You are on page 1of 374

Frequency regulation in wind integrated power system

AUTHOR(S)

A Aziz

PUBLICATION DATE

01-01-2017

HANDLE

10536/DRO/DU:30103451

Downloaded from Deakin University’s Figshare repository

Deakin University CRICOS Provider Code: 00113B


FREQUENCY REGULATION IN WIND
INTEGRATED POWER SYSTEM

ASMA AZIZ

SUBMITTED IN FULFILMENT OF THE REQUIREMENTS FOR THE


DEGREE OF

DOCTOR OF PHILOSOPHY

SCHOOL OF ENGINEERING
FACULTY OF SCIENCE, ENGINEERING AND BUILT ENVIRONMENT
DEAKIN UNIVERSITY

JAN, 2017
Signature Redacted by Library
Signature Redacted by Library
 

© Asma Aziz

2017

1
Dedicated to My Beloved Parents

2
ACKNOWLEDGMENTS

I have great pleasure in expressing my deep sense of appreciation and thankfulness

to my principal supervisor Prof. Aman MTO and Co-supervisor Prof. Alex

Stojcevski, for their invaluable guidance, continual encouragement, and support at

every stage of this work. No amount of words can suffice in expressing my sincere

thanks to them. My special gratitude to Prof. Aman MTO for his assistance and

encouragement during this research and providing me a valuable learning experience.

This work has been carried out at the Department of Electrical and Renewable Energy

Engineering, School of Engineering at Deakin University, Australia. The financial

support by Deakin University is gratefully acknowledged.

I would also like to thanks, Kate Hecimovic for providing valuable administrative

support during this candidature and G. M. Shafiullah for providing valuable feedback

during candidature proposal. 

Finally, I acknowledge my deep gratitude to my loving mother for her endless prayers

and my late father who have always been a source of inspiration to me. Special

gratitude to my husband as this accomplishment was not possible without his

continuous help, fruitful technical discussions, and personal advice. I am very

thankful to my both dear brothers, my sisters-in-law, parents-in-law and all family

members for their continuous support, love, and encouragement for my academic

endeavor.

Last but not least, special thanks to my seven years old son Zayan who always

comfort and console, never complain or interfere and endured all my inattention

during last three years.

3
PUBLICATIONS

Journals

1. Asma Aziz, Amanullah MTO, Alex Stojcevski. “Full converter based wind

turbine generator system generic modeling: Variations and applicability,”

Elsevier’s Sustainable Energy Technologies and Assessments, Volume 14,

April 2016, Pages 46-62, ISSN 2213-1388. [IF=1.089, Q1, CiteScore=3.06].

2. Aziz, Asma, Shafiullah, G. M., Amanullah MTO, and Alex Stojcevski.

“Automatic Generation Control in Wind Integrated Power System: New

Perspectives & Challenges.” Int. J. of Thermal & Environmental Engineering,

Volume 12, No. 1 (2016) 27-38. [Journal Google Scholar h-index=12, i10-

index=17]

3. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. "Automatic Generation

Control of Multi-Generation Power System." Journal of Power and Energy

Engineering 2.04 (2014): 312. [Journal Google Scholar h-index=8, Journal

Google impact factor 0.72]

4. Aziz, Asma, Shafiullah, G. M., Amanullah MTO, and Alex Stojcevski. “Load

Frequency Control in Wind Integrated Multigenerational Power System.”

Elsevier’s Sustainable Energy Technologies and Assessments (Received

Revision and under process for further submission).

5. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “LVRT Capability

Assessment of Grid Code Frequency Responsive VSWTG,” International

Journal of Electrical Energy, ISSN: 2301-3656, DOI: 10.18178/ijoee.

(Accepted)

6. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “Grid Code Frequency

Sensitive Active Power Control of Type 3 and Type 4 Variable Speed Wind
4
Turbine Generator for Improved Frequency Support”. Elsevier’s Sustainable

Energy Technologies and Assessments (Under review)

7. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “AGC in the presence of

Grid Code Frequency Responsive Wind Power Penetration using Intelligent

FGSPID.” Taylor Francis’s Electric Power Components and Systems. (Under

review)

8. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “Wind Integration In

Australian NEM: Issues and Mitigations” Elsevier’s Renewable Energy

Focus (Under review).

9. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “An Analysis of

Frequency Responsive Wind Plants Integration on AGC of Multigenerational

System” (To be Submitted)

Conference

1. Aziz, Asma, Shafiullah, G. M., Amanullah MTO, and Alex Stojcevski.

“Automatic Generation Control of Wind Energy System,” Proceedings of the

9th Annual International Green Energy Conference (IGEC-IX). 25th–28th

May 2014, Tianjin, China.

2. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. "Multi-Generational

Power System Load Frequency Secondary Control,” 6th Asia-Pacific Power

and Energy Engineering Conference (APPEEC 2014), April 14 to 16, 2014,

Shanghai, China.

3. Aziz, Asma, Shafiullah, G.M., Amanullah MTO, and Alex Stojcevski.

“Participation of DFIG Based Wind Energy System in Load Frequency

Control of Interconnected Multigenerational Power System,” IEEE

5
Australasian Universities Power Engineering Conference, (AUPEC) 2014,

Perth, Australia.

4. Aziz, Asma Arangarajan Vinayagam, KSV Swarna, Suiyang Khoo and Alex

Stojcevski. “Power Quality Impacts in a Typical Microgrid,” International

Conference on Sustainable Energy and Environmental Engineering (SEEE

2015), Oct. 2015. DOI: 10.2991/seee-15.2015.20

5. Aziz, Asma, Arangarajan Vinayagam, Amanullah MTO, and Alex Stojcevski.

“Modelling and Comparison of Generic Type 4 WTG with EMT Type 4 WTG

Model”, 12th IEEE India International Conference, 2015 (INDICON 2015),

Delhi, India.

6. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “Modelling and Analysis

of Type 4 Wind Turbine Generator System for Utilization in Frequency

Regulation Studies”, IEEE Australasian Universities Power Engineering

Conference, (AUPEC) 2015, Wollongong, Australia.

7. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “Fuzzy Gain Scheduling

of Load Frequency Controller in the Presence of Frequency Responsive Wind

Power Penetration,” IEEE Australasian Universities Power Engineering

Conference, (AUPEC) 2016, Brisbane, Australia.

8. Aziz, Asma, Amanullah MTO, and Alex Stojcevski. “LVRT Capability

Assessment of Grid Code Frequency Responsive TYPE 3 VSWTG”, 9th

International Conference on Computer and Electrical Engineering (ICCEE

2016), Barcelona, Spain.

Awards

1. Best Paper Presenter Award-IEEE INDICON 2015

2. Travel Award, University of Queensland-AUPEC2016

6
TABLE OF CONTENTS

ACKNOWLEDGMENTS .................................................................................................... 3

PUBLICATIONS ................................................................................................................. 4

TABLE OF CONTENTS ..................................................................................................... 7

LIST OF TABLES ............................................................................................................ 12

LIST OF FIGURES ........................................................................................................... 14

ACRONYMS ...................................................................................................................... 22

NOMENCLATURE .......................................................................................................... 24

EXECUTIVE SUMMARY ................................................................................................ 30

1 CHAPTER 1 INTRODUCTION ............................................................................ 32 

1.1  Current Status of Wind Energy System .............................................................................. 32 

1.2.  Motivations for Research ............................................................................................. 34 

1.3.  Objectives .................................................................................................................... 39 

1.4.  Significance .................................................................................................................. 42 

1.5.  Methodology ............................................................................................................... 43 

1.6.  Contributions ............................................................................................................... 45 
1.6.1.    Frequency Response Investigation in NEM Control Area ................................................. 46 
1.6.2.   Modeling & Performance Assessment of Type 4 Frequency Responsive Wind Turbine ... 47 
1.6.3.    Modeling and Investigation of Grid Code Compatible Active‐Power Controller for Type 3 
& Type 4 VSWTGs .......................................................................................................................... 47 
1.6.4.    Development of Frequency Response Model for Wind Energy Impact Assessment on 
Reserve Response .......................................................................................................................... 47 
1.6.5.    Integration Impact of Grid Code Frequency Responsive VSWTGs on AGC of 
Multigenerational System .............................................................................................................. 48 
1.6.6.    Development of FGSPID Controller for Wind Integrated Power System Load Frequency 
Control ........................................................................................................................................... 48 
1.6.7.    Fault Ride‐through Analysis of Frequency Responsive VSWTGs ....................................... 48 
1.6.8.    Review of Active Power‐Frequency Control in Wind Integrated Power System............... 48 

1.7.  Outline of the Thesis .................................................................................................... 49 

1.8.  References ................................................................................................................... 51 

7
2  CHAPTER 2
WIND INTEGRATION IN AUSTRALIAN NEM ........................................................ 53 

2.1.  National Electricity Market (NEM) Overview ................................................................ 53 

2.2.  NEM Spot Market & AEMO .......................................................................................... 54 

2.3.  DRIVERS OF WIND ENERGY SYSTEM DEVELOPMENT IN AUSTRALIA .............................. 58 
2.3.1.   Renewable Energy Target (RET) ......................................................................................... 58 
2.3.2.   Wind Resources Availability ............................................................................................... 59 
2.3.3.   Cost Effective, Competitive & Established Technology ...................................................... 60 
2.3.4.   Low Carbon Footprint ........................................................................................................ 62 

2.4.  Wind Energy and Australian Market Outlook ................................................................ 63 

2.5.  Challenges to NEM Operation Due to Increased Wind Penetration ............................... 67 
2.5.1.   Merit Order Effect and Reduced Spot Prices ..................................................................... 67 
2.5.2.   Reduced Inertia and High Rate of Change of Frequency Issues ......................................... 70 
2.5.3.   FCAS Issues ......................................................................................................................... 72 
2.5.4.   Network Infrastructure Issues............................................................................................ 74 
2.5.5.   Interconnectors Performance Issues ................................................................................. 75 

2.6.  Wind Integrated Network Frequency Response Assessment ......................................... 77 
2.6.1.   Test Network ...................................................................................................................... 77 
2.6.2. Low Load Scenario ............................................................................................................... 78 
2.6.3.   Assumptions Used in simulations ...................................................................................... 78 
2.6.4.   Simulation Discussions ....................................................................................................... 84 
2.6.5.   Linear Regression Model from Frequency Disturbances ................................................... 89 
2.6.6.   High Load Scenario ............................................................................................................. 93 

2.7.  Possible Frequency Control Solutions ........................................................................... 97 

2.8.  Conclusion ................................................................................................................. 105 

2.9.  References ................................................................................................................. 108 

CHAPTER 3 POWER SYSTEM FREQUENCY CONTROL .................................... 113 

3.1  Introduction ................................................................................................................... 113 

3.2.   Power System Inertia ...................................................................................................... 117 
3.2.1.   Power system Inertial Response ...................................................................................... 119 

3.3.   Primary Frequency Control .............................................................................................. 120 

3.4.   Secondary Frequency Control .......................................................................................... 124 
3.4.1.   Area Control Error ............................................................................................................ 125 

3.5.   Tertiary control ............................................................................................................... 129 

3.6.  Wind Energy System Participation in Frequency‐Power Control .................................. 129 

8
3.6.1.   Synthetic Inertial Control ................................................................................................. 130 
3.6.2.   Primary Control from VSWTG .......................................................................................... 132 
3.6.3.   AGC & Wind Energy ......................................................................................................... 138 
3.6.4.   AGC & Wind Energy Research: New Perspectives and Challenges .................................. 144 

3.7.   Conclusion ...................................................................................................................... 152 

3.8. References ........................................................................................................................ 154 

CHAPTER 4 FREQUENCY RESPONSIVE FULL CONVERTER BASED VSWTG:


MODELLING AND ASSESSMENT ............................................................................. 164 

4.1.  Introduction ............................................................................................................... 164 

4.2.  Type 4 VSWT Model with Synchronous Generator ...................................................... 166 
4.2.1.   Converter Topology ......................................................................................................... 167 
4.2.2.   Synchronous Generator Model ........................................................................................ 169 

4.3.  WECC’s Type 4 WTG Generic Model ........................................................................... 171 

4.4.  Proprietary Type 4 Generic Model .............................................................................. 172 

4.5.  Proposed Type 4 VSWTG Generic Model .................................................................... 174 
4.5.1.   Wind Turbine Mechanical Model ..................................................................................... 176 
4.5.2.   Pitch Control Model ......................................................................................................... 179 
4.5.3.   Electrical Control Model................................................................................................... 181 
4.5.3.   Active Power Control Loop Co‐ordination with Plant Controller ..................................... 183 

4.6.  Model Validation and Analysis ................................................................................... 184 
4.6.1.   Aerodynamic Model Suitability Assessment .................................................................... 185 
4.6.2.   Generator‐Converter Model Comparison with GE’s Generic Model and Detailed Model
 ..................................................................................................................................................... 195 
4.6.3.   Drive Train Mass Model Comparison with GE’s Generic Model & Detailed Simulation 
Model ........................................................................................................................................... 197 
4.6.4.   Virtual Inertia Support Assessment ................................................................................. 198 

4.7.    Conclusion ..................................................................................................................... 203 

4.7.  References ................................................................................................................. 204 

CHAPTER 5 DYNAMIC DEAD‐BAND DEPENDENT GRID CODE FREQUENCY


SENSITIVE ACTIVE POWER CONTROLLER ......................................................... 207 

5.1.  Introduction ............................................................................................................... 207 

5.2.  Grid Code Sensitive Frequency Controller ................................................................... 209 
5.2.1.   Dead‐Band: Static vs. Dynamic ........................................................................................ 209 
5.2.3.   Frequency Processing Model ........................................................................................... 210 

5.3.  Frequency Response Controller Model‐I ..................................................................... 215 

9
5.4.  Frequency Response Controller Model‐II .................................................................... 220 

5.5.  Conclusion ................................................................................................................. 223 

5.6.  References ................................................................................................................. 224 

CHAPTER 6 GRID CODE FREQUENCY SENSITIVE VSWTG: IMPACT


ASSESSMENT ................................................................................................................ 225 

6.1.  Introduction ............................................................................................................... 225 
6.1.1.   Type 3 VSWTG Model ...................................................................................................... 226 
6.1.2.    Applied Commanded Reserve Power Mode ................................................................... 230 
6.1.3.    Wind Profile and Available Power Model ....................................................................... 232 

6.2.   Simulation Setup ............................................................................................................. 233 
6.2.1.   Simulation Assumptions ................................................................................................... 235 

6.3. Discussion ......................................................................................................................... 237 
6.3.1.   Case I: Type 3 VSWTG ...................................................................................................... 237 
6.3.2.   Case II: Type 4 VSWTG ..................................................................................................... 246 

6.4.    Conclusion ..................................................................................................................... 253 

6.4.   References ...................................................................................................................... 255 

CHAPTER 7 AUTOMATIC GENERATION CONTROL IN FREQUENCY


RESPONSIVE WIND PLANT INTEGRATED POWER SYSTEM ......................... 257 

7.1.  AGC / Load Frequency Controllers and Conventional Generation Models .................... 257 
7.1.1.   Classical AGC Controller ................................................................................................... 258 
7.1.2.   Intelligent AGC Controller ................................................................................................ 260 
7.1.3.    Hydro Turbine Governor Model ...................................................................................... 263 
7.1.4.   Gas Power Plant Model .................................................................................................... 268 
7.1.5.   Thermal Power Plant ........................................................................................................ 278 

7.2.  Performance Index ..................................................................................................... 279 

7.3.  Parameter Optimization Algorithm ............................................................................ 281 

7.4.  Control Area Simulation Models ................................................................................. 283 
7.4.1.   A Frequency Regulation Model Incorporating All Three Control Modes in a Single Area 
Power System .............................................................................................................................. 283 
7.4.2.   Decentralized Three Area Interconnected Power System (TAIPS) Model ....................... 287 
7.4.3.   Test System for LVRT Capability Assessment ................................................................... 291 

7.5.   Analysis‐I ........................................................................................................................ 292 

7.6.   Analysis‐II ....................................................................................................................... 306 

7.7.   Analysis‐III ...................................................................................................................... 327 

10
7.8.   Analysis‐IV ...................................................................................................................... 338 

7.9.  References ................................................................................................................. 352 

CHAPTER 8 CONCLUSION AND RECOMMENDATIONS .................................... 358 

8.1.   Conclusion ...................................................................................................................... 358 
8.1.1.   Type 4 Generic VSWTG Model for Frequency Response Studies ..................................... 359 
8.1.2.   Grid Code Sensitive, Frequency Processor Dependent Frequency Controller ................. 360 
8.1.3.   Wind Participation in Classical and Intelligent AGC in Multi‐Generational System ......... 361 
8.1.4.   Challenges and Possible Solution in Wind Integrated Network ....................................... 363 

8.2.   Recommendations .......................................................................................................... 365 

APPENDIX ..................................................................................................................... 366 

11
LIST OF TABLES
TABLE 2‐1 COMPARISON OF RENEWABLE ENERGY TARGET FOR SOME COUNTRIES WITH AUSTRALIA
 .................................................................................................................................................. 59 
TABLE 2‐2 SCENARIO 1: LOW LOAD .................................................................................................... 78 
TABLE 2‐3 LOAD FLOW IN LOW LOAD SCENARIO ............................................................................... 80 
TABLE 2‐4 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 503 (150 MW) IS 
ISOLATED AT 10S AND WIND FARM CONNECTED AT BUS 509 ................................................. 80 
TABLE 2‐5 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 502 (540 MW) IS 
ISOLATED AT 10S AND WIND FARM CONNECTED AT BUS 508 ................................................. 81 
TABLE 2‐6 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 501 (560 MW) IS 
ISOLATED AT 10S AND WIND FARM CONNECTED AT BUS 508 ................................................. 82 
TABLE 2‐7 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 503 (150 MW) IS 
ISOLATED AT THE 20S, AND GENERATOR 501 (560 MW) IS ISOLATED AT 10S, 1146MW WIND 
FARM CONNECTED AT BUS 508 ................................................................................................ 82 
TABLE 2‐8 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 503 (150 MW) IS 
ISOLATED AT 10S AND AREA 3‐5 INTERCONNECTOR IS ISOLATED AT 20 S; WIND FARM 
CONNECTED AT BUS 508 ........................................................................................................... 83 
TABLE 2‐9 AREA FREQUENCY RESPONSE INDICATORS WHEN GENERATOR 502 (540 MW) IS 
ISOLATED AT 10S AND AREA 3‐5 INTERCONNECTOR IS ISOLATED AT 20 S, 1080 MW WIND 
FARM CONNECTED AT BUS 508 ................................................................................................ 84 
TABLE 2‐10 SIMULATION RESULTS FOR LOW LOAD SCENARIO FOR AREA 5 GENERATOR 501 .......... 87 
TABLE 2‐11 SIMULATION RESULTS FOR LOW LOAD SCENARIO FOR AREA 5 GENERATOR 502 .......... 88 
TABLE 2‐12 VALUES FOR LINEAR POLYNOMIAL REGRESSION MODEL ............................................... 90 
TABLE 2‐13 SCENARIO 1: HEAVY LOAD ............................................................................................... 94 
TABLE 2‐14 MODIFIED BUS FOR HIGH LOAD SCENARIO ..................................................................... 94 
TABLE 2‐15 GENERATION OUTPUT IN HIGH LOAD SCENARIO ............................................................ 95 
TABLE 2‐16 AREA FREQUENCY RESPONSE INDICATORS WHEN WIND FARM IS ADDED (109X4 MW)96 
TABLE 2‐17 AREA FREQUENCY RESPONSE INDICATORS WHEN GEN 503 IS ISOLATED AT 10 S AND 
WIND FARM IS ADDED (436 MW) ............................................................................................. 96 
TABLE 2‐18 AREA 5 FREQUENCY RESPONSE INDICATORS WHEN GEN 503 IS ISOLATED AT 5 S, AND 
INTERCONNECTOR IS ISOLATED AT 10 S, WIND FARM ADDED (436 MW) ............................... 97 
TABLE 2‐19 ANTICIPATED TECHNICAL SOLUTIONS FOR FREQUENCY IMPROVEMENT IN WIND 
PENETRATED SYSTEM ............................................................................................................... 98 
TABLE 2‐20 AREA 5 FREQUENCY RESPONSE INDICATORS DURING GENERATOR CONTINGENCY 
(WITHOUT SYNCHRONOUS CONDENSER) ............................................................................... 101 
TABLE 2‐21 AREA 5 FREQUENCY RESPONSE INDICATORS DURING GENERATOR CONTINGENCY (WITH 
ONE SYNCHRONOUS CONDENSER) ......................................................................................... 101 
TABLE 2‐22 AREA 5 FREQUENCY RESPONSE INDICATORS DURING GENERATOR CONTINGENCY (WITH 
BOTH SYNCHRONOUS CONDENSERS) ..................................................................................... 101 
TABLE 3‐1 BASIC FREQUENCY REGULATION SERVICES ..................................................................... 115 
TABLE 3‐2 REGULATION SERVICES NOMENCLATURE ADOPTED BY VARIOUS COUNTRIES .............. 116 
TABLE 4‐1 COMPARISON OF TYPE 4 WTG MODELS SUITABLE FOR FREQUENCY REGULATION 
ANALYSIS ................................................................................................................................. 175 
TABLE 6‐1 COMPARISON OF SIMULATION RESULTS TO SHOW THE EFFECT OF TORQUE CONTROLLER 
RE‐TUNING AND DIFFERENT DROOP VALUES FOR 10% WIND ENERGY PENETRATION ......... 236 
TABLE 6‐2 COMPARISON OF FREQUENCY RESPONSE SIMULATION RESULTS FOR TYPE 4 MODELS 
UNDER 10% WIND ENERGY PENETRATION IN A CONTROLLED AREA ..................................... 252 
TABLE 7‐1 SIMULATION STUDY SET UP CASES ................................................................................. 293 
TABLE 7‐2 OPTIMUM VALUES OF AGC INTEGRAL CONTROLLER ...................................................... 309 

12
TABLE 7‐3 SYSTEM PARAMETERS AND SYSTEM FREQUENCY RESPONSE INDICATORS OBTAINED 
WITHOUT WIND INTEGRATION .............................................................................................. 309 
TABLE 7‐4 WIND PLANT INTEGRATED SYSTEM DATA ....................................................................... 312 
TABLE 7‐5 FACTORS AFFECTING FUZZY PID TUNING ........................................................................ 330 
TABLE 7‐6 FUZZY TUNING RULES FOR KP’ ........................................................................................ 331 
TABLE 7‐7 FUZZY TUNING RULES FOR KD’ ........................................................................................ 331 
TABLE 7‐8 FUZZY TUNING RULES FOR ∝ ........................................................................................... 331 
TABLE 7‐9 SECONDARY FREQUENCY CONTROLLER COMPARISON WITH PERFORMANCE INDEX FOR 
AREA 1 UNDER 10% WIND PENETRATION AND 0.1 PER UNIT LOAD DISTURBANCE IN ALL 
THREE AREAS........................................................................................................................... 336 
TABLE 7‐10 TEST RESPONSE ............................................................................................................. 344 

13
LIST OF FIGURES
FIGURE 1‐1 CUMULATIVE FORECAST BY REGION 2012‐2017 [2] ....................................................... 33 
FIGURE 1‐2 STAKEHOLDERS VIEW’S ON FREQUENCY RESPONSIVE CAPABILITY AS A REQUIREMENT 
FOR WIND ENERGY SYSTEM ...................................................................................................... 39 
FIGURE 1‐3 RESEARCH METHODOLOGY ............................................................................................. 45 
FIGURE 1‐4 CONTRIBUTIONS OF THE STUDY ...................................................................................... 46 
FIGURE 2‐1 NEM REGIONAL BOUNDARIES, THEIR INTERCONNECTIONS AND GENERATION CAPACITY 
[DATA DERIVED FROM [1], FORMULATED BY AUTHOR] ........................................................... 54 
FIGURE 2‐2 AEMO INTERACTIONS WITH OTHER REGULATORS [FORMULATED BY AUTHOR] ........... 57 
FIGURE 2‐3 AUSTRALIA WIND RESOURCES [16] ................................................................................. 59 
FIGURE 2‐4 LEVELISED COST OF ENERGY (LCOE) GENERATION AND CAPACITY FACTOR FOR WIND 
GENERATORS [17] ..................................................................................................................... 61 
FIGURE 2‐5 LCOE COST COMPARISON WITH FUEL TECHNOLOGIES [22] ........................................... 62 
FIGURE 2‐6 PER CAPITA WIND CAPACITY BY COUNTRY AND THE KEY US AND AUSTRALIAN 
STATES/TERRITORIES [23] ......................................................................................................... 63 
FIGURE 2‐7 WIND GENERATION CAPACITY SUMMARY FOR AUSTRALIA 2015 ................................... 64 
FIGURE 2‐8 PERCENTAGE OF WIND CAPACITY BY STATE [MODIFIED FROM 26] ............................... 65 
FIGURE 2‐9 SOUTH AUSTRALIA WIND POWER VS. DEMAND VARIATION ON 22 MAY [28] ............... 66 
FIGURE 2‐10 MERIT ORDER EFFECT [33] ............................................................................................ 69 
FIGURE 2‐11 MARKET PRICE VARIATIONS IN SOUTH AUSTRALIA ON 3‐4 NOV. 2014 WHEN WIND 
GENERATION ENTERS SPOT MARKET [33] ................................................................................ 69 
FIGURE 2‐12 14‐GENERATOR NEM MODEL ........................................................................................ 79 
FIGURE 2‐13 AREA 5 GENERATOR TERMINAL FREQUENCY AFTER ISOLATION OF GEN 503 AT 10 S . 85 
FIGURE 2‐14 AREA 5 GENERATOR TERMINAL FREQUENCY AFTER ISOLATION OF GEN 503 AT 10 S 
AND INTERCONNECTOR ISOLATION AT 20 S ............................................................................. 86 
FIGURE 2‐15 AREA 5 BUS FREQUENCY AFTER ISOLATION OF GEN 503 AT 10 S. ................................ 86 
FIGURE 2‐16 AREA 5 BUS FREQUENCY AFTER ISOLATION OF GEN 502 AT 10 S. ................................ 86 
FIGURE 2‐17 ROCOF OBSERVED AT GENERATOR 503 WHEN GENERATOR 502 IS ISOLATED AT 10 S 87 
FIGURE 2‐18  POLYNOMIAL SURFACE VIEW FOR WIND PENETRATION VARIATION WITH FREQUENCY 
NADIR POINT ............................................................................................................................. 91 
FIGURE 2‐19 POLYNOMIAL SURFACE VIEW FOR MW‐LOSS VARIATION WITH FREQUENCY NADIR 
POINT AND ROCOF AS OBSERVED AT GENERATOR 501 ........................................................... 91 
FIGURE 2‐20 LINEAR POLYNOMIAL REGRESSION SURFACE VIEW FOR WIND PENETRATION 
VARIATION WITH FREQUENCY NADIR POINT AND ROCOF AS OBSERVED AT GENERATOR 502
 .................................................................................................................................................. 92 
FIGURE 2‐21 LINEAR INTERPOLANT SURFACE VIEW FOR MW‐LOSS VARIATION WITH FREQUENCY 
NADIR POINT AND ROCOF AS OBSERVED AT GENERATOR 502 ................................................ 92 
FIGURE 2‐22 POLYNOMIAL REGRESSION SURFACE VIEW FOR ONLINE SYNCHRONOUS GENERATOR 
UNITS VARIATION WITH FREQUENCY NADIR POINT AND ROCOF AS OBSERVED AT 
GENERATOR 502 ....................................................................................................................... 92 
FIGURE 2‐23 LINEAR POLYNOMIAL SURFACE VIEW FOR ONLINE SYNCHRONOUS GENERATOR UNIT 
VARIATION WITH FREQUENCY NADIR POINT AND ROCOF AS OBSERVED AT GENERATOR 501
 .................................................................................................................................................. 93 
FIGURE 2‐24 LINEAR POLYNOMIAL SURFACE VIEW FOR WIND PENETRATION WITH FREQUENCY 
NADIR POINT AND ROCOF AS OBSERVED AT GENERATOR 502 ................................................ 93 
FIGURE 2‐25 ROCOF COMPARISON WITH/WITHOUT SYNCHRONOUS CONDENSERS ...................... 102 
FIGURE 2‐26 FREQUENCY NADIR COMPARISON WITH/WITHOUT SYNCHRONOUS CONDENSERS .. 102 
FIGURE 2‐27  BUSES FREQUENCY DURING GENERATOR CONTINGENCY AND NO SYNCHRONOUS 
CONDENSER ADDITION ........................................................................................................... 103 

14
FIGURE 2‐28 BUSES FREQUENCY DURING GENERATOR CONTINGENCY AND EQUIVALENT 
SYNCHRONOUS CONDENSER ADDED ...................................................................................... 103 
FIGURE 2‐29 FREQUENCY RESPONSIVE WIND TURBINE MODEL...................................................... 104 
FIGURE 2‐30 FREQUENCY OBSERVED AT BUS 506 WHEN NORMAL OPERATED DFIG BASED WIND 
FARM IS ADDED ...................................................................................................................... 104 
FIGURE 2‐31 FREQUENCY OBSERVED AT BUS 506 WHEN FREQUENCY RESPONSIVE DFIG BASED 
WIND FARM IS ADDED ............................................................................................................ 105 
FIGURE 3‐1 FREQUENCY VARIATION ALONG WITH REGULATION SERVICES .................................... 114 
FIGURE 3‐2 FREQUENCY CONTROL DURING CONTINGENCY EVENT [REFORMULATED FROM 4] .... 115 
FIGURE 3‐3 INERTIA CONSTANT VALUES OF VARIOUS PLANTS, DATA DERIVED FROM [70] [45] .... 118 
FIGURE 3‐4 FREQUENCY POWER CHARACTERISTICS THROUGH DROOP CONTROL ......................... 121 
FIGURE 3‐5 PRIMARY FREQUENCY RESPONSE OF UNCONTROLLED THERMAL POWER PLANT WITH 
DIFFERENT DROOP .................................................................................................................. 122 
FIGURE 3‐6 POWER OUTPUT CHANGE OF UNCONTROLLED THERMAL POWER PLANT UNDER 
DIFFERENT DROOP .................................................................................................................. 122 
FIGURE 3‐7 SINGLE AREA PRIMARY FREQUENCY CONTROL MODEL WITH FREQUENCY DEPENDENT 
LOADS [20] .............................................................................................................................. 123 
FIGURE 3‐8 BASIC LOOP OF AGC CONTROL ...................................................................................... 127 
FIGURE 3‐9 AGC CONTROLLED THERMAL HYDRO THERMAL POWER PLANT BASED THREE AREA 
SYSTEM FREQUENCY RESPONSE UNDER 1% LOAD DISTURBANCE ......................................... 128 
FIGURE 3‐10 AGC RESPONSE OF SINGLE AREA THERMAL POWER PLANT UNDER 1% LOAD 
DISTURBANCE ......................................................................................................................... 128 
FIGURE 3‐11 ACTIVE POWER FREQUENCY CONTROL RESEARCH STUDY LINE IN WIND BASED POWER 
PLANTS .................................................................................................................................... 130 
FIGURE 3‐12 VIRTUAL INERTIA MODEL ............................................................................................ 130 
FIGURE 3‐13 INERTIA AND DROOP CONTROL MODEL FOR FREQUENCY RESPONSE FROM VSWTG 132 
FIGURE 3‐14 REPRESENTATION OF OPTIMUM AND DELOADED CURVE FOR VSWTG ...................... 133 
FIGURE 3‐15 CONTROL LOOP FOR DE‐LOADING CONTROL OF VSWTG ........................................... 134 
FIGURE 3‐16 PITCH DROOP DE‐LOADING CONTROL ........................................................................ 135 
FIGURE 3‐17 COEFFICIENT OF PERFORMANCE DE‐LOADING THROUGH PITCH CONTROL .............. 135 
FIGURE 3‐18 PITCH ANGLE OFFSET VARIATION WITH DE‐LOADED POWER RESERVE IN VSWTG .... 135 
FIGURE 3‐19 IMPACT ON SYSTEM FREQUENCY VARIATION WITH PITCH DROOP DE‐LOADED VSWTG
 ................................................................................................................................................ 136 
FIGURE 3‐20 WIND ENERGY STUDY LINE FOR AGC .......................................................................... 139 
FIGURE 3‐21 BASIC SCHEMATIC DIAGRAM FOR WIND FARM FREQUENCY CONTROL ..................... 140 
FIGURE 3‐22 VSWTG BASED WIND FARM CONTROLLER SUPPORTING AGC [58] ............................. 140 
FIGURE 4‐1 SCHEMATIC DIAGRAM OF FULL CONVERTER BASED VSWTG ........................................ 165 
FIGURE 4‐2 DETAILED SIMULATION MODEL OF TYPE 4 VSWTG ....................................................... 166 
FIGURE 4‐3 DC/DC BOOST CONVERTER & DECOUPLED VOC INVERTER INTERFACED SYNCHRONOUS 
GENERATOR WIND ENERGY SYSTEM. ..................................................................................... 168 
FIGURE 4‐4 SIMULATION FRAMEWORK OF SYNCHRONOUS MACHINE IN ROTOR REFERENCE‐FRAME 
[15] .......................................................................................................................................... 169 
FIGURE 4‐5 WECC TYPE 4 WIND TURBINE GENERATOR MODEL [3] ................................................. 172 
FIGURE 4‐6 GE’S TYPE 4 WIND TURBINE GENERATOR GENERIC MODEL (REFORMULATED [5]) ...... 173 
FIGURE 4‐7 PROPOSED TYPE 4 WTG MODEL .................................................................................... 174 
FIGURE 4‐8 SIMPLIFIED AERODYNAMIC SIMULATION MODEL ........................................................ 178 
FIGURE 4‐9 TURBINE POWER CHARACTERISTICS ............................................................................. 179 
FIGURE 4‐10 PITCH CONTROLLER BLOCK ......................................................................................... 179 
FIGURE 4‐11 GENERATOR‐CONVERTER MODEL [20] ........................................................................ 182 
FIGURE 4‐12 ACTIVE POWER CONTROL ............................................................................................ 184 

15
FIGURE 4‐13 TYPE 4 GENERIC MODEL INCORPORATING LINEAR AERODYNAMIC MODEL 
PERFORMANCE CHARACTERISTICS AT CONSTANT WIND SPEED AT 8 M/S, 13 M/S, 18 M/S 
AND 20 M/S ALONG WITH VARYING POWER SETPOINT ......................................................... 188 
FIGURE 4‐14 TYPE 4 GENERIC MODEL INCORPORATING LINEAR AERODYNAMIC MODEL 
PERFORMANCE CHARACTERISTICS UNDER VARYING WIND SPEED & VARYING POWER 
SETPOINT ................................................................................................................................ 188 
FIGURE 4‐15 COEFFICIENT OF POWER CALCULATION MODEL FOR WTG SYSTEM ........................... 189 
FIGURE 4‐16 COEFFICIENT OF PERFORMANCE MODEL 2 FOR WTG SYSTEM [23] ........................... 189 
FIGURE 4‐17 TYPE 4 GENERIC MODEL INCORPORATING DETAILED AERODYNAMIC MODEL 1 
PERFORMANCE CHARACTERISTICS AT WIND SPEED OF 11 M/S, 13 M/S, 18 M/S AND 25 M/S 
ALONG WITH VARIABLE POWER SETPOINT ............................................................................ 191 
FIGURE 4‐18 TYPE 4 GENERIC MODEL INCORPORATING DETAILED AERODYNAMIC MODELS AT 
DIFFERENT CONSTANT WIND SPEED ALONG WITH CONSTANT 0.8 POWER SETPOINT ......... 193 
FIGURE 4‐19 TYPE 4 GENERIC MODEL INCORPORATING DETAILED AERODYNAMIC MODEL 1 
PERFORMANCE CHARACTERISTICS ......................................................................................... 193 
FIGURE 4‐20 TYPE 4 MODEL CHARACTERISTICS WITH DETAILED AERODYNAMIC MODEL 2 
PERFORMANCE CHARACTERISTICS ......................................................................................... 194 
FIGURE 4‐21 GE’S TYPE 4 GENERIC MODEL PERFORMANCE CHARACTERISTICS AT DIFFERENT WIND 
SPEED RANGING FROM 8 M/S – 20 M/S ................................................................................. 196 
FIGURE 4‐22 TYPE 4 DETAILED EMT MODEL INCORPORATING SYNCHRONOUS GENERATOR 
CHARACTERISTICS AT CONSTANT WIND SPEED OF 13 M/S &18 M/S. .................................... 197 
FIGURE 4‐23 UNCONTROLLED THREE AREA INTERCONNECTED SYSTEM FREQUENCY DURING 10% 
LOAD CHANGE ......................................................................................................................... 199 
FIGURE 4‐24 GE MODEL PERFORMANCE WITH/WITHOUT INERTIA SUPPORT ................................ 201 
FIGURE 4‐25 TYPE‐4 MODEL PERFORMANCE WITH/WITHOUT INERTIA SUPPORT .......................... 203 
FIGURE 5‐1 NEM MAINLAND FREQUENCY 1045 HRS. TO 1109 HRS. DURING CONTINGENCY EVENT 
[1] ............................................................................................................................................ 210 
FIGURE 5‐2 BASIC STRUCTURE OF GRID FREQUENCY PROCESSOR .................................................. 211 
FIGURE 5‐3 AN EXAMPLE OF INPUT FREQUENCY OBTAINED FROM NEM‐14 GENERATOR MODEL 
PROVIDED AS INPUT TO FREQUENCY PROCESSOR ................................................................. 213 
FIGURE 5‐4 NEM‐14 GENERATOR MODEL PROCESSED FREQUENCY OBTAINED FROM GRID 
FREQUENCY PROCESSOR ........................................................................................................ 213 
FIGURE 5‐5 MEASURED FREQUENCY VS. GRID PROCESSOR OUTPUT FREQUENCY AT DIFFERENT 
THRESHOLD VALUES ............................................................................................................... 214 
FIGURE 5‐6 MEASURED FREQUENCY VS. GRID PROCESSOR OUTPUT FREQUENCY AT DIFFERENT 
THRESHOLD VALUES (ZOOMED) ............................................................................................. 214 
FIGURE 5‐7 PROCESSED FREQUENCY, TREND FILTER OUTPUT, LOW PASS FREQ. FILTER OUTPUT AND 
DYNAMIC DEAD‐BAND OUTPUT ............................................................................................. 214 
FIGURE 5‐8 PROCESSED FREQUENCY, TREND FILTER OUTPUT, LOW PASS FREQ. FILTER OUTPUT AND 
DYNAMIC DEAD‐BAND OUTPUT WHEN THRESHOLD APPLIED IS >0.06 (ZOOMED) ............... 215 
FIGURE 5‐9 A SCHEMATIC REPRESENTATION OF MASTER‐SLAVE WIND TURBINE CONTROLLER 
INCORPORATING FREQUENCY RESPONSE CONTROLLER MODEL ........................................... 216 
FIGURE 5‐10 FREQUENCY RESPONSE CONTROLLER MODEL‐I .......................................................... 217 
FIGURE 5‐11 FREQUENCY GRID CODE UK [7] ................................................................................... 219 
FIGURE 5‐12 ACTIVE POWER SETPOINT GENERATED WHEN EMPLOYING FREQUENCY RESPONSE 
CONTROLLER‐I IN TWO DIFFERENT MODE ............................................................................. 219 
FIGURE 5‐13 SYSTEM FREQUENCY USED TO DEPICT FREQUENCY RESPONSE CONTROLLER‐I 
OPERATION IN TWO DIFFERENT MODE .................................................................................. 220 
FIGURE 5‐14 FREQUENCY RESPONSE CONTROLLER MODEL‐II ......................................................... 220 
FIGURE 5‐15 NON‐SYMMETRICAL DROOP CURVE FOR FREQUENCY POWER REGULATION ............ 221 

16
FIGURE 5‐16 WTG POWER SETPOINTS WITH LIMITED POWER SETPOINT AS GRID CONTROLLER 
OUTPUT SETPOINT AND CORRESPONDING GENERATED WTG POWER .................................. 222 
FIGURE 6‐1 BASIC COMPONENTS OF GENERIC VSWTG MODEL ....................................................... 227 
FIGURE 6‐2 GENERIC TYPE 3 VSWTG BASIC MODEL WITH AUXILIARY VIRTUAL INERTIA MODEL .... 228 
FIGURE 6‐3 TYPE 3 WTG MODEL WITH ACTIVE POWER CONTROLLER FEEDING POWER SETPOINT 
PAPC TO POWER CONTROLLER LOOP, PITCH CONTROLLER LOOP AND SPEED CONTROLLER 
LOOP ....................................................................................................................................... 229 
FIGURE 6‐4 CHARACTERISTICS CURVE OF TYPE 3 VSWTG UNDER MPPT AND RESERVE POWER MODE
 ................................................................................................................................................ 231 
FIGURE 6‐5 WIND AND AVAILABLE POWER SIMULATION MODEL ................................................... 232 
FIGURE 6‐6 GENERATED WIND SPEED AND AVAILABLE POWER ...................................................... 233 
FIGURE 6‐7 SINGLE AREA HYDRO GOVERNOR TURBINE MODEL WITH INTEGRATED WIND FARM . 234 
FIGURE 6‐8 UNCONTROLLED SINGLE AREA MODEL FREQUENCY DEVIATION FOR 0.1 P.U. LOAD 
DISTURBANCE ......................................................................................................................... 238 
FIGURE 6‐9 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL FOR 0.1 P.U. LOAD 
DISTURBANCE, REQ=0.035 FOR ALL CASES, TORQUE CONTROLLER GAINS (3, 0.6), NO RE‐
TUNING FOR TORQUE CONTROLLER ....................................................................................... 238 
FIGURE 6‐10 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL FOR 0.1 P.U. LOAD 
DISTURBANCE AND   REQ =0.035 IN ALL CASES, TORQUE CONTROLLER RETUNED FOR OTHER 
THREE DROOP BASED VSWTG MODELS .................................................................................. 239 
FIGURE 6‐11 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL FOR 0.1 P.U. LOAD 
DISTURBANCE,                         REQ =0.0556 IN CASE OF FREQUENCY CONTROLLER‐I AND II, 
TORQUE CONTROLLER RETUNED ............................................................................................ 239 
FIGURE 6‐12 RATE OF CHANGE OF FREQUENCY COMPARISON FOR DIFFERENT LOAD CHANGE (P.U.) 
AND 10% VSWTG PENETRATION. ........................................................................................... 240 
FIGURE 6‐13 FREQUENCY NADIR COMPARISON FOR DIFFERENT LOAD CHANGE (P.U.) AND 10% 
VSWTG PENETRATION ............................................................................................................ 241 
FIGURE 6‐14 ELECTRICAL POWER FROM TYPE 3 VSWTG WITH RETUNED TORQUE CONTROLLER IN 
CONTROL AREA. DURING INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE, 
REQ= 0.0556 FOR FREQUENCY CONTROLLER‐I AND II ............................................................ 242 
FIGURE 6‐15 INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE ............................ 242 
FIGURE 6‐16 GENERATED ACTIVE POWER SETPOINT FOR FREQUENCY CONTROLLER‐II IN THE CASE 
OF 0.1 P.U. LOAD DISTURBANCE IN CONTROL AREA .............................................................. 244 
FIGURE 6‐17 GENERATED ACTIVE POWER SETPOINT FOR FREQUENCY CONTROLLER‐I IN THE CASE 
OF 0.1 P.U. LOAD DISTURBANCE IN CONTROL AREA .............................................................. 244 
FIGURE 6‐18 VSWTG ROTOR SPEED VARIATIONS WITH DIFFERENT FREQUENCY DROOP MODELS 
DURING FREQUENCY RESPONSE ............................................................................................. 246 
FIGURE 6‐19 VSWTG ROTOR SPEED VARIATIONS WITH DIFFERENT FREQUENCY CONTROL LOOPS 
[ZOOMED] ............................................................................................................................... 246 
FIGURE 6‐20 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL INTEGRATED WITH 
PROPRIETARY TYPE 4 VSWTG MODEL FOR 0.1 P.U. LOAD DISTURBANCE, REQ = 0.035, NO 
RETUNING APPLIED IN TORQUE CONTROLLER ....................................................................... 247 
FIGURE 6‐21 GENERATED ELECTRICAL POWER FROM PROPRIETARY TYPE 4 VSWTG MODEL DURING 
INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE, NO RETUNING APPLIED IN 
TORQUE CONTROLLER ............................................................................................................ 247 
FIGURE 6‐22 PROPRIETARY TYPE 4 VSWTG MODEL ROTOR SPEED VARIATIONS WITH DIFFERENT 
FREQUENCY DROOP MODELS DURING FREQUENCY RESPONSE ............................................. 248 
FIGURE 6‐23 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL INTEGRATED WITH 
PROPOSED TYPE 4 VSWTG MODEL FOR 0.1 P.U. LOAD DISTURBANCE, REQ=0.035 ............... 250 

17
FIGURE 6‐24 FREQUENCY RESPONSE OF SINGLE AREA LFC CONTROL MODEL INTEGRATED WITH 
PROPOSED TYPE 4 VSWTG MODEL FOR 0.1 P.U. LOAD DISTURBANCE [ZOOMED] ................ 250 
FIGURE 6‐25 GENERATED ELECTRICAL POWER FROM PROPRIETARY TYPE 4 VSWTG MODEL DURING 
INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE, REQ=0.035 .................... 250 
FIGURE 6‐26 GRID CODE SENSITIVE ACTIVE POWER SETPOINT FROM FREQUENCY CONTROLLER‐I 
DURING INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE .......................... 251 
FIGURE 6‐27 GRID CODE SENSITIVE ACTIVE POWER SETPOINT FROM FREQUENCY CONTROLLER‐II 
DURING INITIAL FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE .......................... 251 
FIGURE 6‐28 GENERATED ROTOR SPEED FROM PROPOSED TYPE 4 VSWTG MODEL DURING INITIAL 
FREQUENCY RESPONSE FOR 0.1 P.U. LOAD DISTURBANCE, REQ=0.035 ................................ 251 
FIGURE 7‐1 CONVENTIONAL FUZZY PI CONTROLLER MODEL 1 [REFORMULATED FROM 12] .......... 262 
FIGURE 7‐2 CONVENTIONAL FUZZY PI CONTROLLER MODEL 2 ........................................................ 262 
FIGURE 7‐3 CLASSICAL HYDRO TURBINE GOVERNOR MODEL .......................................................... 264 
FIGURE 7‐4 HYDRO GOVERNOR MODEL ........................................................................................... 265 
FIGURE 7‐5 FREQUENCY RESPONSE FROM HYDRO POWER PLANT ................................................. 267 
FIGURE 7‐6 MECHANICAL POWER OUTPUT FROM HYDROPOWER PLANT DURING FREQUENCY 
RESPONSE ............................................................................................................................... 267 
FIGURE 7‐7 SPLIT SHAFT BASED GAS TURBINE ................................................................................. 268 
FIGURE 7‐8 GAST MODEL ................................................................................................................. 269 
FIGURE 7‐9 GAST MODEL IN THE PRESENCE OF FREQUENCY RESPONSE ......................................... 272 
FIGURE 7‐10 MECHANICAL POWER FROM GAST MODEL DURING FREQUENCY RESPONSE ............ 272 
FIGURE 7‐11 FREQUENCY RESPONSE UNDER AGC CONTROL .......................................................... 273 
FIGURE 7‐12 SIMPLIFIED DCHP MODEL ............................................................................................ 273 
FIGURE 7‐13 DCHP GOVERNOR MODEL ........................................................................................... 274 
FIGURE 7‐14 POWER LIMITATION BLOCK ......................................................................................... 274 
FIGURE 7‐15 POWER DISTRIBUTION BLOCK ..................................................................................... 275 
FIGURE 7‐16 GAS DYNAMIC BLOCK .................................................................................................. 276 
FIGURE 70‐17 DCHP FREQUENCY RESPONSE UNDER 0.01 P.U. LOAD DISTURBANCE ...................... 277 
FIGURE 7‐18 DCHP POWER OUTPUT UNDER 0.01 P.U. LOAD DISTURBANCE .................................. 277 
FIGURE 7‐19 SINGLE REHEAT TURBINE MODEL ................................................................................ 278 
FIGURE 7‐20 SIMPLE GOVERNOR MODEL ........................................................................................ 279 
FIGURE 7‐21 DIRECT SEARCH ALGORITHM ....................................................................................... 282 
FIGURE 7‐22 FREQUENCY REGULATION MODEL FOR A SINGLE CONTROL AREA ............................. 283 
FIGURE 7‐23 PRIMARY FCR SERVICE MODEL WITH WIND PARTICIPATION IN FCR .......................... 285 
FIGURE 7‐24 SECONDARY FRR SERVICES WITH WIND PARTICIPATING IN FRR ................................ 286 
FIGURE 7‐25 RR CONTROL LOOP ...................................................................................................... 286 
FIGURE 7‐26 WTG INERTIA CONTROLLER ......................................................................................... 286 
FIGURE 7‐27 RING TOPOLOGY FOR INTERCONNECTED CONTROL AREA ......................................... 287 
FIGURE 7‐28 TAIPS MODEL WITH INTEGRATED WIND FARM .......................................................... 289 
FIGURE 7‐29 WIND FARM NETWORK CONNECTION FOR LVRT TEST ............................................... 291 
FIGURE 7‐‐30 CONTROL AREA FREQUENCY UNDER NORMAL DISTURBANCE AND NO FREQUENCY 
CONTROL ................................................................................................................................. 293 
FIGURE 7‐‐31 REGULATION RESERVE RESPONSE WITH INTEGRATED WIND PROVIDING NO 
FREQUENCY RESPONSE ........................................................................................................... 294 
FIGURE 7‐32 FCR CONTROL SIGNAL RECEIVED BY WIND PLANT WHEN 40% CONVENTIONAL FCR 
UNITS ARE DISCONNECTED ..................................................................................................... 294 
FIGURE 7‐33 FCR RESPONSE WITH 40 % WIND CONTRIBUTING TO FCR RESERVE .......................... 295 
FIGURE 7‐34 POWER SET‐POINT GENERATED BY FREQUENCY CONTROLLER 2 WHEN WIND PLANT 
PARTICIPATES AS FCR .............................................................................................................. 295 

18
FIGURE 7‐35 GENERATED POWER WITH INERTIAL AND FREQUENCY CONTROLLER 2 ENABLED WHEN 
WIND PLANT IS PARTICIPATING IN FCR .................................................................................. 296 
FIGURE 7‐36 POWER REGULATION RESERVE RESPONSE FOR RANDOM HIGH LOAD WITH 40% 
HYDRO REPLACED WITH DROOP‐INERTIA ENABLED FREQUENCY RESPONSIVE WIND PLANT297 
FIGURE 7‐37 FCR AND FRR COMMAND SIGNAL RECEIVED BY WIND TURBINE UNIT UNDER RANDOM 
HIGH LOAD CONDITION .......................................................................................................... 298 
FIGURE 7‐38 RESERVE RESPONSE DURING RANDOM NORMAL LOAD VARIATIONS WHEN THE WIND 
PARTICIPATES IN FRR ONLY ..................................................................................................... 298 
FIGURE 7‐39 RESERVE RESPONSE DURING RANDOM HIGH LOAD VARIATIONS WHEN THE WIND 
PARTICIPATES IN FRR & FCR .................................................................................................... 299 
FIGURE 7‐40 FREQUENCY RESPONSE INDICATORS COMPARISON WITH WIND PLANT PARTICIPATION 
IN DIFFERENT REGULATION SERVICES AND DIFFERENT LOAD DEMAND ................................ 300 
FIGURE 7‐41 ELECTRICAL POWER UNDER NORMAL LOAD DISTURBANCE (INERTIAL POWER IS ZERO)
 ................................................................................................................................................ 301 
FIGURE 7‐42 ROTOR SPEED UNDER LIMITED POWER SETPOINT ..................................................... 301 
FIGURE 7‐43 WTG ELECTRICAL POWER DURING LOAD DISTURBANCE CONTINGENCY ................... 302 
FIGURE 7‐44 WTG ELECTRICAL POWER DURING LOAD DISTURBANCE CONTINGENCY (ZOOMED) . 302 
FIGURE 7‐45 WTG ROTOR SPEED DURING RANDOM HIGH LOAD .................................................... 303 
FIGURE 7‐46 SYSTEM FREQUENCY WHEN WIND PLANT PARTICIPATES IN DIFFERENT REGULATION 
SERVICES AT HIGH LOAD DISTURBANCE ................................................................................. 303 
FIGURE 7‐47 SYSTEM FREQUENCY WHEN WIND PLANT PARTICIPATES IN DIFFERENT REGULATION 
SERVICES AT NORMAL LOAD DISTURBANCE ........................................................................... 303 
FIGURE 7‐48 SYSTEM FREQUENCY DURING A CONTINGENCY/HIGH LOAD ..................................... 304 
FIGURE 7‐49 WTG ELECTRICAL POWER DURING CONTINGENCY/HIGH LOAD ................................. 304 
FIGURE 7‐50 COMPARISON OF WIND PARTICIPATION IN FCR & FRR RESPONSE IN TERMS OF 
SYSTEM FREQUENCY DURING A RANDOM LOAD ................................................................... 305 
FIGURE 7‐51 COMPARISON OF WIND PARTICIPATION IN FCR & FRR RESPONSE IN TERMS OF WTG 
POWER DURING A RANDOM LOAD ........................................................................................ 305 
FIGURE 7‐52 COMPARISON OF AREA 1‐2‐3 FREQUENCY RESPONSE FOR 0.01 P.U. LOAD 
DISTURBANCE IN AREA 1 ONLY ............................................................................................... 310 
FIGURE 7‐53 COMPARISON OF AREA 1‐2‐3 FREQUENCY RESPONSE FOR 0.01 P.U. LOAD 
DISTURBANCE IN ALL THREE AREA ......................................................................................... 310 
FIGURE 7‐54 MECHANICAL POWER CHANGE OF ALL GENERATING UNIT IN INTERCONNECTED 3 
AREA FOR 0.01 P.U. LOAD IN AREA 1 ONLY ............................................................................ 311 
FIGURE 7‐55 MECHANICAL POWER CHANGE OF ALL GENERATING UNIT IN INTERCONNECTED 3 
AREA FOR 0.01 P.U. LOAD IN ALL THREE AREA ....................................................................... 311 
FIGURE 7‐56 TIE‐LINE POWER CHANGE OF ALL GENERATING UNIT IN INTERCONNECTED 3 AREA FOR 
0.01 P.U. LOAD IN AREA 1 ONLY ............................................................................................. 311 
FIGURE 7‐57 TIE‐LINE POWER CHANGE OF ALL GENERATING UNIT IN INTERCONNECTED 3 AREA FOR 
0.01 P.U. LOAD IN ALL THREE AREA ........................................................................................ 312 
FIGURE 7‐58 FREQUENCY DEVIATION IN AREA 1 FOR DIFFERENT LEVEL OF WIND PLANT FOR 0.01 
P.U. LOAD IN AREA 1 CONSIDERING THE EFFECT OF THE WIND ON R AND B ........................ 313 
FIGURE 7‐59 FREQUENCY DEVIATION IN AREA 1 FOR DIFFERENT LEVEL OF WIND PLANT FOR 0.01 
P.U. LOAD IN AREA 1 WITHOUT CONSIDERING THE EFFECT OF THE WIND ON R AND B ....... 313 
FIGURE 7‐60 FREQUENCY DEVIATION IN AREA 1 FOR DIFFERENT LEVEL OF WIND PLANT FOR 0.01 
P.U. LOAD IN AREA 1 CONSIDERING THE EFFECT OF THE WIND ON R AND B ........................ 314 
FIGURE 7‐61 FREQUENCY DEVIATION IN AREA 1 FOR DIFFERENT LEVEL OF WIND PLANT FOR 0.01 
P.U. LOAD IN AREA 1 WITHOUT CONSIDERING THE EFFECT OF THE WIND ON R AND B ....... 314 

