Frequency Control on an Island Power System

Embed Size (px)

Citation preview

  • Frequency Control on an Island Power

    System

    with Evolving Plant Mix

    by

    Gillian R. Lalor

    A thesis presented to

    The National University of Ireland

    in fulfilment of the

    requirements of the degree of

    Philosophiae Doctor

    in the

    School of Electrical, Electronic and Mechanical Engineering

    University College Dublin

    September 2005

    Supervisor of Research: Professor M.J. OMalleyNominating Professor: Professor A.M. de Paor

  • Abstract

    Continual balancing of active power generated and consumed is vital for power system

    security and stability, and to maintain frequency within an acceptable tolerance around

    nominal system frequency. Due to the large size of individual generators with respect to

    total system size, the loss of a generator in a small island system can cause a large power

    imbalance, and consequently a significant frequency excursion. Low system inertia

    results in high rates of change of frequency when a power imbalance occurs. Therefore,

    system frequency control on an isolated power system is particularly challenging.

    As the generating mix on a power system evolves, moving away from traditional steam

    generating units, the behaviour of the power system in response to a power imbalance

    also changes. Both combined cycle gas turbine (CCGT) and wind turbine generators

    have distinctive effects on system frequency control. As each technology comprises an

    increasing proportion of generation on power systems worldwide, a clear understanding

    of the effects of CCGT and wind turbine generator characteristics on system frequency

    control is required in order to maintain secure and stable power systems.

    A dynamic model of the Ireland electricity system is developed, tuned and validated

    for the purpose of studying short-term frequency control on an island system. Each

    frequency responsive generating unit on the Ireland system is modelled using low order

    models tuned to extensive data from frequency events on the Ireland system. The

    system load is modelled using a single measurement based dynamic load model, which

    incorporates the frequency sensitivity and inertial contribution of the load during power

    imbalances. Frequency control through under-frequency load shedding is also incorpo-

    rated in the model. The system model was subsequently validated through comparison

    with frequency events not previously used for tuning. The resultant system model has

    the ability to predict the under-frequency behaviour of the Ireland power system for

    up to 20 seconds following a loss of generation with a very good level of accuracy.

    i

  • The active power generated by a base loaded CCGT is coupled to system frequency. A

    model suitable for studying the short-term dynamic response of a combined cycle gas

    turbine to a system frequency deviation is developed. The model is tuned and validated

    with event data from combined cycle gas turbines on the Ireland electricity system. This

    model is used in conjunction with the validated system model to study the impact of

    increasing levels of CCGT generation on short-term frequency control of a small island

    system during a loss of generation event. Results indicate that as the number and

    proportion of base loaded combined cycle gas turbines increases, frequency control

    may become more challenging. The magnitude of the system frequency excursion

    increases non-linearly as the proportion of base loaded CCGTs increases. Therefore, if

    the number of CCGTs increases, large frequency excursions will become more likely and

    transmission system operators may need to review their frequency control strategies to

    maintain current security standards and to avoid the shedding of customers.

    Increased system inertia is intrinsically linked to the addition of synchronous genera-

    tion to power systems. However, due to differing electromechanical characteristics, this

    inherent link is not present in wind turbine generators. Dynamic models of two differ-

    ent wind turbine technologies are integrated into the validated system model, which

    is modified to represent the predicted 2010 Ireland electricity system. The effect on

    system frequency during a loss of generation event is examined for varying wind pene-

    trations on the system. The results indicate that regardless of wind turbine technology,

    the displacement of conventional generation with wind will result in increased rates of

    change of system frequency. The magnitude of the frequency excursion following a loss

    of generation may also increase. Amendment of reserve policies or modification of wind

    turbine inertial response characteristics may be necessary to facilitate increased levels

    of wind generation, particularly for an isolated power system.

    In addition to the short-term dynamic effects of wind generation on frequency control,

    longer-term effects as a result of the wind generation characteristics of variability and

    unpredictability need to be taken into account in order to maintain adequate levels

    of system security in all time frames. However, while a clear understanding of the

    technical aspects of frequency control is vital to ensure system security, they comprise

    just one part of the whole frequency control issue. The technical aspects therefore need

    to be put in perspective by examining them as part of the broader picture, which also

    includes economic consequences. With rapidly increasing wind generation on many

    power systems, the effect on all aspects of frequency control is being examined in

    ii

  • detail, to assess and quantify the impact of wind integration. A review of a number

    of previous wind integration studies is carried out and a preliminary methodology

    is proposed for examining the effects of wind integration on all aspects of frequency

    control over a number of time-frames. Some illustrative results are given for a sample

    AC interconnected system.

    iii

  • Acknowledgements

    I would like to thank everybody whose help and support contributed to this thesis. In

    particular, there are some without whom this thesis would not have been possible:

    Professor Mark OMalley, whose guidance, help, expertise and encouragement through-

    out the project was invaluable. Always making time for discussion, his supervision

    throughout has been excellent and I am extremely grateful for everything over the last

    four years. Thank you.

    Dr. Damian Flynn and Julia Ritchie, of the Queens University of Belfast, who collab-

    orated on the development of the Ireland electricity system model and with whom I

    had many useful and informative discussions.

    Dr. Lawrence Jones of Areva T&D, who gave me the opportunity to to spend 3 months

    working with Areva T&D in Bellevue, Washington.

    Professor Chen Ching Liu of the University of Washington, Seattle, without whom the

    trip to Washington would not have been possible.

    Colleagues in ESB National Grid, for many useful discussions and interactions, in

    particular Jonathan OSullivan, Michael Power, Doireann Barry, Kate OConnor, Pat

    McGrath and John Kennedy.

    ESB Power Generation, in particular Michael OMahony, Alan Egan and Nicholas

    Tarrant, for information and advice during the development of the CCGT model.

    Tom Wilson of Viridian, for help and useful discussions during the development of the

    CCGT model.

    Dr. Alan Mullane, for his guidance and expertise in collaboration on the study into

    iv

  • frequency control and wind turbine technology. Also for all the advice and questions

    answered, about LaTex as well as wind, and proof reading this thesis.

    Ronan Doherty, with whom I collaborated on the study into frequency control in com-

    petitive market dispatch in addition to a number of different projects, for the many

    useful and informative discussions throughout.

    All occupants of Room 157 over the course of the last four years. In particular, Shane

    Rourke for his help, advice and the invaluable discussions since I started and Hugh

    Mullany for his advice and proof reading this thesis. Also Tim Hurley, Andy Keane,

    Eleanor Denny, Garth Bryans and Ciara OConner for the constant moral support, tea

    and coffee breaks, and a enjoyable working atmosphere.

    My family, Liz, Pamela, Richard, John, and in particular Mum and Dad. Thank you

    for the constant support and encouragement, not just over the last four years, but in

    everything I do.

    All my friends, whose friendship I value greatly.

    And finally James, for the endless encouragement, support, confidence in me and, not

    least, patience, which have been invaluable over the last number of years. Thank you

    for everything.

    v

  • Publications arising from this thesis

    Journal Papers:

    R. Doherty, G. Lalor and M. OMalley,Frequency Control in Competitive Electricity

    Market Dispatch, IEEE Transactions on Power Systems, August 2005, Vol. 20, No.

    3, pp. 1588-1596. (Appendix C)

    G. Lalor, J. Ritchie, D. Flynn, and M. OMalley, The Impact of Combined Cycle Gas

    Turbine Short Term Dynamics on Frequency Control, IEEE Transactions on Power

    Systems, August 2005, Vol. 20, No. 3, pp. 1456-1464. (Appendix B)

    G. Lalor, A. Mullane and M. OMalley, Frequency Control and Wind Turbine Tech-

    nologies, IEEE Transactions on Power System. In press, 2005. (Appendix D)

    Conference Papers:

    G. Lalor and M. OMalley, Frequency Control on an Island Power System with In-

    creasing Proportions of Combined Cycle Gas Turbines, presented at IEEE Powertech

    Conference, Bologna, June 2003. (Appendix E)

    G. Lalor, J. Ritchie, S. Rourke, D. Flynn, and M. OMalley, Dynamic Frequency Con-

    trol with Increasing Wind Generation, presented at IEEE Power Engineering Society

    General Meeting, Denver, Colorado, June 2004. (Appendix F)

    vi

  • Table of Contents

    1 Introduction 1

    1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.2 Frequency Control of an Island Power System . . . . . . . . . . . . . . 4

    1.2.1 Frequency Control . . . . . . . . . . . . . . . . . . . . . . . . . 4

    1.2.2 Island Power Systems . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.2.3 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    1.3 The Aims and the Scope of this Thesis . . . . . . . . . . . . . . . . . . 12

    2 System Model 14

    2.1 The Ireland Power System . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2.2 Modelling the Ireland Power System. . . . . . . . . . . . . . . . . . . . 16

