. For example, the

ABSOLYTE IIP series of valve regulated lead-acid (VRLA) batteries are rated on a 8

hour rate, with capacities ranging from 105 to 4800 Ah [102], and NP series, rated on a

20 hour rate with capacities ranging from 1.2 to 65 Ah [103]. For a battery, the MSDG

system is a high-rate discharge application. The maximum discharge in MSDG

applications involving micro-turbines typically occurs in 100 s, and therefore discharge

rate of in these applications is about 36C. Battery parameters need to be derated at high

discharge rates, due to increased losses in it under these conditions. Table A-1 gives the

nominal capacity of lead-acid batteries at different rates of discharge. From the table, we

observe that a battery rated on a 10 hour rate has only 44% deliverable capacity if the

discharge is conducted in 0.2 hours (i.e. 12 min.).

Table A-1 Evaluation of voltage regulator performance under single phasing


Discharge time Discharge current (multiple
Actual Capacity (%)
(hours) of C)
0.2 50 44
1.0 10 72
5.0 2 92
10.0 1 100
20.0 ½ 108

Loss of capacity at high discharge rates may also be graphically represented, as is

more common among battery manufacturers’ data-sheets. Figure A-3 shows the cell

discharge voltage as a function of discharge rate. Most battery manufacturers’ data-sheets


212

provide tables containing dc amperes that the battery can source at different final volts

per cell and for different periods of time. These tables incorporate the derating necessary

for high-rate discharge application for values at lower time intervals. It may be noted that

all battery ratings are for a nominal room temperature of 25oC or 77oF, and additional

derating will be necessary for lower temperatures.

Figure A-3 Cell discharge voltage Vs. time for a lead-acid battery [101]

In addition to loss of energy storage capacity, the terminal voltage of the battery

is also strongly affected by the discharge rate. Figure A-4 illustrates the battery current

and the corresponding terminal voltage for two sizes of cell, ‘X’ and ‘D’, at two different

temperatures. In the figure, battery ‘X’ is rated at 3.2 Ah and ‘D’ at 1.8 Ah.

A Thevenin’s equivalent circuit (Figure A-5) of the battery supplying power to a

load can be drawn from the data represented in Figure A-4. The negative slope of the

curves in Figure A-4 gives the Thevenin resistance (Rth) of the battery. Using the

maximum power transfer theorem, we can deduce that, when the load equivalent

resistance equals Rth, maximum instantaneous power gets transmitted. Since Rth
213

corresponds to the battery’s internal resistance, we observe that at this operating point,

losses in the battery internal resistance are equal to the power transferred to load. Hence,

at the maximum power transfer operating point, the power dissipated in the battery would

be enormous. In order to avoid losses in the battery and subsequent deterioration in its

operation, it is necessary for us to select a battery whose internal resistance is very small

as compared to the equivalent resistance of the load during the transient.

Figure A-4 Peak current and voltage per cell of a lead-acid battery for a high-rate discharge [101]

The variation of output power delivered by the batteries as a function of current

drawn for ‘X’ and ‘D’ is shown in Figure A-6. From the figure, it is clearly evident that

the battery has a limit on the maximum instantaneous power that it can support, which is
214

dependent on factors such as temperature and capacity.

Figure A-5 Thevenin equivalent circuit of the battery

Figure A-6 Instantaneous power curves of a lead-acid battery for a high-rate discharge [101]

In a DG system, the battery is also subjected to high-rate charging. This occurs

when the load is suddenly removed and the micro-source continues to supply power until

it is able to respond to the new power set point. The battery has to be capable of

accepting this difference in power for a short duration. Thus it must have a high charge

acceptance.
215

Similar battery demands occur in their application in hybrid electric vehicles

(HEVs). In this application, batteries are primarily used during starting, accelerating and

braking of the vehicle, and no net power is drawn from the battery in normal running

mode of the vehicle. Therefore, desirable attributes of high-power batteries for HEV

applications are high-peak power handling capability during starting and acceleration,

and high charge acceptance to maximize regenerative braking utilization. The basic

requirements for a power buffering battery in a HEV are defined by the Partnership for a

New Generation of Vehicles (PNGV) work group as shown in Table A-2 [104].

Table A-2 PNGV charge sustaining HEV, battery specification [104]


Requirement
Description
(slow response engine)
Energy Cycle Efficiency 90~95 %
+65 ~ +80 kW (18 s)
Power Processing
-70 ~ -150 kW (10 s)
Voltage Variation +/- 15 % from nominal
Battery System Mass < 65 kg (> 1000 W/kg)
Temperature Range -40 oC ~ 52 oC
Minimum Energy Storage >> 0.54 MJ (150 Wh)

Various research activities continue to been conducted for meeting these

requirements by optimal control of battery state using battery management systems [105].