19
FIGURE 7‐62 COMPARISON OF FREQUENCY DEVIATION WHEN ONLY CONVENTIONAL PLANT 
PARTICIPATES IN AGC IN THE PRESENCE OF FREQUENCY RESPONSIVE WIND PLANT AND 0.01 
P.U. LOAD DISTURBANCE. ....................................................................................................... 314 
FIGURE 7‐63 COMPARISON OF FREQUENCY DEVIATION WHEN ONLY CONVENTIONAL PLANT 
PARTICIPATES IN AGC IN THE PRESENCE OF FREQUENCY RESPONSIVE WIND PLANT AND 0.01 
P.U. LOAD DISTURBANCE. ....................................................................................................... 315 
FIGURE 7‐64  INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN GAS BASED AREA 2 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 2. GAST MODEL WITH LOWER FUEL LIMIT OF ‐0.02 IS 
APPLIED FOR GAS PLANT MODEL. .......................................................................................... 315 
FIGURE 7‐65 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN GAS BASED AREA 2 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 2. GAST MODEL WITH LOWER FUEL LIMIT OF ‐1 IS APPLIED 
FOR GAS PLANT MODEL. ......................................................................................................... 315 
FIGURE 7‐66 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN THERMAL‐BASED AREA IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 1. ..................................................................................... 316 
FIGURE 7‐67 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN HYDRO BASED AREA 3 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 3. ..................................................................................... 316 
FIGURE 7‐68 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN DCHP BASED AREA 2 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 2. ONLY CONVENTIONAL GENERATING UNIT PARTICIPATES 
IN AGC. LOAD LIMIT (1.2, ‐0.1) P.U. ........................................................................................ 316 
FIGURE 7‐69 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN DCHP BASED AREA 2 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 2. ..................................................................................... 317 
FIGURE 7‐70 INTERCONNECTED 3‐AREA FREQUENCY RESPONSE WHEN DCHP BASED AREA 2 IS 
INTEGRATED WITH FREQUENCY RESPONSIVE WIND PLANT AND 0.01 P.U. LOAD 
DISTURBANCE IS APPLIED IN AREA 2. ONLY CONVENTIONAL GENERATING UNIT PARTICIPATES 
IN AGC. LOAD LIMIT (1, ‐1) P.U. .............................................................................................. 317 
FIGURE 7‐71 COMPARISON OF FREQUENCY RESPONSE FOR THERMAL‐BASED AREA 1 INTEGRATED 
WITH FREQUENCY RESPONSIVE WIND PLANT WHEN BOTH OF THEM PARTICIPATES IN AGC AS 
PER THE INDICATED PARTICIPATION FACTOR. EFFECT OF WIND PLANT ON R AND B HAS BEEN 
CONSIDERED HERE. ................................................................................................................. 318 
FIGURE 7‐72 COMPARISON OF POWER DEVIATION FOR THERMAL‐BASED AREA 1 INTEGRATED 
WITH FREQUENCY RESPONSIVE WIND PLANT WHEN BOTH OF THEM PARTICIPATES IN AGC 
WITH 0.4 PARTICIPATION FACTOR (PF) FOR THERMAL AND 0.6 PF FOR THE WIND.  EFFECT OF 
WIND PLANT ON R AND B HAS BEEN CONSIDERED HERE. ...................................................... 318 
FIGURE 7‐73 COMPARISON OF FREQUENCY RESPONSE FOR THERMAL‐BASED AREA 1 INTEGRATED 
WITH FREQUENCY RESPONSIVE WIND PLANT WHEN BOTH OF THEM PARTICIPATES IN AGC AS 
PER THE INDICATED PARTICIPATION FACTOR. EFFECT OF WIND PLANT ON R AND B HAS NOT 
BEEN CONSIDERED HERE. ....................................................................................................... 318 
FIGURE 7‐74 COMPARISON OF POWER DEVIATION FOR THERMAL‐BASED AREA 1 INTEGRATED 
WITH FREQUENCY RESPONSIVE WIND PLANT WHEN BOTH OF THEM PARTICIPATES IN AGC 
WITH 0.4 PARTICIPATION FACTOR (PF) FOR THERMAL AND 0.6 PF FOR THE WIND. EFFECT OF 
WIND PLANT ON R AND B HAS NOT BEEN CONSIDERED HERE. .............................................. 319 
FIGURE 7‐75 ACE OBTAINED UNDER DIFFERENT SIMULATION SCENARIOS ..................................... 319 
FIGURE 7‐76 COMPARISON OF FREQUENCY DEVIATION FOR THERMAL‐BASED AREA 1 INTEGRATED 
WITH WIND PLANT. THERMALAGC‐WIND REFERS THAT ONLY THERMAL PLANT PARTICIPATES 

20
IN AGC AND NORMAL OPERATING WIND PLANT IS INTEGRATED INTO SYSTEM. THE WIND (RB) 
INDICATES R AND B VALUES ARE CHANGED AS PER INERTIA REDUCTION IMPACT. FRWIND 
REFERS TO FREQUENCY CONTROLLER‐II BASED FREQUENCY RESPONSIVE WIND PLANT. 
THERMALAGC‐FRWINDAGC REFERS TO PARTICIPATION OF BOTH UNITS IN AGC. ................ 321 
FIGURE 7‐77 COMPARISON OF ROCOF FOR THERMAL‐BASED AREA 1 INTEGRATED WITH 
FREQUENCY RESPONSIVE WIND PLANT. ROCOF VALUES ARE OBSERVED WHEN 0.01 P.U. OF 
LOAD IS APPLIED AT 5 S IN AREA 1 ONLY. LEGENDS ARE SAME AS EXPLAINED ABOVE IN 
EARLIER FIGURE. ..................................................................................................................... 322 
FIGURE 7‐78 COMPARISON OF SETTLING TIME FOR THERMAL‐BASED AREA 1 INTEGRATED WITH 
FREQUENCY RESPONSIVE WIND PLANT. LEGENDS ARE SAME AS EXPLAINED ABOVE. .......... 323 
FIGURE 7‐79 MAIN EFFECT PLOTS FOR AGC PARAMETERS ON SYSTEM FREQUENCY DEVIATION. .. 324 
FIGURE 7‐80 COMPARISON OF FREQUENCY RESPONSE INDICATORS FOR CONVENTIONAL PLANT 
INTEGRATION WITH FREQUENCY RESPONSIVE WIND PLANT WHEN ONLY CONVENTIONAL 
UNIT PARTICIPATES IN AGC AND 0.01 P.U. LOAD IS APPLIED. ................................................ 325 
FIGURE 7‐81 FUZZY GAIN SCHEDULING (FGSPID) STRUCTURE ......................................................... 328 
FIGURE 7‐82 MEMBERSHIP FUNCTIONS FOR  ′ AND  ′ ........................................................... 329 
FIGURE 7‐83 MEMBERSHIP FUNCTIONS FOR ACE AND ∆ACE .......................................................... 329 
FIGURE 7‐84 HYDRO‐TURBINE GOVERNOR RESPONSE .................................................................... 329 
FIGURE 7‐85 FGSPID RULE IMPLICATION PROCESS .......................................................................... 332 
FIGURE 7‐86 FGSPID GAINS SURFACE VIEW ..................................................................................... 333 
FIGURE 7‐87 LFC CONTROLLER COMPARISON WITH 0.1 PER UNIT LOAD DISTURBANCE IN AREA 1 
ONLY AND 90%‐10% HYDRO‐WIND AGC PARTICIPATION. ...................................................... 335 
FIGURE 7‐88 LFC CONTROLLER COMPARISON WITH 0.1 PER UNIT LOAD DISTURBANCE IN ALL 3 
AREAS AND 90%‐10% HYDRO‐WIND AGC PARTICIPATION IN AREA 1. ................................... 335 
FIGURE 7‐89 LFC CONTROLLER COMPARISON WITH 0.1 PER UNIT LOAD DISTURBANCE IN ALL 3 
AREAS AND 80%‐20% HYDRO‐WIND AGC PARTICIPATION IN AREA 1. (ZOOMED TO 40 SEC) 335 
FIGURE 7‐90 THREE AREA FREQUENCY RESPONSE COMPARISON WITH 0.1 PER UNIT LOAD 
DISTURBANCE AND 90%‐10% HYDRO‐WIND AGC PARTICIPATION IN AREA 1. FUZZY GAIN 
SCHEDULED LFC CONTROLLER IS USED. .................................................................................. 337 
FIGURE 7‐91 THREE AREA FREQUENCY RESPONSE COMPARISON WITH 0.1 PER UNIT LOAD 
DISTURBANCE AND 80%‐20% HYDRO‐WIND AGC PARTICIPATION IN AREA 1. FUZZY GAIN 
SCHEDULED LFC CONTROLLER IS USED. .................................................................................. 337 
FIGURE 7‐92 AREA 1 FREQUENCY RESPONSE COMPARISON WITH 0.1 PER UNIT LOAD DISTURBANCE 
AND DIFFERENT WIND INTEGRATION SCENARIO. .................................................................. 338 
FIGURE 7‐93 FAULT RIDE‐THROUGH GRID CODES [48] .................................................................... 340 
FIGURE 7‐94 SIMULATION RESULTS FOR 0.1 VNOM FOR 150 MSEC. .............................................. 342 
FIGURE 7‐95 ROTOR CURRENT DURING VOLTAGE DISTURBANCE ................................................... 343 
FIGURE 7‐96 SIMULATION RESULTS FOR 0.1 V FOR 450 MSEC. ....................................................... 343 
FIGURE 7‐97 SIMULATION RESULTS FOR 0.1 V FOR 150 MSEC FOR NORMAL DFIG OPERATION .... 345 
FIGURE 7‐98 FRT ENABLED ACTIVE POWER CONTROL BLOCK ......................................................... 346 
FIGURE 7‐99 PFRT VARIATION WITH TERMINAL VOLTAGE .............................................................. 347 
FIGURE 7‐100 ELECTRICAL PARAMETERS IN NORMAL VSWTG DURING VOLTAGE DISTURBANCE .. 347 
FIGURE 7‐101 ELECTRICAL PARAMETERS IN FRT ENABLED VSWTG DURING VOLTAGE DISTURBANCE
 ................................................................................................................................................ 347 
FIGURE 7‐102 PRIMARY FREQUENCY RESPONSE COMPARISON DURING 0.01 P.U. LOAD 
DISTURBANCE ......................................................................................................................... 348 
FIGURE 7‐103 AGC FREQUENCY RESPONSE COMPARISON DURING 0.01 P.U. LOAD DISTURBANCE 
WITH AGC CONTROLLER GAIN AS P=‐37.47, I=0.0096 ............................................................ 348 
FIGURE 7‐104 AGC FREQUENCY RESPONSE COMPARISON DURING 0.01 P.U. LOAD DISTURBANCE 
WITH AGC CONTROLLER IN AS P= ‐ 45.63, I = 0.0133 ............................................................. 348 

21
ACRONYMS
Area Control Error ACE

Automatic Generation Control AGC

Australian Energy Market Operator AEMO

Combined Cycle Gas Turbines CCGT

Doubly Fed Induction Generator DFIG

Decentralized Combined Heat Power Plant DCHP

Energy Management System EMS

Fault Ride Through FRT

Frequency Operating Standard FOS

Frequency Control Ancillary Service FCAS

Fast Frequency Response FFR

Frequency Restoration Reserve FRR

Frequency Containment Reserve FCR

Fuzzy Gain Scheduled Proportional Integral


FGSPID
Derivative

Integral Of Time Multiplied By Absolute Error ITAE

Integral of the Squared Errors ISE

Load Frequency Control LFC

Large-scale Renewable Energy Target LRET

Levelised Costs Of Energy Generation LCOG

Maximum Power Point Tracking MPPT

National Electricity Market NEM

National Transmission Network Development Plan NTNDP

Point Of Connection PCC

22
Renewable Energy Sources RES

Rate Of Change Of Frequency ROCOF

Replacement Reserve RR

Short Circuit Ratio SCR

Supervisory Control and Data acquisition SCADA

Transmission System Operator TSO

Three Area Interconnected Power System TAIPS

Variable Speed Wind Turbine System VSWTG

Wind Energy System WES

Western Electricity Coordinating Council WECC

23
NOMENCLATURE

f Frequency

Rate of change of frequency

∆f   Frequency deviation

H Inertia constant of conventional


generating unit

S MVA rating of conventional


generating unit

W   Kinetic energy of rotating mass

IR System Inertia

∆P MW loss from generating unit

Hz Hertz per second


s
F Frequency Nadir Point

P Total Wind generated power

P Synchronous generator power

P Inter area power flow

WP Wind penetration level

F Lowest permissible frequency as per


frequency operating standard

J Moment of inertia of a rotating mass

M Equivalent moment of inertia of all


generators and the motors connected
to the grid

P Mechanical Power

P Electrical Load

H    Equivalent system inertia constant

S    System MVA base

24
S    Power rating of individual generating
unit

∆P    Change in active power output of


synchronous generators in
conventional generating plants

D  Damping constant

db  Deadband

K    1/D

   Angular speed

T    Power system time constant

P    Primary Frequency Control Power

R  Speed regulation or droop of governor

R    Variable droop

R    Transient droop of hydro turbine


model

R    Permanent droop of hydro turbine


model

M    Actual power margin available at a


wind turbine

M   Maximum power margin available


among all wind turbines

T    Turbine time constant

P    Power output from automatic


generation control

K    Proportional time constant of AGC


controller

K    Integral time constant of AGC


controller

NIA  Sum of the actual power flows on all


tie lines or interconnectors

25
NIS  Sum of the scheduled flows on all tie
lines or interconnectors

P Tie-line power

∆P Change in tie-line power between area


1 to area 2

T    Synchronizing coefficient

δ   Relative phase angle between control


areas

B  Frequency bias setting

β System frequency response


characteristic

IME  Interchange (tie line) Metering Error

T‐ω curve  Torque-Speed curve

P‐ω curve  Power-Speed curve

P    Power set point for the de-loaded


wind turbine

ω    Deloaded wind turbine rotor speed

ω    Measured rotor speed for wind turbine

ω    Optimum rotor speed

P    Maximum mechanical power from


turbine

P    Deloaded Power

K    Deloading constant

K    Optimum mechanical power constant

β    Pitch Angle

∆β   Offset pitch angle

Cp   Turbine power coefficient

26
C λ , β    Coefficient of power at optimum
value of tip speed ration and pitch
angle

λi  Tip speed ratio at ith time step

ρ Air density

A  Turbine blades swept area

R Blade length

Vwind Velocity of the incident wind

K   System stiffness

M    Inertia of power supply

Vout    Chopper output voltage

Vin   Chopper input voltage

D  Duty cycle

i , i    2-phase line current in dq synchronous


frame

W    Grid angular frequency as well as


synchronous reference frame speed

WL i & W L i   Induced reference voltages in


synchronous frame

V  and V   Decoupled controller output in dq


reference frame

φ   Flux

K    Park’s transformation matrix

f    Field winding

k    d-axis damper winding

k    q-axis damper winding

k  Number of damper circuits

27
[R]  Winding resistances diagonal matrix
in d q axis

[L] Winding self and mutual inductances


matrix in d q axis

[ω] Rotor speed matrix

T Electrical Torque

P Electrical power output

θ Pitch angle of the turbine blades

θ    Commanded pitch angle

Kaero Aerodynamic gain factor

Pset Power setpoint

∆P Scheduled power setpoint for turbine

∆P Power setpoint imposed by primary


control

i Active current command

IPmax  Upper limit of the active current

Imax  lumped converters ratings

IQmax  Maximum reactive current limit

P    Available wind power

v  Wind speeed

DR    Derating command mode

P    Reference power setpoint

P    Transmission system operator


commanded power set point

ω    Reference angular rotor speed

PAPC  Active power setpoint to turbine


controller

P    Final power command provided to


generator-converter model

28
L    Wind penetration level

H    Frequency responsive wind plant


inertia

T    Total time delay associated with


generating unit model

p.u.  Per unit

e  Error

μ   Membership value in fuzzy control

TW  Water starting time constant

∆P   Gate valve change

∆P Change in power command in terms


of speed change setting

∆ Change or deviation

T    Governor time constant

T Turbine time constant

T Time constant of steam chest and inlet


volume

T Time constant of reheater

F Fraction of total turbine power


generated by high pressure chamber

T Turbine torque

V Control valve position

∆P , , ∆P , , ∆P ,    Power variation from primary,


secondary and tertiary control

P    Inertial Power output from wind


turbine

29
EXECUTIVE SUMMARY

There is coupling between the reliability requirements of maintaining frequency and

the commercial interest in inadvertent energy supply. Modern electric power systems

are facing a critical challenge in the real‐time balancing of demand and supply due to

unpredictability and variability of wind power. Frequency response capability offered

by wind plant is not same as the primary control capability of conventional plants.

Increased penetration of wind generation in interconnected power system thus

intrinsically calls for frequency regulation analysis with present and future wind

technologies. This thesis titled “Frequency Regulation in Wind Integrated Power

System” aims to study the frequency regulation capability of wind energy system and

its integration impact on interconnected power system. Simulations are conducted

under the limitations of generic system modeling with reduced number of parameters

which is an acceptable practice for bulk power system studies like frequency

regulation of large interconnected power system.

First of all, a large interconnected network like Australian NEM is investigated to

assess current problems due to high wind penetration, and few possible solution

discussed. With full converter based wind turbine system being relatively new wind

technology, a simplified yet absolute representation of full converter based wind

turbine generator modeling is presented for frequency regulation related bulk studies.

Response capability of proposed model is verified against a proprietary model and a

detailed model.

This study also explores frequency regulation issues arising from high wind

penetration levels in power system. Given compulsory frequency response

requirement as part of the grid, power electronics based VSWTG requires modified

active power control algorithm to support regulation services. A moving average

30
frequency and the dynamic dead band dependent grid frequency processor has been

implemented with a grid code compatible active power control algorithm for

VSWTG. Grid code compatible frequency response from type-3 and proposed type-

4 VSWTG is investigated and analyzed for improvements with other available

models.

Different generation technologies are dispatched by their capability of responding to

variations in demand and fuel type. Design and modeling aspects of frequency

regulation scheme in the wind integrated multigenerational system are analyzed in

this research work. The frequency response of validated models of thermal, hydro

and gas-based power in the presence of frequency responsive wind plant is

investigated. Special focus is given to secondary frequency regulation or AGC

investigation in the wind integrated power system. Classical and intelligent AGC

controllers are analyzed in the presence of grid code compatible frequency responsive

wind power plant penetration.

Impending requisite for application of frequency responsive wind plants also calls for

fault ride through analysis of such plant. Simulation results for weak network

presented improved LVRT capability of frequency responsive VSWTGs in

comparison to normal VSWTGs. With active power output, well above zero and

turbine speed also remaining under limit during fault conditions, frequency grid code

compatible VSWTGs are found to satisfy FRT grid codes of many countries.

31
1 CHAPTER 1

INTRODUCTION

1.1 Current Status of Wind Energy System

The power grid is an enormous, intricate system and pervasive with interactions

between all the different components and across time. When any new technology gets

connected to this complex grid, its compatibility to rest of grid determined by some

basic characteristics. Renewable as a future fuel security and price offset option have

become an attractive proposition in the current global scenario. There is an eternal

growing obligation for electricity from renewable energy sources (RES). The impact

of the renewable energy sources on the grid and its contribution to the grid is still

relatively unknown with several challenges needing to be overcome before mass

deployment.

Wind power credentials as a rapidly deployable clean technology have put it at the

forefront in the fight against climate change. The Wind is the most cost competitive

renewable source of electricity generation behind hydropower. This proven and

mature technology has highly penetrated energy matrix at the global level, hence

among all available RES, maximum impact potential lies with wind power at largest

scale. There has been a tremendous increase in generated megawatts by the RES all

over the world. Wind energy has optimistic prospects as a partner in energy resolution

which can pave the way for shaping a new sustainable world for future generations.

According to a 2013 published key report [1], wind power deployment has more than

doubled, approaching 300 GW cumulative installed capacities led by China (75 GW),

the United States (60 GW) and Germany (31 GW). Wind power now provides 2.5%

of global electricity demand – and up to 30% in Denmark, 20% in Portugal, 18% in

32
Spain. Figure 1-1 depicts the cumulative wind energy market forecast for different

regions 2012-2017. Wind energy is anticipated as a major contributor for the

foreseeable future with 20% per annum growth forecast for global installed wind

energy capacity.

Figure 1-1 Cumulative Forecast by Region 2012-2017 [2]

Although most wind-power developments still attract financial support: either in the

form of capital subsidies or premium payments for the energy gap between the

generation costs of wind energy and electricity from conventional plants continues to

narrow and are close to cost-competitive with new natural gas generation due to

continuing technological innovation. The onshore wind costs fell in 2014 and, a

realistic best generation cost for onshore wind is now $71/MWh on a global basis [3-

4]. A Bloomberg report has listed the wind as cheaper electricity producing the fuel

than fossil fuels for Australia which is world’s biggest coal exporter. A new wind

farm in Australia supplies electricity at the cost of 80AUD / MWh, while it comes

out to be 143AUD / MWh for a new coal-fired power plant or 116AUD / MWh when

supplied from a new natural gas powered station in 2013 [5]. The prospective plunge

33
in LCOE of wind is expected shortly due to the emergence of low cost but high

technology wind turbine manufacturer.

1.2. Motivations for Research

The present power system is undergoing several changes in its basic structure which

are associated with the adoption of new power production technologies and rapid

integration of renewable energy sources (RES). Integration of renewable energy

resources at large has widespread impacts on power system stability & reliability.

The Wind is the most cost competitive renewable source of electricity generation

behind hydro. Proven and mature wind power technology has highly penetrated

energy matrix at the global level, hence among all available RES, maximum impact

potential lies with wind power at largest scale. In Australia, wind farms have

proliferated in the National Electricity Market (NEM), with some 3,100 MW of

installed capacity entering the market to date. However, it has significantly more

growth potential because of the greater level of as yet unutilized resources. Its cost

competitiveness will be enhanced by a reduction in the cost of turbines, particularly

through low-cost, high volume manufacturing in countries such as India and China,

and to a lesser extent by further efficiency gains through turbine technology

development. Factors that may limit the development of wind energy on a localized

basis are a lack of electricity transmission infrastructure to access remote wind

resources and the intermittency and variability of wind energy. The power output of

a wind turbine, which converts an uncertain energy flux to electricity, has an

additional source of uncertainty. Wind power density is a stochastic process that is a

function of both space and time. Wind power generation introduces uncertainty in

power system operation, and it is continuously variable and difficult to predict.

Modern electric power systems are facing a critical challenge in the real-time

34
balancing of demand and supply due to unpredictability and variability of wind

power.

The Emergence of Type 4 VSWTG Technology: After displacing dominant simple

constant speed turbines, variable speed based Wind Energy Conversion System

(WECS) technology is a pioneer in currently installed wind energy projects. Even

though doubly fed induction generators based type 3 is the predominant technology

nowadays, full converter based type 4 turbine technology is expected to become the

industry standard in medium terms. Joint Research Centre’s (JRC) 2014 report [6]

analyzing wind turbine purchase agreements (TPAs) sample consisting of countries

like China, USA, France, Germany, UK, Brazil and Rest of world identify the ending

of pre-eminence of type 3 technology due to increasing popularity of type 4 wind

turbine system. Considering the connection trends, Australian Electric Market

Operator (AEMO) also assume an increase in type 3 and type 4 wind turbines in NEM

[7].

Variable speed wind turbine generators (VSWTG) having incommensurable dynamic

characteristics than conventional synchronous generators requires precise modeling

as they no longer receive negative load treatment. WECS modeling is now a priority

research topic due to change in perspective of various stakeholders like transmission

system operators, researchers, and wind developers. The perpetual cumulative

obligation for electricity from these vastly penetrated technologically advanced wind

energy system poses an incendiary threat to stability and reliability of modern electric

grid if their dynamic performance aspects are not assessed beforehand. So the

transmission system operators (TSOs) all over the world advocate incoming WES

integrations to follow stringent grid codes in their modeling.

35
The Emergence of Wind-based Ancillary Services Requirement and

Corresponding VSWTG Modelling Modifications: Sophisticated technology has

ushered in improved forecasting of wind energy generation over varying time frames.

These enhancements have initiated the contemplation of wind energy based ancillary

services requirements for improved system scheduling & planning. Though WES

ancillary services like frequency control ancillary services (FCAS) are still in

research & development stage in most of the countries, future WES participation in

electricity economic markets is expected to contribute to enhanced market efficiency,

improved system reliability and macroeconomic benefits to all stakeholders. Even

with proven technology, wind energy generation sector still lacks in aspects like grid

integration assessment. Design and operation of power system in the presence of wind

energy are one of the major issues in wind power integration.

Renewable energy sources like wind having new production and nonlinear control

technology imply new energy management system (EMS) in modern power systems.

Talking about applying energy management in the modern power system, supervisory

control and data acquisition (SCADA), Automatic Generation Control (AGC),

system stability, security control and electrical load management make up the major

units [4]. AGC represents a remarkable economic function of EMS system for

regulating the balance and for distributing the imbalance between designated units.

The sensitivity of any power plant depends upon the response time taken to control

the frequency change due to load variation of governor control. System frequency is

regulated by injecting active power into electric grid through power plants. This

control is indispensable for the stable operation of grid ensuring continuous

adaptation of generation to demand. Under a wide range of ambient conditions,

frequency responsive power plants are expected to provide automatic power variation

36
with frequency drop within a given time frame and ramp limit. The conventional

power plants which remain connected to the grid even in the case of frequency

deviations accomplish active power-frequency control through turbine governor’s

reaction to nominal synchronous speed deviations and the respective boilers-turbine

frequency responsive controls.

Dead-band and speed droop are two significant parameters in conventional turbine

governor action during system event [8]. System frequency regulation is highly

dependent upon implemented dead-band for long-term fidelity. Governor dead band

is detrimental to minimum frequency variation needed before the governor action is

activated. The relative frequency deviation corresponding to the relative change in

power output, defined as speed droop is always positive for stable regulation. Factors

like the available headroom, the maximum-minimum power capacity of generating

units and power setpoint corresponding to enabled frequency operating mode have a

direct effect on total frequency responsive reserve amount. A generator operating at

its maximum generating capacity has almost negligible headroom and therefore

unable to provide any frequency responsive operation irrespective of governor-droop

being enabled. To get emulated synchronous generator response, variable speed wind

turbine generator (VSWTG) modeling should consider parameters like droop, dead-

band, headroom in frequency-active power control.

Integration Impacts of Wind Energy Penetration: Recent studies have found that

the wind energy integration impacts on system frequency and power fluctuation are

nonzero and become more significant at higher sizes of penetrations. For smaller

power systems with low inertia or system with low primary frequency response, a

large amount of wind integration may lead to frequency instability of the system.

Increasing wind penetrations tend to displace (decommit) conventional generation

37
making system less synchronous and more asynchronous. Present wind farms use

induction generators (double-fed or classical) which react weakly to frequency

changes leading to smaller effective inertia. Doubly Fed Induction Generators (DFIG)

without special inertial emulation do not have primary control capability. So

frequency response is degraded along with other effects (e.g., increased dead band,

sliding pressure controls, blocked governor, use of power load controllers, and

change in load frequency response). Fluctuations in the summated output of all wind

farms connected to a power system will cause the frequency to fluctuate and may

change the anticipated power output of dispatchable generation (affecting dispatch in

the next few hours and unit commitment in the next few days). Unexpected sudden

changes in the summated output of wind farms, due to either a widespread change in

wind conditions or response to a power system disturbance are contingencies that

must be assessed for their implications for reserve requirements. With increasing

wind energy penetration level, project participants involving transmission system

operator, the market regulator, manufacturer and wind plant owner presents their

perspective.

As represented in Figure 1-2 by the author, a discussion is going on at different

stakeholders’ level regarding the provision of frequency responsive capability to be

a requirement of all generators or be commercial ancillary service only. TSOs need

frequency control ancillary services (FCAS) from wind plant but doubt their

capability to provide. The market regulator believes that FCAS is easily obtainable

from other energy sources. Wind turbine manufacturer with production capability

assumes that wind FCAS is not anticipated while wind farm owners are not in favor

to provide FCAS without any monetary gain. Such dissimilar perspectives of

stakeholders’ can only be satisfied through by rigorous research studies.

38
Figure 1-2 Stakeholders View’s on Frequency Responsive Capability as a Requirement for Wind Energy

System

1.3. Objectives

A considerable number of studies are available in the literature regarding wind

penetration effects on in an electrical network. This work aims to be distinct due to

the applied models in the study. Proprietary DFIG models [9] [10] [11] were mostly

applied in previous works where some important factors like the incorporation of

continuously varying reference power setpoint are not considered. Frequency

regulation in the wind integrated power system calls for a continuously varying power

setpoint instead of a constant setpoint. With full converter based type 4 VSWTGs

gaining market share, frequency regulation capabilities are expected from them too.

So, a suitable frequency responsive model specific for frequency regulation studies

is highly desired. High penetration of wind energy system has led many of the TSOs

to formulate stringent grid codes for system stable and reliable operation. Previous

39
studies lack the frequency response analysis of grid code compatible wind energy

system on the interconnected system. Maintenance of system frequency in a large

inertia based system within utility prescribed tolerance band is an exclusive

responsibility of the conventional power plants. With increased renewable energy

targets globally, it is usual to find an area regulated by hydro generation

interconnected to another area regulated by thermal generation or gas based

generation or in a combination of both generations along with good penetration of

wind plants. Integration of wind-based generating plant to a system having fixed

governor settings brings wide deviations in system frequency, and further might lead

to system instability. Wind plants with variable output are integrated with peak load

plants like gas based plants & hydro plants rather than with baseload thermal plants.

Dynamics of different combinations of turbine – generators have different effects on

system frequency maintenance. Very little work is available in literature in case of

load frequency response in a multigenerational system with grid code frequency

responsive wind plants.

This research work is proposed to answer some of the following set objectives:

Frequency Control and Impacts of Increased Wind Penetration

• Frequency Control issues arising from low power system inertia levels have

become an important issue. What are the frequency control capabilities and

possibilities in large interconnected wind penetrated systems? A system with

the same amount of spinning reserve but different levels of inertia may exhibit

different frequency response during contingency events.

Investigation of Design Aspects of Load Frequency Scheme in Interconnected Mixed

Generational Power Systems in the Presence Of Wind Power.

40
• Electricity dispatched to the market reflects demand requirements for base

load and peak (or maximum) demand. Different generation technologies may

be grouped by their capability of responding to variations in demand and fuel

type. For generating systems of widely different characteristics, research

needs to be done to study automatic generation controllers in interconnection.

Study of Grid code compatible frequency response capability from VSWTG and the

corresponding effect on Fault Ride-through (FRT) Capabilities.

• Energy withdrawn by FRT action may increase the size of largest contingency

& thus magnitude of frequency excursions. Does frequency responsive wind

plants have improved fault ride through capabilities than normally operative

wind plants?

This work will specifically consider following aspects:

• Design and modeling of type 4 VSWTG for application in frequency

regulation studies.

• Effect of grid code implementation on VSWTGs performance and frequency

response.

• Investigate the effect inertia has on arresting frequency deviations during

frequency events.

• Investigate the effect of reserve distribution and effects of wind generation

participation in a different level of frequency control.

• Study of the characteristics of various frequency control strategies.

• Specify requirements and characteristics of electrical power systems

including thermal, hydro and wind-based power system.

• A power system simulation model for load frequency & tie-line power

variations performance verifications.

41
• Testing and comparison of conventional load frequency controllers with

fuzzy gain scheduled load frequency controllers.

• Application of the performance indicators and a direct search algorithm for

optimum AGC controller gains.

1.4. Significance

The focus of this study is to represent the effects of transmission-connected wind

generation on the power system and some of the improvements that can be made in

wind turbine generator modeling for better system frequency response.

Despite having sufficient feasibilities for wind power integration in Australian NEM,

it still lacks various technical and commercial aspects that require further

investigations to test the system adequacy for reliable and secure operation [12].

Intermittent wind power generation and asynchronous generators are two key

characteristics of wind generation technology affecting its integration in NEM grid

and spot market. Even though AEMO does not identify NEM-wide challenges as

each NEM region has a different generation mix, network configuration, and demand

characteristics, which lead to different challenges or different timing; future increased

wind power penetration will bring some adverse operational challenges in the NEM.

This research is aimed to be in line with the Australian Energy Market Operator

(AEMO) concerns published in latest wind integration-2013 [13]. The main way in

which wind generation can affect NEM frequency control is economic displacement

of synchronous generation. AEMO does not currently control inertia in any way: it is

simply an observed characteristic of the power system, much like load level. AEMO

currently operates the power system around the requirements that arise from the

inertia levels that are present. Participation in frequency control arrangements in the

42
NEM is voluntary, and AEMO’s experience is that wind generators choose not to

participate.

AEMO is investigating the extent to which fast frequency response (FFR) could

provide a substitute for synchronous inertia. Examples include synthetic inertia from

wind turbines, FFR from battery storage, or fast response generation ramping. AEMO

is assessing the capabilities and limitations of technologies that can provide an FFR

service, and whether FFR provides an adequate substitute or supplement to inertia.

Furthermore, the scale and type of response required to make a useful contribution to

the power system are unknown. At present, synthetic inertia has not been

demonstrated to be an exact substitute for mechanical inertia, but could, in

combination with fast-acting responses with the right characteristics, maintain the

frequency operating standard (FOS) with a lower level of synchronous inertia than

otherwise would be required.

There is a considerable body of work to be done to fully understand the dynamic

performance of the large interconnected power system as it goes through gradual

displacement of conventional synchronous plants by low inertia generating plants.

1.5. Methodology

The objectives set in this research study are achieved through advanced simulation

studies and detailed analyses of frequency control in the wind integrated power

system. The complete methodology has been represented in Figure 1-3. It starts with

wind integration assessment in Australian transmission network for problem

identifications and concludes finally at design & analysis of load frequency control

model for the wind-integrated system. An extensive literature review is undertaken

to gather realistic information about wind integration effects on Australian power

network infrastructure. A particular emphasis was given to the South Australia

43
network owing to highest wind penetrated state. This approach helped in developing

the framework of the proposed study to analyze the impacts of integrating wind

energy into the power grid. Frequency response studies require suitable

representation for the wind turbine generator system. Current literature provides

simple proprietary models for type 3 VSWTGs with some of the models not free and

publically available [14]. A Generic model of type 4 is available in public domain

[15], but it was found unsuitable under varying power setpoint requirements for

frequency response. Thus, a suitable model for type 4 VSWTG was developed, and

its performance was validated with a proprietary model performance. Various

unobserved factors which were not considered in previous studies were also

highlighted for generic modeling of type 3 VSWTG. The methodology used for

developing models of frequency control in wind energy interconnected systems

consists of identifying specifies and dynamic characteristics of frequency control in

large wind energy interconnected systems to incorporate them into the models

obtained with the basic theory of frequency control. Indeed, while frequency control

capabilities and possibilities are known and detailed in the literature, wind energy

system based frequency control capabilities and possibilities in large interconnected

systems need more detailed research. The frequency control model suggested in this

work will allow the reserve response simulation for all three types of frequency

control namely primary, secondary and tertiary control. Standard grid codes have

been applied to analyze VSWTGs performance when complying with these grid

codes. MATLAB’s Simulink based SimPower platform has been used for all

simulation studies.

44
Investigation of Frequency Response in NEM

Wind Integration Effect Under Low Load and High  Frequency Response Problem Identifiction and 
Load Scenario Possible Solutions

Development of Frequency Studies Model


Wind Integrated 
Individual Generation Models Frequency Regulation Models Interconnected Power System 
Models

Investigation of Major Challenges in Frequency Control for Wind Energy


System Penetrated Power System

Inertia Effect & Wind Penetration Reserve Distribution & Frequency Control

Analysis Through Design and Simulations

Frequency Regulations, Classical &  Frequency Response of Type 3 
Grid Code Implementation Intelligent Controllers & 4 VSWTGs

Wind Integrated Power System Frequency Response

Figure 1-3 Research Methodology

1.6. Contributions

Wind energy integration is a challenge for energy markets, the transmission systems

as well as for system operators. The collaboration with so many different participants

in power system calls for answers to many issues. An attempt has been made in this

thesis to answer a minor part of one of the big question of frequency dynamics issue.

Main contributions of this thesis are shown in Figure 1-4.

45
Contribution 1
Frequency Response Investigation in NEM Control Area

Contribution 2
Modelling & Performance Assessment Of Type 4 Frequency 
Responsive Wind Turbine

Contribution 3
Modelling & Investigation  of Grid Code Compatible Active‐
Power Controller for Type 4 & Type 3 VSWTGs
Major Contributions

Contribution 4
Development of Frequency Response Model for Wind Energy Impact 
Assessment on Reserve Response

Contribution  5
Integration Impact of Grid code Frequency Responsive 
VSWTGs on AGC of Multigenerational System

Contribution 6
Development of FGSPID Controller for Wind Integrated Power 
System Load Frequency Control

Contribution 7
Fault Ride Through Analysis of Frequency Responsive VSWTGs 

Contribution 8
Review of Active Power‐Frequency Control in Wind Integrated 
Power System

Figure 1-4 Contributions of the Study

1.6.1. Frequency Response Investigation in NEM Control Area

Under high renewable penetration levels, it is conceivable that frequency response of

the system will be lower due to lower inertia of the system and the displacement of

primary frequency control reserves. An empirically-validated model of NEM is

applied to mimic frequency dynamics and analyze frequency indicators. Quantitative

46
influence of system inertia and contingency events on frequency response in a control

area of Australian network is evaluated for low load and high load scenario.

1.6.2. Modeling & Performance Assessment of Type 4 Frequency

Responsive Wind Turbine

Time-frame of different phenomena is always considered in modeling precision and

simulation set up of electrical phenomena. The proprietary generic model exists for

type 3 VSWTG while the public domain available generic model for type 4 is not

very suitable for frequency response studies. A generic model of type 4 VSWTG

model suitable for frequency response is developed, and its performance is analyzed

with different wind turbine generator models.

1.6.3. Modeling and Investigation of Grid Code Compatible Active-

Power Controller for Type 3 & Type 4 VSWTGs

A grid frequency processor dependent upon dynamic dead band is presented. This

grid frequency processor is applied to formulate grid code compatible frequency

controller for active power response from VSWTGs during frequency disturbances.

The performance of this grid code compatible frequency responsive type 3 and type

4 VSWTGs is compared with other frequency responsive models for improved

response.

1.6.4. Development of Frequency Response Model for Wind Energy

Impact Assessment on Reserve Response

A single area frequency control model is presented in which frequency responsive

VSWTG is integrated at primary control level and secondary control level. Reserve

response and frequency indicators are analyzed when wind farm participates in

individual control level or combined in both control level.


47
1.6.5. Integration Impact of Grid Code Frequency Responsive

VSWTGs on AGC of Multigenerational System

Dynamics of different combinations of turbine-generators have different effects on

system frequency maintenance. A Frequency regulation study at primary and

secondary level is conducted to analyze wind plant integration with generating units

of widely different characteristics like a thermal plant, gas plant and hydro plant in

an interconnected control area.

1.6.6. Development of FGSPID Controller for Wind Integrated Power

System Load Frequency Control

Intelligent load frequency controller for wind integrated power system is developed.

Fuzzy gain scheduling is applied for intelligent control and compared with other

prevalent load frequency controllers.

1.6.7. Fault Ride-through Analysis of Frequency Responsive VSWTGs

Conventionally active power is given low priority during voltage disturbance, but

with impending grid codes for FRT and frequency response requirements from

VSWTGs, assessment of grid code compatible VSWTGs for fault ride through

capability is essential.

1.6.8. Review of Active Power-Frequency Control in Wind Integrated

Power System

A comprehensive review of active power-frequency control strategies regarding

primary and secondary frequency control is also presented, and future perspective is

also provided.

48
1.7. Outline of the Thesis

Following a comprehensive literature review of existing power systems, wind energy

technology and their integration with the grid, this report presents the findings and

potential challenges found in the existing literature. A methodological approach has

been used to investigate wind penetration effect on frequency response. After a

literature review, a suitable VSWTG modeling is conducted followed by active

power-frequency response controller development. Wind integration effect is

analyzed for system frequency control and regulation reserves.

Chapter 1 describes the motivation and significance of the study. Key technology and

research contributions are discussed in this chapter. The scope and limitations of the

study are also presented in the chapter.

Chapter 2 reviews the literature associated with Australian electricity network, its

wind resources, and frequency regulation challenge of integrating wind energy with

power grids. 14-generator NEM model is used to investigate frequency response

under contingency events and wind penetration for low load and high load scenario.

Possible solutions are proposed.

Chapter 3 reviews the frequency control and presents new challenges for AGC studies

regarding the issues of wind integration with the power grid.

Chapter 4 presents the modeling of type 4 VSWTGs suitable for frequency response

studies.

Chapter 5 presents the grid code compatible active power frequency controller model.

Chapter 6 presents the investigation of grid code compatible type 4 and type 3

VSWTG model performance and comparison with other frequency responsive model.

Chapter 7 presents frequency regulation in the multigenerational system. Various

models including single area control model for reserve response assessment,

49
intelligent load frequency controller model and fault ride through assessment of type

3 and type 4 VSWTGs suitable for frequency response are discussed.

Chapter 8 concludes this work. The contributions of this work are briefly discussed,

and possibilities of further works are highlighted.

50
1.8. References

2. IEA, (2013). Wind Technology Roadmap to 2050. Available at www.iea.org.

3. GWEC, (2012). Cumulative Forecast by Region 2012-2017. Available at

www.gwec.net.

4. Milborrow David, (2015). Onshore wind is more competitive than ever. Article.

Available at www.windpowermonthly.com.

5. BREE, (2014). Asia Pacific Renewable Energy Assessment. Available at

bree.gov.au.

6. Paton, J. (2013). Australian Wind Energy Now Cheaper Than Coal, Gas. [Online]

Available at www.bloomberg.com/news/articles/2013-02-06/australia-wind-

energy-cheaper-than-coal-natural-gas-bnef-says.

7. JRC, (2014). JRC Wind Status Report. Publications Office of the European

Union. 2015.

8. Kundur P., Balu, N., and Lauby, M. Power system stability, and control. New

York: McGraw-Hill, 1994.

9. W. W. Price, J. J. Sanchez-Gasca (2006). Simplified wind turbine generator

aerodynamic models for transient stability studies, Power Systems Conference and

Exposition, 2006. PSCE '06. 2006 IEEE PES, vol., no., pp.986, 992.

10. M. Singh, M. Vyas, S. Santoso (2010). Using generic wind turbine models to

compare inertial response of wind turbine technologies, in Power and Energy

Society General Meeting, 2010 IEEE, vol., no., and pp.1-7.

11. J. Morren, S. W. H. de Haan, W. L. Kling, J. A. Ferreira (2006). Wind turbines

emulating inertia and supporting primary frequency control, IEEE Transactions on

Power Systems, vol.21, no.1, pp.433-434.

51
12. AEMO, (2017). Black System South Australia 28 September 2016. Available at

www.aemo.com.au

13. AEMO, (2013). Wind Turbine Plant Capabilities Report, 2013. Available at

www.aemo.com.au

14. IEC Standard 61400-27-1, (2015). Wind Turbines—Part 27-1: Electrical

Simulation Models—Wind Turbines. Available at www.collections.iec.ch/std.

15. UWIG. Type 4 - Generic Wind Turbine Generator Model (Phase II), Available at

www.uwig.org:8080

52
2 CHAPTER 2

WIND INTEGRATION IN AUSTRALIAN NEM

This chapter highlights the problems associated with wind integration in the large

interconnected network. Australian electric network has been selected to describe the

wind energy resources integration in it and associated problems with it especially

regarding frequency control response issues. 14-Generator NEM model for frequency

response assessment is applied for investigating wind plant penetration led problems,

and possible mitigation method are also discussed

2.1. National Electricity Market (NEM) Overview

Blessed with diverse & plentiful renewable and non-renewable energy resources,

Australia has the distinction of being world’s ninth largest energy producer country.

Australia under the banner of National Electricity Market (NEM) owns the largest

geographical interconnected power system in the world. Transmission lines and

associated infrastructure extending approximately 51,000 km from Port Douglas in

Queensland to Port Lincoln in South Australia and across the Bass Strait to Tasmania

[1]. NEM spans Australia’s eastern and south-eastern coasts and comprises five

participating states acting as price regions in its jurisdictions – Queensland, New

South Wales (including the Australian Capital Territory), Victoria, South Australia

and Tasmania. There are approximately 270 registered generators in NEM, six state‐

centered transmission networks linked each other by cross-border interconnectors

and 16 major distribution networks for collectively supplying electricity to

consumers. Two regional reference nodes in NEM are interconnected through

transmission flow-paths called as interconnectors and consist of transmission

53
infrastructure located on each side of a regional boundary, connected by a set of high-

voltage transmission lines or cables [2].

Having a total generating capacity of around 45000 MWh, an increase of 0.03% is

forecasted for total electricity consumption in NEM region over the 20-year forecast

period (from 183,258 GWh to 184,467 GWh – giving a flat trend [3]. NEM regional

reference nodes and their interconnections are represented in Figure 2.1 by the author.

Except for Bass Link, all other interconnectors are regulated, deriving revenue by

trading on the spot market.

Figure 2-1 NEM Regional Boundaries, Their Interconnections and Generation Capacity [Data Derived

from [1], Formulated by Author]

2.2. NEM Spot Market & AEMO

NEM initiated operation as a wholesale spot market for electricity in December 1998

[1]. Electricity is exchanged across five interconnected regions through NEM acting

54
as a wholesale commodity exchange. Pool or spot market matches real time

instantaneous demand with power supply through a centralized dispatch process.

Specified amounts of electricity at specified prices is offered by generators to be

supplied to market for agreed time periods. Australian Energy Market Operator

(AEMO) scrutinizes all bids and decides the deployment of specific generators to

produce electricity according to cost-efficient methodology with the dispatching of

the cheapest generator. Spare generating capacity is kept as a reserve after matching

electricity consumption with power generation. Every 5-minute target is applied to

determine a dispatch price based on highest or the marginal bid for electricity

delivery, the spot price is determined for each NEM region by averaging six dispatch

prices over every half hour period. All financial transactions related to electricity

traded in the NEM is settled based on this spot price. A maximum spot price is known

as Market Price Cap, and a minimum spot price, known as Market Floor Price is set

according to the National Electricity Rules. Market Price Cap was set at $13,800 per

megawatt hour, and Market Floor Price was set -$1,000 per megawatt hour in January

2015 [4]. Market Price Cap highly rewards generation supply for meeting demand in

need while negative Market Floor Price strongly encourages power reduction by all

generation, including the wind in case supply exceeds demand. These two price

settings are reviewed every four years by the Australian Energy Market

Commission’s Reliability Panel to safeguard the NEM reliability standard.

Market clearing prices change continuously in different regions with higher cost

generators called upon to meet higher demands which vary throughout the day. This

generator ‘merit order’ means that the last generator required to meet market demand

(called the marginal generator) sets the market clearing price. Generating plants are

paid through customer revenue recovered by AEMO through retailers. A commercial

55
tariff is paid by consumers to the retailers who manages energy purchases on behalf

of their customers. There is are financial risks connected with volatile spot prices

during the trading period which is managed by NEM participants through financial

contracts locking in a fixed electricity price for any production or consumption at a

given time in the future. Besides obligatory regional gross-pool spot energy

operation, NEM operation is also dependent upon Frequency Control Ancillary

Services (FCAS) markets that provide a security-constrained dispatch order every

five minutes for five interconnected regions. NEM’s scheduling and cost allocation

procedures mandate participation of all generators greater than 30MW in size. Both

the transmission system and energy market fall under the direction of the Australian

Energy Market Operator (AEMO).

AEMO as transmission system and retail electricity market operator is responsible

for NEM reliable and secure operation, including maintaining required electricity

reserve. AEMO was established by the Ministerial Council for Energy (MCE), now

the Standing Council for Energy and Resources (SCER), as an independent

operational entity by 1 July 2009. Operating on cost recovery basis, AEMO solely

recovers its operating costs through market participants, and network service

providers paid fees.

As presented by Author in Figure 2-2, AEMO operates alongside three major

organizations. First one is the Australian Energy Regulator (AER), which is

responsible for economic regulation and national rules compliance in the NEM.

Second is the Australian Energy Market Commission (AEMC), which makes the

rules governing the regulation of the energy markets. The third is the Standing

Council for Energy and Resources (SCER) which is responsible for developing

policies related to electricity markets. AEMC is responsible to COAG through SCER

56
while AER is responsible to Commonwealth Government. AER and AEMC interact

with each other through a memorandum of understanding. AEMO and AEMC

collaborate on important issues like NEM future security. Both AEMO and AEMC

can implement policy changes when appropriate and refer them to COAG energy

council. The NEM operates within the framework of National Electricity Rules under

joint legislation enacted by participating states. These rules are maintained and

developed by AEMC and enforced by the AER.

Figure 2-2 AEMO Interactions with Other Regulators [Formulated by Author]

AEMO controls NEM via two identical control centers in different regions overseeing

the performance of all generation connected to the NEM. In the case of any

transmission system limitations or demand variations, AEMO may issue market

notice for generation adjustment. In the case of generation deficiency, AEMO can

instruct for load cut off to some customers to maintain a balance between generation

and consumption. Industry investors and customers can access a range of planning

information documents including benchmark annual National Transmission Network

Development Plan (NTNDP) report published by AEMO.

57
2.3. DRIVERS OF WIND ENERGY SYSTEM
DEVELOPMENT IN AUSTRALIA

2.3.1. Renewable Energy Target (RET)

With increasing environmental change concerns, limited fossil fuel reserves and

increasing electrical energy consumption, renewable energy sources (RES)

development has become a global necessity. There has been a tremendous increase

in generated megawatts by these RES all over the world including Australia. There

were 43 countries with Renewable energy target (RET) in 2005 which increased to

164 countries in 2015 [5]. Australia government is supporting the green revolution

by having set up the RET. Australia’s Renewable Energy Target is practiced as

Large-Scale Target (LRET) is centered on fixed quantity and market price, and

Small-scale Scheme (SRES) centered on a variable amount and fixed price. LRET

mechanism is based on granting certificates to eligible renewable generators that

energy retailers must surrender in proportion to their electricity sales and a fixed

LRET trajectory. Renewable energy participates in the NEM in the same way as the

non-renewable generation, but eligible plants receive additional income that

effectively guarantees a certain volume of large-scale renewable energy in the

market. The new target for large-scale generation of 33,000 GWh in 2020 means

about 23.5 percent of Australia’s electricity generation will be from renewable

sources. It targets to minimize the difference between the renewable energy

generating cost and the cost of generating electricity from traditional fossil fuel

sources and contribute towards government’s emission reduction target. A

comparison of renewable energy target of Australia with some other countries is

presented in Table 2-1.

58
Table 2-1 Comparison of Renewable Energy Target for some Countries with Australia

Country Renewable energy target (share of Achieved so far (share of


electricity in 2020 unless specified) electricity in 2012 unless
specified)
Australia [6] 23.5% 53% (2015)
Brazil [6] 23% by 2030 85%
Canada [7] Province-level targets 65%
(projected national share: 70%)
China [8] 15% final energy by 2015 Tech- 21% (2013)
specific targets (e.g. 200 GW wind by
2020)
Denmark [9] 100% by 2035; 50% by wind by 2020 40% (2016)
France [10] 27% by 2020 and 32 % by 2030 23% (2015)
Germany [6] 40-45% by 2025 28.2% (2015)
India [11] 57% by 2027 33%
S-Africa [12] 21% by 2030 5%
U.K. [13] 30% by 2020 4.23%
USA [14] 30% by 2030 8.3%
EU [15] 20% by 2020 16.7% (2015)

2.3.2. Wind Resources Availability

Figure 2-3 Australia Wind Resources [16]

Wind energy generation at large scale boast of various striking distinctiveness making

it a viable player in Australian energy market. As shown in mesoscale wind map in

Figure 2-3, Australia has some of the world’s best wind resources along its south-

western, southern and south-eastern margins which lie in the path of the westerly

59
wind flow belts between latitudes 35° and 50°known as the ‘roaring 40s’. Westerly

wind belts are characterized by strong, comparatively constant wind conditions, with

average wind speeds in a surplus of 6 meters per second (m/s) and, in places, more

than 9 m/s (shown in orange to red colors).

With first time utilization for electricity generation in 1994, wind power industry has

developed and expanded rapidly over the years with an average annual growth of

69.5 percent. Being source of low to zero-emissions and low cost proven technology,

providing cheap rollout on a large scale, wind power has dominated other renewable

sources for government support under LRET scheme. Approximately 38 percent of

all renewable capacity installed since 2000 is attributed to wind energy which has

engrossed over $5 billion in investment directly in Australia since 2001.

2.3.3. Cost Effective, Competitive & Established Technology

Upfront capital costs and lifetime running costs make up leveled generation cost and

are expressed in cost per unit energy produced. The building of new onshore wind

energy leveled generation Costs lies in-between the cheapest option, Combined Cycle

Gas Turbines (CCGTs) and those of the next cheapest option, that of coal. Offshore

wind farms cost 50% more than that of onshore farms, but experts have predicted cost

reduction with further improvements in technology.

Figure 2-4 shows the relationship between levelised costs of generation (LCOE) and

capacity factor for wind generators as per Australian Energy Technology Assessment

2013 (AETA) resourced values. Distribution of wind resource in the marketable zone

is shown to have capacity factors distributed between 20-50%. An average capacity

factor of 33% makes Australian wind farms an attractive power generation option.

LCOE for wind power is shown to reduce moderately over the projection period,

60
stabilizing from 2020 onwards. The LCOE also is shown to reduce from 153$/MWh

at 20% capacity factor to 70$/MWh at 50%.