    2.2.1 Assumptions of the Model . . . . . . . . . . . . . . . . . . . . . 17

    2.2.2 Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    2.2.3 Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    2.2.4 Connecting System . . . . . . . . . . . . . . . . . . . . . . . . . 32

    2.3 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    2.4 Simulation Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    vii

  • 2.5 Tuning the Ireland System Model . . . . . . . . . . . . . . . . . . . . . 35

    2.5.1 Generating Units . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    2.5.2 Load Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    2.5.3 Connecting System . . . . . . . . . . . . . . . . . . . . . . . . . 44

    2.6 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    2.6.1 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    2.6.2 Frequency Control in Competitive Electricity Market Dispatch . 46

    2.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

    3 Frequency Control with Combined Cycle Gas Turbines 53

    3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    3.2 CCGT Background and Characteristics . . . . . . . . . . . . . . . . . . 54

    3.2.1 Gas Turbine Component . . . . . . . . . . . . . . . . . . . . . . 55

    3.2.2 The Heat Recovery Steam Generator and Steam Turbine Com-ponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    3.3 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

    3.3.1 CCGT Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . 63

    3.4 CCGTs on the Ireland System . . . . . . . . . . . . . . . . . . . . . . . 65

    3.5 CCGT Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    3.5.1 CCGT Model Structure . . . . . . . . . . . . . . . . . . . . . . 66

    3.5.2 Model Tuning and Validation . . . . . . . . . . . . . . . . . . . 70

    3.6 The Impact of CCGT Dynamics on Frequency Control . . . . . . . . . 75

    3.7 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    3.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

    viii

  • 4 Frequency control and Wind Turbine Technology 81

    4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

    4.2 Wind Generation Technology . . . . . . . . . . . . . . . . . . . . . . . 82

    4.3 Wind Turbine Generator Modelling . . . . . . . . . . . . . . . . . . . . 86

    4.3.1 Fixed Speed Wind Turbine Model . . . . . . . . . . . . . . . . . 86

    4.3.2 DFIG Wind Turbine Model . . . . . . . . . . . . . . . . . . . . 89

    4.4 Wind Generation on the Ireland Electricity System . . . . . . . . . . . 92

    4.4.1 Scenarios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    4.4.2 Simulating Procedure . . . . . . . . . . . . . . . . . . . . . . . . 93

    4.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    4.5.1 Response of wind turbine technologies to system frequency devi-ations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    4.5.2 System frequency control with increasing wind penetration . . . 95

    4.5.3 Supplementary response from DFIG . . . . . . . . . . . . . . . . 100

    4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

    5 Supplementary Study: Wind Integration Studies and Frequency Con-trol 105

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

    5.2 Review of Wind Integration Studies . . . . . . . . . . . . . . . . . . . . 107

    5.3 Preliminary Wind Integration Frequency Control Study . . . . . . . . . 114

    5.3.1 e-terra simulator . . . . . . . . . . . . . . . . . . . . . . . . . . 115

    5.3.2 Wind Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

    5.3.3 Determination of Wind Variability Costs . . . . . . . . . . . . . 118

    5.3.4 Determination of Wind Unpredictability Costs . . . . . . . . . . 121

    ix

  • 5.4 Preliminary Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

    5.4.1 Available Wind Data . . . . . . . . . . . . . . . . . . . . . . . . 122

    5.4.2 Scope of the study . . . . . . . . . . . . . . . . . . . . . . . . . 122

    5.4.3 Sample Test System . . . . . . . . . . . . . . . . . . . . . . . . 123

    5.4.4 Scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

    5.4.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

    5.5 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 130

    6 Conclusions 134

    6.1 Synopsis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

    6.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

    6.3 Scope for future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

    References 140

    A Frequency Disturbance Event 152

    B The Impact of Combined Cycle Gas Turbine Short Term Dynamicson Frequency Control 155

    C Frequency Control in Competitive Electricity Market Dispatch 166

    D Frequency Control and Wind Turbine Technologies 176

    E Frequency Control on an Island Power System with Increasing Pro-portions of Combined Cycle Gas Turbines 186

    F Dynamic Frequency Control with Increasing Wind Generation 194

    x

  • List of Figures

    1.1 Operating reserve time-scales . . . . . . . . . . . . . . . . . . . . . . . 6

    1.2 Recorded system frequency on Ireland electricity system . . . . . . . . 9

    2.1 Generation mix on the Ireland electricity system for 1995, 2005 and 2010 15

    2.2 Steam unit model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    2.3 Open cycle gas turbine model . . . . . . . . . . . . . . . . . . . . . . . 24

    2.4 Linear hydroelectric-turbine model . . . . . . . . . . . . . . . . . . . . 28

    2.5 Ireland system model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    2.6 Inertial response control loop . . . . . . . . . . . . . . . . . . . . . . . 37

    2.7 Six frequency events on the Ireland system with corresponding poweroutput of a sample generator . . . . . . . . . . . . . . . . . . . . . . . . 38

    2.8 Comparison between actual and simulated frequency response of a steamunit to a low frequency event . . . . . . . . . . . . . . . . . . . . . . . 40

    2.9 Turlough Hill generating unit response for two low frequency events . . 42

    2.10 Actual and simulated system frequency for a 267 MW generation loss . 46

    2.11 Actual and simulated system frequency for a 277 MW generation loss . 47

    2.12 Actual and simulated system frequency for a 381 MW generation loss . 48

    2.13 Actual and simulated system frequency for a 201 MW generation loss . 49

    2.14 Simplified system frequency model . . . . . . . . . . . . . . . . . . . . 50

    xi

  • 2.15 Comparison of the generation response of the black box model with thevalidated system model . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

    3.1 Single-shaft CCGT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    3.2 Multi-shaft CCGT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    3.3 CCGT model structure . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

    3.4 CCGT ambient temperature dependency . . . . . . . . . . . . . . . . . 71

    3.5 CCGT ambient pressure dependency . . . . . . . . . . . . . . . . . . . 72

    3.6 Change in power output of a typical near base loaded CCGT in responseto a frequency event on the system . . . . . . . . . . . . . . . . . . . . 73

    3.7 Simulated power output of the GT component of a typical CCGT to afrequency drop of 0.5 Hz . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    3.8 Winter peak scenario with 422 MW trip . . . . . . . . . . . . . . . . . 77

    3.9 Summer night valley scenario with 400 MW trip . . . . . . . . . . . . . 78

    3.10 Summer day valley scenario with 400 MW trip . . . . . . . . . . . . . . 79

    3.11 Sensitivity of system frequency nadir to increasing proportions of CCGTs 80

    4.1 Typical Cp curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    4.2 Fixed speed wind turbine Generator . . . . . . . . . . . . . . . . . . . . 84

    4.3 Doubly fed induction generator . . . . . . . . . . . . . . . . . . . . . . 85

    4.4 DFIG model with FOC controller . . . . . . . . . . . . . . . . . . . . . 91

    4.5 Comparison of fixed speed WTG and DFIG WTG responses to the lowfrequency event . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    4.6 Effect of increasing wind penetration on maximum rate of change offrequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    4.7 Simulated system frequency following the trip of largest infeed duringthe SDV scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

    xii

  • 4.8 Frequency nadir and static reserve tripped following the loss of thelargest infeed for increasing wind penetration during the Summer DayValley scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    4.9 Frequency nadir and static reserve tripped following the loss of thelargest infeed for increasing wind penetration during the Summer NightValley scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

    4.10 Supplementary control loop for DFIG WTG controller. . . . . . . . . . 102

    4.11 Comparison of fixed speed WTG and DFIG WTG responses to the lowfrequency event, including supplementary control loop . . . . . . . . . . 103

    4.12 Simulated system frequency following the trip of largest infeed duringthe SDV scenario . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

    5.1 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

    5.2 Wind farm power output time series . . . . . . . . . . . . . . . . . . . 126

    5.3 System frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

    5.4 ACE: Control Area A . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

    5.5 ACE: Control Area B . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

    5.6 ACE: Control Area C . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

    5.7 Generator power output: Control Area A . . . . . . . . . . . . . . . . . 131

    5.8 Generator power output: Control Area B . . . . . . . . . . . . . . . . . 132

    5.9 Generator power output: Control Area C . . . . . . . . . . . . . . . . . 133

    A.1 Recorded system frequency on Ireland electricity system . . . . . . . . 152

    xiii

  • List of Tables

    2.1 Under-frequency setting for Turlough Hill operating modes . . . . . . . 28

    4.1 Comparison of inertial response from various generators . . . . . . . . . 95

    4.2 Maximum ROCOF following loss of largest infeed (422MW) for variousoperating scenarios, wind turbine penetrations and wind turbine tech-nology type. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    5.1 Test system generation capacity . . . . . . . . . . . . . . . . . . . . . . 123