In the near term, the general solution for most recently developed HEVs has been to

oversize the battery capacity.

In summary, battery sizing for a DR system depends on the magnitude and

duration of the mismatch between the real power demand of the load and the power

delivered by the micro-source. Since power drawn from the battery is for a short duration

in this application, lead-acid batteries of high-rate discharge and charge-acceptance


216

capability are suitable. The main property of batteries for high-discharge applications is

their peak power capability, which in itself is a strong function of its capacity and internal

resistance. The construction of these batteries is optimized for peak power, i.e., high

capacity along with ultra-low internal resistance. The batteries must also have a good

charge acceptance to accommodate the high-rate charging upon removal of load. Most

state of the art batteries are hard pressed to meet these stringent requirements. Hence, the

overcapacity of cells through parallel connection reduces the equivalent internal

resistance, thereby allowing them to be applied appropriately for DG applications. A

design procedure is described in the form of a flowchart for sizing these batteries in

[10,15].
217

Appendix B Stability of Complex


Transfer Functions vis-à-vis MIMO
Systems

Consider a transfer function Gcpx(s), represented as a ratio of polynomials in s

with these polynomials having complex coefficients. Such a construction is referred to

here as a “complex transfer function” and it is expressed in rectangular form with real

and imaginary parts as

Gcpx(s) = GRe(s) + j GIm(s) (B-1)

By transforming the input and output three-phase quantities from abc to αβ

reference frame, the complex space-vector input quantity entering the filter is

x(t) = xα(t) + j xβ(t) (B-2)

and the filter output space-vector quantity is represented as

y(t) = yα(t) + j yβ(t) (B-3)

X(s) is the Fourier transform of the input space-vector given by

X(s) = Xα(s) + j Xβ(s) (B-4)

and Y(s) is the Fourier transform of the output space-vector given by

Y(s) = Yα(s) + j Yβ(s) (B-5)

Using this formulation to capture αβ frame information in scalar complex


218

quantities, the goal here is to verify that a complex transfer function may be used to

describe input-output behavior in the Laplace domain as

Y(s) = Gcpx(s) X(s) (B-6)

To verify the validity of this formulation, and to properly interpret the role of

roots of the complex denominator polynomial in determining stability, this transfer

function is expanded into real and imaginary parts. In particular, the input-output relation

can be expressed as a multi-input multi-output (MIMO) system as

⎡ Yα(s) ⎤ ⎡ GRe(s) –GIm(s) ⎤ ⎡ Xα(s) ⎤


⎢ ⎥=⎢ ⎥⎢ ⎥ (B-7)
⎣ Yβ(s) ⎦ ⎣ GIm(s) GRe(s) ⎦ ⎣ Xβ(s) ⎦

The realization of (B-7) is referred to in this thesis as forward coupling form of

complex transfer function and is illustrated in Figure B-1. As seen in Figure B-1, the

forward coupling form is a feedforward form of realization of complex filter. This

realization was proposed in [70] by Lang and Brackett. Thus, the two-input two-output

MIMO system of the form shown in Figure B-1 is transformed to a single-input single-

output (SISO) system, albeit one with complex-coefficient transfer functions.

Let the complex-coefficient transfer function be expressed as

N(s)
Gcpx(s) = D(s) (B-8)

Then, real-coefficient transfer functions of the MIMO realization, GRe(s) and

GIm(s), can be determined as


*
Gcpx(s) + Gcpx(-s) NRe(s)
GRe(s) = 2 = D(s) D*(-s) (B-9)
219
*
Gcpx(s) - Gcpx(-s) NIm(s)
and GIm(s) = 2j = D(s) D*(-s) (B-10)

where real-coefficient polynomials NRe(s) and NIm(s) are respectively given by

N(s) D*(-s) + N*(-s) D(s)


NRe(s) = 2 (B-11)

N(s) D*(-s) - N*(-s) D(s)


and NIm(s) = 2j (B-12)

In the above equations, polynomials N*(-s) and D*(-s) denote the complex

conjugates of polynomials N(s) and D(s), respectively.

xα(t) yα(t)
GRe(s)

GIm(s)
x(t) = xα(t) + jxβ(t) y(t) = yα(t) + jyβ(t)