Figure 2-4 Levelised Cost of Energy (LCOE) Generation and Capacity Factor for Wind Generators [17]

In 2015, Coonooer Bridge Wind Farm near Bendigo in Victoria, Australia supplied

electricity at the cost of 81.50 AUD / MWh, while Ararat Wind Farm (80 MW) agreed

to provide power for 87 $/MWh and Hornsdale Wind Farm (100 MW) agreed to

provide power at 92 $/MW. Power supply cost came out to be 143 $/MWh for a new

coal-fired power plant or 116 $/MWh when supplied from a new natural gas powered

station in 2013[18]. Bloomberg report listed the wind as cheaper electricity producing

the fuel than fossil fuels for Australia which is world’s biggest coal exporter. The

onshore wind generation costs fell in 2014 and, now it is approximate $71/MWh on

a global basis [19-20]. The prospective plunge in LCOE of wind is expected in the

near future due to the emergence of low cost but high technology wind turbine

manufacturer.

61
2.3.4. Low Carbon Footprint

Carbon footprint refers to emissions during the lifecycle of a project consisting of

stages of plant construction; fuel extraction, processing and their transport if

applicable; and plant ultimate decommissioning and disposal. Research data wind

energy projects having very low carbon footprints that too only during the

construction stage only. In comparison to approximately 500 and 900gCO2eq/kWh

for existing gas and coal plant respectively, wind carbon estimate is only

15gCO2eq/kWh [21]. As per CSIRO’s LCOE cost comparison study for the year

2020, the wind is the lowest cost technology with the price of 58.4 $/MWh as shown

in Figure 2-5 Without a carbon price the fossil fuel technologies are lower cost than

solar. With reduced carbon footprint, wind and solar based technologies are less

expensive than the fossil fuel based technologies.

Figure 2-5 LCOE Cost Comparison with Fuel Technologies [22]

62
2.4. Wind Energy and Australian Market Outlook

Figure 2-6 Per Capita Wind Capacity by Country and the Key US and Australian States/Territories [23]

By country, Australia currently ranks 11th in the world for wind generation per capita

ahead of countries like China and France. Denmark leads the world, followed by

Sweden, Spain, and Germany [23]. At a jurisdictional level, Australia’s wind

generation is heavily skewed towards states like South Australia and Tasmania,

which have some of the highest per capita wind generation in the world alongside

leading US states like Iowa and Texas as represented in Figure 2-6. When looking at

a jurisdictional level, Iowa has the highest per capita capacity of wind and solar, with

South Australia ranked second in the world. South Australia currently has around 39

percent of its total generation supplied by the wind and solar. The South Australian

renewables experiment is more significant given most other high renewables

penetration regions and countries - Iowa, Denmark, and Germany – are much more

integrated into larger grids with complementary (dispatchable) generation

technologies. The South Australian grid is partially constrained, connected to

Victoria by two transmission lines which allow it for sourcing a maximum of around

63
20 percent of peak load from Victoria. By contrast, Denmark has interconnections

that allow it for sourcing its entire peak load from other countries.

As per Australia Renewable Agency report [24], the share of wind energy in total

electricity generation is projected to increase by 12.1 percent of Australia’s electricity

generation in 2029–30 after stabilized RET. After hydro with 5% share, wind power

has the second largest share with 2.0% of Australia’s total primary energy

consumption. Among renewable energy sources, wind power had the fastest growth;

increasing on average by 67% per year since 2000 [24]. Australia’s first large-scale

grid-connected wind farm (at Crook well, New South Wales) in 1998 comprised eight

600 kW wind turbines each with a rotor diameter of 44 meters for a combined energy

output of 4.8 MW [24]. Today most onshore wind turbine generators have a capacity

of 1.5 to 2 MW; the largest wind turbines – designed for offshore sites – have a

capacity of 5 MW and rotor blades up to 60 m long (120 m rotor diameter).

Figure 2-7 Wind Generation Capacity Summary for Australia 2015

In 2015, Renewable energy provided 14.6 percent of Australia’s electricity, enough

to provide power for the equivalent of approximately 6.7 million average. The

amount of wind power in Australia has doubled in the past five years [25]. The

contribution of wind power grew by over 20 percent in 2015, while the demand for

64
power across the country increased slightly, the first rise following six years of

decline. The author represents the wind generation capacity summary for Australia

for 2015 in Figure 2-7 while the percentage of wind capacity by the state is shown in

Figure 2-8. In 2015, NEM had around 11802 GWh of wind power generation

accounting for 33.7% of renewable generation and 4.93% of total electricity

generation [25]. By the end of 2013, renewable energy projects constituted 48 percent

of total proposed new capacity in the investment pipeline and 58 percent of total

capital expenditure [20]. Wind-powered generation projects represent the greatest

share of planned capacity for renewable projects accounting for 42 per cent of total

planned capacity and 53 per cent of the value of proposed electricity generation

projects in the investment pipeline.

Figure 2-8 Percentage of Wind Capacity by State [Modified from 26]

Wind energy is projected to be the second fastest growing energy industry after

geothermal over the outlook period to 2029–30, reflecting the relatively low base of

which it is growing and the relative maturity of the technology compared with other

renewable energy sources. It is projected to overtake hydro electricity production

within the outlook period, to become the largest renewable source of electricity

generation in Australia [24]. At the end of 2015, there were 76 operational wind farms

65
in Australia, totaling 4,187 MW in capacity [25]. Approximately 500 MW of wind

capacity is under construction until May 2016, which will come online in 2017. 8057

MW wind capacity has been approved for construction while 6653 MW is in approval

stage [27].

Figure 2-9 South Australia Wind Power vs. Demand Variation on 22 May [28]

Among all states, South Australia owns the leading installed capacity of

approximately 1.5 GW representing highest penetrations in any electricity market.

Wind and solar are becoming new base load power after the closure of South

Australia largest coal-fired generator. The Wind has accounted for up to 49 per cent

of total generation in the South Australia state in May 2016. On 22 May 2016, wind

energy in South Australia hit 100% of demand and exceeded local demand for more

than 10 hours, from 1.40am to just before midday (11.55am), with a peak of 120

percent of demand at 4.30am as shown in Figure 2-9 [28]. In May 2016, the seven

large-scale wind farms in NSW delivered a record output, with a combined capacity

factor of 56 percent, more than two of the biggest coal generators in the state. The

Woodlawn wind farm produced a capacity factor of 62 per cent while only two of the

66
big five coal generators, the Mt Piper and Bayswater coal generators, produced at a

higher capacity factor (67 and 66 per cent respectively) [29].

2.5. Challenges to NEM Operation Due to Increased Wind

Penetration

Despite having sufficient feasibilities for wind power integration in Australian NEM,

it still lacks various technical and commercial aspects that require further

investigations to test the system adequacy for reliable and secure operation.

Intermittent wind power generation and asynchronous generators are two key

characteristics of wind generation technology affecting its integration in NEM grid

and spot market. AEMO does not currently identify NEM-wide challenges. Each

NEM region has a different generation mix, network configuration, and demand

characteristics, which lead to different challenges or different timing. However, the

future increment in wind power penetration could bring some adverse operational

challenges in the NEM. These challenges could be in the form of merit order effect

variation in pricing market, reduced system inertia, reduced interconnector

limitations and retarded network limitation.

2.5.1. Merit Order Effect and Reduced Spot Prices

By 2020, additional wind generation of around 8.88 GW is expected in NEM region.

As per AEMO’s National Transmission Network Development Plan (NTNDP), NEM

will have 11.5 GW of total installed wind generation capacity [30]. The Australian

Wind Energy Forecasting System (AWEFS) is used by AEMO to predict the power

from all the wind farms and has a normalized mean absolute error of less than 1.5%

in the 5-minute look ahead [30]. Currently, maximum wind generation is forecasted

for timeframes ranging from five minutes ahead to two years ahead to include wind

67
generation in the central dispatch process. From being considered as non-scheduled

generators, wind farms in Australia are now both scheduled and semi-scheduled and

included in forecast generation to allow generation output outside the central dispatch

process to be considered in the load–generation balance [31].

Merit order effect in NEM spot market infers that the marginal generator meeting

market demand sets the market clearing price or in simple terms refers to lowering of

wholesale electricity price by subsidized generators by adding reduced short run cost

generation to total supply or merit order. Wind farms bid their output in NEM as price

takers with low marginal cost. Wind farms get a connection to the grid through user

pays scheme while their access to the market is not always guaranteed [32]. Subjected

to the available wind and any network limitation, present wind farms in Australia

always runs at full capacity. Without storage options, wind farm operators typically

dispatch electricity into the market regardless of price. They are they incentivized by

the RET for every unit of electricity they produce. Under favorable wind conditions,

the wind farms output displaces output from coal or gas plants by bidding their low

marginal cost to clear the market, lowering spot prices through the merit order effect.

When the wind is not blowing, electricity can be sourced from hydro, coal, gas plants

or other renewable energy generators. Effect of wind farm effect on merit order of

generators is shown in Figure 2-10. As per a Deloitte study [33] on merit order effect

in South Australia, increased wind penetration in spot market results in backing off

the dispatch of fossil fuel based marginal generator that would have set the marginal

price. Thus, during periods of high wind, the output of marginal generators reduces

significantly resulting in reduced traded electricity which ultimately effects the

economics of generators in the form of reduced prices received for baseload and mid-

merit plant.

68
Figure 2-10 Merit Order Effect [33]

The actual impact of generation technology on prices is driven by a combination of

the demand profile, power supply availability, network topography, gas supply

dependence and carbon pricing. Wind farms quite often oversupply the market and

hence cause downward pressure on the wholesale electricity price as shown in Figure

2-11. During transmission system constraint in the form of generation oversupply,

available wind farm output needs curtailment as all generation cannot be placed on

the market.

Figure 2-11 Market Price Variations in South Australia on 3-4 Nov. 2014 When Wind Generation Enters

Spot Market [33]

69
2.5.2. Reduced Inertia and High Rate of Change of Frequency Issues

All synchronized rotating generators and motors constitute inertia of a power system.

Higher the system inertia, lesser is the frequency volatility due to disturbance. The

amount of a synchronous generation’s inertia depends on its size and design and is

expressed in megawatt seconds (MW.s). It is very difficult to maintain frequency

within acceptable limits for low inertia based power system as it will slow down or

speed up very quickly. Like demand level, power system inertia is also only an

observed characteristic, and currently, AEMO has no control over it in any form.

AEMO presently operates the power system around the requirements that arise from

the inertia levels that are present. With synchronous generators (especially thermal

power plant) providing the majority of energy in NEM, each NEM region had

sufficient inertia adequacy without any effect on system security needs, however

impending renewable energy targets has increased the probability of reduced power

system inertia due to the displacement of conventional synchronous generation,

particularly in South Australia and Tasmania. Current system inertia in South

Australia is around 18725 MW.s, but a low inertia value of around 2000 MW.s has

also been observed [34]. Although some non-synchronous generation, like wind, also

has rotating turbines, these technologies are increasingly connected to the power

system via power electronic converters, so the mechanical movement is decoupled

from the power system. By the year 2010, system inertia would be below acceptable

levels for 30-40% of the time in Tasmania, and 30% of the time in South Australia

[30]. Victoria also experiences low inertia sometimes but can rely on inertia from

other NEM regions due to its strong interconnections.

The initial rate of change of frequency (ROCOF) after generating unit MW loss ∆ )

is given as:

70

∗ 2.1

Or


∗ f 2.2

Where IR is system inertia, H is inertia of conventional generating unit and S is the

MVA rating of the conventional generating unit. ROCOF management is critical to

power system frequency regulation within the frequency operating standard [31] and

for maintaining the power system in a secure operating state. The amount of inertia

required to maintain an ROCOF under different contingency is proportional to the

contingency size. Lower inertia leads to a higher ROCOF than higher inertia system.

That means the frequency changes faster following a disturbance in a power system

with less synchronous generation, and this could result in the loss of additional

generation or load to arrest the frequency deviation when it occurs. Reduced system

inertia can challenge the effectiveness of existing frequency control mechanisms,

which can reduce under high ROCOF. The higher ROCOF will also require

stabilizing control systems to respond more rapidly to contain the change. For

example, for a contingency event resulting in an ROCOF of 1 Hertz per second

(Hz/s), the frequency drop from 50 Hz to 49.5 Hz (the limit of the FOS for credible

contingencies), would take 0.5 seconds. An ROCOF of 2 Hz/s would halve this time,

meaning action would be required within 0.25 seconds to prevent the system

frequency from breaching the FOS for a credible contingency event. High ROCOF

will lead to additional tripping for the same size imbalance. Relays and protection

schemes on generators and feeders have inherent delays and so may not respond

quickly enough to high ROCOF. Critical schemes such as under frequency load

shedding (UFLS) could become compromised in maintaining the frequency operating

standards (FOS). Increased wind penetration would make management of ROCOF

71
after contingency events more challenging. Currently, an inconsistency exists in the

NER requirement for generating units to remain connected through an event where

ROCOF reaches 1 Hz/s, but there is no system standard for power system operation

whereby the ROCOF during any contingency event is to be maintained at 1 Hz/s or

better. Low levels of inertia in the power system are likely to increase the frequency

deviations in the five-minute dispatch period that are controlled by regulation FCAS

in NEM.

2.5.3. FCAS Issues

Frequency control ancillary services (FCAS) in power system is utilized to satisfy

load demand through generation supply. Frequency control ancillary services (FCAS)

is indispensable for the secure operation of large interconnected power system like

NEM. FCAS market also complements AEMO's centrally dispatched energy market.

FCAS services are purchased from market participants by AEMO for adequate power

system frequency control. These services include regulating FCAS services for

normal operating conditions and fast responding contingency FCAS services

following any contingency events. 130 MW is procured for raise FCAS services

while 120 MW is procured for FCAS lower services within a 5-minute dispatch

interval. NEM is managed with this amount of variability between 5-minute

dispatches of the rest of the generators. An accumulated time error of greater than ±

1.5 seconds may require additional support. The current calculation of NEM global

contingency FCAS requirements is determined by the dispatch algorithm and

considers both the size of the largest contingency and the power system demand. All

types of large FCAS requirement is calculated as FCAS requirement= Contingency

risk (MW change due to the generator or load loss) – Load relief (demand change due

to frequency deviation). Load relief factor is 1.5% for the mainland in NEM while it
72
is 1% for Tasmania [35]. The level of fast performing response capabilities or

ancillary service requirements necessary to manage frequency in NEM is determined

by the frequency operating standards. As per NEM frequency operating standards

[35] for normal system operation, the frequency must be maintained within the

normal operating frequency band (49.85 Hz to 50.15 Hz in both Tasmania and the

NEM mainland) for no less than 99 percent of the time. In the case of a credible

contingency event, the system frequency should not exceed the normal operating

frequency excursion band for more than five minutes on any occasion. Following

either a separation or multiple contingency events, the system frequency should not

exceed the normal operating frequency excursion band for more than ten minutes. In

the case of time delay for FCAS activation, synchronous generators inertia is the first

to slow down ROCOF and depending upon contingency sizes, even able to limit

frequency within specified band till FCAS are employed. In a power system with a

greater proportion of variable, non-scheduled generation like wind, operating the

power system over the five-minute cycle would require a greater reliance on

regulation FCAS. In a lower inertia based power system having high ROCOF,

frequency deviations will reach the threshold of the contingency frequency range

more quickly, reducing the required operational response time of stabilizing control

systems, and potentially increase the level of response services required to return to

normal operating conditions [31]. Presently, AEMO always assumes mainland inertia

levels to be high enough not to affect the calculation of mainland Frequency Control

Ancillary Service (FCAS) requirements and ROCOF management. Inertia levels

within Tasmania are much lower than the mainland, so managing Tasmanian

frequency is more technically challenging under the effects of increasing wind

generation. Historically synchronous generations have provided regulation and

73
contingency FCAS in NEM, but with the exit of base load dispatchable generators

from islanding prone South Australia region, FCAS availability at the local level will

be more challenging with increasing wind penetration. FCAS regulation service

required capacity in NEM will increase by approximately 20% by 2020 due to

increased renewable generation [36] which is likely to increase the magnitude of a

minute to minute generation deviations. In the case of insufficient FCAS availability,

NEM frequency maintenance within the required standards will be difficult for

AEMO, and NEM may collapse under big contingency events at worst scenario.

AEMO must review the procurement of Regulation ancillary services, especially in

smaller systems with high wind penetration like in, Tasmania and South Australia to

ensure ongoing control of power system frequency within required limits.

2.5.4. Network Infrastructure Issues

As per NEM rules, connection costs are part of any new generator including wind

farm construction cost while customers fund shared transmission network cost. The

cost of shared transmission system is indirectly affected by supply to the customer

who is dependent upon location and type of generators. Attributing additional shared

transmission cost resulting from increased wind integration in NEM is challenging

for AEMO regarding market benefit economy. Most of the wind farms in Australia

are in remote locations that will require significant transmission investment for

improved transmission network infrastructure otherwise around 35% and 15% of the

wind energy in Victoria and South Australia respectively could be curtailed due to

network limitations [37]. Therefore, balancing wind penetration around the NEM

necessitates robust interconnections to yield geographical diversity benefits. Network

augmentation costs may be deferred by correlating output generation of small wind

farms connected to existing distribution networks with local demand thereby

74
reducing network losses. The cost of connecting a 5 MW wind farm to the NSW 330

kV grid could fall from $2,500/kW to $150/kW to connect four 50MW wind farms

sharing the same substation [38]. This highlights the possibilities for connection costs

reduction for remote wind farms by capturing economies of scale in network

augmentation.

2.5.5. Interconnectors Performance Issues

Power import- export in NEM interconnected regions through interconnectors are

limited by transient or voltage stability due to the contingent trip of the largest

generating unit and potential thermal over-loadings. South Australia’s transmission

network is connected to the rest NEM via the Murraylink (DC link of 220 MW

transfer capability) and Heywood Interconnectors (AC-link upgraded from 460 to

600 MW). The Heywood Interconnector is the only link that provides a synchronous

connection between SA and the rest of the NEM. DC transmission line Basslink

which connects Tasmania to mainland Australia is unable to transfer FCAS between

the mainland and Tasmania under certain conditions most notably when it reaches

the maximum export limit or maximum import limit or when it is transferring power

within a ‘dead-zone’ between -50MW to 50MW. Increased wind generation in states

like South Australia and Tasmania has increased energy imports/export through

interconnectors putting more thermal stress, especially on Heywood interconnectors

during the period of high demand and low wind generation and vice versa. Significant

new wind generation can reduce existing interconnector transfer limits, particularly

under conditions of low demand and high wind speeds when wind generation forms

a large percentage of the generation mix [37]. Any contingency at interconnector

level may have high contingency impact on NEM spot market and FCAS market

operation due to high ROCOF. A combined effect of spot market price variation and

75
interconnectors dependence was observed recently during a contingency event which

occurred on 1st November 2015 when South Australia was islanded for 26 minutes

due to a transmission line tripping and Heywood interconnector being unavailable

due to upgrading works. Sudden fall of interconnector capacity resulted in FCAS

procurement within South Australian region resulting in price hikes.

76
2.6. Wind Integrated Network Frequency Response Assessment

In this section, an augmented form of 14-Gen NEM model is used as the test network

to study wind integration effect on frequency response performance in area 5 which

can be held representative of South Australia (SA). MATLAB’s Simscape has been

used as the simulation software [40]. Several case studies were performed to analyze

the technical problems due to wind farm integration in low load scenario and high

load scenario under low levels of conventional synchronous generation in operation.

Following the frequency response analysis under low and high load scenario, the

solution to improve the frequency response is assessed based on frequency controller

based wind farm and synchronous condensers. One of the key findings is that there

are minimum issues with the SA power system security management if SA is

connected to the remainder of the NEM via the Heywood Interconnector and

sufficient synchronous generation is available in the SA power system. Increased

wind penetration with synchronous generator losses is making frequency operating

standards management difficult due to increasing ROCOF and higher frequency

nadir.

2.6.1. Test Network

A 14-generator NEM model is a simplified model of the eastern and southern 50 Hz

Australian electrical networks, which was originally proposed for small-signal

stability studies [39]. The schematic diagram of the original 14-generator model of

the NEM is shown in Figure 2-12. Areas 1 to 5 represents Tasmania, New South

Wales (NSW), Victoria (VIC), Queensland (QLD) and South Australia (SA),

respectively. In the original model, there are 14 generators, 5 Static VAR

Compensators (SVCs), 59 buses and 104 lines with voltage levels ranging from 15kV

to 500kV. The excitation system (AVR) and power system stabilizer (PSS) of

77
generators are adopted from reference [39]. Also, it is assumed that all thermal and

hydro power plants have a standard steam turbine governor (i.e. IEEEG1) and hydro

turbine governor (i.e. HYGOV), respectively. Test network is studied when DFIG

based wind farms are integrated into Area 5. Area 5 has three lumped generators;

G503 at bus 503, G502 at bus 502 and G501 at bus 501. Generator 503 and 502 are

connected at 15 KV while generator 501 is connected at 20 KV. All related data are

provided in the appendix.

2.6.2. Low Load Scenario

Table 2-2 Scenario 1: Low Load

Summary for low load Converged: The load flow converged in 4 iterations

P (MW) Q (MVAR)

Total generation 15116.22 -569.215

Total PQ load 14807 1595

Total Z shunt 73.06775 -3313.59

Total ASM 0 0

Total losses 236.1522 1149.37

2.6.3. Assumptions Used in simulations

 DFIG based wind turbines are used in wind farm model.

 Aggregated generating machines including wind turbines inside a wind farm are

used.

 Test Network is modeled under light load scenario as given in Table 2-2 and

high load scenario which is given in Table 2-13.

 Generator and interconnector contingency simulations are conducted for area 5

which are held representative of South-Australia.

78
Figure 2-12 14-Generator NEM Model

79
Table 2-3 Load Flow in Low Load Scenario

Power Station/Bus No. of Units P(MW) Q(MVAR)


B101 2 0.001 -208.239
B201 4 2160 -141.833
B202 3 1380 -33.9972
B203 3 1410 30.82429
B204 4 1675.219 -141.814
B301 6 3330 84.09879
B302 2 760 50.08303
B401 3 960 -107.759
B402 2 580 -10.1451
B403 3 960 38.08328
B404 3 651 -20.0589
B501 2 560 -115.507
B502 3 540 19.93927
B503 1 150 -12.8902
Total 41 15116.22 -569.215

Case 1: Generator Contingency Analysis (Low Load Scenario)


Table 2-4 Area Frequency Response Indicators When Generator 503 (150 MW) is Isolated at 10s and Wind
Farm Connected at Bus 509

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

502 -0.5 49.90

501 -0.52 49.93

ROCOF changes from -0.5 to 0.4 Hz/s in 7 seconds soon after


isolation at G502 terminal. For G501, ROCOF changes as -0.52
to 0.15 in 5 seconds.

302 -0.12 49.99 100 MW export


changes to 230 MW
import from area 3-5

301 -0.12 49.98

ROCOF changes from -0.12 to 0.07 Hz/s in 1.2 seconds soon


after isolation at G503 terminal. For G301, ROCOF changes as
-0.125 to 0.07 in 1.2 seconds.

80
Wind farm integration effect on frequency response indicators namely, Rate of

Change of Frequency (ROCOF) (Hz/s) and frequency Nadir at Area 5 generator

terminals is investigated in this study. In various simulation studies, frequency

response indicators are observed at Area 5 and Area 3 generators when one of the

generators are isolated from the network, and equivalent rated DFIG based wind

farms are integrated into area 5. The simulations were performed in MATLAB, and

ROCOF results were determined by taking the derivative of the measured frequency

at generators terminal in the network. Values of ROCOF as observed under different

contingency are indicated in Tables (2-4 – 2-9).

Table 2-5 Area Frequency Response Indicators When Generator 502 (540 MW) is Isolated at 10s and Wind
Farm Connected at Bus 508

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

503 -2.14 49.68

501 -2.75 49.47

ROCOF changes from -2.14 to 2.28 Hz/s in 1.7 seconds soon


after isolation at G503 terminal. For G501, ROCOF changes as
-2.75 to 2.33 in 0.6 seconds.

185 MW import
302 -0.43 49.91
changes to 330 MW
export from area 3-5
301 -0.43 49.99

ROCOF changes from +0.62 to -0.43 Hz/s in 0.75 seconds soon


after isolation at G502 terminal. For G301, ROCOF changes as
-0.437 to 0.37 in 7 seconds.

81
Table 2-6 Area Frequency Response Indicators When Generator 501 (560 MW) is Isolated at 10s and Wind
Farm Connected at Bus 508

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

503 -2.58 49.70

502 -2.73 49.54

ROCOF changes from -2.58 to 0.94 Hz/s in 9 seconds soon


after isolation at G503 terminal. For G502, ROCOF changes as
-2.73 to 1.29 in 0.6 seconds.

200 MW import
302 -0.28 49.90 changes to 350 MW
export from area 3-5

301 -0.3 49.90

ROCOF changes from -0.286 to 0.5 Hz/s in 1.3 seconds soon


after isolation at G501 terminal. For G301, ROCOF changes as
-0.3 to 0.25 in 1 seconds.

Table 2-7 Area Frequency Response Indicators When Generator 503 (150 MW) is Isolated at the 20s, and
Generator 501 (560 MW) is Isolated at 10s, 1146MW Wind Farm Connected at Bus 508

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

502 -2.68 49.61 Export of 500 MW to

50 MW after 1st

301 -0.37 49.97 contingency and then

75 MW after 2nd
302 -0.31 49.95 contingency.

For G502, ROCOF changes from -2.68 to 1.2 Hz/s in 0.6 seconds soon after 1st contingency.

ROCOF changes as -0.85 to 0.4 in 0.6 seconds after 2nd contingency. For G302, ROCOF

changes from -0.3 to 0.2 Hz/s in 0.9 seconds while for G301, ROCOF changes as -0.3 to 0.2

in 0.7

82
Case 2: Generator Contingency and Area 5 Separation Due to Interconnector

Contingency Analysis (Low Load Scenario)

The study investigated the separation of area 5 from rest of NEM due to

interconnector loss. The study considers the frequency stability of area 5 under the

condition of reduced inertia. Frequency response indicators are observed at area 5

generators when one of the generator and area 3-5 interconnector is isolated from the

network, and equivalent rated DFIG based wind farms are integrated into area 5.

Higher values of ROCOF more than 1 Hz/s is possible in case of big contingency. It

is therefore proposed that a higher ROCOF specification is defined for South

Australia which is reflected as area 5 in the current model. Proposed solution should,

of course, be preceded by a study involving the effect of changing ROCOF values on

generating unit characteristics.

Table 2-8 Area Frequency Response Indicators When Generator 503 (150 MW) is Isolated at 10s and area
3-5 interconnector is Isolated at 20 s; Wind Farm Connected at Bus 508

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

501 -0.68 49.27

502 -0.65 49.28

G502 terminal ROCOF changes from -0.3 to 0.05 Hz/s in 3


seconds soon after G503 isolation, then -0.65 to 0.18 Hz/s in 3 s
after interconnector isolation. For G501, ROCOF changes as -
0.33 to 0.05 in 3 s at first isolation and it changes from -0.685 to
0.18 Hz/s in 3 s. 100 MW export from
area 3-5 till
interconnector is
302 -0.28 50.04 isolated.

301 -0.3 50.04

ROCOF changes from +0.125 to -0.05 Hz/s in 0.8 seconds. For


G301, ROCOF changes as +0.125 to -0.03 Hz/s in 1.1 seconds.

83
Table 2-9 Area Frequency Response Indicators When Generator 502 (540 MW) is Isolated at 10s and area
3-5 Interconnector is Isolated at 20 s, 1080 MW Wind Farm Connected at Bus 508

Frequency Nadir
Generator ROCOF (Hz/s) Power Flow
(Hz)

503 -2.18 49.74

501 -2.76 49.5

G503 terminal ROCOF changes from -2.18 to 2.12 Hz/s in 1.7


seconds. For G501, ROCOF changes as -2.76 to 2.6 Hz/s in 0.6 s Initial import of 500
MW, changes to 100

302 -0.5 49.95 MW export and


zero after
interconnector loss.
301 -0.6 49.92

ROCOF changes from +0.6 to -0.5 Hz/s in 0.6 seconds. For


G301, ROCOF changes as +0.5 to -0.6 Hz/s in 0.6 seconds.

2.6.4. Simulation Discussions

The frequency operating standards (FOS) requires the frequency to be maintained

within the range 49.85 to 50.15 Hz for 99% of the time. This standard is valid for

both the mainland regions and the Tasmania region during periods of no contingency

events and no load events. Larger deviations are permitted within the range 49.75 to

50.25 Hz for no more than 1% of the time. For an island within the mainland regions

or an island within the Tasmania power system, the frequency operating standards

require that, during periods when there are no contingency events and no load events,

the frequencies be maintained within the ranges 49.5 to 50.5 Hz and 49.0 and 51.0

Hz respectively.

84
Figure 2-13 Area 5 Generator Terminal Frequency after Isolation of Gen 503 at 10 S

Figure 2-14 shows the frequency observed at three generator terminals when

generator 503 of 150 MW is disconnected at 10 s. An equivalently rated wind farm

is connected to the network. In the case of 150 MW contingency and 9.37% wind

penetration in area 5, frequency nadir observed at generator 502 (540 MW) is more

than generator 501 (560 MW), and frequency at both generators terminal is within

NEM normal frequency range of 49.75-50.25 Hz. With low generator contingency

and interconnector contingency, NEM frequency standard is violated as frequency

goes below 49.5 Hz as shown in Figure 2-15. These results match the South

Australian Jurisdiction which has recommended AEMO for permitting larger

frequency variations (47-52Hz) as frequency can go beyond 49.5 Hz – 50.5 Hz for

the credible loss of the Heywood Interconnector. Frequencies observed at area 5

buses is shown in Figure 2-16. Even though wind farm role is not clear in these

scenarios, frequency at bus bar 508 is observed within limits where 150 MW wind

farm and 490 MW of load is connected.

85
Figure 2-14 Area 5 Generator Terminal Frequency after Isolation of Gen 503 at 10 S and Interconnector

Isolation at 20 s

Figure 2-15 Area 5 Bus Frequency after Isolation of Gen 503 at 10 s.

Figure 2-16 Area 5 Bus frequency after Isolation of Gen 502 at 10 s.

86
Figure 2-17 ROCOF Observed at Generator 503 when Generator 502 is Isolated at 10 s

Figure 2-17 shows the frequency observed at area 5 generators when generator 502

with all units amounting to 540 MW is isolated from area 5, and an equivalent rated

wind farm (540 MW) is also integrated into the area. Other highest rated generator

has frequency drop up to 49.47 Hz for the very small duration, and then it gradually

settles around 49.9 Hz within 10 s. This simulation result again supports the call for

frequency regulation standard modification in regions like South Australia and

Tasmania. It should be noted that these simulation results have been obtained without

implementing secondary control in the area. In all the contingencies above 500 MW,

we observed ROCOF also breaching the standard of 1 Hz/s for around 1-2 s. An

example of ROCOF variation is shown in Figure 2-18. This result again supports the

call for ROCOF standard revision in some of the NEM regions.

Table 2-10 Simulation Results for Low Load Scenario for Area 5 Generator 501

Generator % WIND- ROCOF FNADIR MW LOSS (Gen.+ ACTIVE


Penetration (Hz/s) (Hz) Interconnector) SYN UNIT

G501 0 -0.001 49.91 150 5


G501 9.375 -0.685 49.27 350 5
G501 9.375 -0.52 49.93 150 5
G501 27.135 -2.75 49.47 540 3
G501 42.68 -2.76 49.5 740 3

87
Table 2-11 Simulation Results for Low Load Scenario for Area 5 Generator 502

Generator % WIND- ROCOF FNADIR MW LOSS (Gen.+ ACTIVE


Penetration (Hz/s) (Hz) Interconnector) SYN
UNIT

G502 9.375 -0.65 49.28 350 5

G502 9.375 -0.5 49.9 150 5

G502 27.86 -2.73 49.68 560 4

G502 32.87 -2.75 49.56 710 3

G502 44.14 -2.68 49.61 710 3

Figure 2-19 – 2-25 below shows the polynomial regression surface model for values

of ROCOF, Frequency Nadir Point (FNadir) variation. Table 2-10 and Table 2-11 list

their values for area 5 generators (G503, G502, G501) terminals under various

contingency conditions. We can notice a directly proportional relationship between

ROCOF, FNadir and Wind Penetration as well as between ROCOF, FNadir and MW

Loss. ROCOF and lowest FNadir increase with increasing amount of wind farm in Area

5 and increasing amount of power (MW) Loss. An inverse proportional relationship

is observed with a number of the active synchronous generator in area 5 with ROCOF

and FNadir point. Through simulations, it is observed that ROCOF in area 5 is

maintainable within 1 Hz/s range up to 400 MW generator contingency and increases

sharply with higher MW loss. Duration of this increase in ROCOF is also observed

to be within 2-second range. Increased ROCOF settles to lower ranges as indicated

in tables above. In the Australian NER, no standard is set for a maximum level of

ROCOF on the power system. Generation, on the other hand, is required by their

Access Standards to remain connected through an event where ROCOF reaches ±1

Hz/s. According to new connection standards in NEM, 1 Hz/s is minimum access

standard for ROCOF and 4 Hz/s for 0.25s is automatic access standard for ROCOF

88
where the plant must be capable of continuous, uninterrupted operation [35]. NEM

needs to have clear ROCOF standard for the correct operation of emergency

protection frequency relay that manages multiple contingency events should be based

on maximum ROCOF.

2.6.5. Linear Regression Model from Frequency Disturbances

Wind penetration level (WP) can be defined as ratio of total wind generated power

P by total generated power which includes synchronous generator power P ,

inter-area power flow P and wind power.

WP 2.3

By investigating the development of frequency deviations during a disturbance and

relating the change in frequency to other factors, a relationship can be found. To

generate a linear regression model for wind penetration level, MW loss and number

of the active synchronous generator with respect to ROCOF and frequency nadir,

following steps have been implemented:

1. Set the load level.

2. Maximum wind penetration level is estimated with direct replacement method

where wind power plant has added the equivalent of lowest synchronous

generator rating in area 5.

3. Trip the synchronous generator of lowest rating to initiate generation led

disturbance.

4. Find frequency nadir point and ROCOF value after disturbance from

simulation results. Though ROCOF can be obtained from theoretical equation

2.1 or 2.2, values expressed here are obtained as simulated df/ft soon after

disturbance.

89
5. Step 1-3 are repeated with increasing generation tripping and increased wind

penetration level. MW loss is obtained as generation tripping as well as area

5 to area 3 interconnector tripping.

6. Finally, a mathematical regression model between wind penetration (%) and

ROCOF, FNadir or MW loss and ROCOF, FNadir or Number of active

synchronous units and ROCOF, FNadir is formed.

7. If F F and ROCOF ROCOF ,

Wind penetration level in control area can be increased till F

F and ROCOF ROCOF . Then maximum

wind penetration level is achieved for a control area.

Linear polynomial regression model for variation between wind penetration, MW

loss, ROCOF, Frequency nadir observed for area 5 at generator 501 and 502 is given

below and coefficients given in Table 2-12:

F (x, y) = p00 + p10*FNadir + p01*ROCOF 2.4

Coefficients (with 95% confidence bounds):

Table 2-12 Values for Linear Polynomial Regression Model

Figure F (x, y) Generator p00 p10 p01 R-square

No.

2.19 Wind 501 -110.3 2.241 -12.54 0.8949

Penetration

level (%)

2.20 MW-Loss 501 1.022e+04 -202.4 -158.3 0.9083

2.21 Wind 502 -40.42 0.8702 -11.8 0.8348

Penetration

level (%)

90
2.22 MW Loss 502 1.562e+04 -312.1 -194.3 0.9543

2.23 No. of 502 -1.125 0.1323 0.7738 0.8283

active syn.

units

Figure 2-18 Polynomial Surface View for Wind Penetration Variation with Frequency Nadir Point

and ROCOF as Observed at Generator 501

Figure 2-19 Polynomial Surface View for MW-Loss Variation with Frequency Nadir Point and ROCOF

as Observed at Generator 501

91
Figure 2-20 Linear Polynomial Regression Surface View for Wind Penetration Variation with Frequency
Nadir Point and ROCOF as Observed at Generator 502

Figure 2-21 Linear Interpolant Surface View for MW-Loss Variation with Frequency Nadir Point and
ROCOF as Observed at Generator 502

Figure 2-22 Polynomial Regression Surface View for Online Synchronous Generator Units Variation

with Frequency Nadir Point and ROCOF as Observed at Generator 502

92
Figure 2-23 Linear Polynomial Surface View for Online Synchronous Generator Unit Variation with

Frequency Nadir Point and ROCOF as Observed at Generator 501

Figure 2-24 Linear Polynomial Surface View for Wind Penetration with Frequency Nadir Point and

ROCOF as Observed at Generator 502

2.6.6. High Load Scenario

Test network is studied when DFIG based wind farms are integrated into Area 5. Area

5 has three lumped generators; G503 at bus 503, G502 at bus 502 and G501 at bus

501. Generator 503 and 502 are connected at 15 KV while generator 501 is connected

at 20 KV. Wind farms are connected to bus 506 at 275 KV and bus 508 at 275 KV in

different simulation studies. 1.5 MW, DFIG based is connected at 575 v which is

further stepped up to 275 KV through two transformers. Summary of converged load

flow is given in Table 2-13.


93
Table 2-13 Scenario 1: Heavy Load

Summary for high load Scenario Converged: The load flow converged in 5 iterations!

P (MW) Q (MVAR)

Total generation 23035.66 3795.05

Total PQ load 22300.1 2556.5

Total Z shunt 92.17862 -84791.8

Total ASM 0 0

Total losses 643.3862 86030.31

All system data is described in the [39], and it is available electronically. All the data

will not be reproduced here in this study, but it is important to note that a few changes

have been introduced especially under high load scenario. The number of active

synchronous generators at each plant is different in both low load and high load

scenario. Some modifications in terms of parallel transmission line additions,

additional equivalent MVAR using load blocks and transformers addition were done

in low load scenario model for load flow convergence under high load scenario.

Modifications were done mainly in following bus numbers as given in Table 2-14.

Due to network complexity and longer simulation time, only three simulation test

were conducted for high load scenario.

Table 2-14 Modified Bus for High Load Scenario

B413 B412 B209 B213 B507 B509 B310 B314 B313

Table 2-15 summarizes the generation output in high load scenario, indicating the

active and reactive power output of an individual generation unit and the number of

units in service at each power plant.

94
Table 2-15 Generation Output in High Load Scenario

Power Station/Bus No. of Units P (MW) Q (MVAR)


B101 4 322.7648 161.7217
B201 6 3600 474.2422
B202 5 2500 630.3434
B203 4 1500 501.5524
B204 6 2950.2 672.6861
B301 7 4200 302.7628
B302 3 939.9 -87.7698
B401 4 1400 554.9564
B403 4 1400 212.9104
B404 6 1549.8 335.254
B501 2 600 -52.8549
B502 4 800 -47.5857
B503 4 436 -42.8808
B402 3 837 179.7116
Total 62 23035.66 3795.05

Case 1: Generator-Interconnector Contingency Analysis (High Load Scenario)

The frequency response of a big interconnected transmission system to a disturbance

caused by the loss of connected generation or load depends primarily on the inertial

response of the system, load damping and the action of governors subjected to

frequency errors. However, each of these elements is in turn dependent upon many

other factors and higher order responses. A detailed and complete analysis of the

frequency behavior of a system requires complex and time-consuming modeling of

each element of the system. In the current study, wind farm integration effect on

frequency response indicators namely, Rate of Change of Frequency (ROCOF) (Hz/s)

and frequency Nadir at area 5 generator terminals is analyzed. In the first simulation,

DFIG based wind farm is added, and frequency response indicators are observed at

area 5 generators. In the second simulation, generator G503 is isolated from the

network at 10 seconds, and the equivalent wind farm is present in the network.

ROCOF and Frequency Nadir as observed are given in Table 2-16 – 2-18 below:

95
Table 2-16 Area Frequency Response Indicators When Wind Farm is Added (109x4 MW)

Generator ROCOF (Hz/s) Frequency Nadir (Hz)

503 -1.1 49.97

502 -1.75 49.956

501 -1.64 49.95

302 -0.45 49.99

301 0.4 49.99

Power flow varies from 500 MW and settles around 420 M



Table 2-17 Area Frequency Response Indicators When Gen 503 is Isolated at 10 s and Wind Farm Is Added
(436 MW)

Generator ROCOF (Hz/s) Frequency Nadir (Hz) Power Flow

501 -1.15 49.93

502 -1.3 49.92 Changes from 500

MW to 750 MW
301 -0.11 50.023

302 -0.016 50.026

G502 terminal ROCOF changes from -1.3 to 0.56 Hz/s in 0.6 seconds soon after G503 isolation.

For G501, ROCOF changes as -1.15 to 0.6 Hz/s in 0.55 s. G301 terminal ROCOF changes from

+0.1 to -0.16 Hz/s in 0.5 seconds soon after G503 isolation. For G301, ROCOF changes as +0.1 to

-0.113 Hz/s in 0.7 s.

An improvement in ROCOF values is observed in Table 2-17 after G503 isolation.

This improvement is due to increased power export from area 3 to area 5 from 500

MW to 700 MW while frequency nadir is decreased. We also notice a drop in ROCOF

values at area 3 generator terminals. ROCOF changed from 0.1 Hz/s to -0.1 Hz/s at

96
G302 and G301 both. Lowest frequency point also moved from 49.99 to 50.02 for

both the generator. Similar results are observed in Table 2-18 when interconnector is

also isolated. Generator loss has direct effect on ROCOF changes which are

prominent during high load demand and integration of equivalent wind farm does not

present any improvement in ROCOF values. Test network being a complex

interconnected network a clear conclusion in frequency response behavior due to the

wind farm is difficult and need more investigations.

Table 2-18 Area 5 Frequency Response Indicators When Gen 503 is Isolated at 5 s, and Interconnector is
Isolated at 10 s, Wind Farm Added (436 MW)

Generator ROCOF (Hz/s) Frequency Nadir (Hz) Power Flow

502 -3.37 46.33


Increases to 750 MW
501 -2.85 45.94
after first isolation

302 -0.37 49.99 (at 5 s) and goes zero at


interconnector loss
301 -0.3 50.02 (after 10s)

G502 terminal ROCOF changes from -1.17 to 0.57 Hz/s in 0.6 seconds soon after G503 isolation.
At second contingency of interconnector loss, ROCOF changes as -3.37 to 0.14 Hz/s in 0.9 s. G501
terminal ROCOF changes from -1.09 to 0.57 Hz/s in 0.6 seconds soon after G503 isolation. At
second contingency of interconnector loss, ROCOF changes as -2.85 to 0.25 Hz/s in 0.9 s. G302
terminal ROCOF changes from 0.1 to -0.16 Hz/s in 5 seconds soon after G503 isolation. At second
contingency of interconnector loss, ROCOF changes as 0.4 to -0.37 Hz/s after 10 s.

2.7. Possible Frequency Control Solutions

Currently, AEMO aims to potentially manage the power system impacts with using

constraint equations in the central dispatch process to limit wind generation, or as a

last resort by market intervention through issuing directions to synchronous

generators to ensure sufficient levels of power system inertia are maintained to allow

adequate control of power system frequency. Establishment of improved ancillary

service requirements and control schemes for allowing adequate control of power

97
system frequency under conditions of low power system inertia is another action plan.

Besides these short-term action plans, AEMO intends to improve wind generation

modeling in its operational and planning tools and processes, to improve their

accuracy with increasing levels of wind generation. AEMO is investigating the extent

to which fast frequency response (FFR) could provide a substitute for synchronous

inertia.

Table 2-19 Anticipated Technical Solutions for Frequency Improvement in Wind Penetrated System

Technical Solution High ROCOF Insufficient FCAS Inertia Improvement

Synchronous √ √

Condenser [41]

Frequency responsive √ √ Needs field level testing

Wind Generation for technical feasibility

[42] [43]

Batteries [44] √ √ Needs field level testing

for technical feasibility

Frequency Responsive √ √ Needs field level testing

Photo Voltaic [45] for technical feasibility

Demand Response √ √

[46] [47]

Change in ROCOF Needs field level

protection setting [34] testing for technical

feasibility

New Interconnectors Partial

[34]

Increase Frequency Uncertainty in Uncertainty in

Deviation Limit [34] technical feasibility technical feasibility

Optimize FCAS Partial. Not fully

Requirements [34] address challenge

98
Some of the potential technical solutions and their reach corresponding to technical

frequency regulation challenges are listed in Table 2-19, which need advanced field

level testing and cost-benefit analysis. AEMO is assessing the capabilities and

limitations of technologies that can provide an FFR service, and whether FFR

provides an adequate substitute or supplement to inertia. The scale and type of

response required to make a useful contribution to the power system are still

unknown. There is a considerable body of work to be done to fully understand the

dynamic performance of the Australian power system as conventional synchronous

plants are gradually displaced by low inertia generating plants. Out of the many

possible frequency control solutions, two of them are briefly presented below:

 Synchronous Condensers

The synchronous condenser has been used as a conventional solution for reactive

power regulation, but they are losing value due to growth in power electronics based

reactive power compensation scheme. However, during the global trend towards

renewable energies, the synchronous condenser has been experiencing a renaissance

since last five years as frequency regulation solution. Increasing renewable sources

tends to lower the system inertia thereby introducing frequency regulation problem.

Renewable technology like wind energy systems has very low inertia comparing to

the conventional generating system, so there is a possibility of utilizing retiring

conventional generating units as synchronous condensers. The conversion of non-

profitable or deactivated power station is currently seen as the most cost-effective

option applying synchronous condensers.

Synchronous condensers are synchronous electrical machines attached to the

electricity grid. The machine will be brought up to speed with an electrical motor

attached or via a frequency converter. When the machine is synchronized with the

99
electricity grid, it will act as a motor, turned by the energy taken from the grid.

Because of the nature of the synchronous machine, reactive power can be consumed

and generated by controlling the excitation of the rotor. The generator of the

synchronous condenser with its rotating mass also always acts against a variation of

network frequency, thereby acting as support for stable system frequency. As per

KEMA report, only 1 % – 4 % of the nominal power rating is consumed during inertia

provision of approximately 1 sec. Synchronous condensers can respond immediately

if on otherwise takes less 15 min startup [28]. A synchronous condenser has the

benefit of providing a synchronous inertia response to support ROCOF management

capabilities and providing fault level and voltage support services beneficial to the

power system.

The impact of synchronous condenser effect on the frequency response of 14-

generator NEM model’s area 5 under contingency events was investigated.

Synchronous generator with the same number of units as that of isolated generating

plants is connected to the 275 KV high-voltage transmission network via a step-up

transformer. In practice, the startup and braking system of a Siemens synchronous

condenser is either based on a frequency-controlled pony motor or a starting

frequency converter, the simulation model has turbine governor system removed, and

voltage is controlled through AC4 excitation system. In the first simulation, generator

503 is isolated at 10 s and generator 501 isolated at 20 s from area 5. Total wind farm

penetration in the area is taken as 44% with 585 MW DFIG based wind farm

connected to bus 509 and 561 MW DFIG based wind farm connected to bus 507. In

the second simulation, synchronous condensers of equivalent ratings of generator 503

and generator 501 are introduced in the test network. First, only one synchronous

100
condenser connected at bus 508 is activated to analyze frequency response. Second

simulation test has both synchronous condensers activated in the network.

Table 2-20 Area 5 Frequency Response Indicators during Generator Contingency (Without Synchronous
Condenser)

Active FNadir (Hz) ROCOF (Hz/s) Power Flow


Generator

G502 49.61 -2.68 to 1.2 in 0.6 sec (at 1st Power flow changes from an
Gen. tripping). export of 500 MW to 50 MW after
first generator tripping and then
-0.85 to 0.4 in 0.6 sec. (At 2nd
the import of 75MW after second
Gen. tripping)
generator tripping.

Table 2-21 Area 5 Frequency Response Indicators during Generator Contingency (With One Synchronous
Condenser)

Active
Generator
FNadir (Hz) ROCOF (Hz/s) Power Flow

G502 Power flow changes from an


-2.35 to 0.875 in 0.6 sec
export of 500 MW to 50 MW
& (at 1st Gen. tripping)
49.67 after first generator tripping and
-0.75 to 0.24 in 0.6 sec.
Synchronous then the import of 70MW after
(At 2nd Gen. tripping)
Condenser second generator tripping.

Table 2-22 Area 5 Frequency Response Indicators during Generator Contingency (With Both Synchronous
Condensers)

Active FNadir (Hz) ROCOF (Hz/s) Power Flow


Generator

G502 Power flow changes from an


-1.68 to 0.5 in 0.8 sec (at
export of 500 MW to 50 MW
& 1st Gen. tripping)
49.72 after first generator tripping and
Synchronous -0.5 to 0.14 in 1.1 sec.
then the import of 65 MW after
Condenser (At 2nd Gen. tripping)
second generator tripping.

101
Table 2-20 – 2-22 gave the frequency response indicator values under these two

synchronous condensers application effect and represented in Figure 2-25 – 2-26. It

is observed that with ROCOF improves from -2.68 Hz/s to -1.68 Hz/s with

synchronous condensers rated equivalent to contingency MW loss.

‐0.5
ROCOF (Hz/s)

‐1

‐1.5

‐2

‐2.5

‐3

W/O Syn. Cond Syn. Cond 1 Syn. Cond 2

Figure 2-25 ROCOF Comparison with/without Synchronous Condensers

49.75

49.7
Frequency Nadir

49.65

49.6

49.55

W/O Syn. Cond Syn. Cond 1 Syn. Cond 2

Figure 2-26 Frequency Nadir Comparison with/without Synchronous Condensers

Synchronous condensers of the smaller size have a small effect on improving inertia.

Frequency observed at various buses of area 5 with and without synchronous

condenser is shown in Figure 2-27 – 2-28. Clear improvement in bus 507 and bus 508

frequency can be noticed. Frequency improvement suggests that if large generators

102
of retiring plants are converted as synchronous condensers, the appropriate level of

inertia is possible to achieve with desired ROCOF.

Figure 2-27 Buses Frequency during Generator Contingency and No Synchronous Condenser Addition

Figure 2-28 Buses Frequency during Generator Contingency and Equivalent Synchronous Condenser

Added

 Grid Code Compatible Frequency Controller based Wind Turbines

Frequency-active power control model is an auxiliary control algorithm implemented

in individual wind turbine generator control loop for providing controllable power

reserve on demand in the form of spinning reserve or power ramp rate limit to respond

103
system frequency deviations. A dynamic dead band based droop frequency grid code

compatible wind turbine as shown in Figure 2-29 has been used in this simulation

which will be discussed in detail in coming chapters.


Figure 2-29 Frequency Responsive Wind Turbine Model

Figure 2-30 Frequency Observed at Bus 506 when Normal Operated DFIG Based Wind Farm is added

Figure 2-30 shows the frequency at Bus 506 when DFIG based wind farm is added

to the bus. The Same bus is also receiving power from Generator 503 which is used

in the model with 4 active units, each with 109 MW rated capacity. Same capacity

wind farm without any supplementary droop control is added at Bus 506. It can be

seen from Figure 2-31 that maximum frequency deviation reaches up to 49.1 Hz after

simulation. In the second simulation, grid code compatible frequency controller based

wind farm is added at same Bus 506. A clear improvement in frequency deviation is

104
observed from Figure 2-32. Maximum deviation reached after the simulation is 49.97

in comparison to 49.1 in the previous case. Improvement in frequency deviation

shows that bus frequency can be improved with wind farms addition with some

modified control algorithms making them compatible with country grid code.

Figure 2-31 Frequency Observed at Bus 506 when Frequency Responsive DFIG based Wind Farm is

added

2.8. Conclusion

Based on the low load and high load scenario investigation under varying wind

penetration level, following results can be concluded for area 5 operation:

 Up to 45% wind penetration in area 5 has a negligible impact on strongly

interconnected area 3 regarding changes in ROCOF and frequency nadir.

Thus, a strongly interconnected network with power sharing can be a key to

sustainable wind energy integration.

 NEM has a minimum standard of ±1 Hz/s for one second for generators to

withstand ROCOF. With increasing wind penetration, there is a requirement

for ROCOF dependent relay operation to consider the time duration also for

which there is a change in ROCOF values.

105
 FCAS requirement likely to remain identical with the ramping period under 5

seconds if MW loss and number of the active synchronous generating unit are

same under different wind penetration level.

 ROCOF can be improved to lie under 2 HZ/s for contingency event if

synchronous condensers equivalent to MW loss are added in the network.

 Frequency responsive wind plants are likely to contribute to improving

frequency deviations, but ROCOF variation would need more assessment due

to the involvement of power electronics.

 Islanded network calls for more local regulation FCAS from local generating

units’ including wind energy participation in centrally managed automatic

generation control.

Three tired classification level can be formulated for ROCOF for wind integrated

power system like area 5:

 Green (ROCOF ≤ 1Hz/s) – FOS of 47–52 Hz can be easily met under wind

penetration as high as 40% for both low load and high load scenario.

 Orange (1Hz/s < ROCOF ≤ 4 Hz/s) – FOS of 47–52 Hz cannot be met under

wind penetration more than 20% and high contingency event for both low

load and high load scenario. With higher wind penetration and generation

loss, ROCOF remains under 4 Hz/s, but frequency nadir drops beyond set

level.

 Red (ROCOF > 4Hz/s) – FOS of 47–52 Hz will not be met. Even though with

wind penetration as high as 45% and MW loss as high as 740 MW, ROCOF

remains under 4 Hz/s. If ROCOF goes above 4 Hz/s, FOS standard will be

violated for more than a second leading to generation disconnection.

106
The focus of this research study is to challenge the designs built into the power system

and the regulatory framework within which it operates. Grid code compatible

frequency responsive wind turbine generator modeling is presented in subsequent

chapters, and corresponding impacts on the interconnected network regarding better

system frequency response are discussed.

107
2.9. References
1. AEMO, (2015). National electricity market factsheet. Available at aemo.com.au.

2. AEMC, (2014). Last Resort Planning Power - 2014 Review. Available at

www.aemc.gov.au

3. AEMO, (2016). National Electricity Forecasting Report. Available at

aemo.com.au.

4. AEMC. (2014). National Electricity Rules Version 62. Available at

aemc.gov.au/energy-rules/national-electricity-rules/current-rules.

5. IRENA, (2015). Renewable Energy Target Setting. Available at www.irena.org

6. REN21, 2016. Renewables 2016: Global Status Report. Available at

www.ren21.net.

7. National Energy Board, (2016). Canada Renewable Power Landscape, Energy

Market Analysis 2016. Available at www.publications.gc.ca.

8. Ma Tianjei, (2017). China's Ambitious New Clean Energy Targets, The

Diplomat. Available at www.thediplomat.com.

9. Laurie Guevara-Stone, (2016). A small country goes big with renewables:

Denmark’s goal to be fossil-free. Available at www.reneweconomy.com.au.