    5.2 Test system set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

    5.3 Test case scenarios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

    xiv

  • Nomenclature

    a = Frequency sensitivity of GT exhaust gas flow calculation factor

    A = Area swept by wind-turbine rotor (m2)

    b = Constant, such that a+b=1

    B = Frequency bias setting (MW/0.1Hz)

    Cp = Performance coefficient

    CD = Boiler drum integral coefficient (s)

    CSH = Boiler superheater integral coefficient (s)

    f = System frequency (Hz)

    FA = Actual control area frequency (Hz)

    fgen = Under frequency relay setting for Turlough Hill gen mode (Hz)

    FHP = Fraction of power output from high pressure ST stage

    fint = Under frequency relay setting for interruptible customers (Hz)

    FIP = Fraction of power output from high pressure ST stage

    FLP = Fraction of power output from high pressure ST stage

    fmin = Under frequency relay setting for Turlough Hill min gen mode (Hz)

    fo = Nominal system frequency (Hz)

    fpump = Under frequency relay setting for Turlough Hill pump mode (Hz)

    FS = Scheduled control area frequency (Hz)

    fspin = Under frequency relay setting for Turlough Hill spin mode (Hz)

    fUFLS = Under frequency load shedding relay setting (Hz)

    G = Set of generators

    Gw = Gate position (per unit)

    i = Current (A)

    igv = Inlet guide vane angle ()

    Ij = Inertia of generating unit j (kgm2)

    INT = Logical operator for inertial control loop

    xv

  • J = Polar moment of inertia of wind turbine and rotor (kgm2)

    K = Friction drop coefficient of orifice between drum and superheater

    K1P = Proportional gain for d axis current controller

    K2P = Proportional gain for q axis current controller

    K1I = Integral gain for d axis current controller

    K2I = Integral gain for q axis current controller

    KBB1 = Black box model parameter 1

    KBB2 = Black box model parameter 2

    Ki = IGV controller constant

    kpf = Steady state frequency sensitivity of the load

    kpv = Active power and voltage load model parameter

    kqf = Reactive power and frequency load model parameter

    kqv = Reactive power and voltage load model parameter

    Kscl = Supplementary control loop constant

    KE = Kinetic energy (MWs)

    KEi = Kinetic energy of generator i (MWs)

    KEL = Kinetic energy of system load (MWs)

    KEo = Kinetic energy at nominal frequency fo (MWs)

    Ligv = Inlet guide vane position (per unit)

    Lm = Per phase mutual inductance (H)

    Lr = Per phase rotor inductance (H)

    Ls = Per phase stator inductance (H)

    L = L2

    m LrLs (H)

    m = Steam flow rate out of boiler drum (per unit)

    ms = Steam flow rate into steam turbine (per unit)

    mw = Steam flow rate into boiler drum (per unit)

    N = System speed (per unit)

    NG = Number of generators

    Nref = Reference system speed (per unit)

    NIA = Algebraic sum of the actual flows on all tie lines/interconnectors (MW)

    NIS = Algebraic sum of the scheduled flows on all tie lines/interconnectors (MW)

    p = Differential operator

    P = Active power (MW)

    Pa = Ambient pressure (mbar)

    Paero = Accelerating aerodynamic power (MW)

    xvi

  • PD = Boiler drum pressure (per unit)

    Pelec = Electrical power (MW)

    Pf = Number of machine poles

    PGEN = Active power generated (MW)

    Pk = Amount of generation lost (MW)

    Pload = Active power required by the load (MW)

    Pmax = Maximum rated generator power output (MW)

    Pmech = Mechanical power (MW)

    Pmin = Minimum rated generator power output (MW)

    Po = Steady state system demand (MW)

    Ppu = Active power (per unit)

    PT = Boiler throttle pressure (per unit)

    Q = Heat Energy (per unit)

    Qload = Reactive power required by the load (MVAR)

    R = Resistance ()

    Rd = Droop (%)

    Rr = Radius of rotor (m)

    Rp = Primary reserve available at 5 seconds (MW)

    SP = Generator operating set-point (per unit)

    T1 = Load time constant (T)

    Ta = Ambient temperature (C)

    Tcd = Compressor discharge time constant (s)

    TCH = Steam transport and conversion time constant (s)

    TCO = Steam turbine crossover and conversion time constant (s)

    Tem = Electromagnetic torque (N m)

    Temref = Reference electromagnetic torque (N m)

    Tgf = Gas fuel system time constant (s)

    Ti = IGV controller integration rate (s)

    Tigv = IGV actuator time constant (s)

    Tpf = Ratio of load inertia to system frequency

    Tpv = Active power and voltage load model parameter

    Tqf = Reactive power and frequency load model parameter

    Tpf = Reactive power and voltage load model parameter

    Tr = Gas turbine rated exhaust gas temperature (C)

    Tref = Reference torque (N m)

    xvii

  • TRH = Steam turbine reheater and conversion time constant (s)

    Ts = Droop governor time constant (s)

    Tsc = Supplementary control loop torque (N m)

    Tt = Temperature controller integration rate (s)

    Tv = Valve positioner time constant (s)

    Tw = Water time constant (s)

    Tx = Gas turbine exhaust gas temperature (C)

    Txc = Gas turbine corrected exhaust gas temperature (C)

    Txm = Gas turbine measured exhaust gas temperature (C)

    TDCR = Combustion reaction time delay (s)

    TDTE = Exhaust gas transport delay (s)

    u = wind speed (m/s)

    V ,v = Voltage (V)

    Wf = Gas turbine fuel flow (per unit)

    Wx = Gas turbine exhaust gas flow (per unit)

    = Blade pitch angle (rad)

    Pint,j = Inertial response of generating unit j (MW)

    = Tip-speed ratio

    = Flux linkage (Wb)

    o = Average slip of an average induction machine

    = Supplementary control loop time constant (s)

    i = Inertial control loop time constant (s)

    = Density of air (kg/m3)

    = system speed (rad/s)

    dq = dq reference frame angular velocity (rad/s)

    m = Rotor mechanical angular velocity (rad/s)

    s = Shaft speed (rad/s)

    t = Wind-turbine rotor speed (rads1)

    r = Rotor electrical angular velocity (rad/s)

    Subscripts

    d, q = Direct, Quadrature axis component

    I, P = Integral, Proportional

    r, s = Rotor, Stator

    xviii

  • Acronyms

    AC Alternating Current

    ACE Area Control Error

    AGC Automatic Generation Control

    BPA Bonneville Power Administration

    DC Direct Current

    CCGT Combined Cycle Gas Turbine

    CER Commission for Energy Regulation

    CHP Combined Heat and Power

    CLP China Light & Power Co.

    CO2 Carbon Dioxide

    DFIG Doubly Fed Induction Generator

    EMS Energy Management System

    ESB Electricity Supply Board

    ESBNG ESB National Grid

    FOC Field Orientated Controller

    GT Gas Turbine

    HEC Hong Kong Electric Co.

    HRSG Heat Recovery Steam Generator

    HVDC High Voltage Direct Current

    IEC Israel Electric Corporation

    IGV Inlet Guide Vane

    NIE Northern Ireland Electricity

    NOx Oxides of Nitrogen

    OCGT Open Cycle Gas Turbine

    POR Primary Operating Reserve

    ROCOF Rate Of Change Of Frequency

    SCADA Supervisory Control and Data Acquisition

    SCIG Squirrel Cage Induction Generator

    xix

  • SO System Operator

    SONI System Operator of Northern Ireland

    ST Steam Turbine

    TH Turlough Hill

    TNB Tenaga Nasional Berhad

    UFLS Under Frequency Load Shedding

    WTG Wind Turbine Generator

    xx

  • Chapter 1

    Introduction

    1.1 Background

    The function of a power system is to provide customers with an electricity supply of

    acceptable reliability, where reliability signifies the ability to supply adequate electric

    service on a nearly continuous basis with few interruptions over an extended period of

    time (IEEE/CIGRE, 2004). Therefore, to design and operate a power system within

    adequate reliability margins such that overall costs are minimised is a key objective for

    all system operators.

    Power system security is an indication of the level of robustness of the power system

    at any instant in time to a disturbance (Fink and Carlsen, 1978). When operating

    in a secure state, a power system can withstand most severe disturbances without

    interruption to customer supply. However, if operating in a state with reduced security

    margins, a power system will be more susceptible to disturbance, resulting in a higher

    likelihood of customer supply disruption. To maintain adequate reliability it is desirable

    to maximise the time the power system is operating in a secure state, with frequency

    and voltage levels within acceptable standards. In order for a power system to be

    secure, the power system must be operating in a stable state. Stability indicates

    the ability of the system to return to an equilibrium operating state subsequent to a

    disturbance, and is dependent on both the type of disturbance and the initial power

    system operating conditions (IEEE/CIGRE, 2004). Although the electricity industry

    is undergoing regulatory and organisational changes, the basic concepts and rules for

    1

  • Chapter 1. Introduction 2

    reliable, secure and stable system operation remain unchanged.