GIm(s)

GRe(s)
xβ(t) yβ(t)

Figure B-1 Realization of complex transfer function Gcpx(s) as a two-input two-output MIMO system

Substituting the above expressions for GRe(s) and GIm(s) in (B-7), the transfer

function matrix of the MIMO system is determined in matrix fraction description (MFD)

form [106] as

⎡ GRe(s) –GIm(s) ⎤ -1
H(s) = ⎢ ⎥ = [Nr(s)]2x2 [Dr(s)]2x2 (B-13)
⎣ GIm(s) GRe(s) ⎦
220

⎡ NRe(s) –NIm(s) ⎤
where [Nr(s)]2x2 = ⎢ ⎥ (B-14)
⎣ NIm(s) NRe(s) ⎦
*
⎡ D(s) D (-s) 0 ⎤
and [Dr(s)]2x2 = ⎢ * ⎥ (B-15)
⎣ 0 D(s) D (-s) ⎦

The poles of above MIMO system are given by the roots of

det([Dr(s)]2x2) = 0 (B-16)

which is observed to be the same as roots of

D(s) D*(-s) D(s) D*(-s) = 0 (B-17)

On the other hand, the poles of the complex transfer function Gcpx(s) describing

SISO system are the roots of

D(s) = 0 (B-18)

As compared to the poles of complex transfer function Gcpx(s) those of MIMO

system also include the conjugate pairs. Moreover, the MIMO system also displays

duplication of these complex conjugate pairs. Nevertheless, the two systems are observed

to have poles located in the same half plane of the complex s-plane. As a result, the

bounded input bounded output (BIBO) stability of MIMO system can be deduced from

the compact representation of the complex transfer function Gcpx(s) itself. The MIMO

system is BIBO stable if all the poles of Gcpx(s) lie in the left half of complex s-plane.

As an example, consider the first-order complex transfer function given by

N(s) 1
Gcpx(s) = D(s) = (B-19)
⎛ - j ω60⎞
s
1+⎜ ⎟
⎝ ωo ⎠

The numerator and denominator polynomials of Gcpx(s) are


221

N(s) = ωo (B-20)

and D(s) = s + ωo - j ω60 (B-21)

As a result, the pole of Gcpx(s) is given by the root of D(s) = 0, which is at s = - ωo

+ j ω60.

On the other hand, the transfer function matrix of the MIMO system is obtained

in MFD form as
-1
H(s) = [Nr(s)]2x2 [Dr(s)]2x2 (B-22)

2 2 -1
⎡ ωo(s+ωo) ωoω60 ⎤ ⎡ (s+ωo) + ω60 0 ⎤
=⎢ ⎥⎢ ⎥ (B-23)
⎣ -ωoω60 ωo(s+ωo) ⎦ ⎣ 0 (s+ωo)2 + ω602 ⎦

A state-space [A, B, C, D] realization of the same MIMO system represented by

the above transfer function matrix is observed to have

⎡ A1 0 ⎤ ⎡ 0 1 ⎤
A=⎢ ⎥ where A1 = ⎢ ⎥ , (B-24)
⎣ 0 A1 ⎦ ⎣ -(ωo2 + ω602) -2ωo ⎦

0 0
⎡ω 0

B=⎢
0 ⎥
o
, (B-25)
0
⎣0 ω ⎦
o

⎡ 0 1 ω60 0 ⎤
C=⎢ ⎥ and (B-26)
⎣ -ω60 0 0 1 ⎦

⎡0 0⎤
D=⎢ ⎥ (B-27)
⎣0 0⎦

Matrix A contains two identical principal submatrices (labeled A1) that are

decoupled, and the eigenvalues of this submatrix A1 are determined to be -ωo ± j ω60.
222

Therefore, eigenvalues of A are -ωo ± j ω60 but repeated twice. The same is observed in

the case of poles of transfer function matrix H(s) that are the roots of
2 2
⎛ ⎡ (s+ωo) + ω60 0 ⎤⎞
det⎜ ⎢ 2 2
⎥⎟=0 (B-28)
⎝⎣ 0 (s+ωo) + ω60 ⎦ ⎠

Thus, the poles of MIMO system include the conjugate pairs of the poles of

Gcpx(s), and these get duplicated too. As the two systems are observed to have poles

located in the same half plane of the complex s-plane, information on the BIBO stability

of MIMO system can be determined from that of the compact representation of complex

transfer function Gcpx(s) itself.

View publication stats

Вам также может понравиться