10. France in the United Kingdom, (2016). French renewable energy production up

over 23% in 2015, Energy policy/renewable energy – Communiqué issued

following the Council of Ministers’ meeting. Available at

www.uk.ambafrance.org

11. Tsvetomira Tsanova, (2015). India reached 22% of 175-GW renewable energy

goal, Renewables Now. Available at www.renewablesnow.com/news

12. Oxford Business Group, (2016). South Africa’s push for Renewables-Economic

News. Available at www.oxfordbusinessgroup.com

108
13. Commons Select Committee, Energy and Climate Change Committee (2016).

2020 renewable heat and transport targets report. Available at

www.parliament.uk/ecc

14. Office of Energy Efficiency & Renewable Energy, (2017). Federal Energy

Management Program, Achieving 30% Renewable Electricity Use by 2025.

Available at www.energy.gov.eere

15. Eurostat, (2017). Renewable energy in the EU, Share of renewables in energy

consumption in the EU still on the rise to almost 17% in 2015. Available at

http://ec.europa.eu/eurostat.

16. Geoscience Australia and ABARE, (2010). Australian Energy Resource

Assessment, Based on data from Wind lab Systems Pty Ltd., DEWHA

Renewable Energy Atlas (wind map data); Geoscience Australia. Available at

data.daff.gov.au.

17. Drew, G. (2015). Renewable Energy Superpower. [online] Beyond Zero

Emissions. Available at bze.org.au.

18. Paton, J. (2013). Australian Wind Energy Now Cheaper Than Coal, Gas.

Available at: http://www.bloomberg.com/news/articles/2013-02-06/australia-

wind-energy-cheaper-than-coal-natural-gas-bnef-says.

19. Milborrow David, “Onshore wind is more competitive than ever,” Article.

Available at www.windpowermonthly.com.

20. BREE, (2013). Electricity Generation Major Projects-2013. Available at

bree.com.au.

21. Mott MacDonald. (2010). UK Electricity Generation Costs Update. Available at

www.decc.gov.uk/assets/decc/statistics/projections/71-uk-electricitygeneration-

costs-update-.pdf.

109
22. Brinsmead T.S., J. Hayward and P. Graham (2014) Australian Electricity Market

Analysis report to 2020 and 2030, CSIRO Report No. EP141067

23. Australian Energy Council, (2016). Renewable Energy in Australia. Available

at www.energycouncil.com.au/.

24. ARENA, (2010). Australian Energy Resource Assessment, Chapter 9, Wind

Energy (pp 239-260). Available at www.arena.gov.au/about-renewable-

energy/wind-energy/.

25. CEC, (2015). Clean Energy Council Australia Report-2015. Available at

www.cleanenergycouncil.org.au/policy-advocacy/reports/clean-energy-

australia-report.html.

26. CEC, (2016). Clean Energy Council Australia Report-2016. Available at

www.cleanenergycouncil.org.au/policy-advocacy/reports/clean-energy-

australia-report.html.

27. CEC, (2016). Progress and Status of the Renewable Energy Target Report.

Available at www.cleanenergycouncil.org.au

28. Parkinson, G. (2016). Wind energy hits 100% of South Australia demand on

Sunday. [Online] Available at http://reneweconomy.com.au/2016/south-

australia-graphs-60608

29. Parkinson, G. (2016). Big blow in NSW: Wind farms produce more than coal

generators. [Online]. Available at: http://reneweconomy.com.au/2016/big-blow-

nsw-wind-farms-produce-coal-generators-91159.

30. AEMO, (2013). National Transmission Network Development Plan. 2013.

Available at www.aemo.com.au.

110
31. AEMO, (2015). Report to COAG Energy Council on Security and Reliability in

the Context of Changing Generation Mix. [online] SCER. Available at:

http://www.scer.gov.au.

32. Ceres project, (2013). Wind Farms and the SA Electricity Market Fact Sheet.

Available at: http://www.theceresproject.com.au.

33. Deloitte Access Economics, (2015). Energy markets and the implications of

renewables. Deloitte Access Economics Pty Ltd. Available at

www2.deloitte.com.

34. AEMO, (2016). Future Power System Security Program. Available at

aemo.com.au.

35. AEMO (2012), Frequency Control Ancillary Services. Available at

aemo.com.au.

36. J. Riesz, F.-S. Shiao, J. Gilmore, D. Yeowart, A. Turley and I. Rose, (2011)

Frequency Control Ancillary Service Requirements with Wind Generation -

Australian Projections.

37. AEMO (2013), Integrating Renewable Energy - Wind Integration Studies

Report. Available at aemo.com.au

38. Tanner P (2002), Wind farm grid connection technical issues, Presentation to

Transgrid Country Energy Workshop on Grid Connection of Wind Farms,

Sydney, 11 September.

39. M. Gibbard and D. Vowles, (2008). Simplified 14-Generator Model of the SE

Australian Power System. The University of Adelaide, South Australia.

40. MATLAB, (R2013). Simscape Power System Toolbox. Mathworks. Available

at mathworks.com.

111
41. GE Digital Energy, (2014). Synchronous condenser system. Available at

www.gegridsolutions.com.

42. Nicholas Miller, Kara Clark, and Miaolei Shao, (2010). Impact of frequency

responsive wind plant control on grid performance, GE presentation to MIT’s

wind week. Available at web.mit.edu/windenergy/windweek/Presentations.

43. Antonio Martinez, Kouroush Nayebi, Manoj Gupta and Yi Zhou, (2016). Wind

Power Plant Frequency Control to Support the Penetration of High Levels of

Renewable Sources, Vestas Wind Systems A/S presentation to Wind Industry

Forum.

44. PJM, (2016). Grid-scale energy storage for regulation services, PJM frequency

regulation, Available at www.pjm.com.

45. Clyde Loutan, Vahan Gevorgian, (2017). Using renewables to operate a low

carbon grid, demonstration of advanced reliability services from utility-scale

solar PV plant, California ISO, NREL and First Solar. Available at

www.caiso.com.

46. MacDonald, P. Cappers, D. Callaway, S. Kiliccote, (2012). Demand response

providing ancillary services, Proc. Grid-Interop 2012. Grid Wise Architecture

Council, Irving, TX.

47. AEMC, (2017). Demand Response Mechanism and Ancillary Services

Unbundling Rule. Available at aemc.com.au.

48. Uijlings, W. (2012). System Service Provision An independent view on the

likely costs incurred by potential System Service Providers. [online] London:

DNV KEMA. Available at www.eirgridgroup.com/site-

files/library/EirGrid/System-Service-Provision-DNV-KEMA-Report-2012.pdf.

112
CHAPTER 3

POWER SYSTEM FREQUENCY CONTROL

This chapter presents a brief literature review of active power frequency control

strategies for wind integrated power system model. An attempt has been made to

present new perspectives & challenges for secondary frequency control which need

more advanced investigation for future implementation in the electric grid.

3.1 Introduction

Load/Demand is primary independent control variable in power system while all

other controllable power system elements respond to serve this continuously varying

load. System frequency is one of the dependent variables which indicates system

energy balance and needs to be regulated for stable operation. Low frequency

indicates load increase while high frequency represents a high generation. System

frequency is regulated by injecting active power into electric grid through power

plants. This control is indispensable for the stable operation of grid ensuring

continuous adaptation of generation to demand. Frequency below a certain prescribed

level will deteriorate power quality & may eventually result in a system outage. Any

investigation of system frequency response requires frequency regulation analysis at

three different stages followed lastly by an emergency contingency regulation

control.

Figure 3-1 represents a typical frequency regulation scenario under contingency event

which leads system frequency to cross governor dead band. Under a wide range of

ambient conditions, frequency responsive power plants are expected to provide

automatic power variation with frequency drop within a given time frame and ramp

113
limit. The conventional power plants which remain connected to the grid even in the

case of frequency deviations accomplish active power-frequency control through

turbine governors reaction to nominal synchronous speed deviations and the

respective boilers-turbine frequency responsive controls.

Figure 3-1 Frequency Variation Along with Regulation Services

The normal power imbalance between generation and consumption is instantaneously

covered through a synchronous generators natural inertial response at first stage

followed by system synchronous generator’s governor response to bring system

frequency closer to setpoint. Contingency event leads to immediate governor

response for frequency regulation. A comparatively slower, centralized and

automatic secondary controller termed as automatic generator control (AGC) is

activated to bring frequency to the setpoint. Tertiary control is centralized manual

transmission system operator (TSO) controller utilized in case secondary reserves are

saturated before stabilizing the system frequency.

114
Table 3-1 Basic Frequency Regulation Services

↓Control→ Primary Secondary Tertiary

Impact Limit Deviation & Frequency Mitigate Forecasted


Automatic Frequency Restoration to Imbalance by
Stabilization Setpoint Replacing Primary
and Secondary
Reserves

Controller Position Decentralized, Power Centralized, TSO Manual by Tso


Plant Control Center

Input Frequency Frequency, Power Imbalance

Figure 3-2 Frequency Control during Contingency Event [Reformulated From 4]

Table 3-1 list the various stages involved in system frequency regulation. Basic

control scenario during contingency event is represented in Figure 3-2. Different

countries have different nomenclature for these three basic services. Table 3-2

provides the nomenclature adopted in various countries for generation demand

regulation. European Network of Transmission System Operators for Electricity

(ENTSO-E) [1] has proposed more explicit terminology for these services: Frequency

115
containment reserve services (FCR), Frequency restoration reserve services (FRR)

and Tertiary replacement reserve (RR) while some countries like Australia [2] and

U.K [3] have no explicit representation for tertiary control.

Table 3-2 Regulation Services Nomenclature Adopted by Various Countries

116
In the following next sections, the basic concept of frequency regulation stages is

described along with their presence in the wind integrated power system.

3.2. Power System Inertia

Conventional power system operation is based around traditional synchronous

generators operating in synchronism, i.e. with the same frequency (50/60 Hz), with a

lot of kinetic energy stored in their rotating masses. The angular velocity of these

synchronous generators is proportional to system frequency. Kinetic energy which is

stored in rotating masses will provide a cushion for this momentary imbalance of

power. This stored rotating energy in the synchronous generator is referred as Inertia

having the ability to oppose changes in frequency and represented by inertia constant

as:


H 3.1

W J 2πf 3.2

Where H is inertia constant in MWs/MVA denoting the time duration during which

the generating machine can supply its rated power (S ) in MVA exclusively with its

stored kinetic energy (W) in MJ or MW.s, J is the moment of inertia of rotating mass

117
in kg-m2. The change in kinetic energy leads to classical swing equation as shown in

Equation 3.3, which describes the inertial response of synchronous generator as

change in frequency following a power imbalance between the mechanical power

supplied by the generator P and P as the electric power demand.

J 2π ∗f∗ P P 3.3

On a system MVA base (S , equivalent system inertia constant, H can be

calculated as:

H ∑ H ∗ 3.4

Where H and S are the inertia rating and power rating of individual generating units

and N is the total number of generators.

Figure 3-3 Inertia Constant Values of Various Plants, Data Derived From [70] [45]

The system inertia constant depends on two factors including the number of operating

generators and the inertia constant of each of these generators. A larger H is

preferable as it signifies a higher amount of stored kinetic energy. Inertia constant per

118
unit base MVA for different power plants and synchronous condensers is represented

in Figure 3-3. The single wind turbine has inertia constant value in the range of 4-6 s

which is comparable to hydro and fossil steam based plants in similar MVA rating

range. An inversely proportional trend between inertia constant and machine rating

can also be seen for power plants.

3.2.1. Power system Inertial Response

The frequency of a power system with a large amount of stored kinetic energy tends

to be less sensitive to power imbalances than the frequency of a system with a lower

amount of stored energy. For a synchronous generator, the angular speed is tightly

coupled to the system frequency. When a power system experiences a large

disturbance, such as a sudden loss of generation, the system frequency begins to fall.

Therefore, when the system frequency drops, the generator-angular speed tracks the

falling frequency and, thus, the stored kinetic energy of the power plant is released.

This process is known as the inertial frequency response which acts as an initial

arresting mechanism to the falling system frequency and is determined by the type of

power plants in the system. There is direct coupling between the power system inertia

usually defined using the moment of inertia M, power system frequency and active

power output of synchronous generators in conventional generating plants ∆P as

given below in equation 3.5 and 3.6:

∆F s ∆P s ∆P s 3.5


M 3.6

T 3.7

119
M is the equivalent moment of inertia of all generators and the motors connected to

the grid [MW.s. Hz-1], H is the system equivalent inertia constant expressed in

megawatt second per megavolt ampere [MW.s / MVA], D is load damping constant

and T is defined as power system time constant and K is defined as power system

gain, equal to 1/D.

The generator inertia is, therefore, the first level of frequency control reserve, and it

has the characteristics of a differential controller [16]. Since inertia is proportional to

rotating mass, greater the inertia, the lesser will be acceleration, and the lesser will be

the frequency deviation. Higher system inertia gives a lower rate of change of

frequency and higher frequency nadir. As inertia affects how fast power system

frequency can change, it has implications for calculating frequency control ancillary

service (FCAS) requirements when considering high penetration of wind-based

power system.

3.3. Primary Frequency Control

Following the inertial frequency response of the power system, which reduces the

ROCOF, the primary frequency response stabilizes the frequency at a new value by

increasing (or decreasing) the power generation in proportion to the frequency

variations. Primary frequency control is a proportional control having system

frequency as a control signal and it independent of the rate of change of frequency.

Primary frequency control is represented in terms of change in turbine power ∆P as:

P ∆P ∆f 3.8

Where R is speed regulation or droop of governor and ∆f is frequency deviation. The

contribution of a generator to the correction of a disturbance on the network depends

mainly upon the generator droop value and the primary control reserve of the

120
generator concerned. The amount of frequency change which will cause 100%

change in the generator output is controlled through droop value. Two generators

having same primary control reserves but different droop will have different output

and different frequency offset as shown in Figure 3-4.

Figure 3-4 Frequency Power Characteristics through Droop Control

In case of low-frequency deviation (<∆f ), generator-a having lower droop will

provide more power output in comparison generator-b having higher droop. Even

having same primary control reserve, generator-a will reach maximum power limit

early, and its primary control reserve will be exhausted faster than generator-b. Both

generators will contribute equally to primary control in case of a major contingency

when frequency offset is > ∆f . As an example, 5% droop indicates that there will be

100% output power change for 5% change in frequency.

Author represents the simulated frequency output of thermal power plant based

uncontrolled area in Figure 3-5 when operating with different droop values while

respective power change of power plant is shown in Figure 3-6. Falling frequency

121
indicates more active power requirements, and we can see a plant with lower droop

will provide more output power during frequency disturbance.

Figure 3-5 Primary Frequency response of Uncontrolled Thermal Power Plant with Different Droop

Figure 3-6 Power Output Change of Uncontrolled Thermal Power Plant under Different Droop

Figure 3-7 represents the generic dynamic frequency control model for a single

control area having primary controlled power plants with rotating mass and frequency

dependent loads. ∆ is scheduled power setpoint for turbine while ∆ is the

power setpoint imposed by primary control. The total change in turbine power output

122
can be written in terms of equation 3.8. With ∆ 0, the change in frequency ∆f

with change in load ∆P can be related by equation 3.9. All the parameters in

equation are defined earlier in nomenclature.

∆f ∆P 3.9

Figure 3-7 Single Area Primary Frequency Control Model with Frequency Dependent Loads [20]

Droop values which indicate the required change in power output in response to

changes in frequency are defined in grid codes and vary in different power systems.

In a typical large power system, the power plants allocated to accomplish primary

frequency response should be capable of increasing their output real power within

about 10 seconds of the pre-defined frequency variation and be capable of

maintaining this response for a further 30 seconds. For example, in the Nordic power

system, the frequency controlled normal operation reserve must be at least 600 MW

and fully activated for frequency deviations of ±0.1 Hz. The frequency controlled

disturbance reserve must be at least 1000 MW and fully activated at 49.5 Hz. [17]. In

UCTE, 50% of the primary control reserves must be activated within 15 seconds, and

123
100% within 30 seconds. [18]. In Australian NEM, regulation raise/lower FCAS

of130/120 MW must be available for bringing frequency in restoration range (49.5 –

50.5) within 5 minutes [19].

3.4. Secondary Frequency Control

Maintenance of power system load-generation balance calls for a control system at

generation side as system load is normally uncontrollable and based upon customer’s

preferences. The secondary frequency Control or Automatic Generation Control

(AGC) represents a remarkable economic function of EMS system for energy market

dispatch of generating units and regulating frequency control ancillary service

(FCAS) dispatch. It is an important tool for regulating the balance and distributing

the imbalance between designated units. AGC makes up as part of real-time energy

market dependent upon load forecast. AGC which is slower than the two

aforementioned inertial and primary responses regulates the frequency at its rated

value. It is the IEEE working group which provided the standard definition for power

system AGC [21]. AGC, as defined by IEEE, is the regulation of the power output of

electric generators within a prescribed area in response to changes in system

frequency, tie-line loading, or the regulation of these to each other, to maintain the

scheduled system frequency and the established interchange with other areas within

predetermined limits.

The analysis and design of AGC system in a more deregulated complex environment

plays a vital role in automation & eventually reliable operation of the power system

at a satisfactory level. AGC system cost in modern power system consists of load

frequency control, economic dispatch; interchange scheduling of individual

generators in many generating units. The cost of load following imposed on the

system due to variable generation is also calculated by AGC setpoints and system

124
dispatch requirements. The sensitivity of any power plant depends upon the response

time taken by the automatic generation controller to control the frequency change due

to load variation of governor control. SCADA system periodically sends system

frequency and real power output data to AGC system which intermittently calculates

the area control error to adjust the setpoints of the respective system. If the load is

increased, AGC system increases the mechanical torque by increasing the system

flow rate to the turbine to compensate the load by increasing the output of the

generator in an area. Load-Frequency Control (LFC) is used as part of AGC to

maintain a constant frequency and to regulate tie-line flows. LFC represents the first

implementation of a higher-level control system. It has made the operation of

interconnected systems possible, and today it is still the basis of any advanced

concept for the guidance of large systems. Secondary reserves require a slower initial

response but are maintained for longer periods of time. Secondary control requires

the capability of increasing the output real power within about 30 seconds and

maintaining the response for a further 30 minutes. Secondary LFC or AGC is a central

control technique usually implemented through Proportional-Integral (PI) or Integral

controller as given in equation 3.10 below.

P K ∆ACE 3.10

3.4.1. Area Control Error

Present power systems are divided into various control areas each constituting

coherent group of generators. It is required to control and set the generation output

power according to energy market load-demand economics and area control error

(ACE). ACE is the real time representation of utility’s obligation to the regulatory

authority for maintaining net power inflow/outflow along with a small bias obligation

125
to maintain frequency. Stability of a system or requirement for more generation with

demand variation is indicated through ACE. Minimization of ACE is the main goal

of AGC which is a dynamic study where disturbance is small. For worthy AGC

design, there should not be an enormous increase in ACE signal, and ideally, it should

be zero. ACE should not be allowed to drift or vary much over a certain period. ACE

is allocated to various units of a control area whose participation in AGC depends

upon energy market economic dispatch.

ACE Control Modes

 Flat Frequency Control: When system frequency is utilized as the only

controlling parameter, AGC can be regarded as FFC. In this control, system

generation operations are not altered for any deviations from tie line schedule.

Normally used in the isolated control area.

 Flat Tie Line Control: When interconnection tie line flow is utilized as the only

controlling parameter, AGC can be regarded as FTLC. Limited to special

circumstances when control area frequency is stable and controlled through other

areas, this control mode is utilized to maintain a given net interchange power flow.

 Tie Line Bias Control: Responding to both tie-line power flow deviations and

frequency deviations, TLBC is the most widely utilized mode of system balancing

mode in big interconnected systems.

The ACE is calculated according to following equation 3.11:

ACE = (NIA − NIS) – 10B (FA − FS) - IME 3.11

Where ACE is the Area Control Error in MW. Negative values denote a condition of

under generation, and positive values denote over a generation. NIA is the algebraic

sum of the actual power flows on all tie lines or interconnectors (MW). NIS is the

algebraic sum of the scheduled flows on all tie lines or interconnectors (MW). It is
126
defined as the mutually prearranged intended net MW flow on the area tie lines,

including dynamic schedules. NIA − NIS is tie-line interchange error or flat tie-line

control which compares the net tie-line actual inflow/outflow with power scheduled

to be sold or bought. Flow into a system is defined as negative while flow out as a

positive value. B is the frequency bias setting (MW/Hz).

Frequency bias reflects the ability of any control area to respond to frequency

disturbance through various provisions like generators inertial response, governor’s

response or load shedding response. It is also the utility’s obligation to generate or

absorb power to regulate the total system frequency. FA is the actual frequency (Hz).

FS is the scheduled frequency (Hz). IME is Interchange (tie line) Metering Error which

is accounted from meters installed at the end of tie-lines. Time error is the

accumulation of frequency deviation over a defined period. When time error reaches

±10 seconds, the Interconnection Time Monitor initiates a time correction. The more

advanced formula for ACE can be found from WECC [22].

Figure 3-8 Basic Loop of AGC Control

127
The basic control loop of AGC control in shown in Figure 3-8 while Figure 3-9 gives

the AGC controlled thermal, hydro thermal power plant based three area system

frequency response under 1% load disturbance. The hydro plant has slightly lower

droop value in comparison to thermal plants. Area 1 has highest frequency bias

followed by area 2 and then area 3. The area with the highest bias will bring frequency

to settling values faster in comparison to other two areas. Figure 3-10 gives the AGC

response of single area thermal power plant under 1% load disturbance. It can be

noticed that plant with higher droop will take more time to settle frequency in

comparison to lower droop value based plants.

Figure 3-9 AGC Controlled Thermal Hydro Thermal Power Plant Based Three Area System Frequency
Response Under 1% Load Disturbance

Figure 3-10 AGC Response of Single Area Thermal Power Plant Under 1% Load Disturbance

128
3.5. Tertiary control

Variation of generators setpoints or load participation is done automatically or

manually through tertiary control. Tertiary control is applied to guarantee timely

provision of secondary control reserve and economical distribution of secondary

power control to various generators. Control may involve power connection tripping,

power redistribution, changing of load control and interconnector power variation

through change constraint equations. All tertiary control reserves are used to restore

secondary reserve.

3.6. Wind Energy System Participation in Frequency-Power

Control

Frequency control has become more challenging under conditions of low power

system inertia due to high wind energy penetration. Wind turbine active power

frequency regulation analysis has received considerable research focus in the form of

inertial control and primary frequency control. Figure 3-11 highlights various

concepts available in scientific literature applied for active power-frequency

regulation analysis either at individual wind turbine level or wind farm level [23-28].

Most of the past researchers focus on modifying individual turbine control algorithm

for analyzing primary frequency control capabilities from wind turbines. These

algorithms either focus on achieving de-loaded performance [29-32] to enhance

VSWTG’s frequency regulation capability or investigated the feasibility of constant

droop control [33] or combined inertial-droop control on frequency stability [34-38].

The basic philosophy behind these methods is briefly discussed in upcoming section.

129
Figure 3-11 Active Power Frequency Control Research Study Line in Wind Based Power Plants

3.6.1. Synthetic Inertial Control

Figure 3-12 Virtual Inertia Model

130
The inertial response has been introduced with different phrases such as synthetic

inertial response [39], inertial response [40-41], inertia emulation [36], short-term

frequency regulation [42], and temporary frequency control [30-31]. A segment of

VSWTG’s kinetic energy is transformed into electrical power through the inertial

control and transferred to the grid through converters. The basic idea behind most of

these droop based studies for emulating primary frequency response and inertial

response studies is to add an additional signal as shown in Figure 3-12, sensitive to

the frequency to the speed or torque controller which appear as reference power

setpoint in VSWTG model thereby momentary increasing the wind turbine output

power. The inertia constant H of a wind turbine is in the order of 2-6 s [36] which is

comparable to traditional generation with 2-10 s [43]. Hence the kinetic energy in the

rotor is high enough to support the frequency but must be released using control

actions in the converter active power control to mimic inertia. In the inertial response,

a part of the kinetic energy stored in the turbine is released by temporarily increasing

the output power. When energy is released from the turbine, the rotor speed

decreases. Consequently, the overproduction period has to be followed by a period of

underproduction where the turbine speed is increased, and the energy in the rotating

mass is restored.

Four virtual inertia and droop models are shown in Figure 3-13 which are further used

in coming chapters. Frequency droop model 1 [37], droop model 2 [44] and fourth

virtual inertia model from GE [45] takes absolute variation in frequency as input

while inertia droop model [36] takes absolute variation in frequency as well as the

rate of change of frequency as input. Transmission system operators (TSO) have been

working on grid code requirements for implementation of a generic inertial response

131
from WPPs [46]. Although some reports and recommendations for future grid codes

have been published [39], there are no mandatory requirements at present.

Figure 3-13 Inertia and Droop Control Model for Frequency Response from VSWTG

3.6.2. Primary Control from VSWTG

Similar to conventional generator response, droop characteristic and de-loaded wind

turbine operation have been explored for primary frequency response assessment

from wind turbines. Wind power plants are required to participate in primary

frequency control by ramping their output up and down according to a power–

frequency droop characteristic and during a certain period. De-loaded wind turbine

132
operation provides a power margin for ramping capability of wind turbines. De-

loading enables some reserve power procurement from wind farms for continuous

participation in primary load frequency control during negative frequency deviations.

De-loading through rotational over speed control [47-50] and pitch control [51-53]

has been discussed in various studies. Both methods are based on the idea of

achieving a non-optimal working point in the torque–rotor-speed curve of the turbine

as shown in Figure 3-14. Reference power setpoint to rotor converter control is

retrieved from measured rotor speed, de-loading margin, and T ω or P ω curves.

Figure 3-14 Representation of Optimum and Deloaded Curve for VSWTG

As represented in Figure 3-15 and under the assumption that maximum turbine power

and rotor speed are known, power setpoint for the de-loaded wind turbine can be

calculated as:

P P P P 3.12

P K P 3.13

%
K 1 3.14

133
P K ω 3.15

K ρ πR 3.16

Figure 3-15 Control Loop for De-loading Control of VSWTG

In pitch control de-loaded method shown in Figure 3-16 and represented in Figure 3-

17, pitch angle is regulated as a summation of pitch angle (β from pitch controller

and an offset pitch angle (∆β ; β β ∆β and ∆β K ∆f. Applied pitch offset

value (K can be dependent upon droop characteristics, active power reference or

de-loading power. De-load VSWTG by certain value X can be calculated by

increasing pitch angle as given by equation 3.17:

C λ ,β 100 X %∗C , λ ,β 3.17

Figure 3-18 gives an example of pitch angle offset variation with respect to de-loaded

power serve variation while Figure 3-19 presents the zoomed frequency nadir point

obtained under 10% and 20% pitch de-loaded VSWTG operation where a pitch offset

lookup table similar to Figure 3-18 is applied. An improvement in frequency nadir

can be with increased deloaded turbine and droop control enabled. Besides the offset

values, servo time constants are also important for this method deep analysis.

134
Figure 3-16 Pitch Droop De-loading Control

Figure 3-17 Coefficient of Performance De-loading Through Pitch Control

Figure 3-18 Pitch Angle Offset Variation with De-Loaded Power Reserve in VSWTG

135
Figure 3-19 Impact on System Frequency Variation with Pitch Droop De-loaded VSWTG

Pitch control methods are preferable in medium and high wind speed but lower in

comparison to an over-speeding method which is preferable at low wind speed. The

combination of pitch and over speed control is preferable in medium wind speed. A

hybrid over speeding de-loading control in provided in [54]. Rotor speed is also used

for this selection as over speeding technique is applied for lower rotor speed while

for rotor speed near or above rated value, pitch control is applied. Output reference

power from these control is not grid code compatible without additional algorithm

implementation.

Both static and dynamic droop control concept has been explored in literature for the

primary frequency response of individual wind turbine. Fixed power response can be

achieved by implementing static fixed droop value. The primary frequency

contribution ∆P of the VSWTG corresponding to the system load demand variation

can be computed from equation 3.8.

An important requisite for the stability of VSWTG during primary frequency

response provision is that its primary frequency power contribution should always be

sufficiently lower than the available power margin. To achieve improved stability and

to avoid reserve exhaustion, variable droop has also been studied. Variable power

136
response was shown to be achievable through dynamic droop value which can be

calculated by wind and power reserve conditions [72] or on the basis of ROCOF [71].

Under variable droop setting given by equation 3.18, optimum wind turbine operation

was proposed such that wind turbine having highest power reserve is assigned lowest

droop value so that it gets maximum participation in primary frequency response.

R R R R ∗ 3.18

Here M denotes power margin which is calculated as power margin: M P

P . So M denotes maximum power margin available among all wind turbines

while is the actual power margin available at a wind turbine.

Another study presented a control system for active power control of wind turbine

for tracking a power reference and provides a primary frequency response with a

constant and dynamic droop in the absence of any dead-band with diminishing

turbine structure loading [61]. Simulations were conducted on NREL CART 3 turbine

with a simple generator rather than a variable wind turbine generator model. In

another study, frequency active power set-point to wind turbine is provided as

M K df dP where K represents stiffness and M is inertia of power

supply where the value of M and K is provided by farm controller [56]. These

systems were studied for constant speed generator instead of variable speed

generator, and all of them were based on wind farm controller.

With commercial availability of 8 MW individual VSWTG system, individual

frequency responsive units need to research and investigation for fast primary and

secondary frequency response.

137
3.6.3. AGC & Wind Energy

The AGC problem has been extensively studied during the last four decades, and it

has been one of the most highlighted issues in the design and operation of

independent and interconnected power systems. Preponderance on renewable sources

like wind energy that different demand management as compared to a conventional

system has moved the electricity grid in a transition phase to a new model. These

sources having new production and nonlinear control technology imply new energy

management system (EMS) with Automatic Generation Control (AGC) being one of

the major units of EMS in modern power system [57]. A control area regulated by

hydro generation interconnected to another area regulated by thermal generation or

gas generation, solar generation, wind generation is a usual combination in a mixed

power system. With increasing RESs integration into the grid, it’s a necessity to

explore their impacts on AGC at the planning stage and in real time operation.

Recent investigations have indicated that large penetration of wind energy will have

an impact on modern power system frequency regulation along with AGC systems

and other control operation issues. AGC is expected to be a major player in the

management of short-term uncertainty of renewable power system output as well as

in mitigation of short-term impacts due to variable generation forecast error.

Application of several computational techniques have been researched for the

frequency regulation in the conventional power system; however, the large

integration of wind energy system have created a gap in theory and application with

just a few reports available on the frequency control design in the presence of wind

power units. Thus a lot of research is underway to review and analyze wind energy

association with AGC performance criteria, capabilities, and technologies for

138
ensuring proper system performance. Figure 3-20 illustrates the line of thoughts being

followed for studying AGC impact of wind energy penetration.

Energy Storage
Fuzzy logic
Conventional
Generation
Integration

Genetic Algorithm
Wind Automatic
Generation Control Intelligent
(Wind-AGC) Controllers

Particle swarm

LFC Controller

Figure 3-20 Wind Energy Study Line for AGC

Conventional AGC model assumes a direct physical connection for generators

setpoint change and the secondary control signal provided by the transmission system

operators. AGC implementation in wind energy system has been studied basically at

two levels: at wind farm level and wind turbine level. Figure 3-20 presents the basic

concept of AGC implementation in the wind-based power plant. To bring control area

frequency to required setpoint level, transmission system operator (TSO) sends

reference power set-points to wind farm controller which in turn dispatches

demanded power setpoints to individual wind turbines. A more detailed

representation can be seen in Figure 3-21 where TSO can provide different types of

power setpoint (balance, ramp or delta) as per system need. TSO provides dead band

and droop settings. The basic idea behind this model is to replace MPPT power

setpoint to wind turbine controller with Power setpoint from wind farm controller

while MPPT power output is sent to the controller as instantaneous available power.

In the case of no power limitation from TSO, same MPPT power set-point is provided

139
to turbine controller. Due to the time delay involved at farm controller level,

researchers are focusing on the possibility of implementing AGC control at wind

turbine level itself. This concept of AGC at turbine level has been explored in this

thesis.

Figure 3-21 Basic Schematic Diagram for Wind Farm Frequency Control

Figure 3-22 VSWTG Based Wind Farm Controller Supporting AGC [58]

140
The conventional LFC designs which are based on classical control have less

adaptability, are not very efficient. They are, therefore, not suitable for modern power

systems. The modern power system has a varied structure with large interconnections,

increased penetration of renewable energy sources and exigent demand for power

quality. It is expected that application of modern, robust, automatic and intelligent

control techniques to the LFC schemes will be more adaptive and appealing approach

for the wind-based power system to overcome impositions set up by technical

standards of energy market regulators. Incorporation of energy storage and

coordination of other conventional generation for frequency support in the wind-

integrated system is the other line of thought being followed for wind energy

incorporation in AGC. Table 3-3 provides some of the relevant studies related to

wind-AGC investigations where modifications were done at turbine level or wind

farm controller level. Besides these, there are numerous studies where wind

penetration effect on load frequency performance has been investigated.

Table 3-3: Wind – AGC: A Review

Refs. Methodology Special Findings

[59] Non-Linear differential algebraic equation Presents analytically traceable AGC

(DAE) modeling framework for studying model based on nonlinear differential

wind energy impact on power system algebraic equation (DAE) formalism but

frequency performance metrics utilizes wind energy system based on first

model differential equations and third

order DE for synchronous generators. No

wind participation in AGC was assumed.

[60] Hierarchical architecture of a supervisory Improved AGC performance and better

control system for active and reactive power integration of the wind with the grid when

control at wind farm substation. supervisory system controls active and

141
Supervisory control system active power reactive power while machine system

output is dependent upon TSO provided controls turbine power setpoints.

power setpoint.

[55] Active Power Controller was incorporated Static and dynamic droop curves were

[61] with AGC command tracking from wind implemented, and Field tested for

farm controller NREL’s 550kW 3-bladed Controls

Advanced Research Turbine (CART3)

for frequency regulation participation at

all three levels. Fast Wind power AGC

response especially at high wind speed

and AGC error are low with multiple

turbines in wind farms.

[62] DFIG based wind farm hierarchical control Implemented in DIgSILENT, the central

strategy to regulate the wind farm power wind farm control level controls the

production to the reference power ordered power production of the whole farm by

by the system operators sending out reference power signals to

each wind turbine, while the local wind

turbine control level ensures that the

reference power signal send by the central

control level is reached

[63] Power/load optimization in the presence of Centralized wind farm model and control

ambient wind, wake deficits, wake principles are used for control of a wind

meandering, the wind turbine system, the power to increase energy quality and

wind farm controller, and a network reduce the fatigue loads

operator

[64] Coordinated control strategy for the AGC Proposed control strategy have shown

between combined heat and power plants that the WPPs can actively help the AGC,

(CHPs) and WPPs to enhance the security and reduce the real-time power imbalance

and the reliability of a power system in the power system, by downregulating

142
operation in the case of a large wind power their production when CHPs are unable to

penetration provide the required response

[65] Developed model called (FESTIV) to The model allows for wind energy

mimic operator behavior using security penetration impacts analysis and gives

constraint unit commitment, economic insight to market operation regarding

dispatch and AGC reliability and costs.

[66] Wind farms incorporating energy storage Energy storage devices are used to

systems for AGC participation neutralize the uncertainty of wind power

plant output and maximize the dispatch

ability. With reduced uncertainty, wind

farms are shown to have better AGC

performance.

[67-68] DFIG based wind farm integrated with Operational impacts of the wind power

thermal energy system was used in study fast fluctuations are largely absorbed by

the thermal unit’s large mechanical and

thermal time constants as well as control

dead-bands, and the system dynamics

varies with the size of the AGC time

delay. It is estimated in this study 5% of

rated plant power can be accepted as

perturbation power by thermal plants,

without exceeding the one%frequency

deviation margin.

[69] (PSO-Fuzzy PI) Fuzzy logic load frequency Intelligent computational techniques

controller and Particle swarm optimization were applied to improve AGC controller

technique were applied performance in New England test system

with wind penetration, but no specific

wind farm frequency support was

accounted.

143
3.6.4. AGC & Wind Energy Research: New Perspectives and
Challenges

Wind energy has the potential to maintain and even exceed the dynamic growth rate

of the past several years. It remains important for researchers to develop more reliable

designs and technology to help wind integration with the electric grid and provide a

knowledge base to concerned authorities so as to maintain long-term growth and

innovation in generation, transmission and demand technologies and commercial

arrangements. This section provides some insights into frequency related issues

which need more research for implementation in the wind farm future grid code

development. Implicit implementation of these issues can lead to the development of

wind farm grid codes and implementation policy which will pave the way for wind

farm operation that is optimally balanced between energy cost and technical

performance.

Synthetic Inertia Supporting Wind Farm Participation in AGC

It is well known that AGC action is preceded by primary regulation activity in a

power system. Load frequency control (LFC) action assumes same area control error

(ACE) throughout the concerned power system, but this assumption does not

completely satisfy with wind power based system. High wind penetration in power

system induces large imbalances and brings instability and power quality issues.

With secondary regulation getting exhausted due to wind load variability, it is very

likely that primary control actions of power electronics based generation technologies

also reduce frequency regulation capabilities in case of a fault [73]. Primary

regulation is provided by power system inertia which controls the system frequency

rate of change in case of an imbalance of load and demand in the system. Power

system rotational inertia is normally time-variant and gets reduced non-uniformly


144
with respective grid topology. Primary inertial control can improve frequency nadir

but do little for improving settling time as the inertia of a system is directly related to

the mass of generating machine and respective prime movers. A high stable power

system has high inertia and low ROCOF. Due to energy economics, the number of

synchronous generators operating will be less in the case of low load demand and

high penetration of wind energy, leading to reduced total grid system inertia.

Frequency regulation and hence sensitivity and stability of the total system is largely

dependent upon total system inertia. Managing total system frequency thus stability

without an inherent inertial response will be difficult for wind penetrated power

system.

Studies indicate that power system is having equal value of spinning reserves but

having different inertia levels express varied frequency responses in case of any

contingency. Research indicates that low inertia will have worst effect for power

system having slow frequency response [74]. So modern wind-based systems are

supplementing their inertial response by adopting additional controllers and using

energy storage systems and synchronous condensers. Modern wind plants have the

ability to control active power output in response to grid frequency by providing

inertial and governor response. The wind turbine can fast store or release a large

amount of kinetic energy in the rotating mass because of the power electronic

converter control, large moment of inertia and wide rotational speed range. Wind

turbine inertial response is essentially energy neutral, meaning that the period of

increased power is followed by a period of decreased power. The inherent inertial

response by wind turbines is also not same under all operating conditions but depend

upon active controls. Wind farm operation based on the maximum power capture

curve to extract maximum wind energy may need replacement to de-loading curve to

145
save the available power as reserves for the long-term frequency control. Various

supplementary inertial controls attribute, and designs have been proposed in the

literature to achieve active frequency regulation; inertial control and droop control

being a prominent one. Inertial control acting as extra loop utilize the rate of change

of frequency while droop control utilizes frequency difference between measured

frequency and actual frequency to adjust power to reduce frequency deviations. There

is no standard to quantify the effect of supplementary frequency regulation loop on

the maximum rate of change of frequency (ROCOF).

The wind inertial control increases the system inertia, but the inertial power may

mask the load changes for a few seconds because of the considerably released kinetic

energy from the rotating mass. Therefore, the synchronous generators may delay their

responses to the frequency events. Varied decentralized and distributed control

methodologies to implement inertial and AGC response at turbine level and real scale

power plant level with high ROCOF ride-through capability is still very nascent and

needs more attention as research in this field appears to be sparse.

Intelligent AGC

Conventional optimization techniques based on various mathematical programming

methods have been previously utilized in power system operation and control, but

they are constrained to perform a search operation for a local optimum. Modern smart

grid with multigenerational integrated power system calls for computational

intelligence (CI) and evolutionary computation (EC) techniques for robust and

optimum results in AGC regulation. Usage of an intelligent configuration has been

advocated as the best solution for improving the contribution of a renewable power

plant in AGC function [69]. It has been established that overall deviations in grid

146
frequency and regulation time can be reduced by incorporating intelligent controllers

and distributed AGC in general power systems operating under restrictions and

certain operating ranges.

With the availability of highly capable variable generators, automatic response to

frequency is getting improved for wind plants. Pitch angle control and rotor speed

control have been deduced as popular frequency control techniques at turbine level

[75], but the absence of intelligent controllers in the wind penetrated system can cause

stability problems and penalties causing consequent losses in revenues. Application

of intelligent algorithms incorporating a high degree of optimization and control

strategies can better respond to existing complex variables of power system models

and supports AGC in mitigating the demand and load imbalance in the interconnected

system. Intelligent load frequency controllers will be front runners to enhanced

performance, improved tuning and adaptive competencies. A more advanced

simulation involving high-fidelity system models and control capability is needed to

analyze the full potential and impact of AGC for wind penetrated system. These needs

to be real time verified and investigated for wind-integrated system with application

of intelligent techniques.

Fault Ride-through Capable Wind Farm Effect on AGC

With increasing renewable energy penetrations in the modern power system, all

national TSOs have put stringent ride through capability limits on these power

systems. Ride through capability refers to the ability of generators to remain stable

and connected during normally cleared balanced and unbalanced electrical faults on

a transmission grid. This constraint has been enforced as mandatory to evade

substantial loss of wind energy generation in the occasion of grid faults. Earlier trend

of disconnection of wind farms poses significant damage to frequency and voltage


147
regulation leading to a blackout in the case of a major fault. So as per various national

grid codes, their current wind plants are required to limit their active power as a

function of system frequency in case of over frequency event. Australian grid code

[76] requires their wind plant to provide ride through multiple faults even though it

is not cleared during reclosing sequence of transmission lines. Even though technical

solutions exist now for overcoming fault ride through behavior, there are other rising

issues which need to be further addressed. More involvement of type 3 & 4 wind

turbine by displacing synchronous generators will result in higher ROCOF and

instability in the total system. For longer duration faults, turbine blade pitching

strategies along with turbine converter control become necessary.

FRT compliance is assessed at the point of connection (PCC) while manufacturers

specify operating range of wind turbine generators at its low voltage (LV) terminals.

Depending upon short circuit ratio (SCR) at the point of coupling (PCC) and wind

turbine type, there can be varying stress on different parts of the turbine when it is

subjected to near-to-generator disturbances. Integration of wind energy with a

network having low short circuit ratio (SCR) has increased the risk of system

instability even though with FRT capable wind turbines. Wind farms with such fault

ride through (FRT) capabilities are very likely to deteriorate frequency regulation as

frequency deviations result in voltage instability. As a result, energy withdrawn by

FRT action may increase the size of any contingency and thus the magnitude of

frequency deviations leading to total system instability.

To analyze FRT implementation impact on frequency regulation performance of

wind farms and vice versa; it is important to conduct more research & assessment of

these FRT capable wind system.

148
Spatially Distributed Wind Farm Participation in AGC

A comprehensive literature exists which deals with lots of variability and uncertainty

in wind power output [77-80]. Wind power ramping down event is not as abrupt as

sudden loss of conventional generator and not as smooth as load ramp-down daily

cycles [81]. In spite of technological development and implementation, there is a

limitation on the ability of modern wind energy-based power system in contributing

to regulating frequency under few seconds after a contingency. Most of the solution

proposed in literature mimics the traditional AGC response of synchronous

generators. In a high wind penetrated power system, with varying wind power and

load uncertainties, there is a spatial variation of frequency standards and required

regulation over control area. So for wind energy system (WES), traditional lumped

ACE input changes to distributed ACE and conventional AGC changes to spatially

differentiated frequency regulation. One of the studies indicates that wind farms

spread over the large geographical area have relatively low regulation limits [82]. At

normal operation, the power output of a wind farm can vary up to the 15% of installed

capacity within 15 minutes [83]. The spatial distribution and intermittency of wind

resources affect power system on time scale being followed by AGC and raise the

cost of wind-based electricity. Research from NREL done under German weather

conditions shows that ancillary service requirements decrease substantially for a large

number of turbines and due to the spatial diversity of wind resources. Regulation

analysis from the same report also advocates that the physical separation of wind

plants has a lesser effect on regulation burden than a number of turbines [84].

Conventionally economic dispatch has been applied for generating unit participation

in AGC, but with the emergence of the individual frequency responsive wind energy

system, the advance investigation is needed for such spatially distributed & large

149
aggregation of wind turbines on AGC under various grid codes. AGC system along

with economic dispatch of respective electric network needs to have improved

modeling to account for unique features of wind generation and mitigating its

variability effect.

Frequency Control Ancillary Service (FCAS) As AGC Output

AGC manages continuous correction of frequency deviations as regulation raise or

regulation low service as part of market-based ancillary service in the energy market.

Frequency control, one of the ancillary service requirements are calculated by

contingency size, total inertia of power system and load demand. Frequency control

ancillary service (FCAS) providers bid their services into the FCAS markets in a

similar fashion like generating units bidding into the energy market and receive

financial incentives. Payments for ancillary services include payments for availability

and delivery of the services. Service providers must be able to show their

conformance to the ancillary services specification. FCAS represents a guarantee that

power will be continued to be delivered in the case of contingency.

FCAS availability holds a very important value for high wind penetrated power

system with low inertia. Wind farms which produces electricity at low prices can gain

financial revenues by participating in the provision of regulation services. In a

deregulated environment, energy market participants are required to make

supplementary control contributions to transmission system operator (TSO) for the

stable and reliable power system. To meet control area performance criteria, AGC

participants must track and keep a record of instantaneous fluctuations in demand

load for the continuous duration so as to keep ACE within limits. For a transmission

system operator, assignment of control duties gets constricted with the presence of

150
wind system. System operator aims to minimize the payments to participants for

providing AGC ancillary service while wind participant strives to maximize their

revenues.

Australian energy market operator (AEMO) has the requirement for FCAS to ensure

the frequency to be contained within 49 to 51 Hz in the event of step changes in

supply that result from credible contingencies known as “separation event.”

According to Australia’s national electricity market (NEM), it is compulsory for

generators >=30 MW to participate in FCAS [85] though none of the wind farms

participate in it. An impact assessment of large-scale renewable energy target done

for Australian Energy Market Commission (AEMC) reported an increase in

regulation requirement from ±120MW to ±800MW and increase in frequency

regulation cost from 10 million per annum to 200 million per annum in 2020 due to

projected increase in intermittent wind energy. Causer-pays methodology is being

implemented in Australia for settling regulation costs, with generators or loads paying

for any deviation from their expected dispatch. With increased wind penetration, the

majority of the cost is anticipated to be borne solely by wind farms leading to an

escalation in regulation costs for wind generators from around $0.40 /MWh at present

to $6 - 8 /MWh in 2019-20 [86]. This increase in wind farm costs could have

significant repercussions.

North American Electric Reliability Corporation (NERC) has enacted control

performance standard (CPS) for evaluating frequency control performance in a

control area. Control areas must not be less than 95 % compliant with CPS1 and no

less than 85% compliant with CPS2 [87]. A Study by Wang & McCalley indicates

that CPS1 & CPS2 deteriorate with increasing wind penetration and this effect is

151
observed more for large interconnected systems [88]. Improved load forecasting

along with better AGC is required to improve these control performance standards.

According to European Network for Transmission System Operators for Electricity

(ENTSO-E), all TSOs must ensure regulation service which should be at least 50%

operational within 15 seconds and fully operational within 30 seconds of any

contingency exceeding ±20mHz [89]. A very fast response is needed to seize

frequency decline and control frequency nadir within first 5-10 seconds. Energy

storage technologies have fast response making them suitable for providing

frequency regulation ancillary services.

According to a report by California Energy Commission (CEC), application of energy

storage for the regulation purpose can be around two to three times as effective as

adding a combustion turbine to the system [90].

Besides technology, market policies, as well as system infrastructure, also play vital

role satisfactory AGC performance in a wind energy integrated power system. Wind

farms can boost their revenue and decrease total costs by implementing frequency

regulation technology. More research regarding technology, policies, and

infrastructure is needed for incorporating AGC in wind system to enable their active

participation in FCAS, better energy economics and compliance to control

performance standards. Proper management of FCAS can be a key factor for better

penetration of the wind in a deregulated complex power system.

3.7. Conclusion

This chapter presented present research and perspective challenges for wind

participation in frequency regulation. Effective research and implementation of these

terminologies in AGC will help wind energy market participants to make better

152
contributions to frequency control ancillary services to transmission system operator

(TSOs) and an overall more reliable power system. Therefore, extensive research is

required to review and analyze wind energy association with frequency regulation

criteria, capabilities, and technologies for ensuring cost effective, industry prone

frequency regulation system leading to a stable wind integrated large power system.

With improved turbine and intelligent controller technologies, AGC researchers have

to look deeply into several new aspects for achieving long-term economic

sustainability of wind energy.

153
3.8. References
1. ENTSO-E (2012). Operational reserve ad hoc team report, final version,

Available at www.entsoe.eu.

2. AEMO (2012), Frequency Control Ancillary Services. Available at

aemo.com.au.

3. National Grid (2012). The Grid Code, Issue 4, Revision 10, National Grid

Electricity Transmission plc, UK, January 2012.

4. B. Kirby (2007). Ancillary Services: Technical and Commercial Insights.

Available at www.science.smith.edu.

5. Verband der Netzbetreiber - VDN - e.V. beim VDEW. Transmission code 2007.

Network and system rules of the German transmission system operators.

Available at www.vde.com.

6. National Grid (2016). Mandatory Frequency Response | National Grid. [online]

Available at: http://www2.nationalgrid.com.

7. Christer Bäck (2013). Nordic implementation of Frequency Restoration Reserve

Automatic (FRR-A), Svenska kraftnät (SVK). Available at www.svk.se

8. Ignacio de la Fluente. Ancillary services in Spain: Dealing with high penetration

of RES. Red Eléctrica de España. Available at www.reshaping-res-

policy.eu/downloads/topical%20events/de-la-Fuente_Ancillary-Services-in-

Spain1.pdf.

9. EirGrid. Eirgrid grid code version 4.0, Available at www.eirgrid.com.

10. Hewicker, C., Werner, O. and Zeigler, H. (2013). Qualitative analysis of cross-

border exchange of balancing energy and operational reserves between

Netherlands and Belgium. Available at: http://www.tennet.eu.

154
11. TenneT, The Netherlands, Samenvatting uit “Summary of the Current Operating

Principles of the UCPTE” met betrekking tot de primaire en secundaire regeling,

Dec. 2003.

12. PJM (2003). USA, Manual 10: Pre-Scheduling Operations, Revision: 16.

[Online]. Available at www.pjm.com/contributions/ pjm-manuals/

pdf/m10v161.pdf.

13. PJM (2006). USA, Manual 12: Dispatching Operations, Revision: 13, May 26.

Available at www.pjm.com/contributions/pjmmanuals/pdf/m12v13.pdf.

14. PJM (2005). USA, Manual 35: Definitions and Acronyms, Revision: 05.

Available at www.pjm.com/contributions/pjm-manuals/ pdf/m35v051.pdf.

15. Dan Woodfin (2015). Methodology for Determining Minimum Ancillary

Service Requirements. Available at www.ercot.com.

16. T. Borsche, A. Ulbig, G. Andersson, (2014). A new frequency control reserve

framework based on energy-constrained units, Power Systems Computation

Conference (PSCC) 2014, pp. 1-7, 18-22.

17. Nordel (2006). Agreement regarding operation of the interconnected Nordic

power system (System Operation Agreement). Technical report. Available at

www.entsoe.eu.

18. UCTE. P1 - Policy 1: Load-Frequency Control and Performance [C]. Technical

report, 2009. Available at www.eeh.ee.ethz.ch.

19. AEMO, (2016). Future Power System Security Program. Available at

aemo.com.au.

20. Goran Andersson (2012). Dynamics and Control of Electric Power Systems,

Lecture 227-0528-00, ITET ETH. ETH Zurich. Available at

www.eeh.ee.ethz.ch.

155
21. IEEE (1970). Standard definitions of terms for automatic generation control on

electric power systems, IEEE committee report. IEEE Trans. Power App. Syst.

PAS-89:1356–64

22. Western Power (2011). Technical Rules, Available at

www.westernpower.com.au.

23. R. de Almeida and J. P. Lopes (2007). Participation of Doubly Fed Induction

Wind Generators in System Frequency Regulation. IEEE Transactions on Power

Systems, vol. 22, pp. 944–950.

24. X. Juankorena, I. Esandi, J. Lopez, and L. Marroyo (2009). Method to Enable

Variable Speed Wind Turbine Primary Regulation, in International Conference

on Power Engineering, Energy and Electrical Drives, pp. 495–500.

25. M. Liserre, R. C´ardenas, M. Molinas, and J. Rodriguez (2011). Overview of

Multi-MW Wind Turbines and Wind Parks, IEEE Transactions on Industrial

Electronics, vol. 58, no. 4, pp. 1081–1095.

26. J. Morren, S. de Haan, W. Kling, and J. Ferreira (2006). Wind Turbines

Emulating Inertia and Supporting Primary Frequency Control, IEEE

Transactions on Power Systems, vol. 21, no. 1, pp. 433 – 434.

27. Gowaid, A. El-Zawawi, and M. El-Gammal (2011). Improved Inertia and

Frequency Support from Grid-Connected DFIG Wind Farms, in Power Systems

Conference and Exposition (PSCE), 2011 IEEE/PES, pp. 1–9.

28. H. Ma and B. Chowdhury (2010). Working Towards Frequency Regulation with

Wind Plants: Combined Control Approaches, Renewable Power Generation,

IET, vol. 4, no. 4, and pp. 308–316.

29. Hansen, A. D. (2010). Evaluation of power control with different electrical and

control concepts of wind farms, Tech. rep., Project UpWind, Roskilde, Denmark.

156
30. N. R. Ullah, T. Thiringer and D. Karlsson (2008). Temporary Primary Frequency

Control Support by Variable Speed Wind Turbines— Potential and

Applications, in IEEE Transactions on Power Systems, vol. 23, no. 2, pp. 601-

612.

31. G. C. Tarnowski, P. C. Kjar, P. E. Sorensen and J. Ostergaard (2009). Variable

speed wind turbines capability for temporary over-production, IEEE Power &

Energy Society General Meeting, Calgary, AB, pp. 1-7.

32. Ramtharan G, Ekanayake JB, Jenkins N. (2007). Frequency support from doubly

fed induction generator wind turbines. IET Renew Power Gen; 1: 3–9.

33. A. Aziz, G. M. Shafiullah, A. Stojcevski and A. Mto (2014). Participation of

DFIG based wind energy system in load frequency control of interconnected

multi-generation power system. Power Engineering Conference (AUPEC).

Australasian Universities, Perth, WA, pp. 1-6.

34. A.B. Attya, T. Hartkopf (2013). Wind farms dispatching to manage the

activation of frequency support algorithms embedded in connected wind

turbines, International Journal of Electrical Power & Energy Systems, Volume

53, Pages 923-936.

35. Hughes FM, Anaya-Lara O, Jenkins N, Strbac G. (2005). Control of DFIG-based

wind generation for power network support. IEEE T Power Syst; 20: 1958–1966.