    Ancillary services can be broadly defined as the range of technical services required

    by the system operators to maintain both secure and stable operation of the power

    system. These include operating reserves for frequency control, voltage control and also

    system restoration/black start capability. While the methods by which these services

    are procured may vary and evolve with regulatory structure, the necessity of ancillary

    services is unquestionable. This is highlighted by a number of recent contingencies

    worldwide resulting in severe lapses in the security of power systems including blackouts

    in the Eastern US and Canada, Italy and the UK (NERC, 2004; UCTE, 2003; NGC,

    2003).

    The control of system frequency is a vital aspect of secure and stable power system

    operation. A continuous balance between active power generated and active power

    consumed by the load and losses is required to maintain frequency constant at nominal

    system frequency. Any imbalance in active power will result in a frequency devia-

    tion. While precise instantaneous balancing of active power is not viable, frequency

    control ensures that the system frequency remains within acceptable frequency limits.

    Frequency control can be called upon for a variety of conditions ranging from a grad-

    ual change in load levels over time to a sudden loss of generation or step increase in

    demand.

    A range of power system characteristics including system size, individual generator and

    load frequency response characteristics and plant mix on the system influence frequency

    control. The size and speed of a frequency deviation depends on the magnitude of the

    power imbalance and the power system size. Power system inertia is the resistance

    of the individual rotating masses of the generator and load components synchronised

    to the system to a change in system speed. The greater the inertia of the system,

    the slower the rate of change of frequency in the event of a power imbalance of given

    magnitude. Large interconnected power systems have high system inertia, due to the

    large number of components synchronised to the system. In addition, the size of in-

    dividual components, such as generators, tends to be small in comparison with total

    system size. As such, large frequency excursions from nominal are uncommon, and the

    rate of change of frequency is relatively slow due to high inertia. Small isolated power

    systems, in contrast, have much lower system inertia. Combined with the fact that a

    single generator can comprise a sizeable proportion of total generation, large power im-

  • Chapter 1. Introduction 3

    balances relative to the system size are more frequent and frequency changes are faster.

    Adequate frequency control on such a system is vital to prevent the excursion of system

    frequency beyond limits where interruption to customer supply through load tripping

    starts to occur. Therefore, maintaining system frequency at nominal frequency for a

    small island power system with limited interconnection can be technically challenging.

    Plant mix is continually evolving, for all power systems, large and small. Knowledge

    of the impact that evolving plant mix will have on system frequency control is vital

    to maintain a secure and stable power system with adequate reliability standards.

    Traditionally large coal and oil fuelled thermal plant comprised the majority of the

    generation mix on many power systems. However, due to economic and environmental

    driving forces, increasing proportions of combined cycle gas turbines and open cycle

    gas turbines are now being used to meet increasing demand and to replace older coal

    and oil-fired plants as they are retired.

    Combined cycle gas turbines (CCGTs) offer higher efficiency, greater flexibility and

    lower emissions than many conventional thermal generators, in addition to progressively

    shorter installation times and reducing installation costs. As a result, CCGT generating

    units comprise an ever increasing proportion of generation capacity for many electricity

    systems. The efficiency of combined cycle gas turbines is maximised when operating at

    or near maximum or base load, and declines with decreased loading. The behaviour of

    CCGT generators in response to frequency excursions differs from that of a conventional

    steam turbine, and may have a detrimental effect on the system frequency response

    when the CCGT is run at, or near, base load. This effect will be progressively more

    apparent as CCGTs operating at or near base load comprise increasing proportions of

    the generation.

    In conjunction with the shift towards CCGT plant, many power systems worldwide are

    also experiencing a rapid increase in wind generation. This trend is driven by a variety

    of reasons including environmental concerns, targets for electricity production from

    renewable energy resources, the desire for increased fuel diversity, constant advances

    in technology and economic factors including declining costs. While the addition of

    conventional synchronous generators to a power system will result in an inherent in-

    crease in the system inertial response, this is not necessarily the case with wind turbine

    generators. Therefore, if rapidly increasing levels of wind generation begin to displace

    conventional synchronous generation, erosion of system inertial response may result.

  • Chapter 1. Introduction 4

    This effect will result in increasing rates of change of frequency during power imbal-

    ances, and the magnitude of frequency excursions may also rise. These effects will

    influence small isolated power systems, in particular, where system inertia levels are

    inherently low.

    1.2 Frequency Control of an Island Power System

    1.2.1 Frequency Control

    System frequency provides a instantaneous indication of system operating conditions,

    as any imbalance between active power generated and consumed manifests itself as a

    deviation from nominal system frequency. The magnitude of the frequency excursion

    and the rate of change of frequency are dependent on a number of factors, including

    the size of the power imbalance and the characteristics of the power system. While

    small variations in system frequency will not result in a reduction is system reliability

    or security, large frequency deviations can have a serious impact on power system

    components and power quality is degraded. Damage to generators and transformers

    can result from overheating due to increases in the volts/hertz ratio during times of

    low frequency. In addition, generator damage due to mechanical vibrations can occur if

    frequency deviations greater than 5% of nominal frequency occur (Kirby et al., 2002).

    As a result, most power system components are equipped with protective relays, which

    are triggered if system frequency reaches critical conditions. Therefore, control of

    system frequency is vital for the secure, reliable operation of the power system.

    The objective of frequency control is to maintain adequate balance between active

    power consumed and generated on a power system such that frequency remains within

    acceptable limits around nominal frequency. As the demand of a power system is con-

    stantly changing, frequency control is continuously called upon to fulfil this objective.

    To a large extent, the changing system load is predictable and generators are committed

    and dispatched based on the forecast load levels (Machowski et al., 1997). Therefore,

    under normal operating conditions, the balancing of energy is achieved by adjusting

    generator active power set-points. Signals to generators for such adjustments are ei-

    ther issued by the system operator or automatically generated and issued by automatic

    generation control (AGC).

  • Chapter 1. Introduction 5

    In the event of an unpredicted increase in system load or an unexpected loss of genera-

    tion or transmission line, an imbalance of active power will occur. Every power system

    has stored kinetic energy by virtue of the masses of the generator and load compo-

    nents rotating in synchronism, which is a function of both the system inertia and the

    system frequency. In response to a power imbalance, stored kinetic energy is released

    to redress the imbalance, resulting in an inherent reduction in system frequency. In

    the event of active power generated exceeding demand, kinetic energy is absorbed and

    an increase in system frequency results. However, while frequency control in the event

    of high frequency events is essential and many issues discussed here are relevant, low

    frequency events are the focus of this thesis.

    In order to limit the frequency excursion from nominal system frequency, and to main-

    tain a stable and secure system, action in addition to the inherent system inertial

    response is required, i.e. frequency control. Frequency control may be broadly cat-

    egorised into automatic and manual frequency control. The former responds auto-

    matically to either a deviation from nominal system frequency or a rate of change of

    frequency in excess of a predefined threshold. Sources of automatic frequency control

    are the natural reduction in system load with low frequency, the automatic increase in

    generator active power output activated by the speed droop governor, low frequency

    or rate of change of frequency triggered responses from pumped storage units and the

    automatic shedding of load. Manual frequency control encompasses all instructions

    issued by the system operator to generators (and load if applicable, i.e. in the event

    of load participation) for changes from the reference set point of the generator (or to

    current active power consumption in the case of load).

    Additional active power capacity available (i.e. when compared to steady state oper-

    ation prior to a frequency event) from generation units or through reduction in load

    for the purpose of frequency control is known as operating reserve (ESBNG, 2005b).

    Many different definitions for the categorisation of operating reserve exist. In this the-

    sis, reserve is categorised into primary, secondary and tertiary operating reserve, as

    defined by ESBNG (2005b), and illustrated in Fig. 1.1.

    Primary operating reserve (POR) is the additional active power available from genera-

    tors and through reduction of active power consumption of the load which is available

    between 5 and 15 seconds subsequent to an event on the system. Secondary reserve is

    defined to be the additional active power available and sustainable for the time period

  • Chapter 1. Introduction 6

    Figure 1.1: Operating reserve time-scales (SEI, 2004)

    from 15 to 90 seconds after the event. Tertiary reserve is the additional active power

    available from 90 seconds to 20 minutes subsequent to the event. Finally replacement

    reserve is the additional active power available from 20 minutes to 4 hours after the

    event.

    In the event of a power imbalance, POR automatically responds to arrest the falling

    frequency and initiate recovery towards nominal frequency through the reduction of

    the power imbalance. The predominant source of POR on the majority of systems is

    the automatic droop governor response of generators operating below maximum rated

    active power output to a deviation in speed. Other sources of POR from generators

    can include an increase in active power when under-frequency relays or rate of change

    of frequency (ROCOF) relays are triggered. One example is under-frequency relaying

    triggering a rapid increase in active power generation from a pumped storage generating

    unit.

    System load also contributes to POR. In addition to the natural load reduction due to

    low system frequency, system load can also provide static reserve in the form of either

    interruptible customers or under-frequency load shedding (UFLS). Static reserve is

    defined here as capacity available instantaneously when called upon, with negligible

    dynamics.

    Some customers (interruptible customers) are contracted to make their load available

    for short term interruptions. Specific blocks of load are configured to be tripped by

  • Chapter 1. Introduction 7

    under-frequency relays if frequency falls to a threshold level. UFLS, however, is the

    tripping of uncontracted load at distribution system level and is called upon to pre-

    vent system collapse only when other sources of POR fail to arrest falling frequency.

    Discrete blocks of load are tripped until generation and load are once again in balance

    (Machowski et al., 1997), and frequency decline is arrested.

    Once the system frequency has been arrested and stabilised by the POR, it is the task of

    secondary and tertiary reserve to restore the system frequency to nominal value. This is

    achieved through a combination of automatic droop governor response while frequency

    remains below nominal and through discrete instructions issued by the system operator

    to generators for changes from the reference set point of the generator until power

    is once again balanced, and frequency restored to nominal. Replacement reserve is

    employed to replace operating reserve, restoring the system to a secure operating state.

    Capacity to provide operating reserve is dispatched in conjunction with generation by

    the system operator to ensure availability of adequate operating reserves in the event

    of a power imbalance. For a secure system, adequate operating reserve is required

    so that the power system can withstand most severe frequency disturbances without

    interruption to customer supply. The majority of power system worldwide operate with

    an N-1 security criterion (Bialek, 2003). This criterion states that the power system

    should operate so instability or load shedding do not occur as a result of the most

    severe single contingency. From a frequency control perspective, this entails having

    sufficient reserve to withstand the loss of the large power infeeds to the system.

    1.2.2 Island Power Systems

    Worldwide, power systems have a considerable range of characteristics including size,

    both geographical and electrical, the extent of interconnection to other power systems

    and generation mix. Many formerly isolated power systems with varying characteristics

    have become part of larger synchronous power systems through the use of alternating

    current (AC) interconnection. AC interconnection between power systems yields multi-

    ple advantages, increased system inertia, trading of energy, sharing of spinning reserve

    provision (operating reserve available from online generators) and mutual support dur-

    ing contingencies to name just a few (Mak and Law, 1991). While direct current (DC)

    interconnection allows energy exchange, power systems linked by DC interconnection

  • Chapter 1. Introduction 8

    are not synchronous. Therefore, some advantages of AC interconnection such as in-

    creased system inertia and the sharing of spinning reserves do not inherently occur

    with DC interconnection. However, although frequency control is not inherent, DC

    interconnections may be designed to provide spinning reserve.

    In large interconnected power systems, the size of individual components such as gener-

    ators tends to be small in comparison with the magnitude of the entire system. Power

    imbalances due to the loss of a single component in such systems, when they occur, are

    therefore generally small with respect to the total system size. In addition, the rate at

    which the frequency changes tends to be low due to high system inertia. The construc-

    tion of sizeable, more economically viable generating units is possible with minimal

    risk to system security. In addition, the provision of operating reserve is shared over a

    great number of generators, and may be shared between different systems within the

    larger interconnected power system. Generally, geographical dispersion is a character-

    istic of large interconnected power systems, and can contribute to a reduced capacity

    requirement, as a result of load diversity. One example of the benefits of geographical

    dispersion is the staggered occurrence of peak demand when a power system spans

    different time zones.

    While large strongly interconnected systems comprise a large proportion of power sys-

    tems worldwide, there are nonetheless a sizeable number of small isolated or poorly

    interconnected systems, for example Israel, New Zealand, Crete, Cyprus and Ireland.

    Small power systems that are either isolated or with only DC interconnection have low

    system inertia. A power system with low system inertia is more sensitive to system

    disturbances, due to less stored energy available to redress energy imbalances and to

    slow the rate of change of frequency. In addition, for such power systems, system com-

    ponents such as generators tend to be large in comparison with the total system size.

    In particular at times of low load, a single generator can comprise a large proportion

    of the total system generation. Therefore, in the event of a loss of generation, there is

    a greater likelihood of a large frequency excursion as the power imbalance is large with

    respect to total system size. On the Ireland electricity system, for example, frequency

    deviations of 1% are not uncommon, while larger frequency excursions occur occasion-

    ally, as illustrated by the recorded system frequency in Fig. 1.2. (This frequency event

    is described in more detail in Appendix A.)

    As a consequence of both low system inertia and potential large power imbalances

  • Chapter 1. Introduction 9

    0 100 200 300 400 500 600 700 80048.4

    48.6

    48.8

    49

    49.2

    49.4

    49.6

    49.8

    50

    Time (s)

    Freq

    uenc

    y (H

    z)

    Figure 1.2: Recorded system frequency on Ireland electricity system

    in addition to the relatively small range of generators to provide operating reserve,

    frequency control on small, isolated or DC interconnected systems is particularly chal-

    lenging. Distinctive operating and control strategies are necessary to maintain the

    system within limits of reliability and security. The main dynamic operation problems

    in small power systems relate to frequency control, in particular the behaviour of the

    system in response to large disturbances (Kottick and Or, 1996). Frequency control

    on such systems can in fact cause technical problems an order of magnitude greater

    than those experienced on large interconnected systems (OSullivan et al., 1999). The

    importance of frequency control on island systems during a contingency is evident in

    the considerable volume of relevant literature.

    1.2.3 Literature Review

    The benefits of AC interconnection between power systems are highlighted in Mak and

    Law (1991), where the AC interconnection between The China Light & Power Co.

  • Chapter 1. Introduction 10

    (CLP) and Hong Kong Electric Co. (HEC) are examined. The evolution of CLP from

    an isolated system to one with AC interconnection to other systems was found to have

    beneficial effects on system performance during system contingencies and also resulted

    in a more economical system operation. However, AC interconnection is not always an

    option and therefore a clear understanding of the frequency control dynamics of island

    power system is necessary to ensure optimal system security and reliability.

    Two neural network models that predict the frequency nadir (minimum frequency)

    and calculate the amount of UFLS during a loss of generation on the Israel Electric

    Corporation (IEC) system are developed in Kottick and Or (1996). The IEC operates

    an island system and at the time of the study the installed capacity was approximately

    5050 MW, with a 550 MW unit as the largest infeed. As the transmission system is

    strongly connected, frequency throughout the system is uniform. Therefore, transmis-

    sion effects could be neglected and a single busbar model was employed. Both the

    magnitude of the frequency excursion and the extent of UFLS are indicators of the

    severity of the contingency, and comprise two components of the dynamic security as-

    sessment for the IEC system. The models developed were demonstrated to perform

    well in assessing the UFLS subsequent to a loss of generation. The potential effect on

    frequency regulation (which is the automatic power balancing on a second to second

    basis) of a 25 MWh capacity battery energy storage device on the IEC system was

    investigated in Kottick et al. (1993), where a single busbar model was again used to

    represent the power system. It was demonstrated that simulated frequency deviations

    resulting from sudden demand variations were reduced considerably through the ad-

    dition of the battery energy storage device, which was assumed capable of sustaining

    a power output of 30 MW for 15 minutes. Due to a fast response time, the battery

    energy storage device was found to be potentially useful for regulation and as rapid

    operating reserve on an island system, where the rapid response time is critical due to

    low system inertia.

    The application of UFLS to the isolated power system of Cyprus is considered in Con-

    cordia et al. (1995). At the time of the study, the Cyprus system (with a peak load of

    500 MW) had a largest infeed of 60 MW, which was 12% of peak load and comprised

    a significantly greater proportion of load at times of low demand. While the general

    principles of UFLS are independent of system size, the distinguishing characteristics

    of isolated system must be taken into account when devising the UFLS plan. A well

    devised UFLS schedule results in the system surviving situations that would have oth-

  • Chapter 1. Introduction 11

    erwise resulted in blackouts. In Concordia et al. (1995), a criteria deemed appropriate

    for UFLS on an isolated system was developed and applied to the Cyprus system. It

    was also found that the effectiveness of load shedding increases with increasing system

    load.

    The effects of increasing proportions of renewable generation resources on the system

    frequency control on the island of Crete have been the focus of several studies (Hatziar-

    gyriou et al., 2000, 2002; Papazoglou and Gigandidou, 2003). In particular, Crete has

    experienced a rapid growth in wind generation in recent years. These studies pre-

    dominantly focus on the system under non-contingency operating conditions over the

    economic dispatch and unit commitment time frames. The short-term dynamic effects

    of wind turbine generators on the system frequency during a frequency event are not,

    however, considered.