36. Morren J, de Haan SWH, Kling WL, Ferreira JA. (2006). Wind turbines

emulating inertia and supporting primary frequency control. IEEE T Power Syst;

21: 433–434.

37. Mauricio JM, Marano A, Gomez-Exposito A, Ramos JLM. (2009). Frequency

regulation contribution through variable speed wind energy conversion systems.

IEEE T Power Syst; 24: 173–180.

157
38. Gautam D, Goel L, Ayyanar R, Vittal V, Harbour T. (2011). Control strategy to

mitigate the impact of reduced inertia due to doubly fed induction generators on

large power systems. IEEE T Power Syst; 26: 214–224.

39. National Grid (2010). Requirements for System Inertia, National Grid Code

Frequency Response Working Group, UK.

40. J. Morren, J. Pierik, and S. Dehaan (2006). Inertial response of variable speed

wind turbines, Electric Power Systems Research, vol. 76, no. 11, pp. 980-987.

41. A. Mullane and M. O'Malley (2005). The Inertial Response of Induction-

Machine-Based Wind Turbines, IEEE Trans. Power Systems, vol. 20, no. 3, pp.

1496-1503.

42. O. Anaya-Lara, F. Hughes, N. Jenkins, and G. Strbac (2006). Contribution of

DFIG based wind farms to power system short-term frequency regulation, IEE

Proceedings Generation, Transmission and Distribution, vol. 153, no. 2, pp. 164-

170.

43. Kundur, P., Balu, N. and Lauby, M. (1994). Power system stability and control.

New York: McGraw-Hill.

44. Xue Yingcheng, Tai Nengling (2012). System frequency regulation in doubly

fed induction generators, International Journal of Electrical Power & Energy

Systems, Volume 43, Issue 1, Pages 977-983.

45. K. Clark, N. W. Miller, J. J. Sanchez-Gasca (2010). Modelling of GE Wind

Turbine-Generators for Grid Studies, Version 4.5. General Electric International,

Inc.

46. ENTSO-E (2012). Draft Network Code for Requirements for Grid Connection

applicable to all Generators, European Network for Transmission System

Operators for Electricity.

158
47. R. de Almeida, Castronuovo ED, Lopes JAP. (2006). Optimum generation

control in wind parks when carrying out system operator requests. IEEE

Transactions on Power Systems, 21(2):718-25.

48. Janssens NA, Lambin G, Bragard N. (2007). Active power control strategies of

DFIG wind turbines. In: 2007 IEEE power tech Lausanne; p. 516-21.

49. Kanellos FD, Hatziargyriou ND. (2008). Control of variable speed wind turbines

in islanded mode of operation. IEEE Transactions on Energy Conversion,

23(2):535-43.

50. Courtecuisse V, Robyns B, Francois B, Petit M, Deuse J. (2008). Variable speed

wind generators participation in primary frequency control. Wind Engineering,

32(3):299-318.

51. Holdsworth L, Ekanayake JB, Jenkins N. (2004). Power system frequency

response from fixed speed and doubly fed induction generator-based wind

turbines. Wind Energy; 7 (1):21-35.

52. Conroy JF, Watson R. (2008). Frequency response capability of full converter

wind turbine generators in comparison to conventional generation. IEEE

Transactions on Power Systems; 23 (2):649-56.

53. Moutis P, Loukarakis E, Papathanasiou S, Hatziargyriou ND. (2009). Primary

load frequency control from pitch-controlled wind turbines. In: 2009 IEEE

power tech Bucharest; p. 1-7.

54. P. Moutis, S. A. Papathanassiou, and N. D. Hatziargyriou (2012). Improved

load-frequency control contribution of variable speed variable pitch wind

generators," Renewable Energy, Vol. 48, pp. 514-523.

55. J. Aho, A. Buckspan, J. Laks, P. Fleming, Y. Jeong, F. Dunne, M. Churchfield,

L. Pao and K. Johnson. (2012). A Tutorial of Wind Turbine Control for

159
Supporting Grid Frequency through Active Power Control. Montreal:

Proceedings of the 2012 American Control Conference.

56. Garcia, J. (2010). Frequency Control. 20120035775.

57. Stanton N. K., Giri J. C. and Bose A. (2007). Energy management. In Power

system stability and control, ed. L. L. Grigsby. Boca Raton, FL: CRC Press

58. Poul Sorensen, Anca d. Hansen, Florinn Iov, Frede Blaabjerg and Martin H.

Donovan (2005). Wind farm models and control Strategies, Riso-R Report,

Denmark. Available at orvit.dtu.dk.

59. Garcia A. D. (2012). Models for Impact Assessment of Wind-Based Power

Generation on Frequency Control, In Control and Optimization Theory for

Electric Smart Grids, vol. 3, A. Chakraborty, and M. Ilić (Eds.), Springer-Verlag,

Berlin.

60. Amenedo L. R., Arnalte S., and Burgos J. C. (2002). Automatic generation

control of a wind farm with variable speed wind turbines. IEEE Trans. Energy

Convers. 17(2): 279-284.

61. A. Buckspan, J. Aho, P. Fleming, Y. Jeong and L. Pao (2012). Combining droop

curve concepts with control systems for wind turbine active power control, 2012

IEEE Power Electronics and Machines in Wind Applications, Denver, CO, pp.

1-8.

62. Anca D. Hansen, Poul Sørensen, Florin Iov, Frede Blaabjerg (2006). Centralised

power control of wind farm with doubly fed induction generators, Renewable

Energy, Volume 31, Issue 7, Pages 935-951, ISSN 0960-1481.

63. J. D. Grunnet, M. Soltani, T. Knudsen, M. Kragelund, T. Bak (2010). Aeolus

toolbox for dynamic wind farm model simulation and control", Proc. Eur. Wind

Energy Conf., pp. 3119-3129.

160
64. Basit, Abdul, et al. "Wind power integration into the automatic generation

control of power systems with large-scale wind power." The Journal of

Engineering (2014).

65. E. Ela, M. Milligan, and M. O'Malley (2011). A flexible power system

operations simulation model for assessing wind integration, 2011 IEEE Power

and Energy Society General Meeting. IEEE, pp. 1-8.

66. Michael Antonishen, H. Y. Han, T. K. A. Brekken, A. von Jouanne, Alex

Yokochi, D.A. Halamay, J. Song, D. B. Naviaux, J. D. Davidson, Alex Bistrika.

(2012). A Methodology to Enable Wind Farm Participation in Automatic

Generation Control Using Energy Storage Devices, IEEE Power and Energy

Society General Meeting, IEEE , vol., no., pp.1,7, 22-26.

67. Luo C. and B.T. Ooi. (2006). Frequency deviation of thermal power plants due

to wind farms”, IEEE Trans. Energy Conv., vol. 21, no. 3, pp. 708–716.

68. Luo C., Far H. G., Banakar H., Keung P. K. and Ooi B.T. (2007). Estimation of

wind penetration as limited by frequency deviation”, IEEE Trans. Energy Conv.,

vol. 22, no. 3, pp. 783–791.

69. Hassan Bevrani and Pourya Ranjbar Daneshmand (2012). Fuzzy Logic-Based

Load-Frequency Control Concerning High Penetration of Wind Turbines, EEE

Systems Journal, Vol. 6, No. 1.

70. P. Anderson and A. Fouad, Power System Control and Stability. Piscataway,

US: Wiley - IEEE Press, 2002.

71. M. Hwang, E. Muljadi, J. W. Park, P. Sørensen, and Y. C. Kang, "Dynamic

Droop–Based Inertial Control of a Doubly-Fed Induction Generator," in IEEE

Transactions on Sustainable Energy, vol. 7, no. 3, pp. 924-933, July 2016.

161
72. K. V. Vidyanandan, N. Senroy, "Primary frequency regulation by deloaded wind

turbines using variable droop," IEEE Trans. Power Syst., vol. 28, no. 2, pp. 837-

846, May 2013.

73. Miller N. W., Shao M., and. Venkataraman S (2011). California ISO (CAISO)

frequency response study, GE Energy.

74. Andreas U., Theodor S. B. and Göran A. (2014). Impact of Low Rotational

Inertia on Power System Stability and Operation, Submitted to IFAC World

Congress 2014, Capetown, South Africa

75. Juankorena X., Esandi I., Lopez J. and Marroyo L. (2009). Method to Enable

Variable Speed Wind Turbine Primary Regulation, in International Conference

on Power Engineering, Energy and Electrical Drives, pp. 495–500.

76. Western Power (2011). Technical Rules. Available at

www.westernpower.com.au.

77. Undrill J. (2010). Power and frequency control as it relates to wind-powered

generation, Lawrence Berkeley National Laboratory, Tech. Rep. LBNL-4143E

78. Martinez C., Xue S. and Martinez M. (2010). Review of the recent frequency

performance of the eastern, western and ERCOT interconnections, Lawrence

Berkeley National Laboratory, Tech. Rep. LBNL-4144E

79. Illian H. (2010). Frequency control performance measurement and requirements,

Lawrence Berkeley National Laboratory, Tech. Rep. LBNL-4145E. Available at

www.certs.lbl.gov

80. Mackin P. (2010). Dynamic simulations studies of the frequency response of the

three U.S. Interconnections with increased wind generation, Lawrence Berkeley

National Laboratory, Tech. Rep. LBNL-4146E

162
81. Garcia A. D. (2012). Models for Impact Assessment of Wind-Based Power

Generation on Frequency Control, In Control and Optimization Theory for

Electric Smart Grids, vol. 3, A. Chakraborty, and M. Ilić (Eds.), Springer-Verlag,

Berlin, 2012

82. Holttinen, H. (2005). Impact of hourly wind power variation on the system

operation in the Nordic Countries. Wind Energy, vol. 8, 2, 197–218

83. Ackermann T. (2005). Wind Power in Power System, John Willey & Sons.

84. NREL (1999). Technical report: Analysis of Wind Power Ancillary Services

Characteristics with German 250-MW Wind Data. Available at www.nrel.gov.

85. Thorncraft S. R. and Outhred H. R. (2007). Experience with Market-Based

Ancillary Services in the Australian National Electricity Market, Proceedings of

the IEEE PES General Meeting, Tampa Florida, USA, June 24-28

86. AMEC. (2011). Impact of the LRET on the costs of FCAS, NCAS and

Transmission augmentation. Available at amec.gov.au

87. NERC (2014). Violation Severity level Matrix-Reliability Standard. Available

at www.nerc.com

88. Wang, C. and McCalley, J. D. (2013). Impact of wind power on control

performance standards, International Journal of Electrical Power & Energy

Systems, 47, 225-234

89. ENTSO-E. (2004). UCTE Operational Handbook Policy 1: Load-Frequency

Control and Performance”- Appendix A1, V1. 9. Available at www.entsoe.eu

90. KEMA Inc. (2010). Research Evaluation of Wind Generation, Solar Generation,

and Storage Impact on the California Grid, Available at www.energy.ca.gov

163
CHAPTER 4

FREQUENCY RESPONSIVE FULL CONVERTER

BASED VSWTG: MODELLING AND ASSESSMENT

This chapter presents the generic modeling of type 4 full converter based wind turbine

generator system for application in frequency ancillary service investigations under

varying wind speed and varying reference power setpoint. Prevalent generic model,

manufacturer specific proprietary generic model along with detailed wind turbine

model with the synchronous generator is also provided to highlight various modeling

framework difference. Descriptions of individual sub-models of proposed generic

model are presented in detail and performance results are compared and validated

with GE’s proprietary generic model & detailed WTG model using simulations in the

MATLAB ® Power System Block set.

4.1. Introduction

After displacing dominant simple constant speed turbines, state of the art Variable

Speed Wind Turbine Generator (VSWTG) technology is a pioneer in currently

installed wind energy projects. Even though doubly fed induction generators based

type 3 is the predominant technology nowadays, full converter based type 4 turbine

technology is expected to become the industry standard in medium terms. Joint

Research Centre’s (JRC) 2014 report [1] analyzing wind turbine purchase agreements

identify the ending of pre-eminence of type 3 technology due to increasing popularity

of type 4 wind turbine system. Considering the connection trends, Australian Electric

164
Market Operator (AEMO) also assume an increase in type 3 and type 4 wind turbines

in its Wind Turbine Plant Capability Report 2013[2].

Figure 4-1 Schematic Diagram of Full Converter Based VSWTG

With relatively new wind technology, especially full converter based type 4 VSWTG

as shown in Figure 4-1, development of suitable models to analyze grid code

requirements is a significant subject matter. Not all models are suited for specific

power system phenomenon due to substantial issues. Generalized models having

reduced dataset necessities are employed to attain approximate dynamic wind farm

performance. A simplified yet absolute representation of full converter based wind

turbine generator principles is requisite for various bulk studies. Implementation of

suitable wind turbine model containing dynamic features satisfying ancillary service

requirements is the first step in such case. Very few research papers are available

dealing with generic models of full converter based WTG system which is a relatively

new technology. WECC’s generic model and GE model is the publically available

model while IEC model is not disclosed to the public. Most of the other available

research studies focus on controller development for frequency regulation utilizing

detailed model [13] [14] [15]. With no standard guidelines available, different

165
approaches have been used in adopting the generic model of type 4 WTG [3] [4] [5]

[6] [7] [8] [9], especially in initializing aerodynamic model values and application of

pitch control block. All of these models are aimed at short term simulation of the bulk

network for transient stability analysis under the assumptions of constant wind speed

and constant power setpoints. This chapter proposes a generalized model which

follows the basic structure of Western Electricity Coordinating Council (WECC)

generic type 4 wind turbine generators (WTG) model. Some modifications included

to make it suitable for frequency regulation studies. Model performance is evaluated

and validated with simulation response of GE’s proprietary generic model as well as

detailed type 4 wind turbine model.

4.2. Type 4 VSWT Model with Synchronous Generator

Figure 4-2 Detailed Simulation Model of Type 4 VSWTG

IEEE defines Type 4 wind turbine as a variable speed wind turbine with the

synchronous or asynchronous generator connected to the grid via full-scale power

166
converter [10]. Figure 4-2 shows the detail simulation model of full converter based

type 4 wind turbine generator model. These types of models are suitable for short

circuits analysis and electromagnetic transient investigations. A salient pole

synchronous generator integrated with grid via full capacity converters system is

utilized here in detail model of type 4 WTG systems. It also includes Cp calculation

based aerodynamic model and double mass based gear drive train model. Power

conditioning system consists of a diode rectifier, a DC-DC boost converter, and a

DC/AC IGBT-based pulse width modulated (PWM) inverter. Rotor speed is

completely decoupled from grid frequency through controlled frequency converter.

4.2.1. Converter Topology

Appropriate power converters with a full capacity similar to that of generators

implicitly segregate their dynamics from the main grid. A representation of converter

topology is given in Figure 4-3. With variable input supply, WES generator stator

provides 3-phase AC voltage of variable frequency and variable magnitude which is

later rectified into DC supply through 3-phase diode based Graetz bridge rectifier. A

regulated high DC voltage supply across capacitor suitable for the inverter is

subsequently achieved after being boosted through DC/DC single channel boost

converter. Steady state boost converter operation holds the following relation 4.1:

4.1

Chopper output voltage (Vout coincides with inverter dc bus voltage of inverter while

chopper input voltage (Vin is same as three phase rectifier output voltage. Input dc

voltage varies with generator speed while output dc voltage is maintained constant

higher value. The ratio of a turn on time to switching time period termed as duty cycle

(D) is attuned through a series of PI regulators for boost converter to control the

167
generator active power. IGBT-based PWM inverter using pulse width modulation

(PWM) employing 3 kHz carrier frequency then finally inverts dc voltage to essential

AC voltage and frequency ratings. This grid side converter maintains dc link voltage

by active power injection to the grid. The inductor smoothens any ripple effect in dc

link.

Figure 4-3 DC/DC Boost Converter & Decoupled VOC Inverter Interfaced Synchronous Generator

Wind Energy System.

Grid side connected IGBT inverter is controlled through voltage-oriented control

(VOC) with a decoupled controller. Closed loop current control of transformed 2-

phase line current in dq synchronous frame (i , i provides independent active and

reactive power control. Decoupled controller output can be expressed as [11]:

V K i∗ i WL i V 4.2

V K i∗ i WL i V 4.3

W is grid angular frequency as well as synchronous reference frame speed, W L i

& W L i are induced reference voltages in synchronous frame and K are PI

168
controller transfer function. Assuming balanced DC voltage, DC side voltage and

inverter AC line to neutral voltage can be related as [12] [13] [14]:


V m V 4.4

More details about converter operation can be referred from [11]

4.2.2. Synchronous Generator Model

The electrical part of the salient pole machine is represented by a fifth-order state-

space model considering the dynamics of the stator, field, and damper windings.

Computer simulation of synchronous machine model is implemented utilizing the

approach given in [15] where variables abc are transformed directly to rotor reference

frame using park’s transformation matrix Ks which has arbitrary reference frame

speed equal to rotor speed (ω = ω ) and presented in Figure 4-4.

Figure 4-4 Simulation Framework of Synchronous Machine in Rotor Reference-Frame [15]

Basic voltage and flux linkage are represented by equation 4.5 – 4.10 as given below:

v R ∗i ω ∗ φ, φ L ∗I 4.5

169
Where: fd = field winding, kd= d-axis damper winding, kq = q-axis damper winding, k

= no. of damper circuits, [R] = winding resistances diagonal matrix in d q axis

[L] = winding self and mutual inductances matrix in d q axis

[ω] = rotor speed matrix having all zeros except ω 1,2 ω ; ω (2, 1) =-ω

v = voltage vector = [V V V V V ]'

φ= d- and q- axis flux linkage vector = [φ φ φ φ φ ]'

i = current vector = [i i i i i ]'

With as base angular velocity corresponding to rated or base frequency,

manipulation of voltage and flux linkage equations gives following equations to be

used in simulation [15]:

φ V φ φ φ 4.6

φ V φ φ φ 4.7

φ V φ φ 4.8

φ e φ φ 4.9

φ V φ φ 4.10

i φ φ 4.11

i φ φ 4.12

170
i φ φ 4.13

i φ φ 4.14

i φ φ 4.15

ω T T 4.16

Under balanced operation, per unit electrical torque and electrical power output is

calculated as:

T φ ∗ i φ ∗ i 4.17

P Vi V i 4.18

4.3. WECC’s Type 4 WTG Generic Model

Implemented in Siemens PTI PSS®E and GE PSLF software, Western Electricity

Coordinating Council’s (WECC) Renewable Energy Modeling Task Force (REMTF)

developed reduced-order, positive sequence generic models [3] are appropriate for

dynamic stability evaluation of power system subsequent to system disturbances.

WECC’s generic model of type 4 wind turbines generator system pursues modular

approach where all imperative functions are characterized by individual submodels

working in unison.

Figure 4-5 represents WECC’ generic type 4 model consisting of three respective

submodels as implemented in PSLF. Generator/Converter dynamics block is

implemented with input active and reactive current command yielding real and

reactive power as an output. Second sub-block constitutes a simplified linearized

model for wind aerodynamics and two mass drive train model depicting torsional

171
mode oscillations. The electrical control model incorporating active and reactive

power control uses reference power as an input and current command as output. PSSE

model incorporates only electrical control block and generator-converter block.

Verification of this model’s performance amid unbalanced events is challenging due

to lack of any standard precept from WECC. Additionally, this model is assessed to

be incompatible for investigations associated with volatile wind speed amid dynamic

simulation. This generic model is suitable for bulk power system transient stability

analysis during the early stage of network interconnection planning but not suitable

for investigations like active power-frequency control analysis under varying wind

and changing power demand conditions.

Figure 4-5 WECC Type 4 Wind Turbine Generator Model [3]

4.4. Proprietary Type 4 Generic Model

GE’s type 4 generic model distinguish itself from WECC’s generic model in terms

of electrical control incorporating dynamic braking resistance power consumption,

172
additional torque controller and specific aerodynamic model based on two-

dimensional manufacturer specific curve given below:

C θ, λ ∑ ∑ α,θλ. 4.19

Figure 4-6 GE’s Type 4 Wind Turbine Generator Generic Model (Reformulated [5])

More details about the model can be found in reference work [5]. When power order

generated in this model is higher than the electrical power to be delivered to network,

excess energy is absorbed in braking resistance. So rotor model, as well as reference

speed generator, receives electrical power along with dynamic braking resistor

173
power. During real power priority, real current order, IPCMD is limited by a minimum

of a maximum of temperature dependent converter current and hard active current

limit. GE’s proprietary generic model which is applicable in bulk power system

disturbance studies is represented in Figure 4-6.

4.5. Proposed Type 4 VSWTG Generic Model

Figure 4-7 Proposed Type 4 WTG Model

Pursuing WECC second generation generic modeling framework, some additional

functionality like pitch control and available wind power look up has been

supplemented in proposed model making it admissible for frequency control and

associated investigation. Rotor speed is also incorporated in active power control loop

to account for fluctuating speed. The complete model is shown in Figure 4-7, and its

various sub-modules are discussed in succeeding sections. Salient aspects of WECC

174
generic WTG model and GE type 4 WTG model is distinguished with the proposed

full converter based WTG model and presented in Table 4-1.

Table 4-1 Comparison of Type 4 WTG Models Suitable for Frequency Regulation Analysis

WECC’s Generic Type 4 GE’s Proprietary Generic Modified Generic Type 4

WTG Model Type 4 WTG Model WTG Model

 Applicable for constant  Can be applied to variable  Aerodynamic model & pitch

wind speed evaluation. wind speed related model added making it

 First generation model investigations. suitable for varying winds as

excludes mechanical mass  Includes two-dimensional well as dynamic power

model, pitch model, and aerodynamic model coupled reference setpoints.

aerodynamic model while with double mass model &  Available wind power lookup

second generation model pitch model makes it suitable for primary

includes double mass drive  Normal as well as de-loaded and secondary frequency

train model. operations are possible control assessment

 Applicable to normal  The model utilizes active and investigations. Nominal

operation only. reactive power controller in value or calculated value is

 It utilizes active and reactive the electrical circuit for utilized depending upon wind

power controller in the transient analysis. speed.

electrical circuit for  No study available for  The model utilizes simplified

transient stability analysis. assessment of model for electrical control with only

 The model requires varying reference power active power controller in the

additional modifications for command / setpoints. electrical circuit for

application in frequency frequency control analysis.

regulation investigations. Rotor speed variation is an

additional component added

to the electrical control.

Detailed Model: Mainly applied for short circuit electromagnetic transient analysis over the shorter

time frame. Due to longer simulations times and complex sub-component representation, detail type

4 models are not deemed suitable for primary & secondary frequency control investigations.

175
4.5.1. Wind Turbine Mechanical Model

Wind turbine’s limitation to inertia control contribution can be contemplated through

rotor speed dynamics making the coupling imperative between mechanical drive train

model and the aerodynamic model for accurate speed dynamics reflection. The

mechanical model provides significant turbine dynamics representation associated

with WTG instantaneous shaft speed. Combination of tip-speed ratio calculation

block, power coefficient calculation block and aerodynamic power calculation block

constitute the complete wind turbine model. One mass lumped mechanical shaft

model as given by equation [16] [17] is implemented in this generic WTG model.

ω
T T Dω 4.20

Where H is the total inertia constant given by the ratio of the stored kinetic energy

(MJ) at synchronous speed to the machine rating in MVA. D is the shaft damping

constant in p.u. power / p.u. speed with a default value of 1 while ω denotes the

angular speed. This lumped model assumes constant turbine and generator angular

speed. Aerodynamic model intricacy is highly influenced by the representation of the

power coefficient Cp. Cp is the dimensionless fractional value conveying the amount

of available wind power that can be extracted by the rotor and given as the ratio of

the available wind power to the available wind kinetic power.Betz limit defines the

optimal value of performance coefficient as 59.26% which is applicable to all wind

turbine. WECC proposed a simplified aerodynamic model in generic modeling while

GE’s generic model utilize two-dimensional Cp curve model. WECC proposal of

linear aerodynamic model doesn’t provide any publically available standard

guidelines initializing the model.Initialization of pitch angle and power in linear

aerodynamic model finds scant mentions in current research literature. Suitability of

WECC simplified the linear aerodynamic model, and detailed Cp calculation based
176
model [18] [19] like the one given by 4.21 & 4.22 is assessed in this study before

making final recommendations:

C λ, θ C C θ C e + C λ 4.21

.
4.22
.

The empirical power coefficients c1 to c6 are: c1 = 0.5176, c2 = 116, c3 = 0.4, c4 =

5, c5 = 21 and c6 = 0.0068. λi is tip speed ratio at ith time step. Cp acheives a

maximum value of 0.48 for zero degree pitch angle and nominal value of lambda =

8.1. Maximum extractable wind power defined as turbine mechanical power utilizing

detailed Cp based aerodynamic is given as 4.23 .

1
Pwind = ρ ACp ( , θ V 4.23
2

Where ρ is the air density, A is the turbine blades swept area (πR2 where R is blade

length), Cp is the turbine power coefficient, λ is tip speed ratio defined as rotor

peripheral speed to the wind speed, Vwind is the velocity of the incident wind. At very

high wind speeds, the power is limited to rated power by increased pitch angle. At

very low wind speeds, an increase in pitch angle increases the value of Cp thus

increasing power captured from the wind. Simplified linear aerodynamic model [6]

expressed by 4.24 has reduced parameters leading to lesser simulation time and based

on the assumption of constant wind during initialization without involving power

coefficient curve.

A simulation model of the linear aerodynamic model is shown in Figure 4-8.

P P K ∗ θ 4.24

Where & θ is the initial mechanical power and initial pitch angle respectively.

- Kaero is the aerodynamic gain factor

- θ is the pitch angle of the turbine blades

177
The aerodynamic gain factor Kaero has been given a default value of 0.007 [16-17].

θ ∗ 1
.

The value of Theta2 supplied is 26◦. Vwindpu is the velocity of the wind in per-unit

of the rated wind velocity (13 m/s).

P _ k C _ V _ 4.25

Figure 4-8 Simplified Aerodynamic Simulation Model

Equation 4.25 is the normalized form of equation 4.23, where Pm_pu is per-unit

nominal power for particular values of ρ and A, Cp_pu is per-unitized performance

coefficient of the maximum value of Cp, Vwind_pu is per-unit wind speed in p.u. of the

rated wind speed. Kp is is less than or equal to 1 and denoted as power gain for Cp_pu=1

p.u. and Vwind_pu = 1 p.u. As seen in Figure 4-9, the power generated is proportional

to available wind power i.e.; Ppu = V for underrated wind speed and power

generated is Ppu = 1 for beyond rated wind speed. The actual mechanical power

delivered by the wind turbine to the generator is thus obtained by subtracting per unit

wind power from maximum available wind power.

178
4.5.2. Pitch Control Model

At pitch angle=27 degree At pitch angle=0 degree

Figure 4-9 Turbine Power Characteristics

Figure 4-10 Pitch Controller Block

Figure 4-9 shows the per unitized turbine output power w.r.t to mechanical power at

pitch angle = 0 degrees and pitch angle =27 degree. Pitch control enables optimum

control of aerodynamic wind power by rotor blade pitching to regulate turbine torque.

Maximum power of 1 p.u. is available at zero pitch angle while it is highly reduced

with highest pitch angle. Though pitch control is not adopted in WECC generic type

4 model, a simple variable pitch controller without gain scheduling has been proposed

179
for fluctuating wind speed adaptable WTG model. Simplified aerodynamic model in

the proposed WTG model inputs pitch angle at zero in the case of low wind speed

otherwise amplitude & rate limited pitch angle is initialized through pitch controller

in case of wind speed more than rated speed. Following equations mathematically

express pitch angle obtained through a series of the PI regulator as shown in Figure

4-10:

.
θ K ω ω . K P

P 4.26

Default speed reference calculation according to generated electrical power is done

as per equation:

ω A A P A P 4.27

Where , and are coefficients for turbine operation at maximum power.

λ= 4.28

ω v 4.29

For optimal operation and per unit base:

ω v C v 4.30

Similarly, we can write equation regarding power during optimum operation as:

,
P P , C v 4.31
,

Combining these equations, we can write following equation for three different wind

speeds:

v ,
1 C v , C v , A
C v , 1 C v , C v , A 4.32
v , A
1 C v , C v ,

180
With respect to zero pitch angle, rated wind speed will give rated rotational speed and

rated power thereby converting equation 4.30 and 4.31 as:

C 4.33

C 4.34

Substituting rated wind speed in 4.33-4.34 and other two wind speed values in

medium wind speed range in 4.32, rotor reference speed can be obtained.

Under maximum Cp of 48%, this study uses a reference speed given by:

ω 0.79131P 1.526046P 0.49188 4.35

under the assumption 0.15 p. u. P 0.75 p. u. For P 0.75 p. u, ω

1.2 and P 0.15, ω 0.689 p. u.

Pitch compensation is also introduced in conventional pitch control to vary pitch

angle during active power control so as to aid desired frequency regulator operation.

Required generation MW capacity is adjusted through the variable Pset. It is initialized

through external active power reference control model to provide governor control if

frequency regulation is objective otherwise it is kept fixed at one p.u. The Pset setting

can be used to curtail maximum output of wind turbine such that blade pitch is set to

produce P < Prated under normal situation.

4.5.3. Electrical Control Model

As discussed in the previous section of the detailed model, type 4 VSWTG has very

fast flux dynamics, and machine side converters are completely decoupled from grid

side converters. Hence the type of generator, as well as machine side converter, does

not significantly influence the total system behavior towards grid. Considering the

requirement of an appropriate WTG model for frequency regulations related analysis

rather than short circuit transient analysis, a simplified static generator model is

181
adopted in proposed model. Reduced order model reduction for the full converter is

presented in [21]. This study suggested usage of either equivalent Thevenin voltage

source model or Norton current source model representing generator and converter

which is subsequently followed by a low-pass filter that outputs reference current ip

to network according to input active current command i [21]. Under normal

operation, the active current is given priority for the availability of maximum

available active power to the grid while during voltage disturbances, the reactive

current has high priority for extra reactive current transfer to improve the grid voltage.

Figure 4-11 Generator-Converter Model [20]

During P-priority, the upper limit of the active current IPmax is dependent upon lumped

converters ratings Imax while the maximum reactive current limit IQmax is determined

based on both the lumped converter rating Imax and the active current command.

182
I I , I I I 4.36

With active current priority in frequency regulation mode, all corresponding flux

dynamics have been avoided to reflect swift power command transfers from the

electrical controller to converters. Power from Generator/Converter model is given

as P V ∗ i and corresponding equivalent framework of generator-converter and

their interface to the network is given by Figure 4-11.

Complete electrical model as proposed by WECC and GE generic model includes

reactive power control along with active power control which is achieved by

active/reactive current injection into the grid as per controller commands. With the

assumption that frequency regulation is valid only when active power is priority and

system have near nominal voltage ratings, reactive control with voltage logic blocks

has been ommitted in proposed model for simplicity. Dark blocks show the electrical

model as used in proposed model. The principle feature of the converter model is

incorporated by imposing the max. current limit of 1.1 p.u.

4.5.3. Active Power Control Loop Co-ordination with Plant Controller

Frequency control tends to be operational during normal grid operation with peak

weightage given to real power transfer through activated real power controller.

Real/active power controller exploits power error signal achieved from the difference

of electrical power and reference power setpoint as provided by plant controller.

Generator speed is also included in the loop to elucidate the shaft power oscillation

effects in generated electrical power. WTG system power order from the PI output is

passed through low pass filter and rate limited. Division of power order through

terminal voltage provides desired active current command signal i by a terminal

voltage which is further rate limited by maximum current obtained from the reactive

183
control loop. A reference power setpoint Pset with value varying from 0.5-1 is used

as plant reference in this model, but plant controller can set this value according to

demanded power from system operator as per respective operation like frequency

control, absolute power, Delta power, etc. In the simplest form, it can also be set as a

minimum of nominal WTG power and optimal aerodynamic power achieved from

maximum power point tracking [22]. Virtual inertia control and active power

frequency control can also be added further in this model. Plant controller supporting

active power control is illustrated in Figure 4-12.

Figure 4-12 Active Power Control

4.6. Model Validation and Analysis

Detailed WTG model requires small time step, and this results in large simulation

time consumption. In the current study, different simulation periods have been used

for the generic model assessment. A set of response simulations are performed to

evaluate & compare the proposed generic model in terms suitability of linear

aerodynamic model with the detailed aerodynamic model, suitability of single mass

drive train model in place of double mass drive train model and suitability of static

generator converter model in place of dynamic generator converter model. All results

are compared with proprietary and detail VSWTG model.

184
4.6.1. Aerodynamic Model Suitability Assessment

Mechanical power harvested from the wind depends upon available wind power,

machine power coefficient and ability to respond wind fluctuations. Various

numerical approximations have been proposed in the literature for determining the

coefficient of performance. Simplified aerodynamic model, as well as detailed model

aerodynamic model based on Cp calculations, is assessed for their utilization in

generic type 4 WTG model. Generic model performance characteristics are analyzed

under three major categories: variable wind speed & variable power setpoint, constant

wind speed & constant power setpoint, variable wind speed & constant power

setpoint to assess the performance regarding rotor speed, electrical power generated,

mechanical power and pitch angle.

4.6.1.1. Simplified Linear Aerodynamic Model

WECC proposal of linear aerodynamic model doesn’t provide any publically

available standard guidelines initializing the model. Initialization of pitch angle and

power in linear aerodynamic model finds scant mentions in current research

literature. According to [7], for less than nominal wind speeds, a linear relation exists

between mechanical power & wind speed while a linear relationship between pitch

angle and wind speed can be assumed for wind speeds above nominal value for

initialization of wind turbine model. Another detailed analysis of the simplified

aerodynamic model is provided in [16]. Linear aerodynamic model is dependent upon

initial mechanical power and changing pitch angle to calculate total mechanical

power delivered to the turbine. With dependence on constant wind speed only,

caution is needed in initializing pitch angle in this model to achieve limited accuracy.

185
In the current study, initial mechanical power is initialized according to available

wind power for wind speed less that rated speed while it is limited to rated mechanical

power for higher wind speed. At constant wind speeds, the active power controller

can maintain generated power at 1.0 p.u. and rotor speed at 1.2 p.u. for constant as

well as varying power setpoint which can be observed in Figure 4-13B - 4-13E. In

Figure 4-14, when the wind speed change from rated value of 13m/s to 8 m/s and then

shooting up to the 18m/s and settling at 25 m/s, pitch angle fails to gain as per set

algorithm rapidly. Slow acting pitch control is not able to follow the dynamic, varying

power setpoint and wind speed. Further simulation investigation reveals that transfer

of pitch angle without consideration of rated wind speed dependence in linear

aerodynamic model results in inadequate performance as compared to detail

aerodynamic model.

Since application of the simplified aerodynamic model is limited to constant wind

speed; dynamic studies under stochastic wind conditions like frequency regulations

call for a generic model which is fully adaptable to varying power setpoints and

varying wind conditions.

A. Varying Power Setpoint

186
B. Electrical Power at Different Wind Speeds Ranging from 8-20 m/s and Varying Power Setpoint

C. Mechanical Power at Different Wind Speeds Ranging from 8-20 m/s and Varying Power Setpoint

D. Pitch Angle at Different Wind Speeds Ranging from 8-20 m/s and Varying Power Setpoint

187
E. Rotor Speed at Different Wind Speeds Ranging from 8-20 m/s

Figure 4-13 Type 4 Generic Model Incorporating Linear Aerodynamic Model Performance
Characteristics at Constant Wind Speed at 8 m/s, 13 m/s, 18 m/s and 20 m/s Along with Varying Power
Setpoint

Figure 4-14 Type 4 Generic Model Incorporating Linear Aerodynamic Model Performance

Characteristics under Varying Wind Speed & Varying Power Setpoint

4.6.1.2. Detailed Aerodynamic Model

There are various numerical approximations proposed in literature for calculating C .

Linear aerodynamic model is then replaced with more detailed C calculations based

model in the second stage. Two different detailed aerodynamic models are tested here

in this study. Aerodynamic Model 1 is based on operation at maximum C calculation

model which is highly dependent on wind speed at the nominal speed and at

maximum C along with proper initial wind speed initialization as shown in Figure

4-15. Aerodynamic model 2 is simple C Calculation model as given in Figure 4-16.


188
Figure 4-15 Coefficient of Power Calculation Model for WTG System

These models are simulated and analyzed with constant wind speeds of 9 m/s, 13 m/s,

18 m/s & 25 m/s as well as varying wind speed from initial 8 m/s to final 20 m/s in

various time steps. Power setpoint used in pitch controller and plant controller is also

varied from 1 p.u. to 0.6 p.u. Both C calculation models give acceptable

performances but due to limitations of aerodynamic model 1 dependence upon base

wind speed and proper initial wind speed initialization put some limitation on its

applicability.

Figure 4-16 Coefficient of Performance Model 2 for WTG System [23]

189
A. Rotor Speed at Different Wind Speeds Ranging from 9-25 m/s and Varying Power Setpoint.

B. Electrical Power at Different Wind Speeds Ranging from 9-25 m/s and Varying Power Setpoint

C. Mechanical Power at Wind Speeds Ranging from 9-25 m/s

190
D. Pitch Angle at Different Wind Speeds Ranging from 9-25 m/s

Figure 4-17 Type 4 Generic Model Incorporating Detailed Aerodynamic Model 1 Performance

Characteristics at Wind Speed of 11 m/s, 13 m/s, 18 m/s and 25 m/s along with Variable Power Setpoint

Figure 4-17 shows various performance results achieved by applying aerodynamic

model 1 with an initial wind speed of 8 m/s. WTG performance results in changes

with a change in initial and base wind speed. It is also observed that this model fails

to provide proper results in mechanical power and corresponding rotor speed and

pitch angle if input wind speed is less than base wind speed and initial wind speed so

proper wind speed initialization is required in adopting aerodynamic model 1.

Figure 4-18 shows mechanical power and pitch angle variation for both aerodynamic

models under constant wind speed and constant power set-point. It can be seen from

these Figures that both aerodynamic models provide similar performance for

mechanical power. A slight variation in speed of pitch angle variation response is

observed with aerodynamic model 2 giving faster response at higher wind speeds.

Type 4 WTG with both aerodynamic models were also simulated at variable wind

speed and variable power set-points as shown in Figure 4-19 – 4-20. The close match

between mechanical power and electrical power is observed when applying

aerodynamic model 1 while the slight oscillatory response is observed in

aerodynamic model 2.

191
A. Mechanical Power Variation for Aerodynamic Model 1

B. Mechanical Power Variation for Aerodynamic Model 2

C. Pitch Angle from Aerodynamic Model 1

192
D. Pitch Angle from Aerodynamic Model 2

Figure 4-18 Type 4 Generic Model Incorporating Detailed Aerodynamic Models at Different Constant

Wind Speed along with Constant 0.8 Power Setpoint

A. Type 4 WTG Characteristics at Variable Wind and Variable Power Setpoint with Aerodynamic
Model 1

B. Type 4 Characteristics at variable wind and variable Power Setpoint [Zoomed]

Figure 4-19 Type 4 Generic Model Incorporating Detailed Aerodynamic Model 1 Performance
Characteristics

193
A. Type 4 Characteristics at Variable Wind and Variable Power Setpoint with Aerodynamic

Model 2

B. Type 4 Characteristics at variable wind and variable Power Setpoint with Aerodynamic

model 2 [Zoomed]

Figure 4-20 Type 4 Model Characteristics with Detailed Aerodynamic Model 2 Performance

Characteristics

Comparing overall performance, detailed aerodynamic model combined with single

mass drive train model show the best combination for dynamic studies like active

power frequency regulation analysis under varying wind and power setpoints. This

result contrasts with the simplified aerodynamic model as proposed by WECC

generic model for dynamic studies of longer time frames.

194
4.6.2. Generator-Converter Model Comparison with GE’s Generic

Model and Detailed Model

Results of proposed model with both types of aerodynamic models are compared with

detailed WTG model and GE’s generic model performance in next stage. It is to be

noted that detailed model includes a synchronous generator with dynamic converter

topology while GE’s model includes braking resistance absorbed power in

calculating reference speed and corresponding power order. Static generator

converter model has been used in proposed type 4 generic model. Generated electrical

power in per unit quantities obtained from proposed model with both types of

aerodynamic models under different wind conditions are comparable with other two

WTG models. It can be seen from Figure 4-21B that 1 p.u. electrical power is

achieved in 10-15 sec for wind speed equal or greater than rated wind speed from GE

WTG model while Figure 4-22A & 4-22 B show that same 1 p.u. electrical power is

achieved within 15 sec. for detailed WTG model due to faster converter dynamics.

Comparable electrical power in p.u. is also achievable from proposed type 4 model

within 15-20 sec. as seen in Figures 4-17, 4-19 & 4-20.

A. Rotor Speed

195
B. Electrical Power at Different Wind Speeds Ranging from 8-20 m/s

C. Pitch Angle

D. Mechanical Power

Figure 4-21 GE’s Type 4 Generic Model Performance Characteristics at Different Wind Speed Ranging

From 8 m/s – 20 m/s

196
A. Type 4 WTG Detailed Model Characteristics at Wind Speed= 13 m/s

B. Type 4 Detailed Model Characteristics at Wind Speed= 18 m/s

Figure 4-22 Type 4 Detailed EMT Model Incorporating Synchronous Generator Characteristics at

Constant Wind Speed of 13 m/s &18 m/s.

It can be established from the simulation results that static generator-converter model

with a detailed aerodynamic model based on pitch control block makes the best

combination to achieve desirable results under the varying wind, and varying power

reference setpoints, hence making complete model valid and suitable for power

frequency regulation analysis.

4.6.3. Drive Train Mass Model Comparison with GE’s Generic Model

& Detailed Simulation Model

Drivetrain comprising rotating parts (Turbine-Generator rotor and gearbox)

constitutes one of the mechanical sub-block of WTG system. Implementation of the

197
detailed mass model is a topic of discussion for WECS modeling. Detailed

mechanical mass model is important for analyzing transient loads while reduced order

model can be applied in bulk network analysis to speed up the simulation. Third stage

analysis was regarding mass model applicability in the generic model. Single mass

model for drive train oscillations has been used both in proposed model as well as in

GE’s model while two mass drive train model has been used in detailed` model.

Figures 4-13E, 4-17A, 4-19 & 4-20 shows a variation of rotor speed under different

wind speeds with satisfactory results in comparison to rotor speeds obtained from GE

model and detailed model. Some oscillations are observed in mechanical power at the

higher cut-off wind speed of 25 m/s when rotor speed goes up to 1.85 p.u.

With converters isolating machines from the grid, one mass model is sufficient to

analyze the performance of type 4 wind turbine system in long time frame dynamic

studies.

4.6.4. Virtual Inertia Support Assessment

Wind turbines provide an artificial inertial response through additional control feature

called virtual inertia as represented in Figure. 3-12, Chapter 3. Virtual inertia controller

introduces an additional power signal to reference power setpoint upon detection of

system frequency deviation thereby momentary increase in wind turbine output power.

Four virtual inertia models are tested for proposed type-4 model under 0.8 power

setpoint with satisfactory results. Three of these inertia models are provided in Figure

3-13 of Chapter 3. Inertia Model 1 incorporates virtual inertia model from GE with a

dead band of 1 which is denoted as inertia model in Figure. 3-13 while basic model

[24] is incorporated into inertia model 2 providing supplementary inertial control

through frequency deviation and rate of change of frequency. This model is referred

198
as inertia droop model in Figure 3-13. Inertial model 3 (referred as droop model 2 in

Figure 3-13) takes frequency deviation as input and has only droop control enabled

while inertial model 4 incorporates only inertial control with the rate of change of

frequency as input. Uncontrolled three area power system model under 10 % load

disturbance is utilized for providing test frequency deviation signal as given in Figure

4-23 for analyzing virtual inertia model responses. A constant wind speed of 13 m/s

and constant power setpoint of 0.8 p.u. is applied in the simulation. It should be noted

that wind turbine model is not feeding electrical power to the network during the

simulation.

Figure 4-24 presents the response of GE type-4 model where a momentary increase in

electrical power and mechanical power is observed with a very small decrease in rotor

speed of around 0.002pu. Results of four virtual inertia models in proposed type 4

model are compared without virtual inertia support and given in Figure 4-25. All

virtual inertia models can provide a higher electrical output during frequency

deviations with highest electrical power change of around 0.08 p.u. obtained from

applying GE wind inertia model. Control features in proposed model can maintain

rotor speed within the limit of 1.2 p.u. after a small speed drop during the frequency

event showing the sufficiency of the proposed type 4 model.

Figure 4-23 Uncontrolled Three Area Interconnected System Frequency During 10% Load Change

199
A. Electrical Power

B. Rotor Speed

C. Mechanical Power at 13 m/s Wind Speed

200
D. Mechanical Power at 13 m/s Wind Speed (Zoomed)

Figure 4-24 GE Model Performance with/without Inertia Support

A. Rotor Speed

B. Rotor speed (Zoomed)

201
C. Electrical Power

D. Electrical Power (Zoomed)

E. Mechanical Power (Zoomed)

202
F. Pitch Angle (Zoomed)

Figure 4-25 Type-4 Model performance with/without Inertia Support

4.7. Conclusion

Most of the frequency regulation research studies to date have been done utilizing an

only generic model of type 3 doubly fed induction generator. A simulation model

modified from the basic structure of WECC’s type-4 wind turbine generator has been

proposed here which can be utilized in wind penetration impact studies as well as

frequency regulation studies. The performance of this generic model is evaluated

through various simulation sets under different wind speeds and power set-points

combinations for rotor speed variations, generated electrical power, developed

mechanical power, pitch angle variation, and virtual inertia support. Simulation

results are also compared with GE’s proprietary generic WTG model and detail

model to reflect the adequacy of proposed model.

Simulation results show that detail aerodynamic model along with single mass drive

train model gives acceptable performance highlighting their adequacy and best

combination in generic models for frequency control studies.

The generator converter effect of designed model is also found to be comparable with

manufacturer specific generic model disabling dependence on propriety information.

203
4.7. References

1. JRC (2014). JRC Wind Status Report, Publications Office of the European Union.

2015

2. AEMO (2013). Wind Turbine Plant Capabilities Report. Available at

aemo.com.au

3. Pouyan Pourbeik (2014). Specification of second generation generic models for

wind turbine generators, EPRI-WECC.

4. UWIG. Type 4 - Generic Wind Turbine Generator Model (Phase II), Available at

www.uwig.org:8080.

5. K. Clark, N. W. Miller, J. J. Sanchez-Gasca (2010). Modeling of GE Wind

Turbine-Generators for Grid Studies, Version 4.5, General Electric International,

Inc.

6. W. W. Price, J. J. Sanchez-Gasca (2006). Simplified wind turbine generator

aerodynamic models for transient stability studies, Power Systems Conference

and Exposition, 2006. PSCE '06. 2006 IEEE PES, vol., no., pp.986, 992.

7. M. Singh, M. Vyas, S. Santoso (2010). Using generic wind turbine models to

compare inertial response of wind turbine technologies, in Power and Energy

Society General Meeting, 2010 IEEE, vol., no., and pp.1-7.

8. Hansen AD, Margaris ID. (2014). Type IV Wind Turbine Model. DTU Wind

Energy.

9. M. Singh and S. Santoso (2011). Dynamic Models for Wind Turbines and Wind

Power Plants.

204
10. IEEE (2009). Tutorial on Wind Generation Modeling and Controls, IEEE PES

PSCE.

11. Bin Wu, Yongqiang Lang, Navid Zargari, Samir Kouro (2011). Power

Conversion and Control of Wind Energy Systems, John Wiley & Sons.

12. R. Pena, J. C. Clare, G. M. Asher (1996). Doubly fed induction generator using

back-to-back PWM converters and its application to variable-speed wind-energy

generation, in Electric Power Applications, IEE Proceedings -, vol.143, no.3, pp.

231-241.

13. Juan W. Dixon and Boon-Teck Ooi (1988). Indirect Current Control of a Unity

Power Factor sinusoidal Current Boost Type Three-phase Rectifier, IEEE

Transactions on Industrial Electronics, Vol. 35, No. 4, pp. 508-515.

14. Marcelo Gustavo Molina and Pedro Enrique Mercado (2011). Modeling and

Control Design of Pitch-Controlled Variable Speed Wind Turbines, Wind

Turbines, Dr. Ibrahim Al-Bahadly (Ed.), InTech.

15. P. C. Krause (1986). Analysis of Electric Machinery, McGraw-Hill, Section 12.5

16. Jens Fortmann (2014). Modeling of Wind Turbines with Doubly Fed Generator

System, Springer.

17. Generic Type-3 Wind Turbine-Generator Model for Grid Studies (2006). WECC

Wind Generator Modeling Group, Version 1.1, Sept. 14.

18. Siegfried Heier (1998). Grid Integration of Wind Energy Conversion Systems,

John Wiley & Sons Ltd, ISBN 0-471-97143-X

19. J. G. Slootweg, S. W. H. de Haan, H. Polinder and W. L. Kling (2003). General

Model for Representing Variable Speed Wind Turbines in Power System

205
Dynamics Simulations, IEEE Transactions on Power Systems, Vol.18, No.1,

pp.144-151.

20. Seman, Slavomir (2011). Need for confidentiality. A converter manufacturer’s

view, Wind Integration Symposium, 16-17 June.

21. Fortmann, Jens et al. (2010). Generic simulation model for DFIG and full size

converter based wind turbines. 9th International Workshop on Large-Scale

Integration of Wind Power into Power Systems as well as on Transmission

Networks for Offshore Wind Power Plants, Quebec City.

22. Spudic, Vedrana et al. (2010). Hierarchical wind farm control for power/load

optimization. The Science of making Torque from Wind (Torque2010).

23. J. G. Slootweg, H. Polinder, W. L. Kling (2001). Dynamic modeling of a wind

turbine with direct drive synchronous generator and back to back voltage source

converter and its control, Proc. of the European Wind Energy Conference, pp.

1014-1017.

24. J. Morren, S. W. H. de Haan, W. L. Kling, J. A. Ferreira (2006). Wind turbines

emulating inertia and supporting primary frequency control, IEEE Transactions

on Power Systems, vol.21, no.1, pp.433-434.

206
CHAPTER 5

DYNAMIC DEAD-BAND DEPENDENT GRID CODE

FREQUENCY SENSITIVE ACTIVE POWER

CONTROLLER

Adding to the study line of active power frequency control in wind turbines which

was presented in Chapter 3, a grid frequency processor scheme based on dynamic

dead-band around moving averaged frequency instead of conventional static dead-

band for generation of continuously varying, frequency sensitive active power

reference setpoint for VSWTG is presented in this chapter. Grid frequency processor

algorithm based on dynamic dead-band is implemented for two different grid codes

to generate active power setpoints which are supplied to a modified speed–pitch

control loop based VSWTG model. Area frequency response along with VSWTG

electrical power support will be compared with other droop based VSWTG model in

next chapter 6 to establish the superiority of proposed frequency-active power

controller based VSWTG over other droop based VSWTG models.

5.1. Introduction

Frequency responsive active power control has advantages in the form of minimized

transmission lines overloading, improved power quality standard, power ramps

limitation can avoid power surges or reduced power unavailability. The conventional

power plants which remain connected to the grid even in the case of frequency

deviations accomplish active power-frequency control through turbine governors

reaction to nominal synchronous speed deviations and the respective boilers-turbine

frequency responsive controls. Dead-band and speed droop are two significant
207
parameters in conventional turbine governor action during system event [1]. Long-

term system frequency regulation reliability is highly dependent upon implemented

dead-band. Governor dead band is detrimental to minimum frequency variation

needed before the governor action is activated. The relative frequency deviation with

respect to the relative change in power output defined as speed droop is always

positive for stable regulation. Factors like the available head room, the maximum-

minimum power capacity of generating units and power setpoint corresponding to

enabled frequency operating mode have a direct effect on total frequency responsive

reserve amount. A generator operating at its maximum generating capacity has almost

negligible headroom and therefore unable to provide any frequency responsive

operation irrespective of governor-droop being enabled.

Some of the above-mentioned parameters are needed to be considered in VSWTG

modeling for frequency-active power control similar to that provided by synchronous

generator. Active power contribution to frequency deviation was earlier only

acknowledged in trivial islanded power systems, but the impending system inertia

reduction with increased integration of power electronics based wind power plants

has led transmission system operators (TSO) to establish new grid codes for

frequency based ancillary services from wind farms in the big interconnected

electrical grid. Given compulsory grid code frequency support, these power

electronics based VSWTG requires modified active power control algorithm to

support regulation services. Frequency-active power control model is an auxiliary

control algorithm implemented in individual WTG control loop for providing

controllable power reserve on demand in the form of spinning reserve or power ramp

rate limit to respond system frequency deviations.

208
The operation of the VSWTGs in maximum power tracking mode results in zero

spinning reserve for utilization in frequency regulation. A brief discussion about wind

turbine participation in primary frequency control through de-loading was presented

in chapter 3. Wind turbines are required to operate at reduced power output to

establish power reserve. As discussed earlier, various methods are present in literature

for the de-loading operation of wind turbines. Separate pitch control or rotor over

speeding de-loaded control are most common approach discussed in previous studies.

Owing to high dependence on available power reserves upon maximum allowable

rotor over-speed, de-loading of wind turbine finds the limited practical application.

Moreover, dependence upon wind speed measurement still needs more scientific

investigations as inaccurate measurement may highly affect wind turbine output and

turbine life.

5.2. Grid Code Sensitive Frequency Controller

5.2.1. Dead-Band: Static vs. Dynamic

Generator droop characteristic means that generating unit will inversely change power

output level in proportion to system frequency. Responding to frequency deviations,

generating units depart from their dispatch targets according to the set droop

characteristics. This droop behavior can be avoided by implementing a dead-band on

each unit. Static dead-band is thus asymmetrical buffer zone on either side of 50/60

Hz frequency that compensates for frequency noise. Droop response of generating

machine from controlling the power output of the machine is avoided when frequency

lies within the upper and lower dead-band range. Dynamic dead-band incorporates a

buffer zone on either side of moving averaged frequency signal instead of the nominal

frequency signal. Both these dead bands are represented in Figure 5.1.

209
5.2.3. Frequency Processing Model

Figure 5-1 NEM Mainland Frequency 1045 hrs. To 1109 hrs. During Contingency Event [1] 

Providing grid frequency directly for winding turbine controller will generate noise

induced frequency sensitive active power setpoint. This will result in noisy output

power thus providing frequency response for any frequency error even due to grid

noise. Noisy power output will have an adverse effect on supplied power quality and

turbine life. An analysis of electrical grid dynamics can serve as the base of grid

processing system for distinguishing different types of frequency signals. Figure 5-1

shows the NEM mainland real frequency trace during 2009 contingency event when

3205 MW of total generation disconnected automatically resulting in under frequency

load shedding across the Australian NEM interconnected system [1]. Though we

could not get enough data points to represent a very clear picture here still three

different frequency signals 1, 2 and 3 as defined below can be distinguished in this

Figure. Electrical grid frequency signal can be considered as composed of three types

of signals [2]:

210
1. Low ramping signals arising due to normal trend of generation-demand and

grid dynamics over the long term of several minutes.