    Another example of an island electricity system is the Ireland power system, which con-

    sists of two synchronous power systems. Before interconnection between the Northern

    Ireland Electricity (NIE) system and the Electricity Supply Board (ESB) system of the

    Republic of Ireland, each system on the island of Ireland operated as a small isolated

    system. With peak loads of approximately 1650 MW and 3300 MW respectively before

    interconnection, each system had low system inertia and as a result emergency control,

    i.e. frequency control in the event of a contingency, was critical.

    The strategies of the NIE system for emergency control of frequency when operating

    as an isolated system are outlined in Fox and McCartney (1988). Several obstacles

    such as inaccurate unit response information and difficulties with the coordination of

    under-frequency relay settings with system dynamics are also discussed. Limited UFLS

    was tolerated as a likely necessity on the NIE system at the time of the study in the

    event of the loss of a major infeed. Further studies into the control and proper design

    of UFLS arrangements are carried out in Fox et al. (1989) and Thompson and Fox

    (1994). The use of rate of change of frequency as an activating signal for UFLS was

    used in Fox et al. (1989), and found to provide more accurate load shedding than

    the use of under frequency relays. System frequency was simulated using a single

    busbar model. This approach was expanded in Thompson and Fox (1994), where each

    UFLS relay uses system demand, spinning reserve, system inertia and the amount of

    low priority load available for shedding elsewhere in conjunction with the local rate of

    change of frequency to assess whether to operate. This approach resulted in a significant

  • Chapter 1. Introduction 12

    reduction in the amount of excessive UFLS when compared to the fixed rate of change

    of frequency scheme of Fox et al. (1989). The effect of flywheel energy injection on

    emergency control of frequency on the NIE system has also been studied (Hampton

    et al., 1991). Once again a single busbar system model to predict system frequency

    following a unscheduled generation outage was used, and the model was validated by

    comparison to actual power system measurements. It was found that the use of the

    flywheel energy storage and retrieval scheme can contribute to considerable savings

    through spinning reserve replacement if correctly designed and scheduled.

    An emergency reserve model of the ESB system was developed and implemented to

    study frequency control on an island power system in OSullivan and OMalley (1996),

    OSullivan (1996) and OSullivan et al. (1999). The single busbar model was tuned

    using actual frequency events on the ESB system to accurately account for the dy-

    namic system characteristics following a loss of generation. The provision of frequency

    control was shown to be a critical issue as electricity markets emerge for island systems

    (OSullivan et al., 1999). As a result, the above model was subsequently incorporated

    into a new methodology for the provision of reserve in a competitive market (OSullivan

    and OMalley, 1999).

    1.3 The Aims and the Scope of this Thesis

    The objective of this thesis is to examine frequency control on an island system with

    evolving plant mix. In particular, the influence of the characteristics of CCGTs and

    wind turbine generators on system frequency control will be examined, and the Ireland

    electricity system is used as an illustration.

    Simulation, using validated models, is a good first step in understanding frequency

    control in the context of evolving plant mix. A single busbar model of the Ireland

    electricity system is developed, tuned and validated in Chapter 2. This model is suitable

    for the study of frequency response behaviour of an island system for up to 20 seconds

    after a power imbalance occurs. This system model is subsequently employed to tune a

    black-box model, which is used as the basis for the derivation of a minimum frequency

    control constraint (Appendix C).

    A model suitable for studying the short-term dynamic response of a combined cycle gas

  • Chapter 1. Introduction 13

    turbine to a system frequency deviation is developed, tuned and validated in Chapter

    3. This model is then used in conjunction with the Ireland system model of Chapter 2

    to study the impact of increasing levels of CCGT generation on frequency control of a

    small island system (Appendix B).

    Models for two different wind turbine technologies are presented in Chapter 4. To

    examine the short-term dynamic response of an island power system to sudden power

    imbalances with increasing proportions of wind generation, these models are integrated

    into the Ireland system model of Chapter 2, which is modified to represent the proposed

    2010 system model (Appendix D).

    The impact of wind generation on both short-term and long-term frequency control

    are assessed in Chapter 5. A review of a number of wind integration studies is carried

    out. Consequently, a preliminary methodology for a wind integration frequency control

    study using the Areva T&D e-terra simulator is proposed, which is applicable to both

    island and interconnected power systems. This work was carried out during a three

    month industry placement with Areva T&D in Bellevue, Washington.

  • Chapter 2

    System Model

    2.1 The Ireland Power System

    The electricity system on the island of Ireland operates at 50 Hz, with a current peak

    load of approximately 6100 MW (ESBNG, 2004b; SONI, 2003b). The Ireland electricity

    system consists of two power systems: the NIE system, operated by System Operator

    for Northern Ireland (SONI) and the ESB system, operated by ESB National Grid

    (ESBNG). Prior to 1995, the ESB and NIE power systems operated in isolation, with

    the limited connection between the two systems generally out of service and, as a

    consequence, unreliable (OSullivan, 1996). In 1995, however, the two systems were

    reconnected, and now comprise a single synchronous system, connected to each other

    through a number of AC lines. The main connection between the NIE and ESB systems

    consists of two 275 kV circuits, each with a capacity of 600 MW and of length 50 km.

    There are also two additional 110 kV lines, with capacity of 120 MW, connecting

    the systems at two separate locations along the interface between Northern Ireland

    and the Republic of Ireland (ESBNG, 2004a). A single high voltage direct current

    (HVDC) interconnection is in operation between Northern Ireland and Scotland with

    a capacity of 500 MW (ESBNG, 2004b). However, this HVDC interconnection is not

    currently configured to provide frequency response in the short time frame. With no

    AC interconnection to other systems to increase the inertia of the system and share

    reserve provision requirements, the Ireland electricity system is essentially an isolated

    system.

  • Chapter 2. System Model 15

    The generating capacity of the Ireland electricity system consists of a combination of

    reheat and non reheat fossil fuelled steam turbine generators, open cycle gas turbines

    (OCGTs), combined cycle gas turbines (CCGTs), hydroelectric generators, a single

    pumped storage station and wind turbine generators. In addition, other resources such

    as biomass generators, combined heat and power (CHP) and other renewables also

    provide limited generating capacity. Generation mix is constantly evolving, with both

    CCGTs and wind turbine generators in particular comprising increasing proportions of

    generation on the system. A comparison of the generation mix on the Ireland system

    in 2005 with the ESB and NIE systems in 1995 is illustrated in Fig. 2.1.

    Figure 2.1: Generation mix on the Ireland electricity system for 1995, 2005 and 2010((SONI, 2003b; ESBNG, 2004b))

    The reduction in the proportion of steam units from 1995 to 2005, alongside the increase

    in proportions of both CCGT and wind generation is clearly illustrated. The predicted

    Ireland generation mix in 2010 is also included in Fig. 2.1, to illustrate the forecast

    changing proportions of generation on the system.

    The system operator (SO) of each system performs scheduling and dispatch indepen-

    dently, while incorporating contracted flows on the interconnections between the two

    systems. The provision of primary operating reserve, however, is shared between the

    two systems.

  • Chapter 2. System Model 16

    In accordance with the system grid codes (ESBNG, 2005b; SONI, 2003a) frequency

    regulation is provided by each generator on the system, by virtue of a droop governor,

    with a compulsory droop setting of 4%. Operating reserve is divided into several

    categories according to the timescale within which it is available in response to an event,

    as described in Chapter 1. On the Ireland system, the primary operating reserve (POR)

    requirement corresponds to 75% of the largest infeed onto the system. At present, the

    largest infeed is 422 MW (the 500 MW HVDC interconnection to Scotland is operated

    with a maximum limit of 400 MW for system security reasons i.e. to limit the size

    of the largest infeed), thus making the primary reserve requirement 317 MW. The

    availability of primary reserve to meet this requirement is divided such that the ESB

    and NIE systems provide 67% (211 MW) and 33% (105 MW) respectively. Sources of

    POR include spinning reserve from generating units online and static reserve, such as

    interruptible load. Static reserve consists of blocks of reserve that are available almost

    instantaneously when tripped by the system frequency falling below the predetermined

    frequency setting of each block. Interruptible load is a form of static reserve, whereby

    certain load on the power system has an agreement with the SO that some or all of

    the load may be tripped during certain hours when the frequency falls below 49.3 Hz.

    The proportion of the POR provided by spinning and static reserve sources varies with

    time of day. The contribution of the pumped storage station to POR depends on the

    operational mode in which it is running, as described later in Section 2.2.2.

    2.2 Modelling the Ireland Power System.

    The dynamic model of the Ireland system used in this thesis is based on two previous

    models (OSullivan, 1996; Fox et al., 1989), with considerable enhancements introduced.

    An emergency reserve model of the ESB electricity system circa 1995 was developed

    by OSullivan (1996). The objective of this single busbar model was to accurately

    predict the system frequency following a contingency, by simulating the primary reserve

    response of the system. Installed generators are represented using low order models,

    and include dynamics for prime mover, turbine and governor valve characteristics where

    appropriate. A low order model, derived based on consideration of resistive and motor

    loads, is used to represent the system load. A similar single busbar model of the isolated

    NIE system circa 1989 was applied in Fox et al. (1989) for the evaluation of emergency

    load shedding schemes.

  • Chapter 2. System Model 17

    The present Ireland electricity system has evolved and developed from the ESB and NIE

    systems represented in OSullivan (1996) and Fox et al. (1989) respectively. A sizeable

    growth in system load has occured and new generating plants have been introduced onto

    the system and older plants decommissioned and removed. In particular, proportions

    of both combined cycle gas turbines and wind generation have increased significantly,

    and are predicted to comprise increasingly large proportions of system generation mix

    in the future, as illustrated in Fig. 2.1.

    The dynamic model representative of the Ireland electricity system is developed based

    on OSullivan (1996), and also Fox et al. (1989), augmented with more detailed and

    additional models where necessary. Each generating unit on the Ireland system is

    individually modelled, with the exception of wind, small hydro and other generators

    subject to a de minimis level of 10 MW. The details of the individual unit models are

    given in Section 2.2.2. The load model is presented in Section 2.2.3, and an overview

    of the entire Ireland model, with details of the connecting system and is presented in

    Section 2.2.4. The development and tuning of the Ireland electricity system model

    was carried out in collaboration with Dr. Damian and Flynn and Julia Ritchie of the

    Queens University of Belfast.

    2.2.1 Assumptions of the Model

    A fundamental assumption made in the development of the system model is that fre-

    quency is uniform throughout the system. This assumption is made on the basis that

    the electricity system of Ireland is tightly meshed and electrically short, with the rel-

    ative impedances between nodes quite small. Therefore, during a major contingency

    involving the loss of significant generation, the system will remain in synchronism and

    the frequency deviation will be very similar at all points on the system. This is borne

    out by system studies carried out by the Transmission System Operators and by fre-

    quency measurements during major events with frequency deviations of up to 0.8 Hz

    during sudden generation deficits. The system is designed and operated so that in

    the event of the loss of the largest infeed there should be no consequential events (i.e.

    the protection does not trip out any other devices, for example lines). Historical data

    shows that the loss of a transmission line has never occurred during a loss of generation

    event. In particular, during a major loss of generation there are noticeable changes in

    power flow across the AC lines between the NIE and ESB systems, due to the shar-

  • Chapter 2. System Model 18

    ing of reserve (DETINI, 2003; ESBNG, 2005c). These rapidly changing power flows

    have never caused any additional tripping of lines. Therefore, for frequency control

    studies on the Ireland electricity system, a uniform system frequency is assumed and

    a single busbar model has traditionally been employed (OSullivan, 1996; OSullivan

    and OMalley, 1996; OSullivan et al., 1999) and is appropriate for this study. It is

    also assumed that changes in voltage have negligible effect on real power balance on

    the system. In the event of a loss of generation on the system, voltage deviations will

    occur around the site of the contingency. However, these effects are only local and will

    not manifest themselves globally (OSullivan, 1996).

    The objective of the development of the Ireland electricity system model is to examine

    the effect of evolving plant mix on frequency control during a frequency event. The

    timescale of interest is the time immediately prior to and the first 20 seconds of a

    frequency disturbance event. For the purpose of short-term frequency control, as of

    interest in the system model, dynamics outside the timescale of interest are neglected.

    In addition, the system is assumed to be in steady state at nominal frequency prior to

    any frequency event.

    On the Ireland electricity system, if the frequency falls below 49.7 Hz (99.4% of nom-

    inal), it is deemed to be a significant frequency disturbance event (ESBNG, 2005b).

    Below 47.5 Hz (95% of nominal), the system will lose synchronism as generating units

    and load trip off the system. The system model is designed for relatively small fre-

    quency deviations of less than 3% ( 1.5 Hz). Therefore, system frequency changes

    can be assumed to be small with respect to nominal system frequency.

    2.2.2 Generation

    The majority of electricity generating units consist of a prime mover to produce me-

    chanical energy and a generator to convert this mechanical energy into electrical energy

    suitable for supply to the power system (Machowski et al., 1997). A variety of energy

    resources may be used to impart energy to the prime mover, resulting in a number

    of different prime movers types. These can be broadly categorised as steam turbines,

    combustion turbines and turbines which are driven directly by the energy resource,

    such as hydroelectric turbines, wind turbines and tidal energy turbines.

  • Chapter 2. System Model 19

    The energy resources associated with steam units are generally fossil fuels and nu-

    clear fission. The energy from combustion of the fuel (or heat energy resulting from

    the nuclear fission) is used to produce steam in a boiler, which drives the steam tur-

    bine. Combustion turbines, alternatively, use the exhaust gases from the combustion

    reaction to drive the prime mover. Hydroelectric turbines can have many different con-

    figurations, but water is used to drive the prime mover in all cases. Similarly, energy

    from wind and tides can be extracted in appropriately designed turbines to produce

    mechanical energy to convert into electrical energy in the generator.

    Steam Units

    Fossil fuelled steam units have been in use for over 120 years and can have a number of

    different configurations. However, the basic principle remains the same: the combustion

    of the fuel in the furnace heats the water in the boiler to produce steam, which is used

    to drive the steam turbine. Thus, the working fluid for a steam unit is water. The

    rotational energy of the steam turbine is then converted to electrical energy in the

    synchronous generator.

    The boiler can be either drum or once through configuration. In a drum boiler, water

    enters the waterwalls from the reservoir of water in the drum. Heat energy from the

    fuel combustion process is transferred through conduction, convection and radiation to

    the water in the waterwalls. The resultant steam and water mixture then re-enters the

    drum. In the drum, steam separates from water and travels to the steam turbine, via

    the superheater, with steam flowrate dependent on the pressure differential between

    drum and turbine. The pressure in the drum is dependent on the fuel firing rate, and

    steam flowrate can therefore be controlled by adjusting the firing rate. Superheated

    steam temperature is controlled by means of spray water attemperation (Flynn, 2003).

    Drum boilers may use either natural or forced circulation, but natural circulation is

    the more common configuration. Although higher pressure, and thus efficiency, can be

    achieved through the use of pumps, it is generally only viable in very large plants.

    The once through boiler configuration contains no internal reservoir and the water is

    forced through a continuous pipe to the steam turbine by means of pumping. As a

    result, steam flowrate is determined by the boiler feed pump and steam temperature

    is controlled by adjusting the fuel firing rate. Once through boilers are designed to

  • Chapter 2. System Model 20

    operate at supercritical temperatures, yielding higher efficiency. However, although

    increased efficiency results in reduced operating costs, the increased installation costs

    of such boilers generally make drum boilers more economically viable (Flynn, 2003).

    Boiler dynamics are highly complex, and over long time frames, detailed models are

    required to accurately capture dynamic behaviour. Complex models derived from phys-

    ical principles have been developed (Chien et al., 1958; McDonald and Kwatny, 1970;

    Kwan and Anderson, 1970; Flynn and OMalley, 1999) to capture the dynamics of the

    boiler over different timescales. A low order nonlinear boiler model was developed in

    IEEE (1973b). In this model, pressure and steam flowrate are defined as functions of

    the energy input to the boiler and the turbine control valve area.

    DeMello (1991) examined and justified the use of simplified boiler models (IEEE,

    1973b) for power system modelling. Various simplified models, which captured the

    essential nonlinear characteristics of the boiler, were compared to a detailed model

    based on mass, volume and energy balance equations and found to be adequate for use

    in power system dynamic performance studies. The simplified model of IEEE (1973b)

    and DeMello (1991) was adopted for the emergency reserve model of OSullivan (1996),

    and is employed in the Ireland system model of this thesis to model the boiler compo-

    nent of the steam units on the Ireland system.

    The boiler model is illustrated in the schematic of the steam turbine model in Fig. 2.2,

    which also includes the steam turbine model and the governor model.

    There are time delays in a boiler associated with the fuel dynamics and the transfer of

    heat energy to the water in the waterwalls when generating steam. The time constant

    for the fuel dynamics varies from 20 to 40 seconds depending on the fuel type in use.

    The waterwall lag has a time constant of approximately 5 seconds. Due to the relatively

    long time constant for the fuel dynamics in comparison with the time scale of interest

    here, the heat energy, Q, can therefore be assumed constant. However, as the heat

    energy Q remains constant, the waterwall lag component, a first order lag, may be

    neglected. Therefore, steam generation mw is equal to the heat energy Q.