2. High amplitude, high-frequency range during contingency which is normally

low in occurrence.

3. Low amplitude signals in high-frequency range due to stochastic grid noise

with high occurrence rate.

Figure 5-2 Basic Structure of Grid Frequency Processor

First two signals have a very low effect on turbine life while highly occurring type 3

noise signals have ample effect on turbine lifetime and should be suppressed. A closer

look at Figure 5-1 shows that usual trend frequency signal 1 normally lies outside

static dead-band and high amplitude signal 2 always lies outside static dead-band.

Discerning these signals just by frequency through filters is not a viable option as

they are present in the whole frequency signal over the time period. Implementing a

dynamic dead band centered around the moving averaged control frequency signal

named as trend frequency signal instead of using a static dead-band in a grid

processing system would distinguish these types of signals. Given the above-

mentioned problem, a signal processing system centered around moving averaged

grid frequency signal was proposed in [2]. A modified version of the frequency
211
processor is implemented in the current study for utilization in active power setpoint

controller for VSWTGs operating under frequency responsive mode.

A grid frequency processor block which provides dynamic dead-band based

processed frequency output is shown in Figure 5-2. Electric grid frequency fluctuates

around the nominal frequency of 50 Hz or 60 Hz. Moving average of this measured

grid frequency can be termed as a trend frequency which characterizes the long-term

behavior of grid frequency. Frequency processor takes measured grid frequency and

nominal frequency as an input signal. The moving average filter (MAF) averages

some points from the measured frequency thereby producing every point in the output

frequency. Mathematically expressed as

yi ∑ yi j , 5.1

Where y is output frequency signal, M is a number of points in average [3].

MAF which can be simulated in various ways like finite impulse response filter

produces lowest noise component in the output signal by equally treating all incoming

signal. A low pass filter with a suitable time constant has been used to act as MAF to

construct trend frequency in this study. Resulting trend frequency signal is subtracted

from the filtered measured frequency and resulting absolute frequency difference is

then passed through dynamic dead-band.

Frequency responsive mode operates by replacing static dead-band and implying

dynamic dead band centered around the control trend frequency signal instead of set

frequency signal. With a suitable threshold value, the output frequency is passed out

as control trend signal if input measured frequency signal follows the control trend

signal within dynamic dead-band otherwise input measured signal is passed out as an

output signal. In this way, noise induced frequency signals are filtered out, and high

amplitude frequency signals, as well as normal frequency signal, can be passed to

212
next block for generating active power set-point as per implemented grid code

algorithm.

An example of frequency which is obtained from NEM 14-bus model and provided

as input signal to frequency processor is shown in Figure. 5-3 while output processed

frequency signal is shown in Figure. 5-4. Frequency disturbances are removed

applying moving averaged dead-band provided by TSO. An example of a measured

frequency and processed out frequency from grid processor for different threshold

values for selector switch is shown in Figure 5-5, and its zoomed version is shown in

Figure 5-6. Output processed frequency is the composition of trend frequency and

measured filtered frequency as seen in Figure 5-7 and Figure 5-8.

Figure 5-3 An Example of Input Frequency Obtained from NEM-14 Generator Model Provided as Input
to Frequency Processor

Figure 5-4 NEM-14 Generator Model Processed Frequency Obtained from Grid Frequency Processor

213
Figure 5-5 Measured Frequency vs. Grid Processor Output Frequency at Different Threshold Values

Figure 5-6 Measured Frequency vs. Grid Processor Output Frequency at Different Threshold Values
(Zoomed)

Figure 5-7 Processed Frequency, Trend Filter Output, Low Pass Freq. Filter Output and Dynamic Dead-
band Output

214
Figure 5-8 Processed Frequency, Trend Filter Output, Low Pass Freq. Filter Output and Dynamic Dead-

band Output When Threshold Applied Is >0.06 (Zoomed)

We can notice output processed frequency in black color following input measured

frequency in red color most of the time when dead-band output is greater than selected

threshold while processed frequency follows trend frequency in green color when it

is less than selected threshold value. This response changes with different threshold

limits. Output frequency follows the measured input frequency most of the time until

input frequency exceeds set static dead-band value.

As stated previously, frequency processor is dependent upon implemented threshold

value and dead-band limits which can be set as per TSO requirements. Upper Dead-

band of 0.015 which is the standard value for most of the TSO [4-5] and a lower dead-

band of 0.011 has been used in this simulation. As per NERC policy, total dead–band

applied should be limited to 0.035 [6].

5.3. Frequency Response Controller Model-I

A schematic representation of a master-slave agent based frequency response

controller model is shown in Figure 5-9. This configuration can be implemented

either at wind turbine level or wind farm level, but author recommends it for large

215
individual frequency responsive wind turbine. Three conditions can be applied for

selecting normal operation. If either wind speed is at a low level or rotor speed is less

than cut-in speed or rotor speed is more than cut-off speed, then wind turbine operates

at normal operating mode insensitive to any frequency variations. If the wind speed

is in medium or high-speed range and rotor speed lies between cut-in and cut-off

speed, the wind turbine can be made to operate at frequency control mode by setting

frequency FLAG.

Figure 5-9 A Schematic Representation of Master-Slave Wind Turbine Controller Incorporating

Frequency Response Controller Model

Though frequency response controller can be implemented at de-loaded power,

frequency response controller is modeled at available power level only in this work.

Frequency response is provided as automatic power variation according to system

216
frequency through a predefined grid code compatible droop. Master control agent

receives droop value and demand power setpoint from TSO and transfers it to slave

control agent. Slave control agent selects the turbine operating mode as per detected

wind speed, or rotor speed explained previously. Earlier studies didn’t consider load

setpoint, but total commanded power in this study is provided as a sum of frequency

response power and demand setpoint. Final power set-point to turbine controller is

limited according to the maximum power rating of the turbine. Figure 5-10 shows

frequency response controller model-I which incorporates basic frequency processor

in its structure to provide frequency response operation. This frequency controller

block can provide two types of a frequency response according to grid code as shown

in Figure 5.11. Working of this frequency response controller is explained below:

Figure 5-10 Frequency Response Controller Model-I

 Frequency Sensitive Response Mode:

By setting FLAG equal to 1, frequency sensitive response mode can be activated.

Power set-point will change proportionally to both up and down frequency deviation

from reference frequency signal. The processed output frequency is generated

through the coordination of the trend control frequency

217
signal and dynamic dead band. If measured frequency follows trend control

frequency signal within dynamic dead-band, trend frequency signal is forwarded for

generating droop power response otherwise measured frequency signal is forwarded.

VSWTG final active power setpoint is generated through an algorithm implemented

on frequency response power and power demand setpoint.

 Frequency Limited Sensitive Mode:

By setting FLAG equal to 0, limited frequency sensitive response mode can be

activated to provide a high-frequency response. There should not any power variation

corresponding to any frequency changes except when the system frequency exceeds

set an upper limit. 50.4 Hz is used in the current simulation. Insensitive mode

incorporates a static dead band around the frequency setpoint. If frequency error

remains within the static band, trend frequency signal and dynamic dead-band have

no influence on output power.

Frequency response operation from this controller can be understood by Figure 5-12

and Figure 5-13. Under frequency sensitive mode, wind frequency response

controller-I provides high active power setpoint in the case of low system frequency

and low active power setpoint in the case of high frequency. In limited frequency

response mode, frequency response controller-I provides low active power setpoint

when system frequency is more than 50.4 Hz for reduced power generation. In the

case of the frequency being lower than 50.4Hz, wind turbine follows the demand

setpoint. Wind availability can also be considered in this controller, but it is not

applied in this study. This type of frequency controller is highly compatible with grid

codes like that of U.K. and Australia which requires a full response from wind farms

218
under normal conditions and limited up/down response under high-frequency

conditions.

Figure 5-11 Frequency Grid Code UK [7]

Figure 5-12 Active Power Setpoint Generated when Employing Frequency Response Controller-I in Two

Different Mode

219
Figure 5-13 System frequency Used to Depict Frequency Response Controller-I Operation in Two

Different Mode

5.4. Frequency Response Controller Model-II

Figure 5-14 Frequency Response Controller Model-II

Model as shown in Figure 5-14 implements the setting of active power set-point as

per typical non-symmetrical droop curve very similar to Irish grid code as shown in

Figure 5-15. Wind speed and processed frequency from basic frequency processor

block are its two inputs. The available wind power in per unit is calculated as a

function of the cube of incoming wind speed:

P ρAV 5.2

220
Where P is the wind power [W], ρ is air density [kg/m3], and v is wind speed

[m/s], and A is the swept area A [m ] of rotor disk that is perpendicular to the wind

flow [8].

Figure 5-15 Non-Symmetrical Droop Curve for Frequency Power Regulation

Active power set-point is primarily constrained by the available wind and different

power commands regarding de-rating power or reserve power from TSO. Aho et al.

have provided three de-rating command mode [9] as defined below:

DR P ,P 1 DR P and DR P 5.3

In another study, three different operating modes (de-rated, absolute reserve and

spinning reserve) [10] are defined such that operating mode is de-rated or normal

where:

P P , if P P 5.4

P P if P P 5.5

These command modes can readily be implemented in proposed model.

A wind power plant (WPP) running as a spinning reserve will produce less power at

all wind speeds thereby always providing a power reserve. Grid code controller will

act as spinning reserve controller or de-rating controller to vary the WPP grid power

production as per transmission system operator request. In the case of wind farm

221
operation, turbine set-point can be multiplied by active power demand provided by

farm controller to generated frequency responsive power demand.

In the current model, filtered available wind power is multiplied with the power set-

point received from algorithms defining respective grid codes for incoming

frequency. Power setpoint is then compared with maximum and minimum load

setpoint restrictions to generate final active power setpoint. Limited set-point is

achieved when additional condition is imposed as IF Frequency Deviation < 0.996

AND Frequency Deviation > 1.004, P EQUALS TO P x 0.95 per unit of

rated power. Implementation of this restriction helps in maintaining maximum output

from a wind turbine in case of limited deviation. Even though wind farm operators

strive for maximum power output, wind-AGC especially downward regulation is

highly dependent upon grid code implementation. In the case of TSO commanded

restriction on active power output, Irish grid code imply all grid integrated generating

units to can operate at a reduced output level. All interconnected grid code compatible

wind farms are required to ensure a fast pitch control as well as other mechanical

controller to be able to participate in a frequency response services in acceptable time.

Figure 5-16 WTG Power Setpoints with Limited Power Setpoint As Grid Controller Output Setpoint and

Corresponding Generated WTG Power

222
Figure 5-16 shows various power set-point generated through grid code compatible

frequency controller-II at variable wind speed and fluctuating grid frequency. The

blue color is the power setpoint required according to the grid code while green is the

available wind power. This available power is multiplied by grid code power and

limited according to desired command mode as explained above thereby providing a

reserve power to be used in frequency deviations. We can notice an increase in

generated electrical power from 1 per unit to 1.1 per unit when the turbine is working

in reserve mode. The Same type of response can be implemented with wind turbine

working in de-loaded mode. In the current study, frequency response controller-I

works independently of wind speed measurement while frequency response

controller-II includes wind speed and corresponding available power in its algorithm.

5.5. Conclusion

This chapter presented a modeling framework for frequency dependent active power

setpoint generation in variable speed wind turbines. An individual wind turbine can

be made grid code frequency compatible by including additional active power

setpoint generator output to the modified torque control loop of the respective turbine.

Active power setpoint generator applies designed power limitation on available wind

power, rated turbine power, and TSO commanded power and provides setpoint to the

turbine. A grid frequency processor based on dynamic dead band and moving

averaged frequency filter is used to suppress noise frequency signals from passing to

the active power setpoint generator. In next chapter, the performance of frequency

controller with and without including wind power in the algorithm will be analyzed

for Type-3 and Type-4 wind turbine response.

223
5.6. References
 
1. AEMO. (2009). Power system operations event report: Multiple generator

disconnection and under frequency load shedding Thursday 2nd July 2009.

Website: aemo.com.au.

2. Hepner, S. and Wihler, A., (2000). Method for controlling an output of an

electrical power plant. US6118187 A.

3. Smith, S., (1997). The scientist and engineer's guide to digital signal

processing. San Diego, Calif.: California Technical Pub.

4. EIRGRID, (2013). Indication of Dynamic Model Validation Process.

5. EY, (2014). Ancillary service standards and requirements study. Perth:

Available at wa.aemo.com.au.

6. North American Electric Reliability Council, (2004). NERC OPERATING

MANUAL. NERC.

7. Manwell, J. F., McGowan, J. G., and Rogers, A. L., (2009). Wind Energy

Explained: Theory, Design and Application, Wiley.

8. Manwell, J. F., McGowan, J. G., and Rogers, A. L., Wind Energy Explained:

Theory, Design and Application, Wiley, 2009.

9. Janssens NA, Lambin G, Bragard N. (2007). Active power control strategies

of DFIG wind turbines. In: 2007 IEEE power tech Lausanne; p. 516-21.

10. Tarnowski, G.C.; Kjær, P.C.; Dalsgaard, S.; Nyborg, (2010). A. Regulation

and frequency response service capability of modern wind power plants. In

Proc. Power and Energy Society General Meeting, 2010 IEEE, 25-29 July,

Minneapolis, USA.

224
CHAPTER 6

GRID CODE FREQUENCY SENSITIVE VSWTG:

IMPACT ASSESSMENT

Second generation WECC VSWTG models were released in 2014, but their adoption

has not been up to the satisfactory level. With these existing models, plant controllers,

especially for frequency regulation, are still in research and development stage. A

wind turbine active power setpoint controller architecture based on the dynamic dead-

band dependent grid frequency processor was presented in chapter 5. Area frequency

response along with VSWTG electrical power support is investigated and compared

to establish the superiority of proposed grid code frequency-active power controller

based VSWTG model over other droop based VSWTG models in this chapter.

Application of type 3 VSWTG generic model and earlier proposed type 4 VSWTG

model in frequency regulation studies is presented in this chapter. With simplified

generator converter model implemented in investigated generic VSWTG model,

distinguishing electrical response for various droop inertia configuration is achieved.

Consequent electrical output has a remarkable effect on system frequency response

so generic modeling requires detailed investigation when implemented in frequency

regulation studies.

6.1. Introduction

Short-term dynamics have little effect on system frequency stability mechanisms. So

frequency regulation investigation can be satisfactorily achieved with generic

component modeling. The generic models are functional models appropriate for the

investigation with lower simulation period for large-scale power systems. These
225
generic models can sufficiently represent all dynamics associated with the impact of

active power variations in the time range of 50 ms to 100 min. Electromagnetic

transients are neglected in these simulation models.

6.1.1. Type 3 VSWTG Model

Different type 3 VSWTG models [1-3] exists in literature studies, but wind turbine

and wind power plants model still lack standardization and works are in progress.

Various software like PSLF, PSS-E, DlgSILENT, etc. have been used for

implementing these models. An official version of second generation WECC models

was released in 2014 but their adoption has not been up to the desired level as not

every type of VSWTG model can be applied in every study [4]. Long term system

dynamics requires simplified generic modeling with fewer parameters as detailed

WTG modeling with low simulation time step takes much longer simulation time.

Even though WTG model characterizations development has been fundamentally

accomplished, however, development of plant controllers especially regarding

frequency response control is still under developing stage.

Block diagram in Figure 6-1 shows the basic components of generic VSWTG model.

A basic schematic diagram implementing these basic modules for type 3 (Doubly fed

asynchronous generator) based WTG control structure is presented in Figure 6-2. The

model has three basic blocks: turbine model, Generator/Converter model, and

Electrical control model. Type 3 electrical control model is represented by active

current command only with all flux dynamics neglected. During frequency response,

real power has priority, hence detailed reactive power loop in not applied in the

current study. The aerodynamic model given in chapter 3 has been applied for type 3

226
VSWTG model while single mass drive train is utilized for drive train modelling in

this study.

Figure 6-1 Basic Components of Generic VSWTG Model

Turbine control system involves two control loops having speed error as input and

two control outputs: 1. Wind turbine reference power provided to the converter

control. 2. Pitch reference value. Rotor speed is controlled at optimum level as per

reference power command through torque limitation in speed control loop while

mechanical power and a corresponding shaft speed of wind turbine are controlled

through pitch controller and pitch compensation loop when rotor speed exceeds

optimum level. It should be noted that frequency responsive power setpoint is

subtracted from commanded power setpoint to achieve required power order to

generator/converter model while a fixed rated power command is applied to pitch

compensator. Pitch control enables optimum control of aerodynamic wind power by

rotor blade pitching to regulate turbine torque. Maximum power of 1.2 per unit is

available at zero pitch angle while it is highly reduced with highest pitch angle.

Similar to proposed type 4 VSWTG model in chapter 4, the pitch angle is obtained

227
through a series of the PI regulator is mathematically expressed by following

equations:

.
θ K ω ω K P P 6.1

Where angular reference speed is approximated as:

ω 0.79131P 1.526046P 0.49188 6.2

Under the assumption 0.15 p. u. P 0.75 p. u. For P

0.75 p. u. , ω 1.2 and P 0.15, ω 0.689 p. u.

Figure 6-2 Generic Type 3 VSWTG Basic Model with Auxiliary Virtual Inertia Model

Most of the past research studies have considered the basic type 3 model as shown in

Figure 6-2 and basic control loop for frequency response studies shown as an inset in

Figure 6-3. Current study applies a modified wind turbine control loop adopted from

[5] which is shown in red dotted line. Pitch compensation block takes frequency

228
responsive active power set-point PAPC generated from frequency responsive

controller instead of rated 1 p.u. reference power. Power order from torque controller

is altered by passing active power set-point PAPC through the power response rate

limiting block.

Figure 6-3 Type 3 WTG Model with Active Power Controller Feeding Power Setpoint PAPC to Power

Controller Loop, Pitch Controller Loop and Speed Controller Loop 

Rotor speed error is given as input to both these controls. Final power order provided

to generator converter model is generated by adding limited active power setpoint

229
from frequency response controller block to difference between active power limited

setpoint power and power order from speed controller.

P P P P 6.3

The ramp rate is implemented by including a washout filter whose time constant is

detrimental to rate limit imposed on changes in power order. Power order provided

to converter control can also be transformed by inertia controller to exploit the

rotating masses stored energy along with frequency responsive active power

command.

6.1.2. Applied Commanded Reserve Power Mode

In reserve power mode as applied to this study, wind turbine regulates electrical

power according to frequency responsive final power order. Rotor speed and pitch

based de-loading methods were discussed in chapter 3. Prior de-loading of turbine

through C table inversion [6], over speed and under speed operation has been

suggested in many research studies [7-10] for enabling wind turbine frequency

responsive support. A combined torque-pitch control method utilizing PAPC is applied

in this study to obtain reserve power mode operation. Optimum speed is maintained

corresponding to active power set-point through the torque-speed controller. Pitch

controller initiates blade pitching in case speed increases and reduces mechanical

power thereby reducing speed. All interconnected grid code compatible wind farms

are required to ensure a fast pitch control as well as other mechanical controller to be

able to participate in a frequency response services in acceptable time.

In MPPT operating mode, turbine power set-point is determined such that:

.
P K ω , where K ,β 6.4

230
Let us assume wind turbine to be operating at MPPT mode initially such that P

P and W W and pitch angle β β 0 . When turbine switches from

MPPT mode to commanded power mode, MPPT power setpoint P changes

to P P ; VSWTG power drops below mechanical power thereby

increasing rotor speed to W . This increase in rotor speed is controlled through

increasing pitch angle β to β provided through pitch controller in coordination with

pitch compensator. Consequently, power-speed characteristic of wind turbine is

varied which is shown in red in Figure 6-4 and rotor speed settles at ω . To

guarantee a safe switching between MPPT mode and commanded mode and

safeguard turbine operation beyond rated values, final power order is limited to the

optimum power value (P K ω . The turbine will continue to run in MPPT

mode unless there is a change in wind speed or change in active power setpoint from

frequency controller.

Figure 6-4 Characteristics Curve Of Type 3 VSWTG under MPPT and Reserve Power Mode

231
It should be noted that active power setpoint from frequency controller is already

limited as per desired reserve power of 10%. To explain reserve power mode

operation, assume steady state condition represented by point a for blue curve at some

wind speed such that turbine mechanical power equal to VSWTG electrical

powerP P . In the case of frequency deviation due to generation loss,

frequency responsive VSWTG’s power will increase to compensate for this loss and

may reach VSWTG rated capacity limit shown by point b in Figure 6-4. Power

extraction due to kinetic energy will continue until rotor speed hits the minimum limit

at point d. After this point, mechanical power will be more than electrical power and

rotor will again speed up to finally settle at point e with speed W .

6.1.3. Wind Profile and Available Power Model

Figure 6-5 Wind and Available Power Simulation Model

A wind profile and corresponding available power simulation model are presented in

Figure 6-5. This wind model is used to input wind speed shown in Figure 6-6 to both

frequency controller presented in chapter 5 while available power is also applied to

232
frequency controller-II where active power set-point was dependent upon available

wind power. Wind speed very similar to Avalon wind speed data obtained from [11]

was applied to a lookup table, and final wind speed was limited by a maximum wind

speed and nominal speed through SR flip-flop logic. Electrical power is produced as

the cube of available wind, and nominal power is produced at nominal wind speed.

Figure 6-6 Generated Wind Speed and Available Power

6.2. Simulation Setup

A single area controlled model incorporating hydropower plant is used to test the

frequency responsive VSWTGs. Hydro turbine governor is modeled according to

Figure 6-7 and discussed further in Chapter 7. On System base value of 5000 MW

and assuming 0.05 p.u. extra support from VSWTG during frequency excursions,

wind-integrated system modified inertia constant, H is calculated as [12]:

233
H H 1 L H L 6.5

∑ ∗
H , where S ∑ S 6.6

Where H and S are the inertia rating and apparent power rating of individual

generating units. L is wind penetration level. For frequency responsive wind plant,

wind inertia contribution H to power system by providing ∆P extra active power

when the system is subjected to step load disturbance ∆P is calculated as:

H ; 6.7

∆ ∆
X e 1 6.8
∆ ∆

Figure 6-7 Single Area Hydro Governor Turbine Model with Integrated Wind Farm 

Total time delay T associated with hydro governor turbine delay model is calculated

by delay theory originally given in [13]. Summation of governor time constant, valve

motion delay time and turbine response delay time results in total time delay at which

minimum frequency deviation occurs after system disturbance Value of T used in

the simulation is calculated as 3.7814 seconds. To derive dynamic response for the

uncontrolled system, let us assume hydro turbine governor time constant equal to zero

in Figure 6.7, we can write following equation 6.9 - 6.10.

234

∆F s |∆ ∗ 6.9
,

∆f t ,
1 exp t ,
∆P 6.10
, T ,

Where K is defined earlier as . The minimum frequency occurring at T after load

disturbance can be calculated as equation 6.11, where the hydro turbine response

delayed by T covers the combined load disturbance in the presence of wind energy.

T
∆f t K 1 exp T ∆P 6.11


In case of integral controlled system with ∆P s , frequency
,

deviation can be expressed according to equation 6.12 as:


∆F s 6.12

,
.
,

6.2.1. Simulation Assumptions

System simulations are performed and compared with different frequency responsive

type 3 and type 4 wind plant model. A comparative study is made of frequency

response and its indicators, the effect on VSWTG electrical support and

corresponding rotor speed and hydraulic mechanical power change. Following

assumptions are used in simulation:

1. A constant load disturbance is applied.

2. All VSWTG models are provided same wind speed as shown in Figure 6.6.

3. Frequency response controller-I (Figure 5.8) and frequency response

controller-II (Figure 5.12) is implemented for VSWTG model as given in

chapter 5. Parameters are given in the appendix.

235
4. Frequency droop model 1, droop model 2 and inertia droop model as given in

Figure 3.13 of Chapter 3 are incorporated in basic torque control loop based

VSWTG model [12] which is shown as an inset in Figure 6.3.

5. Droop setting used in simulations are given in Table 6.1.

6. Equivalent droop setting of area 1 changes with wind penetration as R

. A 10% wind penetration is used in the simulation.

System parameters used in the simulation are given in the appendix. Results obtained

were assessed for effect of frequency responsive VSWTG model integration on:

 Area frequency response (Settling time, frequency nadir and rate of change of

frequency).

 VSWTG electrical output and VSWTG rotor speed variation during and after

frequency support.

Table 6-1 Comparison of Simulation Results to Show the Effect of Torque Controller Re-Tuning and

Different Droop Values for 10% Wind Energy Penetration

Controlled Uncontrolled Control Uncontrol


Torque
Area Area Area Area
Controller
Models Settles
Gain Settling Settling
(Hz) at
( PI ) Time (s) Time (s)
. .

Case A Torque Controller Re-tuned in basic droop-inertia models Only

Droop/Inertia (-271.55,
0.035 0.0315 49.96 49.981 0.4 30 30
Models -310.70)

Freq. Cont.-I 0.035 0.0315 (3, 0.6) 49.98 49.988 1.01 50 25

Freq. Cont.-II 0.035 Var. (3, 0.6) 49.99 49.994 1.1 50 30

Case B No tuning of Torque Controller in basic droop-inertia models.

Droop/Inertia
0.035 0.0315 (3, 0.6) 49.982 49.986 0.04 40 35
Models

Freq. Cont.-I 0.035 0.0315 (3, 0.6) 49.987 49.987 1.01 40 25

Freq. Cont.-II 0.035 Var. (3, 0.6) 49.991 49.995 1.1 40 30

236
Droop values changed for Hydro and Wind power plants; No tuning
Case C
applied in torque controller

Droop/Inertia
0.0556 0.05 (3, 0.6) 49.979 50.005 0.04 40 30
Models

Freq. Cont.-I 0.0556 0.05 (3, 0.6) 49.990 49.993 1.01 20 10

Freq. Cont.-II 0.0556 Var. (3, 0.6) 49.984 49.994 1.1 30 25

Droop values changed for Hydro and Wind power plants; No Tuning
Case D
Applied in Torque Controller

Droop/Inertia
0.035 0.0315 (3, 0.6) 49.982 49.995 0.04 40 20
Models

Freq. Cont.-I 0.0556 0.0315 (3, 0.6) 49.991 49.993 1.01 20 10

Freq. Cont.-II 0.0556 Var. (3, 0.6) 49.984 49.994 1.1 30 25

Case E Different Droop Values for Hydro and Wind Power Plants

Droop/Inertia
0.0556 0.0315 (3, 0.6) 49.979 50.005 0.04 40 20
Models

Freq. Cont.-I 0.0556 0.0315 (3, 0.6) 49.990 49.993 1.01 20 10

Freq. Cont.-II 0.0556 Var. (3, 0.6) 49.984 49.994 1.1 30 25

6.3. Discussion

6.3.1. Case I: Type 3 VSWTG

Summary of 5 test cases is provided below in Table 6-1 in which proposed frequency

controller based type 3 VSWTG models are compared with other droop inertia based

type 3 VSWTG model. Important points from these simulation results are discussed

in below sections.

 Effect on Control Area Frequency Response

Uncontrolled and controlled control area were simulated with different droop values.

Effect of torque re-tuning is also conducted to highlight the difference in generated

power and corresponding frequency response. Figure 6-8 gives the frequency

deviation obtained from single area uncontrolled model with 0.1 per unit load

disturbance. An example simulation result of area frequency response for controlled-

237
single area LFC model with integrated type-3 WTG is shown in Figure 6-9 where

basic type 3 VSWTG model’s torque controller is not retuned.

Figure 6-8 Uncontrolled Single Area Model Frequency Deviation for 0.1 P.U. Load Disturbance 

Figure 6-9 Frequency Response of Single Area LFC Control Model for 0.1 P.U. Load Disturbance,
Req=0.035 for All Cases, Torque Controller Gains (3, 0.6), No Re-Tuning For Torque Controller

Similarly, frequency response for a control system with re-tuned torque controller is

shown in Figure 6-10 and 6-11. Effect of this re-tuning will be explained in section.

From these Figures, a clear improvement in area frequency deviation can be observed

for both controlled and uncontrolled system with proposed frequency responsive

VSWTG model integration. Respective minimum frequency and settling time for

these simulation test are provided in Table 6-1. Except for re-tuned model, improved

settling frequency in the range of 20s-30s is observed for proposed frequency

responsive VSWTG model. With changed in droop value from 0.0315 to 0.05 in all

238
models, an improvement in settling frequency is observed for both uncontrolled and

controlled system. Comparing case C and case E, the effect of system droop change

is more visible over wind droop value. With constant system droop and changed wind

energy droop from 0.05 to 0.0315, the same settling time is observed. With other

three droop controlled models, more overshoots and undershoots are observed a with

a slight increase in settling time. Though the inclusion of wind power in governing

active power setpoint has a role in system frequency response bringing more

oscillations in terms of undershoots and overshoots, yet superiority of grid code

compatible frequency response controllers is established through simulation results

regarding reduced settling time, improved ROCOF and frequency nadir point.

Figure 6-10 Frequency Response of Single Area LFC Control Model for 0.1 p.u. Load Disturbance and
Req =0.035 in All Cases, Torque Controller Retuned for Other Three Droop Based VSWTG Models

              

Figure 6-11 Frequency Response of Single Area LFC Control Model for 0.1 p.u. load Disturbance,
Req =0.0556 in Case of Frequency Controller-I and II, Torque Controller Retuned 

239
 Effect on Rate of Change of Frequency and Frequency Nadir

The management of ROCOF is critical to maintaining power system frequency within

the frequency operating standard and to maintaining the power system in a secure

operating state. We discussed system inertia and the corresponding effect on ROCOF

in Chapter 2. Lower inertia leads to a higher ROCOF and power electronics based

wind turbines decoupled from grid do not contribute to any system inertia without

additional modifications in control loops. Even TSOs have no control over minimum

system inertia, but ROCOF has to be within certain limits as per grid code. Currently

installed wind turbines have maximum acceptance ROCOF within the range of 2–4

Hz/s for 200–300 milliseconds and 1 Hz/s is accepted in a long range of around 1

second [14].

Figure 6-12 Rate of Change of Frequency Comparison for Different Load Change (p.u.) and 10%
VSWTG Penetration.

In another comparative simulation study, the rate of change of frequency (ROCOF)

obtained under different load disturbance scenario, and 10% wind penetration is

given in Figure 6-12. Though the values obtained can not be held indicative of real

time scenario which has interaction between components of varying electrical

characteristics, a clear improvement is observed in ROCOF values for control area

when proposed frequency based controller based Type 3 VSWTG is integrated with
240
the hydro power plant. A more detailed system will be required to analyze actual

ROCOF changes.

Figure 6-13 Frequency Nadir Comparison for Different Load Change (p.u.) and 10% VSWTG

Penetration

Frequency nadir after any contingency is detrimental to primary frequency regulation

for maintaining system stability. The primary frequency reserve adequacy criterion

can be expressed as [15]:

Freq ∆ P , Primary Frequency Reserve, F Freq , 6.18

where ∆P is the maximum power loss during contingency, F (Hz) is the

maximum value for governor’s dead band. Freq is the frequency nadir after the

loss of ∆ P , and Freq denotes the minimum frequency required as per

operating standard. Wind power plant integration and its participation in frequency

regulation services can be analyzed through frequency nadir points achieved under

different combinations. Frequency nadir obtained for different frequency responsive

VSWTG models integrated into control area is shown in Figure 6-13. Frequency nadir

point as achieved under different load disturbance scenario indicates the superiority

of grid code compatible frequency controllers over normal droop based VSWTG

model.

241
 Effect on VSWTG Electrical Output

An important point to notice about other three droop controlled model is regarding

wind plant power output during and after frequency response in LFC model. With

simplified generator converter model implemented in currently investigated generic

VSWTG model, we observed distinguishing electrical response for various droop

inertia configuration implemented. Consequent electrical output has a remarkable

effect on system frequency response so care should be taken while implementing

generic models.

Figure 6-14 Electrical power from Type 3 VSWTG with Retuned Torque Controller in Control Area.

During Initial Frequency Response for 0.1 p.u. load Disturbance, Req= 0.0556 for Frequency Controller-I

and II

Figure 6-15 Initial Frequency Response for 0.1 p.u. Load Disturbance 

242
Figure 6-14 shows the zoomed version of electrical power output from VSWTG with

different droop controlled models with retuned torque controller during the initial

frequency deviation in control area. Zoomed version of control area frequency is

shown in Figure 6-15. Even though a variable wind speed is applied in simulation, a

constant wind speed of 11.3 m/s is observed during approximate 40 seconds of

frequency deviation. In the case of a 0.1 p.u. load disturbance and 0.05 p.u. active

power support from VSWTG, control area sees this frequency support from a

machine with 3.4 s inertia constant. This inertia constant will change to 7.17 s if 0.1

p.u. of extra support from VSWTG is assumed. VSWTG will replace other

conventional generation (having equivalent inertia 6 s) if they provide an active

power support of approximately 0.8 p.u. during frequency excursions.

It can be observed that even with same load disturbance and same system inertia

constant for all control area, the different frequency response is obtained. Due to

anomalies in applying basic VSWTG model and auxiliary droop models, we notice a

slow response in generated electrical power. Electrical power from basic VSWTG

model with retuned torque controller finally settles at 0.4 p.u. as soon as frequency

improves otherwise power drops to 0.04 p.u. All these three droop based models

attempt to provide increased electrical power up to 1.2 p.u. soon after detecting the

frequency drop but fails to maintain net power output after frequency disturbance

correction. Drop in electrical power output during and after frequency response also

earlier confirmed in [16] and requires torque controller parameter retuning for

increased electrical power after temporary frequency deviations. Simulation tests

were conducted with retuned torque controller as given in Table 6-1 but still proposed

frequency controller based VSWTG outperforms other models electrical output. A

243
clear improvement in electrical power temporary support from WTG model

incorporating frequency controller-I and II can be observed from above Figure 6-14.

Figure 6-16 Generated Active Power Setpoint for Frequency Controller-II in the case of 0.1 p.u. Load
Disturbance in Control Area

Figure 6-17 Generated Active Power Setpoint for Frequency Controller-I in the case of 0.1 p.u. Load
Disturbance in Control Area

Following the active power setpoint generation algorithm, a constant active power

setpoint up to 1.01 p.u. is provided by frequency controller-I to ramp up the power

during the moments of frequency deviations. VSWTG model as used in this study

has 10% spinning reserve so a maximum electrical power of 1.1 p.u. is available

during frequency disturbance. Frequency controller-II also provides an increasing

power reference setpoint of around 0.85 p.u. during the initial 20 seconds of

frequency deviation and changes according to grid requirements and available power.

244
The active power setpoint provided by frequency response controller-I and II to the

VSWTG is shown in Figure 6-16 - Figure 6-17. Effect of these reference setpoint is

reflected in generated electrical power. WTG model with frequency controller-I can

maintain electrical power around 1.1 p.u. during the moments of initial frequency

deviations while WTG with frequency controller 2 has an initial power drop up to 0.3

p.u. and then starts increasing and settles at around 1.1 p.u.

 Effect on VSWTG Rotor Speed

A representative rotor speed variation is shown in Figure 6-18. The change in rotor

speed for type 3 VSWTG model during frequency response in control area has been

same for all test cases. A minimum limit of 0.75 p.u. and a maximum limit of 1.2 p.u.

has been imposed in currently investigated frequency responsive VSWTG models. It

can be observed from Figure 6-19 that all VSWTG models show reduced rotor speed

maintained approximately 0.87 p.u. during those initial moments of frequency

deviations till 20 seconds when extra electrical power from VSWTG is expected.

Some fluctuations in rotor speed can be seen after 5 seconds when frequency drop is

at peak position, and VSWTG control loop tries to increase power. Rotor speed from

basic type 3 VSWTG with other droop models finally settles at around 0.95 p.u. with

consequent reduced electrical output power to 0.4 p.u. If proper speed limits are not

imposed in droop model 1, rapid variations in rotor speed are observed till 250

seconds before settling smoothly around 0.95 p.u. In contrast to these basic VSWTG

model, more adaptive frequency responsive operation is seen in type 3 VSWTG with

frequency controller-I and II where we can notice a change in electrical power with

a corresponding change in rotor speed during the moment of another frequency

deviation at around 235 seconds. Drop in rotor speed is observed due to an imbalance

in mechanical and electromechanical torque when electrical power is increasing for

245
frequency controller-I and II, but as soon as frequency deviation settles, an increase

in rotor speed is also observed which settles at around 1.2 p.u.

Figure 6-18 VSWTG Rotor Speed Variations with Different Frequency Droop Models during Frequency
Response

                   

Figure 6-19 VSWTG Rotor Speed Variations with Different Frequency Control Loops [Zoomed] 

6.3.2. Case II: Type 4 VSWTG

Various generic models of full converter based type 4 VSWTG models were

presented in Chapter 4. The inertial response of proprietary model and proposed type

4 model were presented under constant wind speed and constant power setpoint of

0.8 p.u. Both the models were shown to provide inertial support for frequency

deviation which was provided to models through an auxiliary signal. No wind power

participation in supplied power was considered. In this section, simulation results of

earlier proposed type 4 VSWTG model and proprietary model are presented when

246
working in frequency responsive mode. Different inertia and droop models presented

earlier in chapter 3 are integrated with these models. Results are compared when grid

code sensitive active power setpoint from proposed frequency controller in Chapter

5 is provided with these models. The aim of these simulation results is to show the

improvement of frequency response and corresponding generated power supported

by rotor speed when working under grid code sensitive power setpoint.

Figure 6-20 Frequency Response of Single Area LFC Control Model Integrated with Proprietary Type 4

VSWTG Model for 0.1 p.u. Load Disturbance, Req = 0.035, No Retuning Applied in Torque Controller

Figure 6-21 Generated Electrical Power from Proprietary Type 4 VSWTG Model during Initial

Frequency Response for 0.1 p.u. Load Disturbance, No Retuning Applied in Torque Controller

Figure 6-20 presents the frequency response obtained for single control area when

model similar to proprietary type 4 VSWTG model shown in Figure 4-6 is integrated

into frequency responsive mode. No tuning of torque controller is applied in these

247
models. Recalling that frequency controller-II works according to grid code similar

to Irish code and frequency controller-I works according to British grid code. Wind

available power is considered in frequency controller-II while the frequency

controller-I is independent of available power. It should be noted again that during

the initial frequency deviation of around 40 s, wind speed is assumed constant. An

improved frequency response along with corresponding electrical power variation

and rotor speed can be observed in Figure 6-18 – Figure 6.20 when proposed

frequency controller based grid code sensitive power set-point is provided to it in

comparison with other droop inertia models.

Figure 6-22 Proprietary type 4 VSWTG Model Rotor Speed Variations with Different Frequency Droop

Models during Frequency Response

In similarity with type 3 VSWTG droop based models, a drop in electrical power to

0.04 p.u. is observed during and after frequency deviation for other three droop inertia

model as shown in Figure 6-21. Type 4 rotor speed also increases sharply above rated

value during frequency deviation and fails to recover back as seen in Figure 6-22. It

is to be noted that power setpoint from these auxiliary models was added to power

order from torque controller loop in the similar manner as shown in the inset of Figure

6-3 while a rated power of 1 p.u. is provided to pitch compensator. Increased power

support from proprietary type 4 model requires torque retuning which is difficult
248
during real time operation. Active power setpoint from proposed frequency

controllers was added in the modified loop similar to shown in Figure 4-6 of Chapter

4. Improved variation of electrical power during and after frequency deviation can be

observed from Figure 6-21. Even rotor speed remains under rated limit after the

frequency responsive operation. These simulation results confirm that by proper

modeling of power control loop, improved power response can be obtained during

frequency disturbances.

Impact assessment of previously proposed type 4 VSWTG model in chapter 4 is

represented in upcoming simulation results where all three droop based auxiliary

model and proposed frequency controller-I and II are integrated.

Under same droop settings and all similar system parameter, frequency response as

shown in Figure 6-23 is obtained for proposed type 4 VSWTG model. Minimum

nadir point is achieved with frequency controller-II while frequency response from

frequency controller-I in both modes is nearly similar to droop model 1 as represented

in zoomed version in Figure 6-24. It should be noted that other droop models are not

grid code compatible while an upper dead-band limit of 0.0315 and lower dead-band

limit of 0.001 has been imposed on frequency controllers to provide active power

setpoint. Improvement in system frequency response in case of frequency controller-

II can be matched with corresponding active power setpoint as shown in Figure 6-27

and respective electrical power shown in Figure 6-25. Electrical power output as

shown in Figure 6-25 is also found to be varying according to another active power

setpoint from frequency controller-I which is shown in Figure 6-26.

249
Figure 6-23 Frequency Response of Single Area LFC Control Model Integrated with Proposed Type 4

VSWTG Model for 0.1 p.u. Load Disturbance, Req=0.035

Figure 6-24 Frequency Response of Single Area LFC Control Model Integrated with Proposed Type 4

VSWTG Model for 0.1 p.u. Load Disturbance [Zoomed]

Figure 6-25 Generated Electrical Power from Proprietary Type 4 VSWTG Model during Initial

Frequency Response for 0.1 p.u. Load Disturbance, Req=0.035

250
Figure 6-26 Grid code Sensitive Active Power Setpoint from Frequency Controller-I during Initial

Frequency Response for 0.1 p.u. Load Disturbance

Figure 6-27 Grid Code Sensitive Active Power Setpoint from Frequency Controller-II during Initial

Frequency Response for 0.1 p.u. Load Disturbance

Figure 6-28 Generated Rotor Speed from Proposed Type 4 VSWTG Model during Initial Frequency

Response for 0.1 p.u. Load Disturbance, Req=0.035

251
Table 6-2 Comparison of Frequency Response Simulation Results for Type 4 Models Under 10% Wind

Energy Penetration in a Controlled Area

, ,
, (Hz) Settling Time (s)
(p.u Hz / p.u. MW) (p.u. Hz / p.u. MW)

Case A Proprietary Type 4 VSWTG model

Droop/Inertia models 0.035 0.0315 49.982 45

Freq. Cont.-I 0.035 0.0315 49.987 40

Freq. Cont.-II 0.035 Var 49.991 40

Droop/Inertia models 0.0556 0.05 49.982 45

Freq. Cont.-I 0.0556 0.05 49.991 20

Freq. Cont.-II 0.0556 Var. 49.985 30

Case B Proposed Type 4 VSWTG model

Droop/Inertia models 0.035 0.0315 49.98 100

Freq. Cont.-I 0.035 0.0315 49.977 80

Freq. Cont.-II 0.035 Var. 49.973 80

Droop/Inertia models 0.0556 0.05 49.972 35

Freq. Cont.-I 0.0556 0.05 49.972 35

Freq. Cont.-II 0.0556 Var. 49.977 45

A comparison of both type 4 model is given in Table 6-2. Even though frequency

nadir point is slightly less than the proprietary model and more settling time more,

proposed type 4 model presents improved electrical power and rotor speed. Except

for droop model 2, proposed type 4 VSWTG can provide improved power response

with other inertia droop model though not according to grid code. Proposed type 4

VSWTG model can provide rotor speed within rated limit during and after frequency

disturbance as shown in Figure 6.28 while it overshoots to high speed for droop model

2 where we see a drop in electrical power. Increased in system droop value is found

effective in reducing the settling time. Basic Type 3 and Type 4 VSWTG models call

for torque controller gains re-tuning for improved participation in frequency response

while this can be avoided with frequency response controllers. These comparative

values are obtained for 10% wind penetration at 5000 MW system base value. The

252
more effective comparison requires higher wind penetration in the system, but overall

objective of this study was to show the improvement in frequency response and

electrical output under proposed frequency controllers and modified torque speed

loop of VSWTG.

6.4. Conclusion

An individual wind turbine can be made grid code frequency compatible by including

additional active power setpoint generator output to the modified torque control loop

of the respective turbine. Active power setpoint generator applies designed power

limitation on available wind power, rated turbine power, and TSO commanded power

and provides setpoint to the turbine. A grid frequency processor based on dynamic

dead band and moving averaged frequency filter is used to suppress noise frequency

signals from passing to active power setpoint generator. The simulation was carried

out for a single area control model with 10% wind penetration. Control area

frequency response, as well as integrated VSWTG's electrical and rotor speed

responses, were compared with other frequency droop based VSWTG model

responses. Following conclusions can be based on these simulations.

 Promising results regarding improved settling time, better electrical power

and rotor speed variations during frequency deviations were obtained for

proposed frequency responsive grid code compatible VSWTG model in

comparison to other common droop-inertia based model.

 Basic VSWTG model requires retuning of torque controller to provide

electrical power output of 0.4 p.u. soon after frequency disurbances while

electrical power drops to 0.04 p.u during frequency response. Proposed

frequency responsive grid code compatible VSWTG model does not requre

253
any tuning of torque controller and provides electrical power as per power

setpoint for frequency response and rotor speed being under rated limit.

 Improved system frequency is obtained when grid code frequency responsive

type 3 VSWTG is integrated in comparison to type 4 grid code frequency

responsive VSWTG.

 With suitable modeling, full converter based type 4 VSWTG can participate

in frequency response. A simulation test at 10% load disturbance with type 4

VSWTG is integrated to the system provided approximately same frequency

nadir but slightly increased settling time and increased ROCOF when

compared with type 3 VSWTG integrated system response.

 The effect of system droop change is more visible over wind droop value in

terms of frequency deviation.

 Frequency response for area 1 was determined using classical linear hydraulic

turbine model. It would be interesting to investigate the area frequency

response with advanced inelastic non-linear hydraulic turbine model

interaction with frequency responsive wind power plants.

254
6.4. References

1. Jonkman, J., Butterfield, S., Musial, W., and Scott, G. (2009). Definition of a 5-

MW Reference Wind Turbine for Offshore System Development, NREL/TP-

500-38060, Golden, CO: National Renewable Energy Laboratory.

2. Pourbeik, P.; Ellis, A.; Sanchez-Gasca, J.; Kazachkov, Y.; Muljadi, E.; Senthil,

J.; Davies, D. (2013). Generic stability models for Type 3 & 4 wind turbine

generators for WECC, Power and Energy Society General Meeting (PES), 2013

IEEE, pp.1, 5, and 21-2.

3. B. Sørensen, B. Andersen, J. Fortmann, K. Johansen, and P. Pourbeik (2011).

Overview, status, and outline of the new IEC 61400 -27 – electrical simulation

models for wind power generation, International Workshop on Large-Scale

Integration of Wind Power into Power Systems as well as on Transmission

Networks for Offshore Wind Power Farms.

4. WECC (2010). WECC Wind Power Plant Dynamic Modeling Guide, WECC

Renewable Energy Modeling Task Force. Available at renew-ne.org.

5. K. Clark, N. W. Miller, J. J. Sanchez-Gasca (2010). Modelling of GE Wind

Turbine-Generators for Grid Studies, Version 4.5. General Electric International,

Inc.

6. P. Venne and X. Guillaud (2009). Impact of wind turbine controller strategy on

deloaded operation. Integration of Wide-Scale Renewable Resources In to the

Power Delivery System, 2009 CIGRE/IEEE PES Joint Symposium, Calgary,

AB, pp. 1-1.

7. N.A. Janssens, G. Lambin, N. Bragard (2007). Active power control strategies

of DFIG wind turbines. IEEE power tech Lausanne, pp. 516–521.

255
8. Ramtharan G, Ekanayake JB, Jenkins N. (2007). Support for spinning reserve

from DFIG based wind turbines. In: 2007 international conference on industrial

and information systems. p. 111–116.

9. F.D. Kanellos, N.D. Hatziargyriou (2008). Control of variable speed wind

turbines in islanded mode of operation. IEEE Transactions on Energy

Conversion, 23 (2) (June), pp. 535–543.

10. V. Courtecuisse, B. Robyns, B. Francois, M. Petit, J. Deuse (2008). Variable

speed wind generators participation in primary frequency control. Wind

Engineering, 32 (3), pp. 299–318.

11. Wind speed data. Available at www.bom.gov.au.

12. N. R. Ullah, T. Thiringer and D. Karlsson (2008). Temporary Primary Frequency

Control Support by Variable Speed Wind Turbines— Potential and

Applications, in IEEE Transactions on Power Systems, vol. 23, no. 2, pp. 601-

612.

13. M. L. Chan, R. D. Dunlop and F. Schweppe (1972). Dynamic equivalents for

average system frequency behavior following major disturbances, IEEE Trans.

Power App. Syst., vol. PAS-91, no. 4, pp. 1637-1642.

14. AEMO (2013). Wind Turbine Plant Capabilities Report. Available at

aemo.com.au

15. H. Chavez, R. Baldick and J. Matevosyan (2015). The joint adequacy of AGC

and primary frequency response in single balancing authority systems, IEEE

Trans. Sustain. Energy, vol. 6, no. 3, pp. 959-966.

 
 

256
CHAPTER 7

AUTOMATIC GENERATION CONTROL IN

FREQUENCY RESPONSIVE WIND PLANT

INTEGRATED POWER SYSTEM

This chapter presents integration effect of frequency responsive wind plants effect on

load frequency control in a multigenerational interconnected power system, with

special emphasis on some of the challenges presented in chapter 3. Validated models

of conventional power plants are applied to analyze frequency responsive wind plant

integration impact on classical and FGSPID AGC performance. Effect of wind

participation in frequency regulation is analyzed regarding reserve response and

corresponding impacts on wind plant performance. FRT capability of frequency

responsive wind plant is also investigated in this chapter.

7.1. AGC / Load Frequency Controllers and Conventional

Generation Models

The AGC problem has been extensively studied during the last four decades, and it

has been one of the most highlighted issues in the design and operation of

independent and interconnected power systems. Load Frequency Control (LFC) is

used for many years as part of Automatic Generation Control (AGC) in power system

around the world. Maintenance of system frequency within utility prescribed

tolerance band was considered as the exclusive responsibility of the conventional

power plants. With increased renewable energy targets, globally, it is usual to find an

area regulated by hydro generation interconnected to another area regulated by

thermal generation or gas based generation or in a combination of both generations

257
along with good penetration of wind plants. Frequency response capability offered

by a wind plant is not the same as the control capability of conventional plants. While

frequency response capability offered by a conventional plant can be counted on with

good assurance once the appropriate control modes are on, the response capability of

a wind plant will always be conditional on the statistics availability of the wind.

Increased penetration of wind generation in interconnected power system thus

intrinsically calls for LFC analysis with present and future wind technologies. This

chapter presents integration effect of frequency responsive wind plants effect on load

frequency control in a multigenerational interconnected power system.

7.1.1. Classical AGC Controller

In case of free governor operation with fixed speed changer setting in an isolated

system, steady state frequency change due to load disturbance (∆P can be given as

[1]:

∆f ∆P 7.1

The free governor operation gives a reasonable performance with a frequency drop

of 4-5 % between no load and full load. A steady state frequency is reached when the

sum of online generation is equal to system load. For better frequency control some

other mean should be there, as a large change in frequency cannot be tolerated, and

the steady change in frequency should be zero. While steady state frequency can be

brought back to the scheduled value by adjusting speed changer setting, the system

could undergo intolerable dynamic frequency changes with changes in load. The

speed changer setting should be adjusted automatically by monitoring the frequency.

So, setpoint frequency is changed through a centralized controller which takes a

258
signal normally referred as area control error (ACE). An example of classical AGC

controller is integral controller given as:

∆P K ∆f dt 7.2

As an example, steady state frequency deviation ∆F under integral controller for a

thermal plant based system can be written as:


∆F s 7.3

∆f| s ∆F s 0

Where T is governor time constant, T is turbine time constant for thermal power

plant. K is power system gain and T is power system time constant as defined in

Chapter 3.

If the frequency drops by 1 Hz (∆f = -1) then the integrator calls for an increase in

power, with the call increasing at the initial rate of K p.u. MW/s. Negative polarity

must be chosen to cause a positive frequency error to give rise to a negative or

decrease command. From a control system point of view, the integrator has the merit

that it drives the ACE to zero in steady state. Of course, we are implicitly assuming

the system is stable, so the steady state is achievable. So, the gain of the controller

should be optimized to meet high accuracy requirements as the output response

depends in some manner upon the integral of actuating signal. A significant

contribution of this controller to the steady-state performance is however obvious as

the additional integration in the forward loop changes the system order and that the

error of the input is eliminated or considerably reduced, as the practical integration

may not be perfect. For integral type control:

1. If we use subcritical gain settings, we obtain the sluggish non-oscillatory

response of the control loop. This means that the integral of ∆f t , and thus

259
the time error, will be relatively large. In a practical situation, this setting is

most often used. The advantage is that the generator now will not necessarily

“chase” rapid load fluctuations, causing equipment wear.

2. As the sudden load increase sets in, the frequency starts falling off at some

exponential rate. During the first instants, the integral controller has not yet

had time to go into action, and the system response is determined by the

primary automatic load frequency control loop. After a certain time (the

shorter the time, the higher integral gain Ki), the integral controller comes

into action and eventually lifts the frequency back to its original value.

7.1.2. Intelligent AGC Controller

Industrial proportional-integral (PI) or proportional-integral – derivative (PID)

controllers parameters are tuned either manually depending upon operator’s

experience or in an automatic way depending upon process period of several cycles

of operation before getting related parameter gains. This tuning procedure is typically

not applicable in a multigenerational power system with fluctuating renewable power

output serving continuous load demand. For more than four decades, considerable

interest has been shown towards the application of different control strategy for

arriving at more efficient load frequency control for the interconnected power system.

Conventional LFC controllers are simple but may present derogatory control

performance under varying parameters and online frequency tuning. Moreover,

conventional LFC controllers are more model specific failing to provide a satisfactory

response in case of a large non-linear system with stochastic parameters [2-5]. On the

other hand, robust and dynamic LFC controllers involve complex control structures

which are not very much preferred in practices [6]. Motivated by its ability to handle

260
imperfect information and intrinsic uncertainties in the knowledge base, Fuzzy logic

controllers (FLCs) have gained wide popularity for analysis and control of nonlinear

systems [7-9]. Originally developed in 1965 by Zadeh, Fuzzy Logic is now a common

technology at the industrial level. The first application of fuzzy logic to control theory

dates back to 1975 [10]. Fuzzy logic is fundamentally management of non-numeric

and linguistic definite values ranging amid absolutely true and false values, such as

high, low, medium, and often Zero [11].

All fuzzy logic based controller technique comprises three basic components:

Fuzzification: In fuzzification, the input parameter values are fuzzified by some

predefined membership functions, and crisp values are converted to linguistic values

via membership functions. Different membership functions can take different shapes

like trapezoidal, Gaussian, and triangular. However, large membership quantity

rather than type and shape brings more effectiveness in control action regarding speed

and accuracy. Normally recommended for more complex engineering problems,

increased membership functions result in higher accuracy and longer processing time.

By contrast, lesser membership functions provide faster processing time and the

advanced probability of diversity.

Rule base interference: Output parameters are controlled through an inference

engine in this stage. A rule base table is used to apply rules to membership functions.

System knowledge is applied to design these rules by “if–then” concept. Mamdani's

inference method is the most common method that uses the max–min operation

approach.