    Pressure in the drum, PD, is proportional to the integral of the difference between the

    steam generation mw (i.e. steam flowing into the drum) and the steam flowrate out of

    the drum, m. The throttle pressure at the entrance to the steam turbine valve, PT , is

    propotional to the integral of the difference between the steam flowrate from the drum,

  • Chapter 2. System Model 21

    Figure 2.2: Steam unit model (See text for details)

    m and the steam flowrate exiting the valve into the steam turbine, ms. The non-linear

    nature of the boiler process is due to the steam flowrate from the drum to the throttle,

    which is proportional to the square root of the pressure difference between the two.

    From the boiler, the steam enters the steam turbine, where the heat energy in the steam

    is converted to rotational energy to turn the turbine. When a gas passes through

    a nozzle, it expands converting heat energy into mechanical energy. Therefore, as

    the steam expands through the stages of the steam turbine blades, temperature and

    pressure drop as the energy is imparted to the rotating turbine.

    A model which can represent both reheat and non-reheat turbine systems is presented

    in IEEE (1973a). This model, also used in IEEE (1991) and OSullivan (1996), is

    employed to model the steam turbine component of the steam units in this thesis, as

    illustrated in Fig. 2.2.

    The steam flow entering from the boiler, ms, directly determines the power generated

    in the turbine, with appropriate time delays incorporated. In a reheat steam turbine,

    three time delays are required. The time delay for transport of steam from the boiler

    to the first turbine stage and also the conversion of steam to rotational energy in the

    first or high pressure turbine stage is represented using a first order lag with a time

  • Chapter 2. System Model 22

    constant TCH . The time delay due to the reheater and conversion of heat energy to

    rotational energy in the intermediate stage is captured by the first order lag with time

    constant TRH . Finally, the transport delay to the final or low pressure turbine stage as

    well as the conversion of heat to rotational energy in this stage is captured by the a first

    order lag with time constant TCO. The fractions of the total power output generated

    at each stage are represented by FHP , FIP and FLP . A non-reheat unit will result in

    all power generation coming from the high pressure stage and therefore FIP and FLP ,

    the fractions of the total power generated in intermediate and low pressure stages of

    the steam turbine, will be zero.

    All dispatchable generating units on the Ireland system are fitted with droop governors

    and required by regulations (ESBNG, 2005b; SONI, 2003a) to have a droop setting of

    4%. The governor model used for the steam turbine units is the simplified speed

    governor of Elgerd (1982), and is illustrated in Fig. 2.2. Using the difference between

    actual and nominal frequency as input, the signal to the control valve is determined,

    which is proportional to the inverse of the droop, Rd. A time delay with time constant

    Ts due to the action of the hydraulics associated with the turbine valves is incorporated,

    along with maximum and minimum limits (Pmax and Pmin) to maintain the valve area

    within actual limits.

    Open Cycle Gas Turbine Units

    In the open cycle gas turbine, the working fluid is air. Air at atmospheric conditions

    enters the compressor, where energy is imparted to the air from the spinning compressor

    stages. On exiting the compressor, the air enters the combustion chamber where fuel is

    added and combustion occurs, resulting in the conversion of chemical energy in the fuel

    into heat energy. The hot gases produced by the combustion process then enter the

    gas turbine under high pressure and are expanded as they pass through the different

    turbine stages, converting the heat energy of the gases into rotational energy of the

    gas turbine. The gas turbine, compressor and generator rotate on a single shaft. The

    mechanical energy produced by the turbine, less the mechanical energy required by the

    compressor, is the net mechanical energy and this is converted to electrical energy in

    the synchronous generator.

    Numerous models of gas turbines have been developed, with varying degrees of com-

  • Chapter 2. System Model 23

    plexity and detail (Rowen, 1983b; Hung, 1991; Cohen et al., 1996; Kunitomi et al.,

    2001; Pourbeik, 2002). The dynamic model developed by Rowen (1983b) incorporates

    the main dynamics of a gas turbine for a wide range of operating speeds (95 - 107%)

    and conditions, and is suitable for use in power system stability studies. As a result,

    this model is used as the basis for many gas turbine and combined cycle gas turbine

    modelling applications (Bagnasco et al., 1998; Kunitomi et al., 2001; Lalor et al., 2005).

    The model developed in Rowen (1983b) includes the relationships between pertinent

    gas turbine components and also the control systems for speed, temperature and ac-

    celeration in addition to maximum and minimum fuel limits. Further simplifications

    of the model are also outlined.

    At the time of development of the emergency reserve model (OSullivan, 1996), a num-

    ber of OCGTs existed on the ESB system. However, due to an absence of data for any

    gas turbines during frequency events, identification of the dynamic parameters for the

    gas turbine model of Rowen (1983b) was not possible (OSullivan, 1996). As a result,

    a simple ramp based model was implemented, using the approximate reserve charac-

    teristics supplied by the system operator to provide limited identification (OSullivan,

    1996).

    For the Ireland system model developed here, OCGT generating units are explicitly

    modelled, using a model adapted from that of Rowen (1983b). The OCGT model

    structure is illustrated in Fig. 2.3.

    As acceleration control generally only comes into play during start-up and shut-down,

    the acceleration controller is neglected in this model, due to the assumption that the

    model is initially generating in steady state. The output from the speed and temper-

    ature controllers feeds into a minimum selector, where the lower of the two signals

    determines the fuel flow. Under normal operating conditions, the fuel flow is under

    the control of the speed controller (the governor). The speed controller consists of a

    simple droop governor. The temperature control loop compares the measured temper-

    ature of the exhaust gases, Txm, to the rated exhaust temperature, Tr. If measured

    exhaust temperature exceeds rated exhaust temperature, the temperature controller

    signal falls below unity, and the temperature controller takes over the fuel flow control.

    The exhaust temperature Tx is calculated using the equation (2.1):

  • Chapter 2. System Model 24

    Figure 2.3: Open cycle gas turbine model (See text for details)

    Tx = Tr 700 (1Wf) + 550 (1N) (2.1)

    where Wf is the gas turbine fuel flow (per unit) and N is the system speed (per unit).

    Power output is the product of the torque and the system speed, where the torque

    produced by the gas turbine is determined by the equation (2.2):

    Torque = 1.3 (Wf 0.23) + 0.5 (1N

    Nref) (2.2)

    where Nref is the reference system speed (per unit).

    Combined Cycle Gas Turbine Units

    Combined cycle gas turbines fuse the technologies of both steam and gas turbine units,

    resulting in a unit capable higher efficiency than either of the two individual compo-

    nents. CCGTs are not included in the emergency reserve model of OSullivan (1996).

  • Chapter 2. System Model 25

    However, due to a significant increase in CCGT generation on the ESB system, these

    units comprise a large proportion of total generating capacity. The details of the

    CCGT model structure used for the system model in this thesis, together with the

    CCGT model tuning and validation methodology are given in Chapter 3, where the

    impact of CCGT short term dynamics on frequency control is studied.

    Hydroelectric Units

    Hydroelectric generation uses the potential energy contained in water moving from a

    higher to a lower height to turn a turbine. Thus the potential energy is converted into

    kinetic or rotational energy of the turbine, which in turn is connected to an electric

    generator where kinetic energy is converted to electrical energy. There are three prin-

    cipal forms of hydro-turbines available and choice of turbine will depend on both the

    potential energy or head of the water supply and also the volume of water available.

    The three types are the Francis turbine, the Kaplin turbine and the Pelton wheel (Kun-

    dur, 1994). Both Francis and Kaplin hydraulic turbines are reaction turbines, relying

    on the weight of the water column to react against blades of the rotating element of

    the turbine, the runner. The runner is fully immersed in water and is enclosed in a

    pressure casing. The pressure differences across the runner blades imposes lift forces,

    which cause the runner to rotate. Francis turbines are the most widely used turbine

    type and designed to operate under medium heads of water, while the Kaplan turbine

    generally operates under lower heads of water with higher flow rates. Pelton wheels are

    impulse type turbines where the runner is spinning in air and rotated by the impact of

    a water jet on the blades. These are generally used only with high heads of water.

    One distinctive property characteristic of hydroelectric generation is the initial drop in

    power output that occurs when the opening of the intake gate is increased. This non-

    minimum phase characteristic is caused by the inertia of the water and the momentary

    reduction in head that occurs when the gate initially opens further, before the water

    flow through the turbine increases again (Elgerd, 1982; OSullivan, 1996). Therefore

    adequate representation of this phenomenon is required for accurate dynamic models

    of hydroelectric generating units, in particular for frequency control.

    The two hydro-turbine types used for generation on the Ireland system are Francis and

    Kaplan hydraulic turbines (OSullivan, 1996; ESBNG, 2004). A number of different

  • Chapter 2. System Model 26

    models for hydraulic hydro-turbines and their speed controllers