Defuzzification: Output linguistic values are again converted into relevant crisp and

numerical values using membership functions with the range varying according to

system designer knowledge. Several de-fuzzification techniques exist, such as the

261
center of the area and mean of maxima; the latter is more common in control

applications [13].

Figure 7-1 Conventional Fuzzy PI Controller Model 1 [Reformulated from 12]

Figure 7-2 Conventional Fuzzy PI Controller Model 2

Conventional fuzzy PI controller use K and K parameter as two input and controlled

reference power signal as one output [12]. Area control error and its derivative are

fuzzified using membership functions in different ranges. The conventional fuzzy

logic PI controller is implemented in two different ways as depicted in Figure 7-1 and

Figure 7-2. In Figure 7.1, PI controller output and rate of change of controller output

are used as fuzzy controller input to determine the corresponding output. A

262
membership value for the various linguistic variables is calculated by the rule given

by:

μ e t , ce t , where e t K ACE K ACE dt 7.4

The grade of membership values is determined through the equation of the triangular

membership function given as:

| |
A x 7.5

Where A(x) is the membership grade value, ‘b’ is the width, ‘a’ is the point coordinate

at which the grade of membership is 1, and x is the input variables values.

Conventional fuzzy PI output performance is based on selected membership

functions, corresponding base, and de-fuzzification method. Output power reference

from the model in Figure 7.1 is obtained as:

∑ μu
∆P ∑ μ 7.6

In Figure 7.2, ACE and are two fuzzy input variables with five triangular

membership functions. K′ and K′ are output variables after defuzzification. Output

power reference signal obtained is given as:

∆P K ACE K ACE /s 7.7

7.1.3. Hydro Turbine Governor Model

The representation of hydraulic turbine and water column in load frequency control

study is selected based on the following assumptions:

1. The hydraulic resistance is negligible.

2. The penstock pipe is inelastic, and water is incompressible.

3. The velocity of water varies directly with the gate opening and with the

square root of the net head.

263
4. Turbine output power is proportional to the product of head and volume flow.

Figure 7-3 Classical Hydro Turbine Governor Model

The Classical model of hydro turbine and water column is represented by a single

pole and single zero. For an ideal lossless turbine, change in turbine power output to

change in gate opening is represented in Figure 7-3 and mathematically given as


7.8
∆ .

Water starting time TW represents the time required to accelerate water in penstock

from standstill to the required velocity, and it varies with load. Mechanical power

change (∆ with gate valve change (∆P is given as:

∆P t 1 3e ∆P 7.9

Classic hydraulic turbine linear model being non-minimum phase system has an

important characteristic representing initial mechanical power flow opposite to the

direction of change in gate position. Under sudden gate opening, water inertia

restricts immediate change in flow while pressure across turbine is reduced resulting

in reduced output power. Final steady state output is established only at a steady flow

determined by water starting time.

Hydro Governor Model

As stated above, water inertia brings a non-minimum phase system characteristics to

hydro plant. A large transient droop with a large resetting time is essential for stable

operation of such system, and it is accomplished as transient droop gain feedback as

264
shown in Figure 7-4. This governor model provides high droop during large speed

deviation while exhibits low droop during steady state operation. A linearized model

of this turbine governor was represented in Figure 6-7 of Chapter 6. According to

[14], this speed regulating model will be stable in case of low phase lag contributions

from governor and if


1.5 ∗ 7.10

For transient droop compensator in current hydro model G s ,

transient droop and reset time is calculated according to [15] as

R 2.3 T 1.0 0.015 7.11

T 5.0 T 1.0 0.5 T 7.12

The permanent droop R governs the p.u. frequency deviation in steady state with

p.u. change in valve position. This classic linear model being non-minimum phase

system has an important characteristic representing initial mechanical power flow

opposite to the direction of change in gate position.

Figure 7-4 Hydro Governor Model

Mechanical power change (∆ with gate valve change (∆P is given as:

265
∆P t 1 3e ∆P . 7.13

Hydro turbine rotor speed after a load step ∆ is given as:

∆ω ∆P 7.14

Neglecting transfer function , gate valve position change is given as:

∆ ∆ ∆ ∆
. 7.15

Rate of change of turbine gate valve position just after the load step when ∆ 0

is given as:

∆ ∆
. 7.16

With these equations, it can be concluded that initial mechanical power surge lasts

for 1-2 seconds depending upon load step change, inertia constant and water starting

time. Water inertia restricts immediate water flow during sudden gate opening

thereby reducing pressure across turbine and power reduction. A new steady state

power flow is finally established depending upon water starting time, TW.

Model Validation

Frequency response capability of hydro plant model is evaluated for a hydro unit with

inertia constant of 3 s, load damping constant equal to 1 and droop value of 4.0 %.

Under the assumption of balanced initial load, model performance for primary

frequency response and secondary frequency response was then evaluated under 0.01

p.u. load disturbance. Damping constant of the remaining load of 0.99 p.u. is given

as:

D= ∗ 0.99 ∗ 0.0198 p.u. MW / p.u. Hz


266
Now a regulation of 4.0% means that 4.0 % change in frequency causes a 100 %

change in power generation. So

0.5 p.u. MW / p.u. Hz


Figure 7-5 Frequency Response from Hydro Power Plant

Figure 7-6 Mechanical Power Output from Hydropower Plant during Frequency Response

Composite system frequency response characteristic in this case is:

β D 0.5 0.0198 0.5198 p.u. MW/ p.u. Hz

Steady state frequency response can be calculated as:

∆ .
∆f 0.0192 p.u. Hz or -0.00038 Hz
.

Thus, steady state frequency obtained is 49.9996 Hz

Generation contribution to regulation is ∆P ∆f ∗ β 0.0095 p.u. MW


267
Governor droop response is considered as an uncontrolled primary response while

controlled response refers to AGC response of the unit. The dynamic response of this

model under 0.01 p.u. load disturbance calculated above is shown in Figure 7-5–7-6.

7.1.4. Gas Power Plant Model

The study of a gas based power plant in AGC is important from the viewpoint that

gas-based plants are increasingly being added to the grid system and gas turbines

construction and dynamics are very different from hydro and steam turbines. The

generators are of round-rotor type, have relatively low inertia compared to hydraulic

units, and they spin at higher speeds. A simple cycle gas turbine cycle consists of

three parts. First is the axial compression of gas (typically air). Second is the addition

of heat energy into the compressed gas by either directly firing or combusting the fuel

in the compressed air or transferring the heat through a heat exchanger into the

compressed gas. This stage is followed by expansion of the hot pressurized gases in

a turbine to produce useful work under Brayton cycle. The work required by the

compression is supplied by the turbine with the remaining being available for useful

work. The useful work developed by the turbine may be used directly as mechanical

energy or may be converted into electricity by turning a generator.

Figure 7-7 Split Shaft Based Gas Turbine

268
There exist several dynamic models of turbine-governors with varying degrees of

complexity to represent different makes and models of gas turbine units. There are

essentially two types of gas turbine designs. One is a high-speed single shaft design

with the compressor and turbine mounted on the same shaft as the alternator. Another

is a split shaft design that uses a power turbine connected via a gearbox. In the split

shaft design as shown in Figure 7-7, although there are two turbines, one is gasifier

turbine driving a compressor, and another is a free power turbine driving a generator,

and there are only one combustion section and one gasifier. OCGTs are used mainly

as peaking units due to their operational flexibility, but their relatively low efficiency

means they are typically too expensive for base load operation. CCGTs, which have

relatively high efficiency but reduced flexibility, are more suited for base load

operation. For a base load operation, the gas powered generator is usually more

expensive than coal. Two gas turbine models of varying complexity have been used

in this study for assessing their frequency reserve response.

GAST Model

Figure 7-8 GAST Model

269
The GAST model (based on the split shaft), developed by GE, USA and shown in

Figure 7-8 is one of the models applied for simulation studies for representing the

dynamic behavior of open cycle gas powered turbine governor systems. It is the

simplest representation of the gas turbine model still used in WECC and Western

Power, Australia. The model structure has been selected from the model of reference

[16-17] for simulation studies. Parameters values are in accordance with WSCC

testing programs for gas turbine unit [17].

Figure 7-8 shows three control loop involving fuel control, temperature control loop

and speed feedback loop. A Low-Value Select block is used to determine which

control loop is asking for minimum fuel. This control is then given priority and drives

the fuel valve actuator to control the combustor output. This model is deficient in

predicting accurate operation at high load level. Gas turbine units have relatively low

inertia compared to hydraulic units. They spin at higher speeds, and the generators

are of round-rotor type. Following a sudden load rejection event, a low inertia

machine can experience excessive over-speed, which can be harmful if over speed

protection fails to operate. If over speed protection does operate to arrest the speed,

it will introduce non-linearity in the frequency versus time characteristics. Therefore,

the amount of load rejected should be low to limit over-speed and therefore the

possible impact of non-linearities. The maximum opening by the governor must

permit short time load of at least 104 %. The load level 100 % describes the situation

that the unit is running on maximum load setting and 50 HZ. Model parameters values

are provided in the appendix.

Model Validation

Active power control capability of GAST model based gas turbine for frequency

response performance is evaluated for 95 MW machine with inertia constant of 6.5 s


270
and droop value of 4.2 %. External grid inertia is not considered here, and it is

assumed the system frequency is balance with 1 p.u. of load initially. Model

performance for primary frequency response and secondary frequency response was

then evaluated under 0.01 p.u. load disturbance. Primary frequency response is highly

dependent upon implemented droop and generator characteristics. Damping constant

of a remaining load of 0.99 is given as:

D= ∗ 0.99 ∗ 0.0198 p.u. MW / p.u. Hz


Now a regulation of 4.2% means that 4.2 % change in frequency causes a 100 %

change in power generation. So

. 0.4761 p.u. MW / p.u. Hz


Composite system frequency response characteristic in this case is:

β D 0.4761 0.0198 0.49599 p.u. MW/ p.u. Hz

Steady state frequency response can be calculated as:

∆ .
∆f 0.0201 p.u. Hz or -0.00040 Hz
.

Thus, steady state frequency obtained is 49.9996 Hz

Generation contribution to regulation is ∆P ∆f ∗ β 0.0099 p.u. MW

Dynamic response of this model under 0.01 p.u. load disturbance and calculated

above is shown in Figure 7-9 - 7-10. It can be seen from Figure 7-9 that as soon as

frequency drop occurs, governor droop instantly activates primary responsive FCR

to generate power to bring the frequency to a steady state value of 49.9996 Hz. Flat

frequency control is used for AGC response where the integral controller brings the

steady state frequency to the setpoint frequency of 50 Hz. AGC response when

implemented without system bias or network frequency control characteristics

271
β gives increased settling time to 60 sec while it is only 25 sec. when

implemented with bias as shown in Figure 7-11. β equals to the sum of

individual primary frequency responses of all participating generating units within

the power system.

β ∑ β

In current case, as only one gas turbine based unit is there in network, so

β β 0.4912 p.u. MW / p.u. Hz

Figure 7-9 GAST Model in the Presence of Frequency Response

Figure 7-10 Mechanical Power from GAST Model during Frequency Response

272
Figure 7-11 Frequency Response under AGC Control

Decentralized Combined Heat and Power Plant (DCHP) Model

Simplified combined heat and power plant model based on gas turbine is investigated

for its dynamic response in frequency regulation studies. DHCP model adopted in

this study is based on [18]. The basic configuration of gas turbine governor comprises

of four basic blocks as represented in Figure 7-12.

Figure 7-12 Simplified DCHP Model

Speed Governor Model

System frequency control is provided through governor droop model by activating

corresponding FCRs. A low pass filter and the static dead band are applied to avoid

any undesired frequency response and any frequency deviation above the dead band

273
will be changed into corresponding power command (∆Pc) through implemented

droop characteristic. This frequency-power command is passed to the power

limitations block of gas turbine which is shown in Figure 7-13.

Figure 7-13 DCHP Governor Model

Power Limitation

Figure 7-14 Power Limitation Block

Gas turbine power response is restricted under physical constraints imposed by

combustion technology through this limiter block as shown in Figure 7-14. Power

level corresponding to maximum and minimum load is applied for frequency

response limitation. The lowest value of power demand and limited power is passed

as commanded power. Highest value among this commanded power and reference

power setpoint serves as input to power limiter block which limits the commanded

power between the maximum and minimum load setpoint. Rate limiter block prevents

burner flame pulsation during ramping events like loading and unloading. This load

variation commanded power setpoint is then forwarded to power distribution block

for fuel and air flow management.

274
Power Distribution

Two combustion chamber are used in gas turbine: environment combustion chamber

(EV) and sequential environmental combustor (SEV). The environmental capacity is

60% that of SEV burners capacity. Compressed air is mixed with half of fuel and

heated in first EV chamber. Heated mixture after expansion through high-pressure

turbine moves to SEV chamber where it is mixed with remaining fuel. Efficiency and

emission level of the gas turbine is dependent upon air to fuel ratio. Respective

combustion power contribution factors (CEV & CSEV) and variable inlet guide vane

position compressor factor (CVGV) are used in the model shown in Figure 7-15 to

reflect physical characteristics of air-fuel flow and temperature allowance. CFM as

the function of base load, frequency, and the ambient temperature is considered as 1

in the current study.

Figure 7-15 Power Distribution Block

Gas Dynamics

Combustor and compressor dynamics of the gas turbine are represented by this

module shown in Figure 7-16. First order transfer function models the EV/SEV

environmental combustors dynamics while second order transfer function as

275
represented in Figure represents the compressor/VIGV unit dynamics. A limited

mechanical power output is finally obtained from this module to be passed to the

generator.

Figure 7-16 Gas Dynamic Block

Model Validation

Active power control capability of DCHP model for frequency response performance

is evaluated for 265 MW machine with inertia constant of 8.6 s and droop value of

4.0 %. External grid inertia is not considered here, and it is assumed the system

frequency is balance with 1 p.u. of load initially. Under varying load condition, it was

difficult to validate results analytically. The current study evaluates model

performance for primary frequency response under 0.01 p.u. step load disturbance

and initial DHCP reference power setup as 1 p.u. MW. Output power is limited

between 0 - 1.1 p.u. MW.

Primary frequency response is highly dependent upon reference power setpoint,

implemented droop and generator characteristics. Damping constant is taken as 0.

The primary response is provided through speed governor who provides up/down

power change command as shown in Figure to reduce frequency deviation. For a

276
regulation of 4.0% means that 4.0 % change in frequency causes a 100 % change in

power generation. So, Composite system frequency response characteristic for this

plant is β .

. 0.5 p.u. MW / p.u. Hz


Steady state frequency response can be calculated as:

∆ .
∆f 0.02 p.u. Hz or -0.0004 Hz
.

Thus steady state frequency obtained is 49.999 Hz. Generation contribution is ∆

∆ ∗ β 0.01 p.u. MW. These can be observed in Figure 7-17 - 7-18.

Figure 70-17 DCHP Frequency Response under 0.01 P.U. Load Disturbance

Figure 7-18 DCHP Power Output Under 0.01 P.U. Load Disturbance

277
7.1.5. Thermal Power Plant

Figure 7-19 Single Reheat Turbine Model

Single reheat steam turbine model is shown in Figure 7-19. Steam enters the high

pressure (HP) section through the control valve. With substantial steam remaining in

inlet chest, exhaust steam is passed through reheater. The reheat steam flows into

intermediate pressure (IP) section through intercept valve. A path is provided for IP

section exhaust steam to low pressure (LP) inlet through crossover piping. Time

constants involve in each of these respective stages is shown in Figure 7.19.

Approximated transfer function model of the thermal system is applied under the

simplistic assumption of constant steam pressure and temperature and unlimited

steam supply. Neglecting crossover piping time constant (T and under the

assumption of control valve linear characteristics, simplified transfer function of the

turbine which relates turbine torque ∆T to changes in control valve (∆V ) is

expressed as:


7.17


7.18

278
Here, F refers to fraction of total turbine power generated by HP section, T refers

to time constant of inlet main volume and steam chest, T refers to time constant of

reheater. In the case of the non-reheat turbine, as applied in the current study, the

turbine is represented by single gain and single time constant. With R representing

droop, K as governor gain and T as governor time constant, speed governor is

modeled as simple transfer function represented in Figure 7-20 and expressed as:


∆X s ∆P s 7.19

Figure 7-20 Simple Governor Model

7.2. Performance Index

The design of a control system is an attempt to meet a set of specifications, which

define the overall performance of the system in terms of certain measurable

quantities. Several performance measures have been introduced so far in respect of

dynamic response to step input (t , t , t ) and steady state error e to both step and

higher order inputs. These measures must be satisfied simultaneously in design, and

hence the design necessarily becomes a trial and error procedure. If, however, a single

performance index could be established based on which one may describe the

goodness of the system response, then the design procedure will become logical and

straightforward.

Furthermore, in many of the modern control schemes, the system parameters are

automatically adjusted to keep the system at an optimum level of performance under

279
varying inputs and varying conditions of operation. Such part of the system is called

adaptive control system. These systems require a performance index, which is a

function of the variable system parameters. Extremum (minimum or maximum) value

of this index then corresponds to the optimum set of parameters values. Other

desirable features of a performance index are its sensitivity, i.e. its power to clearly

distinguish between an optimum and non-optimum system, its sensitivity to

parameter variations and the ease of its analytical computation.

Several such performance indices are used in practice, the most common being the

Integral of the Squared Errors (ISE), given by:

SE ACE t dt 7.20

This criterion has good mathematical trackability properties penalizes positive and

negative errors equally. It penalizes heavily on large errors; hence a small ISE usually

results in a system with small overshoot. Since the integration is carried out over (0,

∞), a small ISE limits the effect of small error lasting for a long time and thus results

in small settling time. Also, a finite ISE implies that steady-state error is zero.

Minimization of ISE by adjusting system parameters is a good compromise between

reduction of rising time to limit the effect of the large initial error, reduction of peak

overshoot and reduction of settling time to limit the effect of small error lasting for a

long time.

Another similar index is the integral of time multiplied by absolute error (ITAE)

which exhibits the additional useful features that the large initial error (unavoidable

for a step input) is not heavily weighted. More sensitive than ISE, it is useful to those

practical applications which require very sensitive criteria [19]. ITAE produces

smaller overshoots and oscillations than ISE, and it is given as:

TAE t|ACE t |dt 7.21

280
These two performance indexes are used in the current study to investigate the

superior performance of different controllers. A control system is optimum when the

selected performance index is minimized. MATLAB’s Simulink is used in

conjunction with the optimization algorithm to determine the optimum parameters of

the AGC for the performance index considered.

7.3. Parameter Optimization Algorithm

There are two aspects of the optimization problem, namely, parameter optimization

and optimal system control. This project report deals with parameter optimization of

load frequency controllers. This means that optimum values of the controller

parameters are to be found for minimum deviations in the frequency and tie-line power

oscillations. To achieve optimal performance, various state feedback, state adaptive

optimal controllers, ANN, Fuzzy based controllers have been proposed all of which in

general suffer from complicated computational burden and memory.

A general pattern search algorithm for optimization has been utilized for obtaining

optimal integral gain for nominal input parameters for AGC controllers. In this

method, a combination of exploratory moves and heuristic pattern moves is made

iteratively. An exploratory move is performed in the vicinity of the current point

systematically to find the best point around the current point. After that, two such

points are used to make a pattern move. In the exploratory move, the current point is

perturbed in positive and negative directions along each variable one at a time, and

the best point is recorded.

The current point is changed to the best point at the end of each variable perturbation.

If the point found at the end of all variable perturbations is different than the original

point, the exploratory move is a success. Otherwise, the exploratory move is a failure.

In any case, the best point is the outcome of the exploratory move.

281
Figure 7-21 Direct Search Algorithm

In pattern move, a new point is found by jumping from the current best point x c

along a direction connecting the previous best point and the current base point. The

optimization method is summarized below in the flow chart in Figure 7-21.

282
7.4. Control Area Simulation Models

7.4.1. A Frequency Regulation Model Incorporating All Three Control

Modes in a Single Area Power System

Figure 7-22 Frequency Regulation Model for a Single Control Area

Figure 7.22 represents a hierarchical frequency regulation MATLAB implemented a

model for single area control formulated based on generating unit power reserve

classification given by European Network of Transmission System Operators for

Electricity (ENTSO-E) [20]. The single control area system dynamic is represented

by following differential equation:

∆f s ∆P s ∆P s 7.22


Where T is defined as power system time constant; K , power

system gain and frequency deviation are represented by ∆f. Combined mechanical

power variation from primary, secondary and tertiary control is represented by ∆P

as:

∆P ∆P , ∆P , ∆P , 7.23

283
The load imbalance is represented by ∆P , H is the total system inertia and D the load

damping coefficient of the network. Hydro turbine governor model discussed earlier

has been used for primary control service while thermal turbine governor model also

discussed earlier is applied for secondary control services. Single area control model

consists of frequency containment reserve services (FCR) for primary frequency

regulation, frequency restoration reserve services (FRR) for secondary load

frequency regulation and tertiary replacement reserve (RR) services. Thermal turbine

unit or gas turbine unit can be implemented as RR units to represent tertiary control

dynamics. Proportional controller ∆P 1


R ∆f is utilized as local primary FCR

controllers where R is the TSO provided droop. PID control action on secondary

control command signal is implemented as a tertiary response. The secondary

controller incorporates a centralized tuned proportional-integral (PI) controller to

respond to area control error. All three control loops are shown in Figure 7-22. Single

area system representation by nonlinear equation can be given as:

x f x, u, w 7.24

Complete system model can be defined by 7m+1 state variables:

x ∆f x x … x 7.25

where m is a number of turbines-governors used in each control level and

x ∆P , ∆P , ∆P , ∆P , ∆P , ∆P , ∆P , 7.26

2m controlled inputs u ∆P , ∆P , 7.27

Disturbance w ∆P 7.28

Wind plant state variables have not been considered in above representation. This

representation can be appended depending upon of level of complexity of wind

turbine model and its participation in respective regulation services. According to

284
several of grid codes, primary control should start as early as 5 seconds, and

secondary control should start between 30-60 seconds and dispatched every 2-5

seconds. Hence, the delay associated with the secondary and tertiary controller in this

model has been set as 60 s and 120 s and is dispatched every 10 and 60 seconds

respectively. Primary control is required and modeled to act instantaneously.

Participation factor of individual units in FCR has been calculated as:


PF 7.29

Where predefined reserve power is denoted by FCR is. B is

grid code defined frequency control bandwidth half (Hz) for FCR activation, and R

is droop characteristics in Hz/per unit. Wind participation in FCR model is shown in

Figure 7-23. Wind plant participates in FRR services by taking automatic generation

control (AGC) power set-point from PI controller as represented in Figure 7-24.

Participation factor of individual units in FRR including wind plant has been

implemented as

PF 7.30

Figure 7-23 Primary FCR Service Model with Wind Participation in FCR

285
Figure 7-24 Secondary FRR Services with Wind Participating In FRR

Figure 7-25 RR Control Loop

Figure 7-26 WTG Inertia Controller

RR control loop is shown in Figure 7-25. Instantaneous power reserve from wind

plant is provided through synthetic inertia controller. Input frequency deviation to

synthetic inertia controller as implemented for this study and shown in Figure 7-26 is

provided through frequency processor block, and their output is considered as an

additional power reference point to modify the final power order of wind turbine.

When individual wind turbine participates in FCR control, final power order received

by electrical control is modified asP P P P . When wind

plant participates in an FCR providing group, it receives an extra FCR command P

286
which gets added up in final power order. In FRR participation, additional reference

point P is added. Economic dispatch for the generators is excluded in this study

and 50% wind participation is assumed.

7.4.2. Decentralized Three Area Interconnected Power System (TAIPS)

Model

TAIPS model interconnected in a ring topology as shown in Figure 7-27 has also

been considered to investigate AGC performance. It is assumed that same plant is

participating in primary and secondary control in the individual area. For example,

hydro plant is participating in primary as well as secondary control in area 3 while

gas plant operates these two services in area 2 and thermal plant operates these

services in area 1. An example of such TAIPS model is given in Figure 7-28.

Figure 7-27 Ring Topology for Interconnected Control Area

For a simple model of any area i with an aggregate governor, turbine generator unit

(turbine time constant T , governor time constant T , governor gate opening P ,

mechanical power output P , Power command in terms of speed change setting P )

overall generation demand dynamics can be written in terms of speed change in the

rotating mass with moment of inertia M and damping D as:

∆ω ∆P ∆P D ∆ω ∆P 7.31

Turbine model dynamic can be written as

287
∆P ∆P ∆P 7.32

Governor dynamic can be written as

∆P ∆P ∆ω ∆P 7.33

Each of the control areas is interconnected to another area through tie-lines.

Summated tie-line power gives the total power flow over the transmission line. The

relative phase angle between control areas which further depends upon relative speed

deviations in control area units dictates the direction of transmission line net power

flow. The tie-line power with small deviations in the angles is given as:

∆P cos δ δ ∆δ ∆δ MW. 7.34

With T defined as synchronizing coefficient, tie-line power deviation can be written

as:

∆P T ∆δ ∆δ MW 7.35

where,T cos δ δ and ∆δ 2π ∆f dt rad 7.36

Tie-line power in terms of frequency deviation is given as:

∆P t T ∆f t ∆f t MW; i, j 1,2,3. i j. 7.37

∆P t ∆P t 7.38

Total tie-line power flow between areas-i and other areas can be calculated as

∆P t ∑ ∆P t 7.39

Load frequency control (LFC) operation is accomplished through individual LFC

controller implementing tie-line bias control which implies area control error (ACE)

as a linear combination of frequency deviations and tie-line power. With wind plant

integration to an area, corresponding area control error is calculated as:

ACE ∆P t b , ∆f ; i 1,2,3. 7.40

288
Where ∆P is tie-line power deviations and ∆f is the frequency deviations.

Figure 7-28 TAIPS Model with Integrated Wind Farm

It was shown in equation 6.15-6.17 in Chapter 6; maximum frequency deviation is

dependent upon system inertia and the equivalent regulation constant. In the presence

of power electronics based wind plant increased integration, system inertia reduces

while the equivalent regulation constant increases. Now as presented in chapter 5 and

6, wind plant participation in frequency regulation is possible in current and future

scenario.

There are three equations in literature for calculating system inertia in the presence

of wind plant which has been applied in this chapter.

289
1. With H and S representing inertia rating and apparent power rating of

individual generating units, wind integrated system inertia constant, H

can be calculated as:

∑ ∗
H , where S ∑ S S 7.41

2. Assuming a fraction of frequency responsive wind power plant z

contribution to system inertia by displacing a fraction of conventional

generation z , new system inertia and droop can be expressed as

H H 1 z z 7.42

R R / 1 z z 7.43

3. Wind integrated system modified inertia constant, H using delay theory of

turbine governor [21] [22]:

H H 1 L H L 7.44

∑ ∗
H , where S ∑ S 7.45

Where H and S are the inertia rating and apparent power rating of individual

generating units. L is wind penetration level.

For frequency responsive wind plant, wind inertia contribution H to power system

by providing ∆P extra active power when the system is subjected to step load

disturbance ∆P is calculated as:

H 7.46

∆ ∆
X e 1 7.47
∆ ∆

Recalling from chapter 6, the minimum frequency occurring after load disturbance

with new system inertia and droop can be calculated as:

290
∆f t 1 exp t ∆P 7.48
T

A sensitivity analysis investigation showed that maximum frequency deviation to

changes in D is very small in comparison to R [23] so load damping constant being

very low can be assumed constant and it does not have a remarkable impact on the

exponential term.

7.4.3. Test System for LVRT Capability Assessment

A typical network connection of a wind farm with electrical grid is shown in Figure

7-29 which is used for LVRT simulation tests. Six DFIG based turbines, each rated

1.5 MW are used in 9 MW wind farm which is connected to the 25KV distribution

network. Wind farm exports power to 120 KV electrical grid through 25 kV feeder

which is 30 km long. A constant wind speed at 11 m/s is applied while power

electronics based IGBTs are used for converters. The wound rotor induction machine

with a maximum operating slip of ±20% produce rated converter voltage. The dc-link

capacitors with nominal dc voltage rating of 1100 V are used for their additional surge

capacity. The impact of pulse width modulation (PWM) switching is minimized by

incorporating a three-phase choke. Voltage disturbances are simulated at bus 120, so

that near to zero voltage is experienced at bus 120 until the line is cleared of

disturbances.

Figure 7-29 Wind Farm Network Connection for LVRT Test

291
7.5. Analysis-I

Synthetic Inertia Enabled Grid Code Compatible Wind Plant Participation

Capability in System Frequency Regulation and Its Effects on System Reserve

Response.

The main aim of this study is to show the design possibility of earlier presented grid

code compatible wind plant participation in various level of frequency regulation.

Simulation test was conducted to assess the frequency responsive wind turbine effect

on system frequency control response and the corresponding effect on wind turbine

output. Changes in reserve power response along with frequency response indicators

like the rate of change of frequency (ROCOF) and frequency nadir are analyzed under

generation disconnection condition as well as load disturbance condition. Normal

load disturbance and contingency load disruption conditions have been used in the

analysis. Normal load disturbance given as [0 0.025 0.04 0.05 0.2] p.u. and

contingency load disruption conditions given as [0 0.2 0.25 0.050 0.200] p.u. have

been used in power imbalance function given in [24] for load disturbance generation.

All these simulations are conducted under frequency controller-II based VSWTG.

This simulation model can be readily applied to any combination of frequency reserve

services along with wind plant integration to determine frequency response

indicators.

Continuous primary response with FCR availability along with timely activation is

very crucial for grid stability. Wind farms can participate as a module within an FCR

providing group and the minimum requirement for activation time can be distributed

among all units in FCR group. Following sections discuss wind integration and

corresponding control area response. Single area simulations are mainly classified

into following categories given in Table 7-1:

292
Table 7-1 Simulation Study Set Up Cases

Case FCR FRR WPP Remark


(MW) (MW) (MW)
A 330 500 0 0.0833 No Wind integration
B 330 500 133.2 0.082 Wind plant integrated to system but no
contribution to frequency control services
C 197 500 133.2 0.07 Wind plant is providing a contribution to
FCR services when 40% conventional
hydro FCR generating units are displaced
and FRR provided by thermal based
generating units alone.
D 330 300 200 0.0633 Wind plant providing contribution to FRR
services when 40% conventional thermal
FRR generating units are displaced and
FCR provided by hydro generating units
alone
E 197 300 333.3 0.05 Wind plant is providing a contribution to
FCR and FRR services with 40%
conventional generating units displaced in
respective services.

Case B

Figure 7--30 Control Area Frequency under Normal Disturbance and No Frequency Control

The control model is first tested for frequency variation and respective changes in

reserve power under normal per unit load change conditions as shown in Figure 7-30

– 7-31. Wind plant is just feeding generated power to the system without participating

293
in any frequency responsive services. FCR service is activated to regulate the system

frequency quickly, and their response is observed as a symmetric version of

frequency. FRR service response which is a resemblance of load disturbance profile

curve in red helps in restoring frequency to the nominal value. RR service is required

to be operational only in case of severe frequency deviation.

Figure 7--31 Regulation Reserve Response with Integrated Wind Providing No Frequency Response

Case C

Figure 7-32 FCR Control Signal Received by Wind Plant When 40% Conventional FCR Units are

Disconnected

As per Nordic grid code, primary Frequency Containment Reserve (FCR) should be

activated as soon as frequency reaches 49.9 Hz. The Same condition is valid for

NEM, Australia which requires regulation FCAS to be activated as soon as frequency

294
reaches 49.9 HZ. In second simulation condition with a random contingency load

disturbance condition and 40% of hydro FCR units disconnected, frequency

controller 2 based wind plant is used to participate in FCR services. When wind farm

participates in FCR group, FCR command signal as per the participation factor

defined previously is sent to wind farm. This FCR command signal as shown in

Figure 7-32 is further distributed to individual wind turbines through farm controller.

An aggregate wind farm model has been used in current study. Total FCR power

change as shown in Figure 7-33, consisting of 40% wind power and 60% thermal

power is initially responsible for frequency deviations reduction.

Figure 7-33 FCR Response with 40 % Wind Contributing to FCR Reserve

Figure 7-34 Power Set-Point Generated by Frequency Controller 2 when Wind Plant Participates as FCR

295
Various power setpoint generated during contingency load condition is presented in

Figure 7-34 where it can be noticed that during upward high frequency, power

setpoint drops to 0.6-0.59. Grid controller block generates frequency responsive

active power setpoint varying between 0.97 p.u. - 0.6 p.u. according to frequency

deviation and available power. Limited power setpoint as per grid code condition is

forwarded to turbine controller which further generates total power command for

generating electrical power. An increase in wind turbine power output is observed

when frequency increases above set dead band resulting in total power to approx.

1.01 p.u. including inertial power as shown in Figure 7-35.

Figure 7-35 Generated Power with Inertial and Frequency Controller 2 Enabled when Wind Plant is

participating in FCR

Power reserve response during the loss of 40% of the conventional generating unit

under random high load demand conditions is shown in Figure 7-36. FCR units

though provide support to arrest initial frequency overshoots but are not able to

regulate the frequency variations which activates FRR services which provide

comparatively fast response to follow the high demand change while RR services

activation is quite slow. Activation of FRR and RR service can be made faster by

controller tuning. If the wind plant is made to operate in FCR control with normal

296
droop-inertia controller enabled but no frequency responsive active power setpoint,

wind output is observed dropping to 0.35-0.4 per unit, and no support is noticed in

terms of total FCR power change. Increasing the wind turbine output, in that case,

requires changing the torque controller parameters as suggested in [25] and in the

previous chapter.

Figure 7-36 Power Regulation Reserve Response for Random High Load With 40% Hydro Replaced

With Droop-Inertia Enabled Frequency Responsive Wind Plant

Case D

Frequency responsive wind plant participation in secondary frequency control is

assessed by providing FRR power setpoint for winding turbine controller loop. Two

different sets of simulations were conducted for secondary control response

investigation under normal and contingency load disturbance. System regulatory

power change response is analyzed when 40% of wind based plant capacity

participates in secondary control with 60% of thermal based plant and 100% hydro

based FCR control and 40% Wind plant with 60% hydro based plant FCR

combination. FRR power reserve AGC signal from PI controller is distributed to all

participating units. A clear increase in FCR and FRR power reserve activation

response is observed with wind participation in these services due to the high ramping

capability of the wind plant. An example of FCR and FRR command signals received
297
by individual turbines when there is a random high load variation is shown in Figure

7-37, and corresponding power reserve change is given in Figure 7-38.

Figure 7-37 FCR and FRR Command Signal Received by Wind Turbine Unit under Random High Load
Condition

Figure 7-38 Reserve Response during Random Normal Load Variations When The Wind Participates in

FRR Only

Case E

In this case, the wind farm is required to provide more response to FCR services

command in comparison to FRR services. Comparing reserves response from Figures

7-36, 7-38 and 7-39, we can observe that FRR was comparatively fast and matches

closely with load disturbance while a derogatory FRR is achieved in the case of

incorporation of wind plant in both services as noticed in Figure 7-39. The addition

of wind plant output in both FCR and FRR services reduces considerable system

298
inertia and adds extra stress on automatic generation controller requiring parameter

tuning for better system frequency settling response. It is to be noted that PI controller

has been used in wind-based FRR service controller while integral controller has been

used in conventional plant-based FRR services when wind plant participates in FCR

services only. Proper controller selection and tuning is another important topic which

is out of scope for current study.

Figure 7-39 Reserve Response during Random High Load Variations When The Wind Participates in
FRR & FCR

Frequency nadir after any contingency is detrimental to primary FCR performance

for maintaining system stability [26]. Wind plant integration and its participation in

regulation services can be analyzed through frequency nadir points achieved under

different combinations. In the current study, highest drop in frequency is observed

during contingency load condition indicating insufficient frequency control power

reserves while no wind integration gives minimum frequency nadir. Figure 7-40

shows the minimum frequency nadir and ROCOF under varying wind conditions and

different integration scheme. Lowest frequency nadir point is obtained when the wind

is participating in FCR and FRR services both during under during contingency high

load and during normal load conditions. These dips in frequency nadir points are due

299
to generation disruption resulting in lower system inertia. 40% of conventional units

in each service are replaced with wind plant resulting in lower system inertia.

Figure 7-40 Frequency response Indicators Comparison with Wind Plant Participation in Different
Regulation Services and Different Load Demand

ROCOF is lowest without any wind integration. ROCOF is highest when wind plant

is integrated with network and participating in secondary control (FRR) and primary

control (FCR) both. When 40% wind plant is contributing to FCR along with 60%

conventional hydro plant, the rate of change of frequency (ROCOF) increases in

comparison to that of 100% conventional system FCR support without wind power

addition to the network. This ROCOF change from 0.005 Hz/s to 0.5 Hz/s is due to

change in total inertia of the system. An approximate increase of 50% in ROCOF is

observed during contingency load condition when primary FCR and secondary FRR

are not able to avoid sustained frequency deviations. The values obtained suggest that

no relay tripping is possible with around 40 % wind penetration in regulatory

services.

Impact of Wind Plant Participation in Regulation Services on Generated

Electrical Power, Rotor Speed, and System Frequency

Wind electrical power output is limited to 1.1 per unit in this study. Under normal

load disturbance and sufficient power reserves, electrical power of 1 per unit without

any inertial power is produced as shown in Figure 7-41. Effect of the drop in wind

300
speed change can be noticed with a drop in electrical power and rotor speed during

1800-2300 s which is shown in Figure 7-42. Figure 7-43 and Figure 7-45 shows the

increased electrical power output and rotor speed of wind plant during random high

load conditions and participation in regulation services. Zoomed version of electrical

power for a time duration of 500-1000 second is shown in Figure 7.44.

Figure 7-41 Electrical Power under Normal Load Disturbance (Inertial Power is Zero)

Figure 7-42 Rotor Speed under Limited Power Setpoint

With the increase of power from 1 p.u to 1.1 p.u. during time step of 550-650 sec, a

relatively small drop in rotor speed from 1.2 p.u. is obtained during the same time

duration. Under extreme frequency conditions on the downward side, inertia

controller provides 0.1 p.u. extra power to the total wind generated power as shown

in Figure 7-44. Wind turbine inertia controller gets activated if input frequency is

more than the frequency dead-band limit and inertia controller power output being

301
limited to 0.1 p.u. which is set in accordance with a change in rotor speed to avoid

any mechanical damage. With set power limitation and reference speed, wind turbine

operation is in a stable state during the whole operation. The torque-speed controller

can maintain rotor speed up to 1.2 p.u. with a minor drop of 0.01 p.u. during the

inertial response. Simulation results show that with modified control loop, the wind

turbine can implement frequency responsive power setpoint by generating frequency

responsive power and maintain rotor speed up to a maximum of 1.5 p.u. for wind

speed varying up to a maximum limit of 20 m/s. Wind turbine protection loops are

enabled at high wind and high-frequency deviations.

Figure 7-43 WTG Electrical Power during Load Disturbance Contingency

Figure 7-44 WTG Electrical Power during Load Disturbance Contingency (Zoomed)

302
Figure 7-45 WTG Rotor Speed during Random High Load

Figure 7-46 System Frequency when Wind Plant Participates in Different Regulation Services at High

Load Disturbance

Figure 7-47 System Frequency When Wind Plant Participates in Different Regulation Services at normal

Load Disturbance

303
Control area system frequency achieved with wind integration and participation in

regulation services is shown in Figure 7-46 - 7-47. With evident inertia changes, best

regulation of frequency to the setpoint is seen with wind plant participation in

individual control service like FCR services only while frequency overshoots at some

instances when the wind participates in both primary and secondary control services.

It is to be noticed that wind plant can respond to AGC signal for FRR participation if

frequency nadir goes beyond a minimum threshold and all FCR reserves are

exhausted.

Figure 7-48 System Frequency during a Contingency/High Load

Figure 7-49 WTG Electrical Power during Contingency/High Load

FRR capable wind plant closely follows the system frequency in terms of generating

power as per their capacity. It can be seen in Figure 7-49 that electrical power is 1.1

per unit power at a time step of 650 s when corresponding frequency drop is lowest

304
as shown in Figure 7-48. As soon frequency starts increasing, wind turbine adjusts its

power out as per grid code active power set-point. More clear response for wind

participation in FRR services at a random load can be noticed in Figure 7-51 where

WTG increases its output to 1.1 per unit at around 400 sec when system frequency

crosses the threshold of 49.5 Hz as shown in Figure 7-50. Wind FRR response is

seized as soon as frequency improves and starts generating as per given active power

setpoint. More overshoots and undershoots in frequency and corresponding

fluctuations in WTG power outputs are observed during a random load when the wind

participates in FCR services only or combination with FRR services.

Figure 7-50 Comparison of Wind Participation in FCR & FRR Response In Terms Of System Frequency

during a Random Load

Figure 7-51 Comparison of Wind Participation in FCR & FRR Response In Terms Of WTG Power

during a Random Load

305
7.6. Analysis-II

Frequency Responsive VSWTG Integration Effect on Multi-Generational

System AGC Control

For generating systems of widely different characteristics, research needs to be done

to study automatic generation controllers in interconnection. A steam turbine converts

the potential energy of high temperature and pressure steam into mechanical

(rotating) energy, which then is converted into electrical energy in the generator.

Compared with steam turbines, hydro turbines are easier and cheaper to control, while

gas turbine model is quite complex with respect to earlier two as this system must

take care of different control loops and actuate only one as per requirement. The

hydropower systems have the relatively large inertia of the water being used as the

source of energy which causes a significantly larger time lag in response to changes

of the prime mover torque, due to changes in gate position. Also, the response may

contain oscillating components caused by the compressibility of the water or by surge

tanks while gas turbines have relatively low inertia compared to hydraulic units. They

spin at higher speeds, and the generators are of round rotor type. Following a sudden

load rejection event, a low inertia machine can experience excessive over-speed,

which can be harmful if over speed protection fails to operate. Modern WTGs have

inertia constants which are comparable to those of conventional turbine-alternators.

Also, VSWTs can operate in a wide range of speed changes. The generator speed can

drop to as low as 0.7 p.u. speed, while conventional unit speed can only drop to as

low as 0.95 p.u. speed. Wind kinetic energy could be utilized to provide temporary

primary frequency control support to the grid in the event of a load/generation

mismatch. However, the power is limited by the operating conditions and the power

rating of the VSWT.

306
As presented in Chapter 5, an important feature of VSWT is the possibility for their

active power outputs to be controlled as required by system frequency and operators

command. VSWT can increase their output power almost instantaneously. This is an

important feature, although the steady-state active power delivered to the grid by a

VSWT depends on the mechanical energy transferred from the wind, the electric

power can be transiently controlled, to a certain extent, by resorting to the mechanical

system kinetic energy. With advances in control techniques, wind-based plants are

expected to participate in secondary frequency regulation. Wind plants with variable

output are integrated with peak load plants like gas based plants & hydro plants rather

than with baseload thermal plants. Dynamics of different combinations of turbine-

generators have different effects on system frequency maintenance.

This frequency regulation study is conducted to analyze wind plant integration with

thermal plant, gas plant and hydro plant in an interconnected control area and

discusses the tradeoff in selecting a hydro or thermal plant with comparatively low

inertia gas plant. Three area control model incorporating thermal, hydro and GAST

based gas power plant model presented earlier has been used for this investigation.

Hydro and thermal generating unit’s parameters are applied according to [27] [1] and

other system parameters are calculated as per the equations given earlier. Though

several simulations were conducted for different configurations of conventional

generating units, the interaction of thermal-gas-hydro implemented in area 1-area 2-

area 3 are discussed here.

Simulation test scenarios are constructed as follows:

1. Area 1 includes a thermal based generator which is combined as a single unit

participating in both primary and secondary regulation. Area 2 includes gas

based power plant while area 3 has a hydro based power plant.

307
2. Wind power is included in one area only.

3. Load disturbance is simulated by 0.01 p.u. step function at 5 s.

4. Level of wind penetration is adjusted as per variation of inertia reduction: 0%,

10%, 20%, 40% and 60%.

5. Normally operated wind plant and grid code compatible frequency responsive

wind plant are applied. Rated wind speed of 15.5 m/s is applied in this study.

6. The impact of wind power on AGC performance is analyzed with and without

considering the variation in frequency bias and equivalent system droop.

7. Area bias factor B is taken equal to power plant frequency response

characteristic .

8. Frequency responsive wind plant participation in Area 1 AGC is assessed and

simulated for different participation factor of thermal-wind as 0.9-0.1, 0.8-

0.2, 0.6-0.4 and 0.4-0.6.

9. Both equations of system inertia defined earlier are applied for frequency

responsive wind plant to analyze the impact of R, H, and B on system

frequency deviation.

10. A comparison of frequency deviation for GAST based gas turbine model is

made with DCHP based gas power plant. Effect on frequency response due to

the variation in fuel limit in GAST model is analyzed. DHCP operates at an

initial power of 0.75 p.u. All other values are provided in Appendix.

With Step reduction factor α = 2, termination parameter ε = 10 –4 & step size Δ = 0.1,

optimum values of integral controller gain were found as given below. The integral

of time multiplied by absolute error ITAE has been used for calculating performance

index of the system in this study and provided in Table 7-2:

308
Table 7-2 Optimum Values of AGC Integral Controller

Optimum Value KI / ITAE KI/ ITAE KI/ ITAE

Thermal-Gas-Hydro -0.3502/ 5.3857 -.0865/34.6329 -0.85/46.5892

Scenario 1: No Wind Integration

Three area control model is simulated to assess frequency response without wind

plant integration. System parameters and corresponding frequency response

indicators are given in Table 7-3.

Table 7-3 System Parameters and System Frequency Response Indicators Obtained Without Wind

Integration

Inertia D R 2H B ∆
Reduction
(p u)
%

0 0.01 4 10 100 20 0.26 0.15 0.01

Control Area / Generating Unit ∆ (Hz) Settling Time (s) ROCOF (Hz/s)

Load Disturbance of 0.01 p.u. is given at 5 s in all three areas.

1 – (Thermal) -0.063 60 0.05

2 – Gas -0.062 60 0.05

3 – Hydro -0.085 60 0.05

Load Disturbance of 0.01 p.u. is given at 5 s in Area 1 only.

1 – (Thermal) -0.027 50 0.05

2 – (Gas) -0.020 60 0.015

3 – (Hydro) -0.026 60 0.017

Load Disturbance of 0.01 p.u. is given at 5 s in Area 2 only.

1 – (Thermal) -0.025 50 0.017

2 – (Gas) -0.034 70 0.05

309
3 – (Hydro) -0.036 70 0.023

Load Disturbance of 0.01 p.u. is given at 5 s in Area 3 only.

1 – (Thermal) -0.034 60 0.02

2 – (Gas) -0.038 70 0.025

3 – (Hydro) -0.043 60 0.05

Figure 7-52 Comparison of Area 1-2-3 Frequency Response for 0.01 p.u. Load Disturbance in Area 1

Only

Figure 7-53 Comparison of Area 1-2-3 Frequency Response for 0.01 p.u. Load Disturbance in All Three
Area

310
Figure 7-54 Mechanical power Change of all Generating Unit in Interconnected 3 Area for 0.01 p.u. Load
in Area 1 Only

Figure 7-55 Mechanical Power Change of All Generating Unit in Interconnected 3 Area for 0.01 p.u.
Load in all Three Area

Figure 7-56 Tie-Line Power Change of All Generating Unit in Interconnected 3 Area for 0.01 p.u. Load

In Area 1 Only

311
Figure 7-57 Tie-line Power Change of All Generating Unit in Interconnected 3 Area For 0.01 p.u. Load in

All Three Area

Scenario 2: Wind plant integration without any frequency response

Type 3 based aggregate wind plant model is integrated with each conventional plant

in the different area and frequency response indicators investigated. Table 7-4

provides the value of R, B, and H with and without considering wind integration

impact. These values are applied only to wind integrated area. Equation 7.44 has been

applied to calculate system inertia for frequency responsive wind plant.

Table 7-4 Wind Plant Integrated System Data

Inertia
Frequency Responsive Wind Plant
Without Considering Wind Effect

Reduction R H B R H B H
(%)

0% 4 5 0.26 4 5 0.26 5
on R & B

10% 4.45 4.5 0.235 4 4.5 0.26 4.76

20% 5 4 0.21 4 4 0.26 4.50

40% 6.67 3 0.159 4 3 0.26 3.88

60% 10 2 0.11 4 2 0.26 3.10

312
Figure 7-58 Frequency Deviation in Area 1 for Different Level of Wind Plant for 0.01 p.u. Load in Area 1

Considering the Effect of The Wind on R and B

Figure 7-59 Frequency Deviation in Area 1 for Different Level of Wind Plant for 0.01 p.u. Load in Area 1

without Considering the Effect of The Wind on R and B

Scenario 3: Frequency responsive wind plant integration

Grid code frequency responsive wind plant presented in Chapter 5 is integrated with

a thermal plant in area 1. In the second case, same wind plant is integrated with

thermal, gas and hydro for comparing their response.

313
Figure 7-60 Frequency Deviation in Area 1 for Different Level of Wind Plant for 0.01 p.u. Load in Area 1

Considering the Effect of The Wind on R and B

Figure 7-61 Frequency Deviation in Area 1 for Different Level of Wind Plant for 0.01 p.u. Load in Area 1

without Considering the Effect of The Wind on R and B

Figure 7-62 Comparison of Frequency Deviation When Only Conventional Plant Participates In AGC in

the presence of Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance.

GAST Model with Lower Fuel Limit of -0.02 is applied for Gas Plant Model

314
Figure 7-63 Comparison of frequency deviation When Only Conventional Plant Participates in AGC in
the presence of Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance.

Figure 7-64 Interconnected 3-area Frequency Response When Gas Based Area 2 is Integrated with
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is Applied in area 2. GAST Model
with Lower Fuel Limit of -0.02 is applied for Gas Plant Model.

Figure 7-65 Interconnected 3-area Frequency Response When Gas Based Area 2 Is Integrated With
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is Applied in Area 2. GAST Model
with Lower Fuel Limit of -1 is applied for Gas Plant Model.

315
.
Figure 7-66 Interconnected 3-area Frequency Response when Thermal-Based Area is Integrated with
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is Applied in Area 1.
Only Conventional Generating Unit Participates in AGC.

Figure 7-67 Interconnected 3-area Frequency Response When Hydro Based Area 3 is Integrated with
Frequency Responsive Wind Plant and 0.01 p.u. load Disturbance is applied in Area 3.
Only conventional Generating Unit Participates in AGC.

Figure 7-68 Interconnected 3-area Frequency Response when DCHP Based Area 2 is integrated with
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is applied in Area 2. Only
Conventional Generating Unit Participates in AGC. Load Limit (1.2, -0.1) p.u.

316
Figure 7-69 Interconnected 3-area Frequency Response when DCHP Based Area 2 is Integrated with
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is applied in Area 2.
Only Conventional Generating Unit Participates in AGC. Load Limit (1, 0.02) p.u.

Figure 7-70 Interconnected 3-area Frequency Response when DCHP Based Area 2 is Integrated with
Frequency Responsive Wind Plant and 0.01 p.u. Load Disturbance is applied in Area 2. Only
Conventional Generating Unit Participates in AGC. Load limit (1, -1) p.u.

Scenario 4: Frequency responsive wind plant participation in AGC

Grid code frequency responsive wind plant with frequency controller-II participates

in AGC with each conventional plant under different participation factor which is

decided as per their penetration level.

317
Figure 7-71 Comparison of Frequency Response for Thermal-Based Area 1 Integrated with Frequency
Responsive Wind Plant when Both of Them Participates in AGC as Per the Indicated Participation
Factor. Effect of Wind Plant on R and B has been considered here.

Figure 7-72 Comparison of Power Deviation for Thermal-Based Area 1 Integrated with frequency
Responsive Wind Plant when Both of Them Participates in AGC with 0.4 Participation Factor (PF) for
Thermal and 0.6 PF for The Wind. Effect of Wind Plant on R and B has been considered here.

Figure 7-73 Comparison of Frequency Response for Thermal-Based Area 1 Integrated with Frequency
Responsive Wind Plant when Both of Them Participates in AGC as per the Indicated Participation
Factor. Effect of Wind Plant on R and B has not been considered here.

318
Figure 7-74 Comparison of Power Deviation for Thermal-Based Area 1 Integrated with Frequency
Responsive Wind Plant when Both of Them Participates in AGC with 0.4 Participation Factor (PF) for
Thermal and 0.6 PF for The Wind. Effect of Wind Plant on R and B has not been considered here.

Discussion:

Figure 7-75 ACE obtained Under Different Simulation Scenarios

319
Each power network has a specific characteristic based upon its components and

parameter involved still some noteworthy observations can be made from these

simulation tests:

1. Any positive values of area bias factor B along with an integral controller

guarantees that all ACEs will be brought to zero eventually [28]. Also, it has

been suggested earlier in [29] that selecting area frequency response B equal

to system bias factor β in all areas fulfills the non-interaction period between

frequency and tie-line power as well as interconnected area for low-frequency

approximation. Simulation results show that this rule remains valid for low

load scenario without any wind power integration giving minute ACE for all

three areas. In the case of wind integration with system inertia reduction;

setting B= β fails the non-interaction criteria for pool operation of the

interconnected network as shown in Figure 7-75. Under all the same

parameters, ACE in wind integrated area increases under the rule of B= β as

is evident from the Figure 7-75. With frequency responsive wind plant, ACE

decreases in comparison to normal wind plant integrated the system.

Incomplete speed droop adjustment of conventional units and additional

droop response of wind plant along with any surplus generation can lead to a

different bias factor. Any change in scheduled transmission line power needs

to be considered with wind power integration.

2. From scenario 1, Table 7-3, it can be noticed that frequency deviation is much

smaller in the absence of any wind plant and it increases when load

disturbance is given in all three areas. With same generating capacity and

under low load disturbance, highest and fastest power is provided by thermal

plant followed by gas and then by the hydro plant as shown in Figure 7-54. In

320
the case of high load disturbance, thermal plant response is followed by hydro

and then by the gas plant as shown in Figure 7-55. Viewing the tie-line power

response in Figure 7-56 - 7-57, it can be observed that highest tie-line power

deviation under low load is for thermal based area 1 followed by a gas plant

in area 2 while hydro based area 3 has an excess of power. This scenario

changes under high load in all three area where highest tie-line power

variation is observed for hydro based area 3 making it suitable as base load

plant.

Figure 7-76 Comparison of Frequency Deviation for Thermal-Based Area 1 Integrated with Wind
Plant. ThermalAGC-Wind Refers that only Thermal Plant Participates in AGC and Normal
Operating Wind Plant is integrated into System. The Wind (RB) Indicates R and B Values are changed
as Per Inertia Reduction Impact. FRWind Refers to Frequency Controller-II Based Frequency
Responsive Wind Plant. ThermalAGC-FRWINDAGC Refers to Participation of both Units in AGC.

3. With increasing amount of wind power in an interconnected network, there is

an increase in frequency deviation due to a proportional reduction in inertia

and higher amount of equivalent system droop. This can be observed in

321
Figures 7-58 - 7-61. A comparison graph for frequency deviation for thermal

based area 1 integrated with wind plant is shown in Figure 7-76 where we can

see that frequency deviation observed in area 1 under the effect of wind

increases with increasing amount of the wind and corresponding inertia

reduction. Frequency deviation is -0.025 for area 1 when no wind power is

added while it increases to -0.55 Hz for 60% normally operated wind power

penetration in the network when no effect on R and B is considered.

Frequency deviation worsens further to -1.35 if values of R and B are also

changed correspondingly. Frequency nadir is sensitive to system droop and

power generated in a control area while it shows low sensitivity to system

inertia. Time to reach frequency nadir being sensitive to system inertia

increases with reduction in inertia due to increasing wind penetration.

Figure 7-77 Comparison of ROCOF for Thermal-Based Area 1 Integrated with Frequency Responsive
Wind Plant. ROCOF values are observed when 0.01 p.u. of Load is applied at 5 s in Area 1 Only. Legends
are same as explained above in earlier Figure.

322
4. Even though only generic models are applied in this study, still some

significant observations can be made for ROCOF. Comparing the values of

rate of change of frequency which is measured at 5 s when 0.01 p.u. step load

is applied; it can be seen in Figure 7-77 that ROCOF increases with increasing

wind penetration. This effect is more prevalent when the effect of the wind on

values of R and B are considered in simulations. At higher penetration level

of 60%, ROCOF values improve to -0.4 Hz/s when frequency responsive

wind participation is more than the conventional unit in comparison to the

ROCOF value of 0.9 Hz/s when normally operated wind plant is integrated.

This behavior calls for a deeper analysis for a high wind penetrated system as

protective relays based on the rate of change of frequency are highly

vulnerable to ROCOF values. Further to this, the trend of increasing the

ROCOF values due to the consideration of variation in R and B values

changes for 60% wind penetration. We observed that ROCOF values improve

for higher wind penetration both normally operated and frequency responsive.

Figure 7-78 Comparison of Settling Time for Thermal-Based Area 1 Integrated with Frequency

Responsive Wind Plant. Legends are same as explained above.

323
5. Similar to ROCOF values, settling time also increases with penetration level

as observed in Figure 7-78. A decreasing trend in settling time is observed

when the wind participates in AGC along with the conventional unit. This

happens due to higher ramping of wind power with faster frequency response

controller. Figure 7-72 and 7-74 shows the power variation in area 1 when

thermal-wind plant participates in AGC with participation factor (PF) of 0.4-

0.6. With higher PF, wind plant with faster power response provides more

power than a thermal plant having low PF. Under low load disturbance of 0.01

p.u. in area 1 only, tie-line power support from other area is low. High

deviation in frequency (-0.9 Hz) and increased settling time (120 s) is

observed for changed values of R and B under wind impact while this

deviation is highly reduced to -0.2 Hz and settling time is reduced to 70 s if R

and B are not changed as per inertia reduction.

Figure 7-79 Main Effect Plots for AGC Parameters on System Frequency Deviation.

6. Area 1 was integrated with frequency responsive wind plant in different

penetration level and system inertia calculated according to both equations

7.42 and 7.44. R and B values were also varied as per Table 7-4. Frequency

324
deviations obtained was plotted against H, R and B. It can be observed from

Figure 7-79 that frequency deviation is least affected by system bias value

while it is highly affected by R variation.

7. Contingency reserves in Australia are divided into six categories as defined

in chapter 3 earlier. Four of these services are fast raise (R6) and fast lower

(L6) which provides a response within 6 seconds while slow raise (R60) and

slow lower (L60) provide stabilization response within 60 seconds.

Conventional generating unit frequency response was investigated when

respective control area has 10% inertia reduction due to wind integration.

Thermal plant is integrated into area 1 with 10% frequency responsive wind

plant and load disturbance of 0.01 p.u. is given in area 1 only while other areas

support the area 1 during this disturbance. Similarly, it is done for area 2 and

area 3. Mechanical power of thermal generating unit changes rapidly in

comparison to two other units.

Figure 7-80 Comparison of Frequency Response Indicators for Conventional Plant Integration with
Frequency Responsive Wind Plant when Only Conventional Unit Participates in AGC and 0.01 p.u. Load
is applied.

Comparing the individual response of conventional units from Figure 7-63 -

7-68 and 7-80, we can observe that thermal plant gives fastest and best

combination for integration with wind plant followed by gas and then by the

325
hydro plant. Change in initial opposite transient power due to gate opening at

penstock, hydro turbine governors require transient droop compensation with

long resetting time for stable speed operation and give a relatively slow

response. Thermal and gas plant with their normal speed droop gives a faster

ramping response, but their response decreases over time due to fuel

limitation of gas and steam. Their turbine output is highly dependent upon set

fuel and load limit. Thermal and gas based units though have relatively lower

efficiency, but their combination with fast, responsive units like frequency

responsive wind can reduce regulation procurement in an area. So, in the

Australian context, thermal of gas based units are more suitable for R6 FCAS

in the wind-integrated area while the hydro plant is better suited for R60

services.

8. A comparison of GAST model with DCHP plant model when integrated with

frequency responsive wind plant is conducted for frequency deviation under

0.01 p.u. load disturbance. The use of negative value for the low limit in

GAST model represents negative mechanical power. At high loading levels,

the high exhaust temperatures can damage the turbine blades. Therefore,

when the exhaust temperature exceeds a set temperature, gas turbine output

transfers to temperature control. The GAST model does not give an adequate

representation of the temperature control loop. Though GAST model has fuel

limitation, it was observed that by changing the GAST fuel limit from 0.02 to

-1 p.u., there is no effect on system frequency deviation but settling time

shows improvement in the interconnected system as shown in Figure 7-62 -

7-65. It can be seen from Figure 7-68 – 7-70 that much higher deviations are

observed for DCHP plant in comparison to simple GAST model. DCHP

326
model output is highly dependent upon power limitation and maximum-

minimum load limitation for frequency response. If the turbine is initially

below full-load (i.e. below its temperature limit), then in the event of a

frequency excursion there will be a transient overshoot in power output as the

governor acts to increase power. After a few seconds (the time constant

associated with the thermocouples measuring exhaust temperature) the

temperature control loop will take control and start to decrease turbine output

to maintain the exhaust temperature limit if the governor attempts to push it

beyond this limit. GTs, whether stand-alone or as part of a DCHP, can make

some change in mechanical output affecting the frequency response. The

increase in speed following a frequency disturbance should result in some

response by the speed governor to decrease turbine power during the first

swing in machine speed. This initial governor response, even if small, will

improve the ability of the unit or plant to remain synchronized. It is important

that a detailed representation including temperature control and maximum

rate of fuel valve closing be included in models.

7.7. Analysis-III

Development of FGSPID Based Intelligent AGC and Analysis in the presence of

Grid Code Frequency Responsive Wind Power Penetration

Gain scheduling is a distinct non-linear control technique to alter linear controller

parameters in a non-linear mode depending upon process state and scheduling

variable. With changing process operating points, scheduling variables also changes

its values. Fuzzy gain scheduling works on the principle where PID controller gains

are altered during the online process through fuzzy logic and reasoning. Many

researchers consider conventional fuzzy PI controller as gain scheduled controller,


327
but the original concept of fuzzy gain scheduling proposed in [30] differs from

conventional fuzzy PI techniques as the controller parameters obtained after de-

fuzzification are further normalized between a maximum and minimum range

depending upon stable system response. This concept is applied in the current study

and explained briefly in this section. The transfer function of PID controller is given

as:

G s K 1 T s 7.49

where T = K /K , T , K , K and K are proportional, integral and derivative

gains. T & T are integral and derivative time constants and related as: T ∝T .

So K 7.50
∝ ∝

Figure 7-81 Fuzzy Gain Scheduling (FGSPID) Structure

Proportional and integral derivative gains are assumed to lie within a prescribed

range, and gain scheduling method has been applied to adjust K , K and ∝

parameters conferring to area control error signal (ACE) and its first difference

changes, as illustrated in Figure 7-81 and represented by membership functions which

are shown in Figure 7-82 – 7-83. Two input ACE and ∆ACE with A and B being

corresponding fuzzy sets are applied while we have three outputs K , K with C and

328
D corresponding fuzzy sets while last output ∝ being held constant. Fuzzy rules are

selected on the generalized theory behind the step response of the process thereby

bringing area control error to zero and system frequency to set value.

Figure 7-82 Membership Functions for and

Figure 7-83 Membership Functions for ACE and ∆ACE

Figure 7-84 Hydro-Turbine Governor Response

329
Figure 7-84 shows the step response of hydro turbine governor system. It can be

noticed that at point A, the large rise time is desired implying large proportional and

integral gain and small derivative gain from LFC controller. Small T or small ∝ will

result in a strong integral gain. Rule around point A can be (Positive big), AND ∆ACE

is ZO (Zero), THEN K is Big, K is Small, and ∝= 2. Similarly, at point B large

overshoot is to be avoided thereby implying small LFC control signal which means

small Kp and Ki and big Kd values. Thus, the rules for other points can be described

as per Table 7-5.

Table 7-5 Factors Affecting fuzzy PID Tuning

Action Rise time Overshoot Steady state Settling System stability

error Time

Increase Kp Fast Increase Decrease Minor Deteriorate

Changes

Increase Kd Slow Decrease Minor Changes Decrease Improve

Increase 1/Ti Fast Increase Large Decrease Increase Deteriorate

Fuzzy rules applied in this study for K , K and ∝ are given in Table 7-6 – 7-8. It

should be noted from original work that Ziegler-Nichols PID tuning rule was

compared before selecting values of ∝ which are selected as varying from 2-5 as

single membership function varying from small to big. In all IF-THEN statement

rules, the predecessor portion of the rules consists of two parts joined with “AND”

operators which are inferred by Mamdani product [31]. Considering IF-THEN rules,

the predecessor portion of rules may be written as: μ μ ACE]. μ [∆ACE] where

μ , is the membership function value of the input fuzzy set A for a given value of

ACE, and μ is membership function value of the input fuzzy set B for a given value

of ∆ACE.

330
Table 7-6 Fuzzy Tuning Rules for Kp’

∆ACE

NB NM NS ZO PS PM PB

NB B B B B B B B

ACE NM S B B B B B S

NS S S B B B S S

ZO S S S B S S S

PS S S B B B S S

PM S B B B B B S

PB B B B B B B B

Table 7-7 Fuzzy Tuning Rules for Kd’

∆ACE

NB NM NS ZO PS PM PB

NB S S S S S S S

ACE NM B B S S S B B

NS B B B S B S B

ZO B B B B B S B

PS B B B S B S B

PM B B S S S B B

PB S S S S S S S

Table 7-8 Fuzzy Tuning Rules for ∝

∆ACE

NB NM NS ZO PS PM PB

NB 2 2 2 2 2 2 2

ACE NM 3 3 2 2 2 3 3

NS 4 3 3 2 3 3 4

ZO 5 4 3 3 3 4 5

PS 4 3 3 2 3 3 4

PM 3 3 2 2 2 3 3

PB 2 2 2 2 2 2 2

331
Figure 7-85 FGSPID Rule Implication Process

Implication process of fuzzy rule is shown in Figure 7-85 where values of K and

K for each rule are determined from corresponding membership functions. Input

membership functions as depicted in Figure 7-85 are such selected that sum of

membership grades for a particular input is unity, i.e., ∑ 1. Fuzzy output of

inference engine is converted to crisp by centroid method of de-fuzzification.

Following output parameters are obtained after de-fuzzification:

K ∑ μ K 7.51

K ∑ μ K 7.52

∝ ∑ μ ∝ 7.53

Where K and K are respective grade values corresponding to the grade μ for the

ith rule. Surface view of obtained parameters gains is shown in Figure 7-86.

After de-fuzzified gain parameters, post processing gain scheduling is performed for

normalizing the K and K between the range of K and K through following

equations:

K K K K ∗K 7.54

332
K K K K ∗K 7.55

K 7.56

The minimum and maximum ranges for process parameter gains are normalized using

suitable normalization factors. The normalized gains are calculated as: K

0.32K , K = 0.6K , K 0.08K T , K = 0.15K T , where K is the

ultimate gain that can make process unstable under proportional feedback control and

T is the ultimate period under stability limit [32]. K is taken as 5.1278 while T is

taken as 2.72 in the current study.

Figure 7-86 FGSPID Gains Surface View

333
Comparison of Fuzzy Gain Scheduled Secondary Frequency Controller with

Various Secondary Controllers

Three area conventional LFC model is utilized with each area implementing hydro

turbine governor dynamics and frequency responsive generic VSWTG model

integrated to area 1. Three area controlled power system model discussed earlier is

utilized such that only area 1 has frequency responsive wind farm and conventional

generating unit of hydro plant represented by hydro turbine governor dynamics while

other areas have hydro plant only. Wind integrated system modified inertia

constant, H is calculated according to equation 7.44. System parameters are given

in the Appendix which is used in LFC simulations. Fuzzy gain scheduled controller

performance is evaluated by performing simulations using MATLAB Simulink

software under the following scenario.

Following load disturbance scenario is used in the current study:

1. 0.1 per unit load disturbance in area 1 only.

2. 0.1 per unit load disturbance in all three area.

3. 10% and 20% wind penetration in Area 1.

Reduced number of rules with five membership functions are applied for the

conventional fuzzy PI controller. Simulation results of fuzzy gain scheduled PID

controller and fuzzy gain scheduled I controller are compared with conventional

fuzzy LFC PI controller with fixed gain and conventional PID LFC controller as

shown in Figure 7-87 - 7-89.

334
Figure 7-87 LFC Controller Comparison with 0.1 Per Unit Load Disturbance in Area 1 only and 90%-

10% Hydro-Wind AGC Participation.

Figure 7-88 LFC Controller Comparison with 0.1 Per Unit Load Disturbance in all 3 Areas and 90%-

10% Hydro-Wind AGC Participation in Area 1.

Figure 7-89 LFC controller Comparison with 0.1 Per Unit Load Disturbance in all 3 Areas and 80%-20%

Hydro-Wind AGC Participation in Area 1. (Zoomed to 40 sec)

335
Gain-scheduled fuzzy controller presents better performance in comparison to other

two controllers when control areas are subjected to different load perturbations and

different wind penetration. These controllers are observed to produce better settling

time and reduced overshoots. In the case of different load disturbance, fuzzy gain

scheduled PID as well as fuzzy gain scheduled I controller shows better performance

than conventional PID and fuzzy controller in terms of better settling time and

reduced overshoots. Though some overshoots are observed in gain scheduled fuzzy

PID, it can be further reduced with optimized gain values which are not considered

in the current study.

Table 7-9 Secondary Frequency Controller Comparison with Performance Index for Area 1 Under 10%

Wind Penetration and 0.1 per Unit Load Disturbance in All Three Areas

Secondary-Frequency Controller ISE ITAE

FGSPID 0.0113 56.82

FGPI 0.015 68.40

PID 0.061 167.52

FGSI 0.021 81.65

Table 7-9 shows performance index values observed through simulation for area 1

with different load frequency controllers. Fuzzy gain scheduled produces the lowest

values among all four controllers for the performance index. Even though fuzzy

controllers show better performance than conventional PID, best results with lowest

ISE and ITAE are observed by using gain scheduling in the fuzzy controller.

336
Figure 7-90 Three Area Frequency Response Comparison with 0.1 Per Unit Load Disturbance and 90%-

10% Hydro-Wind AGC Participation in Area 1. Fuzzy Gain Scheduled LFC Controller is used.

Figure 7-91 Three Area Frequency Response Comparison with 0.1 Per Unit Load Disturbance and 80%-

20% Hydro-Wind AGC Participation in Area 1. Fuzzy Gain Scheduled LFC Controller is used.

Area frequency responses are compared with wind penetration varying from 10-20%

in area 1. It can be seen from Figure 7-90 – 7-91 that wind-integrated area has more

initial frequency deviation with increasing wind penetration. But with frequency

responsive wind plants, frequency settling time is lower in area 1 (< 40 seconds) in

comparison to other two areas (≅60 seconds). Area 1 frequency settles at around 40

sec. in comparison to around 60 seconds for other two area.

337
Case C. Comparison of Fuzzy Gain Scheduled based LFC Controller With and

Without Wind Participation in Secondary Control

Figure 7-92 Area 1 Frequency Response Comparison with 0.1 Per Unit Load Disturbance and Different

Wind Integration Scenario.

Figure 7-92 shows area 1 frequency response with different integrated wind plant

configurations. Highest deviation and longest settling time are observed when wind

plant is just feeding power to control area and smallest deviation with lowest settling

time is observed for control area without wind plant. Grid code responsive wind plant

with automatic generation control (AGC) participation shows the best performance

with frequency response comparable to control area without wind plant with identical

settling time (≅35 seconds). Integration of simplified droop based VSWPP [25]

produces highest frequency deviations and longer settling time ((≅60 seconds). Thus,

the superior performance of grid code frequency responsive WPP with gain scheduled

fuzzy LFC controller is easily acceptable from Figure 7-91.

7.8. Analysis-IV

LVRT Capability of Grid Code Frequency Responsive VSWTG and its Effect

on Frequency Response

338
Besides having maximum power extraction capability with variable speed operation,

type 3 and type 4 variable speed wind turbine generators (VSWTG) have four-

quadrant active and reactive power capability. However, these VSWTGs exhibit a

very sensitive response to electrical network disturbances, predominantly to voltage

sags. Voltage loss leads to magnetization loss and consequent high current flow

through stator and rotor circuits of induction generator based VSWTG. These high

inrush currents can damage power electronics based converters necessitating

disconnection or current bypassing to ride through disturbance and avoid damages.

Voltage disturbance due to grid faults leads to transient overcurrents in rotor circuit

and temporary power control loss in DFIG based VSWTG. Protecting methods are

imperative during transient fault periods to save the system from damage due to

overcurrent or dc overvoltage. Type of protective methods, application duration, and

efficiency, are widely researched topics for DFIG FRT response. [33-39]. In normal

operation, active power support is a priority. Low voltage ride through capability

(LVRT) threshold can vary from 0.8-0.9 p.u. contingent on the wind turbine design.

Normal wind turbine operation is suspended whenever the turbine voltage drops

below this threshold, and reactive current is injected into the grid. Active power

control is a lower priority during this disturbance period, and reference active power

is set to zero [40]. Voluminous literature exists for individual synthetic inertial

support and centralized controlled low-frequency response schemes [41-46] from

power electronics based wind farms during generation-demand mismatch. However,

their proficiency throughout the grid faults is ambiguous owing to the ride-through

capability of power electronics based converter topology during severe disturbances.

While frequency response competencies are typically evaluated as a consequence of

generator outage events in a network, LVRT capability of frequency sensitive type 3

339
VSWTG is evaluated in this section with preliminary findings showing better fault

ride through capability without incorporation of any additional hardware like crow

circuit.

FRT Grid Code Requirement

Figure 7-93 Fault Ride-through Grid Codes [48]

FRT requirements defined in various countries grid codes for connection and

operation of transmission systems is shown in Figure 7-92. The joint consensus of

these grid codes demands stable operation from all new generators under a range of

defined credible grid faults and time duration. As represented in Figure 7-93, voltage

drop level varies between 20%-0% of nominal voltage value while minimum fault

duration during this voltage drop varies from 100 msec (in Denmark) to 625 msec (in

Ireland, USA and Canada. The most stringent requirement for FRT is in Germany

which seeks 0% volts for 150 msec (Germany) while Ireland’s code is challenging

regarding fault duration of 100ms for 15% volt. Except for Australia, technical

standards defined in almost all these countries grid code are either at a fixed level or

a level determined by the transmission system operator. The Australian NEM

Standard for FRT [47] is classified as the automatic standard which requires voltage

ride through capability in the range of 80%-90% of nominal voltage for 10 seconds.

340
Most wind generators cannot meet this and therefore comply with a lesser minimum

standard of 0% volt for 430msec. Generating unit reactive power support to local

voltage is expected during fault period and fault recovery period. Reactive current

injection up to 100% during a fault is also desired in grid codes of countries like

Spain, Germany, and Great Britain.

The test system was investigated for different voltage disturbance range. Simulation

results for two of the most onerous cases are presented here: a 10% voltage enduring

for 150 msec & 450 msec and a 0% voltage enduring for 150 msec. Test cases were

chosen to reflect grid code requirements, as discussed previously. Any

comprehensive low voltage ride through technique must be capable of satisfying

following three points.

 The reduced voltage drop at the generator.

 Divert or cancel high rotor currents below the surge capability of the

converter.

 Produce appropriate power output during faults.

Simulation results for LVRT capability of FR-VSWTG show improvement over

conventional VSWTG in all the three aspects mentioned above.

Case A: 0 Voltage Disturbance

It was observed that frequency responsive VSWTG was unable to sustain 0V

disturbance for 450 msec. and exhibits un-stability in all parameters. On the other

hand, they can sustain 0V disturbance for 150 msec. without any instability.

Case B: 0.1 Voltage Disturbance

Figure 7-94 shows the voltage and current measured at different point of test system

when voltage disturbance of 0.1 V is induced at bus 120 for 150 msec. starting at

341
2 secs. As soon as voltage at point of common coupling drops to 0.1 V , active

power starts decreasing and it is at minimum level but above zero and sustain this

voltage disturbance thereby avoiding a shutdown. A slight variation (4-8% of

nominal dc voltage) of dc link voltage is observed at the start of disturbance and after

clearing of disturbance. Figure 7-95 shows rotor current where we can notice high

current up to 2 p.u. for very small duration around 5 msec. during and after voltage

disturbance while it is around 0.9 p.u. (reference rotor voltage is 0.9 p.u.) during the

rest of time of voltage disturbance.

Figure 7-94 Simulation Results for 0.1 Vnom for 150 msec.

342
Figure 7-95 Rotor Current during Voltage Disturbance

Figure 7-96 Simulation Results for 0.1 v for 450 msec.

343
When 0.1 V appears at PCC at bus 120, around 0.3 p.u. voltage drop is observed

at generator terminal. Figure 7-96 presents the simulation results for 0.1 V voltage

drop for 450 msec. Even though rotor currents are sustained within 0.9 p.u. during

the voltage disturbance duration, active power recovery time is more than the

previous case of 150 msec. Rotor currents increase to 1.8 p.u. for 17 msec. after

removal of voltage disturbance. Even though total currents are under 2 p.u. of

maximum converter limit, additional protection device would be beneficial for this

duration of disturbance. Swift active and reactive power restoration following fault

initiation and clearance is expected from any proposed FRT scheme to improve the

fault response of other locally connected equipment, ultimately contributing to

system frequency stability. Minimum active power during 0 voltage drop for 150

msec is 0.3-0.4 p.u. while it around 1.3 p.u. - 1.4 p.u. for 0.1 V drop for the same

duration. Active power changes from 1.5 p.u. to 0.5 p.u. during voltage disturbance

of 0.1 V for 450 msec. Respective test results are given in Table 7-10.

Table 7-10 Test Response

Voltage drop, Disturbance duration, Time to ramp up active Max. dc link voltage, V

p.u msec power, s

0.1 150 0.02 1240

0.1 450 0.35 1950

0 150 0.02 1270

0 450 Unstable unstable

Figure 7-97 shows the DFIG response during 0.1 V voltage disturbance when it

is not working in frequency controlled mode and rated 1 p.u. power is applied to pitch

compensator. Differences were observed in rotor current, power ramp rate, reactive

power change and rotor speed change during the voltage disturbance.

344
Figure 7-97 Simulation Results for 0.1 v for 150 msec for Normal DFIG Operation

Current at generator terminal is initially more than 2 p.u. and 1 p.u. during voltage

disturbance, while it is 0.9 p.u. with frequency responsive operation. Under similar

test conditions, lower active power is achieved both before and after the disturbance

(varying from 8 p.u. - 6.5 p.u.) in comparison to frequency responsive VSWTG

(delivering 9.5 p.u. before and after disturbance). Reactive power also takes more

time to settle to zero. It is also observed that rotor speed goes into super synchronous

speed mode i.e.; above rated 1.2 p.u. during the disturbance, while it moves into

subsynchronous speed mode when operating in frequency responsive mode. DC link

345
voltage remains the same in both cases. If reference power is set to zero during

voltage disturbance, normal DFIG operation fails to ride through, and imbalance rotor

current and high variation in dc link voltage are observed.

FRT Effect on System Frequency Response

Figure 7-98 FRT Enabled Active Power Control Block

To provide a realistic FRT capable VSWTG model and assess its effect on frequency

response, an FRT block is included in the generic model as shown in Figure 7-98.

This FRT block determines the reduction in active current based on terminal voltage.

During frequency response, active power control is in priority and current is limited

by maximum converter current. During FRT mode, reactive and active currents are

to be regulated to avoid converter overloading. To emulate reduction of active current

during voltage disturbance, P is reduced by multiplying by a factor P

[f V which is depicted in Figure 7-99.

To test the effect of this low voltage / FRT logic enabled VSWTG response on system

frequency response, a test voltage is provided to VSWTG model such that voltage

drops to 0.6 p.u. from 1 p.u. at 11 s and remains at this level till 11.2 s. FRT enabled

VSWTG, and normal VSWTG are integrated into a single control area dominated by

hydro power plant which was defined in chapter 6 earlier. A load disturbance of 0.01

p.u. is applied in the area at 5 s. Figure 7-100 showed the applied voltage and

346
generated current command for normally operated VSWTG while Figure 7-101

represents the same parameters along with PFRT for FRT enabled VSWTG. It can be

seen current and power rises with a decrease in voltage in the normal case while

representing realistic scenario during voltage transient; they decrease as per applied

PFRT factor in FRT enabled VSWTG.

Figure 7-99 PFRT Variation with Terminal Voltage

Figure 7-100 Electrical Parameters in Normal VSWTG during Voltage Disturbance

Figure 7-101 Electrical Parameters in FRT Enabled VSWTG during Voltage Disturbance

347
Figure 7-102 Primary Frequency Response Comparison During 0.01 p.u. load Disturbance

Figure 7-103 AGC frequency response Comparison during 0.01 p.u. Load Disturbance with AGC

Controller Gain as P=-37.47, I=0.0096

Figure 7-104 AGC Frequency Response Comparison during 0.01 p.u. Load Disturbance with AGC

Controller in as P= - 45.63, I = 0.0133

348
Figure 7-102 represents system primary frequency response when the voltage of 1

p.u. is applied to normally operated wind turbine while the voltage drops up to 0.6

p.u. is applied in FRT enabled VSWTG. A drop in frequency can be observed when

electrical power drops due to voltage disturbance. Though the frequency deviation is

small in this case, it might be significant during large amount wind power dropping

during voltage disturbance. Figure 7-103 - 7-104 shows the AGC based frequency

response, and it can be observed that frequency deviation in control area with FRT

enabled VSWTG can aggravate under different AGC gain. This significant drop in

frequency can trigger frequency deviation based relays and might further escalate the

problem. The inclusion of such FRT logic in generic modeling can be useful in

analyzing frequency response situations like South Australian blackout during 28

Sep. 2016, where the number of FRT enabled wind farms reduced their power output

during voltage disturbance and finally situation worsen to the level of complete

system blackout.

Conclusion

The imbalance between energy production and its consumption is best indicated by

frequency in a power system. Primary and secondary regulation comes into the act to

return the balance in the system. Higher wind penetration induces large variations in

system frequency and reduces regulation capability to large contingencies. In this

chapter, it is shown that frequency responsive grid code compatible wind power plant

participation in frequency control services, especially AGC can have a major role in

mitigation of short-term impacts due to variable generation. It is also shown that

application of modern robust, automatic and intelligent control techniques to the

AGC scheme will be more adaptive and appealing approach for the wind-based

349
power system to overcome impositions set up by technical standards of energy market

regulators. Following are the specific conclusions derived in this chapter:

Analysis-I: Synthetic Inertia Enabled Frequency Responsive Wind Plant

Participation Capability in System Frequency Regulation and Its Effects on

System Reserve Response.

Frequency response indicator results from this simulation study and tuning

requirements of system AGC controller suggest that wind participation in frequency

regulation is better in the case of singular services like FCR rather than providing its

services in both control regulation method like FCR and FRR. Besides system inertia

reduction, load change also directly impacts system frequency response. Synthetic

inertia boost power support from VSWTGs and with proper turbine power control

loop modeling, net power output can be limited as per the TSO and demand

requirement under favorable wind conditions. Frequency nadir point as obtained from

different simulations show that no relay tripping is possible with around 40 % wind

penetration in regulatory services.

Analysis-II: Frequency Responsive VSWTG Integration Effect on Multi-

Generational System AGC Control

It can be concluded from this study that a higher wind penetration level up to 40% is

a possibility in a strongly interconnected network and reduced load disturbance while

wind participation in AGC operation under supporting wind conditions can take this

penetration level up to 60%. System planners need to consider the proper system bias

value by inertia reduction and corresponding system equivalent droop to analyze

wind integration impact.

350
Analysis-III: FGSPID Based Intelligent AGC in the Presence of Grid Code

Frequency Responsive Wind Power Penetration

Extensive digital simulations with increased membership values and rule base have

verified the effectiveness of Fuzzy gain scheduled PID controller performance in

comparison to other LFC controllers when applied to a non-linear system containing

stochastic wind power. Gain-scheduled LFC controllers are found to provide

effective compensation for nonlinearities induced due to predictable wind power

variations in frequency control. Despite the promising results, this controller

performance can be improved with optimized gain values and the analysis extended

to analyze other power plants dynamic response characterization for more robust

controller design with improved wind plant technology integration to the system.

Analysis-IV: LVRT Capability of Grid Code Frequency Responsive VSWTG

and Its Effect on Frequency Response

Even though low priority is considered in active power control in conventional DFIG

operation, simulation results show improved LVRT response, when it is under

frequency responsive mode with reference power varying under grid frequency and

TSO limitations. Frequency responsive VSWTG can sustain 0 voltage drop for

150ms which is most stringent FRT requirement among all grid codes. A more detail

analysis must be conducted for concluding and finalizing change in active current

priority threshold in frequency responsive VSWTGs. A generic modeling for

frequency response assessment is also presented for FRT enabled VSWTG in a weak

grid.

351
7.9. References

1. D.P. Kothari and I.J. Nagrath (2003). Modern Power System Analysis, Tata

McGraw Hill, New Delhi.

2. Nanda J., Kaul B. (1978). Automatic generation control of an interconnected

power system, IEE Proceedings 125 (5) 385/ 390.

3. Elgerd O.I., Fosha C.E. (1970). Optimum megawatt frequency control of

multiarea electric energy systems, IEEE Transactions on PAS, PAS-89, 556/563.

4. Hiyama T. (1989). Applications of rule based stabilizing controller to electrical

power systems, IEE Proceedings C 136 (3), 175/ 181.

5. Dash P.K., Liew A.C., Mishra B.R. (1998). An adaptive PID stabilizer for power

systems using fuzzy logic, Electric Power Systems Research 44, 213/222.

6. C. C. Lee. (1990). Fuzzy logic in control systems: Fuzzy logic controller, Part I,

IEEE Trans. Syst., Man, Cybern, vol. SMC-20. pp. 404- 418.

7. M. Sugeno, ed. (1985). Industrial Applications of Fuzzy Control. Amsterdam,

The Netherlands: North-Holland.

8. Takagi T. and Sugeno M. (1985). Fuzzy identification of systems and its

applications to modeling and control, IEEE Trans. Syst., Man, Cybem., vol. 15

pp-116-132.

9. Zhao Z. Y., Tomizuka M. and Sagara S. (1992). A fuzzy tuner for fuzzy logic

controllers, in Proc. 1992 Amer. Control Conf., Chicago, IL, June 24-26, pp.

2268-2272.

10. Mamdani E, Assilian S. (1975). An experiment in linguistic synthesis with a

fuzzy logic controller. International Journal of Man-Machine Studies. 7(1):1-13.

352
11. Karatepe E. Syafaruddin, T. Hiyama (2009). Artificial neural network-polar

coordinated fuzzy controller based maximum power point tracking control under

partially shaded conditions, IET Renew Power Gener, 3, pp. 239–253

12. M. R. I. Sheikh, S. M. Muyeen, R. Takahashi, T. Murata and J. Tamura (2009).

Application of self-tuning FPIC to AGC for load frequency control in multi-area

power system, PowerTech, 2009 IEEE Bucharest, Bucharest, pp. 1-7.

13. M. A. S. Masoum and M. Sarvi (2005). A new fuzzy-based maximum power

point tracker for photovoltaic applications, Iranian Journal of Electrical &

Electronic Engineering, 1:28-35

14. IEEE Working Group on Prime Mover and Energy Supply Models for System

Dynamic Performance Studies, (1992). Hydraulic Turbine and Turbine Control

Models for Dynamic Studies, IEEE® Transactions on Power Systems, Vol. 7,

No. 1, pp. 167-179.

15. P. L. Dandeno, P. Kundur, and J. P. Bayne (1978). Hydraulic Unit Dynamic

Performance Under Normal and Islanding Conditions -- Analysis and

Validation, in IEEE Transactions on Power Apparatus and Systems, vol. PAS-

97, no. 6, pp. 2134-2143.

16. L. M. Hajagos, G. R. Berube (1999). Utility experience with gas turbine testing

and modeling, IEEE Symposium on frequency control requirements, New York.

17. M. Nagpal, A. Moshref, G. K. Morison and P. Kundur (2001). Experience with

Testing and Modelling of Gas Turbines, IEEE Power Engineering Society Entity

Annual Report No.7802-6672-7/01.

18. CIGRE Task Force 25 of Advisory Group 02 of Study Committee C4 (2003).

Modeling of Gas Turbines and Steam Turbines in Combined Cycle Power

Plants, CIGRE, 2003.

353
19. Stanley M. Shinners (1998). Modern Control System Theory and Design, 2nd

Edition, Wiley – Interscience.

20. European Network of Transmission System Operators for Electricity, (2012).

Operational Reserve Ad Hoc Team Report. Belgium: ENTSO-E. Website:

entsoe.eu.

21. N. R. Ullah, T. Thiringer, and D. Karlsson (2008). Temporary Primary

Frequency Control Support by Variable Speed Wind Turbines— Potential and

Applications, IEEE Transactions on Power Systems, vol. 23, no. 2, pp. 601-612.

22. M. L. Chan, R. D. Dunlop and F. Schweppe. (1972). Dynamic Equivalents for

Average System Frequency Behavior Following Major Distribances, IEEE

Transactions on Power Apparatus and Systems, vol. PAS-91, no. 4, pp. 1637-

1642.

23. H. Ahmadi, H. Ghasemi (2014). Security-Constrained Unit Commitment With

Linearized System Frequency Limit Constraints, Power Systems, IEEE

Transactions on, vol. 29, no. 4, pp. 1536–1545.

24. I. M. Cecilio, A. M. Ersdal, D. Fabozzi and N. F. Thornhill (2013). An open-

source educational toolbox for power system frequency control tuning and

optimization. Proceedings of the 4th IEEE PES Innovative Smart Grid

Technologies (ISGT) Europe Conference.

25. A. Aziz, G. M. Shafiullah, A. Stojcevski and A. Mto. (2014). Participation of

DFIG based wind energy system in load frequency control of interconnected

multi generation power system. Power Engineering Conference (AUPEC).

Australasian Universities, Perth, WA, 2014, pp. 1-6.

354
26. H. Chavez, R. Baldick and J. Matevosyan (2015). The joint adequacy of AGC

and primary frequency response in single balancing authority systems, IEEE

Trans. Sustain. Energy, vol. 6, no. 3, pp. 959-966.

27. Kundur, P., Balu, N. and Lauby, M. (1994). Power system stability and control.

New York: McGraw-Hill.

28. G. Andersson (2010). Dynamics, and control of electric power system. EEH –

Power system laboratory.

29. G. Quazza (1996). Noninteracting Controls of Interconnected Electric Power

Systems, IEEE Transactions on Power Apparatus and Systems, vol. PAS-85, no.

7, pp. 727-741.

30. Zhen-Yu Zhao, Masayoshi Tomizuka and Satoru Isaka (1993). Fuzzy gain

scheduling of a PID controllers, IEEE Trans. Syst., Man Cybern., 23 (5).

31. Mamdani, E. H. (1977). Application of fuzzy logic to approximate reasoning

using linguistic synthesis, IEEE Transactions on Computers 26(12): 1182–1191

32. Zhen-Yu Zhao, Masayoshi Tomizuka and Satoru Isaka (1993). Fuzzy gain

scheduling of a PID controllers, IEEE Trans. Syst., Man Cybern., 23 (5).

33. Z. Peng, and H. Yikang (2007). Control strategy of an active crowbar for DFIG

based wind turbine under grid voltage dips, International Conference on

Electrical Machines and Systems, pp. 259-264.

34. L. Peng, B. Francois, and Y. Li. (2009). Improved crowbar control strategy of

DFIG based wind turbines for grid fault ride-through, APEC 2009, pp. 1932-

1938, Washington, USA.

35. I. Erlich, H. Wrede, and C. Feltes (2007). Dynamic behavior of DFIG-based

wind turbines during grid faults, Power Conversion Conference – Nagoya 2007,

pp. 1195-1200, Nagoya, Japan.

355
36. J. Yang, J.E. Fletcher, and J. O'Reilly. (2009). A series dynamic resistor based

converter protection scheme for doubly-fed induction generator during various

fault conditions, IEEE PES GM 2009, Calgary, Canada..

37. J. Lopez, P. Sanchis, X. Roboam, and L. Marroyo (2007). Dynamic behavior of

the doubly-fed induction generator during three-phase voltage dips, IEEE Trans.

Energy Conv., vol. 22, no. 3, pp. 709-717.

38. A.H. Kasem, E.F. El-Saadany, and H.H. El-Tamaly (2008). An improved fault-

ride through strategy for doubly fed induction generator based wind turbines,”

IET RPG, vol. 2, no. 4, pp. 201-214.

39. H.N.D. Le, and S. Islam (2008). Substantial control strategies of DFIG wind

power system during grid transient faults, IEEE/PES T&D, Chicago, USA.

40. K. Okedu (2012). Stability enhancement of DFIG-based variable speed wind

turbine with crowbar by FACTS device as per grid requirement, International

Journal of Renewable Energy Research, vol. 2, pp. 431-439.

41. G. Lalor, A. Mullane, and M. O’Malley (2005). Frequency control and wind

turbine technologies, IEEE Trans. Power Syst., vol. 20, no. 4, pp. 19051913.

42. G. Ramtharan, J.B. Ekanayake, and N. Jenkins (2007). Frequency support from

doubly fed induction generator wind turbines, IET RPG, vol. 1, no. 1, pp. 3-9.

43. L. Holdsworth, J.B. Ekanayake, and N. Jenkins (2004). Power system frequency

response from fixed speed and doubly fed induction generator-based wind

turbines, Wind Energy, vol. 7, no. 1, pp. 21-35.

44. J. Morren, J. Pierik, and S.W.H. de Haan (2006). Inertial response of variable

speed wind turbines, Electric Power Systems Research, vol. 76, no. 11, pp. 980-

987.

356
45. M. Kayikci, and J.V. Milanovic (2009). Dynamic contribution of DFIG-based

wind plants to system frequency disturbances, IEEE Trans. Power Syst., vol. 24,

no. 2, pp. 859-867.

46. R.G. de Almeida, J.A.P. Lopes (2007). Participation of doubly fed induction

wind generators in system frequency regulation, IEEE Trans. Power Syst., vol.

22, no. 3, pp. 944-950.

47. Australian Energy Market Commission (AEMC) (2009). National Electricity

Rules Version 63. Available at aemc.gov.au

48. F. Iov, A. D. Hansen, P. Sørensen and N. A. Cutululis (2007). Mapping of grid

faults and grid codes, Tech. Rep. Risø-R-1617(EN), Risø Nat. Lab., Tech. Univ.

Denmark, Roskilde, Denmark.

357
CHAPTER 8

CONCLUSION AND RECOMMENDATIONS

Power systems around the world are undergoing the continuous shift from centrally

dispatched large scale synchronous generation towards scheduled and non-scheduled

intermittent distributed generation. This change in generation mix due to high

renewable energy penetration challenges the interconnected system designs built

under different network configuration, operational strategies and the regulatory

framework within which it operates. Though wind-based ancillary services like

frequency control ancillary services (FCAS) are still in research & development stage

in most of the countries, future WES participation is expected to contribute to

enhanced market efficiency, improved system reliability and macroeconomic

benefits to all stakeholders.

8.1. Conclusion

Design and operation of power system in presence of wind energy are one of the

major issues in wind power integration. Variable speed wind turbine generators

(VSWTG) having incommensurable dynamic characteristics than conventional

synchronous generators requires precise modeling as they no longer receive negative

load treatment. In view of increasing renewable penetration and compulsory grid

code frequency support, power electronics based VSWTG requires modified active

power control algorithm to support regulation services. This thesis mainly deals with

modelling issues encountered in wind integrated power system for frequency

regulation under the limitations of generic modelling with reduced parameters. This

thesis investigated the grid code compatible frequency response capability from

358
variable speed wind turbines in primary and secondary control. Important modelling

modifications needed for frequency response studies are presented specific to

generating units and system control area. Topics covered in this thesis are frequency

responsive VSWTG modelling especially type 4 VSWTG model, multi-generational

interconnected system AGC control, intelligent AGC control and fault ride through

behavior of frequency responsive wind plant.

8.1.1. Type 4 Generic VSWTG Model for Frequency Response Studies

With relatively new wind technology, especially full converter based type 4 WTG,

development of suitable models to analyze grid code requirements is a significant

subject matter. A simplified yet absolute representation of full converter based wind

turbine generator principles is requisite for various bulk studies. A simulation model

modified from the basic structure of WECC’s type-4 wind turbine generator has been

proposed here which can be utilized in wind penetration impact studies as well as

frequency regulation studies. The performance of this generic model is evaluated

through various simulation sets under different wind speeds and power set-points

combinations and compared with proprietary model and detailed model. Simulation

results show that detail aerodynamic model along with single mass drive train model

gives acceptable performance highlighting their adequacy and best combination in

generic models for frequency control studies. This result contrasts with the simplified

aerodynamic model as proposed by WECC generic model for dynamic studies of

longer time frames. The generator converter effect of designed model is also found

to be comparable with manufacturer specific generic model reducing dependence on

propriety information. Proposed model shows electrical power ramping to 1 p.u.

within 15-20 seconds which is nearly equal to ramping time of 15 seconds as achieved

by detailed WTG model.

359
8.1.2. Grid Code Sensitive, Frequency Processor Dependent Frequency

Controller

Transient low-frequency deviations are handled by inertial control while active power

frequency response controller is needed for high-frequency control. To emulate

synchronous generator like response; dynamic dead-band, speed droop and available

headroom is considered in variable speed wind turbine generator (VSWTG) modelling

for frequency-active power control. The individual wind turbine is made grid code

frequency compatible by including additional active power setpoint generator output

to the modified torque control loop of respective turbine. A modelling framework for

frequency dependent active power setpoint generation in variable speed wind turbines

is presented and its corresponding effects on system frequency regulation response is

investigated. Active power setpoint generator applies designed power limitation on

available wind power, rated turbine power and TSO commanded power and provides

setpoint to turbine. A grid frequency processor based on dynamic dead band and

moving averaged frequency processor is used to suppress noise frequency signals

from passing to active power setpoint generator. Incorporation of frequency processor

helps in avoiding unnecessary changes for frequency response from wind turbines.

Frequency controller is developed for grid codes very similar to British/NEM and Irish

grid. Promising results in terms of improved settling time and better electrical power

and rotor speed variations during frequency deviations were obtained for proposed

frequency responsive type 3 and type 4 VSWTG model in comparison to other

common droop-inertia based model. Proposed grid code compatible VSWTG type 3

model presents setting time reduced to 20-30 seconds from 40-50 s for other

frequency responsive models. Both VSWTG type-3 and type-4 models with

frequency controller-I can maintain electrical power around 1.1 p.u. during the

360
moments of initial frequency deviations while frequency controller 2 presents initial

power drop up to 0.3 p.u. before ramping and settling at around 1.1 p.u. Basic Type

3 and Type 4 VSWTG models calls for torque controller gains re-tuning for improved

participation in frequency response while this can be avoided with proposed

frequency response controllers.

8.1.3. Wind Participation in Classical and Intelligent AGC in Multi-

Generational System

Generic conventional generating unit models representing relevant dynamic features

are developed in MATLAB and numerically validated individually. Automatic

generation control is then investigated for interconnection of these generating

systems of widely different characteristics. Design aspects of AGC scheme for

minimization of system frequency of interconnected multi-generation power systems

with/without the presence of wind power is analyzed. A single area control model

incorporating all three-frequency regulation mode is presented. Control loop of

primary and secondary regulation with specific integration of frequency responsive

VSTG is presented. Performance of AGC for multigenerational power system is

examined, including how frequency bias setting, equivalent droop setting influences

AGC response and inadvertent tie-line interchange and corresponding effects of

inertia change. It is shown through simulation that both type 3 and type 4 VSWTG

can participate in frequency regulation services with suitable modelling

modifications. System frequency control shows better improvement (in terms of

frequency nadir and settling time) when grid code compatible type 3 VSWTG is

integrated to network instead of type 4 VSWTG integration into the network.

Simulation result show that frequency deviation is least affected by system bias value

while it is highly affected by droop variation in AGC. It is also observed that control

361
area having grid code frequency responsive wind integrated with base load plant

gives better AGC performance than the control area having wind integrated with peak

load plant. Simulation result shows that thermal plant gives fastest and best

combination for integration with wind plant followed by gas and then by hydro plant.

In Australian context, thermal or gas based units are more suitable for R6 FCAS in

wind integrated area while hydro plant is better suited at R60 services. Frequency

responsive wind plant participation in FCR and FRR services in a control area

presents improvement in power reserve activation response due to high ramping

capability. With higher participation factor, frequency responsive wind plant with

faster power response provides more power than thermal plant having low

participation factor in a control area AGC.

Modern smart grid with multigenerational integrated power system calls for

computational intelligence techniques for robust and optimum results in AGC

regulation. Fuzzy gain scheduled rule-based proportional integral and derivative

(FGSPID) AGC control system is analyzed for an improved solution for an overall

reduction in grid frequency deviations and settling time in general power systems

operating under restrictions and certain operating ranges. An improvement of settling

time from 60 s to 40 s is observed with FGSPID AGC. Gain-scheduled LFC

controllers are found to provide effective compensation for nonlinearities induced

due to predictable wind power variations in frequency control. Gain scheduling in

fuzzy controller provides lowest performance index values for ISE and ITAE in

comparison to other controllers.

362
8.1.4. Challenges and Possible Solution in Wind Integrated Network

Australia national electricity model is studied for real-time challenges associated with

the large-scale integration of wind plant. Even though NEM-wide challenges are not

identified as each NEM region has a different generation mix, network configuration,

and demand characteristics, which lead to different challenges or different timing;

future increased wind power penetration will bring some adverse operational

challenges in the NEM. 14-Gen NEM model is used to study wind integration effect

on frequency response performance under low load scenario and high load scenario

in area 5 which can be held representative of South Australia (SA). A linear

polynomial regression model for variation between wind penetrations, MW loss,

ROCOF, Frequency nadir is developed. The key finding shows that frequency

operating standard of NEM can easily be maintained for up to 40% of wind

penetration under all load scenario and ROCOF limited to less than 1 Hz/s. Frequency

operating standards management is difficult under wind penetration more than 20%

and high synchronous generator contingency due to increasing ROCOF and higher

frequency nadir. Simulation results show that integration of synchronous condensers

and frequency responsive wind plants can contribute to improvement in system

frequency response.

The results reported in this thesis can be summarized as follows:

 Frequency responsive type 4 model is developed for application in frequency

regulation studies. The accuracy of the developed models has been validated

through the performed case studies and comparison with detailed and a

proprietary model.

363
 It is possible to achieve frequency responsive improved power output from

VSWTGs through a modified torque speed loop and operating in spinning

mode.

 Assessment of grid code sensitive frequency responsive wind plants

integration under different penetration level is performed in a multi-

generational system with improved frequency response in an interconnected

system.

 Integration of wind plants into AGC system can be used in active power

imbalances where wind plant can ramp up and ramp down its power output as

per grid code requirement. Conventional generating units are still required for

system frequency regulation.

 A higher wind penetration level up to 40% is a possibility in a strongly

interconnected network and reduced load disturbance while wind

participation in AGC operation under supporting wind conditions can take

this penetration level up to 60%.

 Frequency Nadir shows improvement if grid code compatible wind plant

participates in either FCR or FRR services than participating in both services.

 Even though low priority is given to active power control in conventional

DFIG operation consider, simulation results show improved LVRT response

when it is under frequency responsive mode with reference power varying

under grid frequency and TSO limitations. Frequency responsive VSWTG

can sustain 0 voltage drop for 150ms which is most stringent FRT requirement

among all grid codes.

364
8.2. Recommendations

The objectives stated for the thesis are achieved within the scope and limitations of

generic modeling for frequency response studies. However, further works can be

done as follows:

 Frequency responsive VSWTG model has been simulated in MATLAB

software. The control strategy can be further implemented and tested in an

experimental setup to study the performance in practice.

 ROCOF values have been obtained under limited parameters of the

interconnected system. A detailed modeling including all system components

is required to quantify realistic ROCOF values.

 Aggregate models of generating units, both conventional and wind plants

have been considered in this study. This study can be extended to spatially

distributed wind farm model and more advanced grid code responsive

conventional generation units representation.

 Economic dispatch process could be implemented in AGC study to achieve

more realistic results.

 Step load disturbance has been considered in this study. Controllable loads

and demand response needs to be implemented in the interconnected system

to study fast frequency response in the presence of frequency responsive wind

plants.

 Even though technical features have been highlighted for wind integrated

power system model, the study is incomplete without economic implications.

Frequency bias criteria are shown to deviate from standard in the presence of

wind penetration. Future work should investigate system frequency bias

criteria for wind integrated power system.

365
APPENDIX

All data related to 14-Gen NEM model is taken from this reference:

M. Gibbard and D. Vowles, (2008). Simplified 14-Generator Model of the SE


Australian Power System. The University of Adelaide, South Australia.

Normal Steady State Analysis Data without Wind Integration

Light Load Case Heavy Load Case

Total Generation MW 15050 23030


Total Load MW 14810 22300
Inter-Area Flows
Area 4-Area 2 MW -200 500
Area 2-Area 1 MW 470 1134
Area 1-Area 3 MW 200 1000
Area 3-Area 5 MW 200 500

Frequency Grid Processor Parameters

Pre filter Constant : 0.5s, Trend Generator/Filter Constant: 8s,


Upper dead-band: 0.015
Lower Dead-band: 0.001, Post Filter Constant: 0.5s, Droop: 0.0315

Droop Model Parameters

Inertia Droop Model: Low Pass Filter Constant: 0.1s,


K inertia: 2*H wind, High Pass Filter Constant: 1s, K droop: 0.0315
Droop Model 1: Low Pass Filter Constant: 0.01s, Washout Filter Constant: 6s,
Droop: 0.0315
Droop Model 2: Low Pass Filter Constant: 0.2s, K droop: 0.0315

366
VSWTG Parameters (Differences are highlighted in bold)

Type 3 VSWTG (Modified Torque


Type 3 VSWTG (Basic Torque
Loop) with Frequency Response
Loop) with droop controllers
Controller

Aerodynamic Model: Aerodynamic Model:

-Same
,

.
.

c1 = 0.5176, c2 = 116, c3 = 0.4, c4 = 5,


c5 = 21 and c6 = 0.0068.

Pitch Controller: Kpp = 150, Kip = 25, Pitch Controller: Same

Pitch Angle Limitation: 00-270 Angle Limitation: Same

Pitch Compensation: Kpc = 3, Kic = 30, Pitch Compensation: Same Gains

Inputs: PAPC, Pinp. Inputs: Pref=1, Pinp

Torque Controller: Kptrq = 3, Kitrq = Torque Controller: Kptrq: -271.5,


0.6 Kitrq : -310.7

Low Pass filter (TC) Constant: 60s Low Pass filter (TC) Constant: 5s

Power Limitation: 0.04-1.1 p.u. Power Limitation: Same

Washout Filter Constant: 1.0 Washout filter not present.

Gen/Converter Model: Gen/Converter Model:

Maximum current limitation: 1.1 Maximum current limitation: 1.1

Low Pass Filter: 0.02 Low Pass Filter: 0.02

Type 4 basic VSWTG model has all the


 Type 4 VSWTG proprietary model has same parameters as that of modified
all the same parameters as that of Type loop based type 4 VSWTG model
3 VSWTG. Electrical Control: Ebst (pu): except Pitch compensator taking
0.2, Kdbr: 10 Pref=1 as input. Torque Controller:
Kptrq = 3, Kitrq = 0.6

367
LFC System parameters used in Single Area LFC simulation

H system T R
6 3.7814 0.0315, 0.05
L (Wind
K P (p.u.)
Penetration)
-3.88 1% 10%

System Parameters for LFC 3-Area model with hydro plant only used in
FGSPID

H D B
6 1 2.1
S P L
5000 MW 0.1 10%
T T T T T T
0.06 0.06 0.08
T 0.2 s T 1 s T 3.7814
R , 0.05 R 0.38 T 5 s

GAST Parameters

Governor Droop R 4%
Fuel system lag time constant 1 T1 1.5
Fuel system lag time constant 2 T2 0.1
Load limiter time constant T3 3
Load limit Lmax 1
Temperature control loop gain KT 1
Maximum value position Vmax 1
Minimum value position Vmin -0.02
Turbine damping Dt 0

368
DCHP Parameters

Parameters Description Variables


ω Reference frequency 1 p.u.
Pre-filter time constant 0.05 s
Governor
Dead-band Dead Band
R Governor droop 4%
maximum power level
FR max 1 p.u.
for frequency response
the minimum power level
FR min 0.04 p.u.
for frequency response
Rate
Power ramp rate 0.014 p.u./s
Limiter
CFM Base load function 1s
Environmental burner
CEV 0.15 s
capacity
Sequential
CSEV Environmental burner 0.25 s
capacity
Variable inlet guide vane
CVGV position compressor 0.60 s
Turbine capacity
Environmental burner
TEV 5s
capacity time constant
Sequential
TSEV Environmental burner 5s
capacity time constant
Un-damped natural
0.22 rad/s
frequency
Damping ratio of the
0.8 p.u.
compressor
Pmax Maximum power 1.1 p.u.
Pmin Minimum Power 0.04 p.u.
Lmax Maximum Load Setpoint 1 p.u.
Lmin Minimum Load Setpoint 0.05 p.u.

369
Non-Reheat Thermal Plant Parameters

K Governor Gain Constant 1.0 s


Governor T Governor Time Constant 0.08 s
R Droop 4%
K Turbine Gain Constant 1.0
Turbine
T Turbine Time Constant 0.4 s

Hydro Plant Parameters

T Filter time constant 0.05 s


T Servo time constant 0.2 s
R Permanent droop 0.05
Governor
R Temporary droop 0.38
Temporary droop
T 5s
time constant
Turbine T Water time constant 1.0

370

You might also like