Modeling and Control of VSC-HVDC Links Connected to ...322475/...2.2 HVDC transmission using...

232
Modeling and Control of VSC-HVDC Links Connected to Weak AC Systems Lidong Zhang ROYAL I NSTITUTE OF T ECHNOLOGY S CHOOL OF E LECTRICAL E NGINEERING E LECTRICAL M ACHINES AND P OWER E LECTRONICS Stockholm 2010

Transcript of Modeling and Control of VSC-HVDC Links Connected to ...322475/...2.2 HVDC transmission using...

  • Modeling and Control of VSC-HVDC

    Links Connected to Weak AC Systems

    Lidong Zhang

    ROYAL INSTITUTE OF TECHNOLOGY

    SCHOOL OF ELECTRICAL ENGINEERING

    ELECTRICAL MACHINES AND POWER ELECTRONICS

    Stockholm 2010

  • Submitted to the School of Electrical Engineering in partial fulfillment of the

    requirements for the degree of Doctor of Philosophy.

    Stockholm 2010

    TRITA–EE 2010:022

    ISSN 1653-5146

    ISRN KTH-EE–10/22–SE

    ISBN 978-91-7415-640-9

    This document was prepared using LATEX.

  • To Yibin, Karin, Vivianne

    and my parents

    and my sister Lixia

  • iv

  • Abstract

    For high-voltage direct-current (HVDC) transmission, the strength of the ac system is

    important for normal operation. An ac system can be considered as weak either because its

    impedance is high or its inertia is low. A typical high-impedance system is when an HVDC

    link is terminated at a weak point of a large ac system where the short-circuit capacity

    of the ac system is low. Low-inertia systems are considered to have limited number of

    rotating machines, or no machines at all. Examples of such applications can be found

    when an HVDC link is powering an island system, or if it is connected to a windfarm.

    One of the advantages of applying a voltage-source converter (VSC) based HVDC system

    is its potential to be connected to very weak ac systems where the conventional line-

    commutated converter (LCC) based HVDC system has difficulties.

    In this thesis, the modeling and control issues for VSC-HVDC links connected

    to weak ac systems are investigated. In order to fully utilize the potential of the VSC-

    HVDC system for weak-ac-system connections, a novel control method, i.e., power-

    synchronization control, is proposed. By using power-synchronization control, the VSC

    resembles the dynamic behavior of a synchronous machine. Several additional functions,

    such as high-pass current control, current limitation, etc. are proposed to deal with various

    practical issues during operation.

    For modeling of ac/dc systems, the Jacobian transfer matrix is proposed as a uni-

    fied modeling approach. With the ac Jacobian transfer matrix concept, a synchronous ac

    system is viewed upon as one multivariable feedback system. In the thesis, it is shown

    that the transmission zeros and poles of the Jacobian transfer matrix are closely related to

    several power-system stability phenomena. The similar modeling concept is extended to

    model a dc system with multiple VSCs. It is mathematically proven that the dc system is

    an inherently unstable process, which requires feedback controllers to be stabilized.

    For VSC-HVDC links using power-synchronization control, the short-circuit ratio

    (SCR) of the ac system is no longer a limiting factor, but rather the load angles. The right-

    half plane (RHP) transmission zero of the ac Jacobian transfer matrix moves closer to the

    origin with larger load angles, which imposes a fundamental limitation on the achievable

    bandwidth of the VSC. As an example, it is shown that a VSC-HVDC link using power-

    synchronization control enables a power transmission of 0.86 p.u. from a system with an

    SCR of 1.2 to a system with an SCR of 1.0. For low-inertia system connections, simulation

    studies show that power-synchronization control is flexible for various operation modes

    v

  • related to island operation and handles the mode shifts seamlessly.

    Keywords: Control, modeling, multivariable feedback control, HVDC, power systems,

    stability, subsynchronous torsional interaction, voltage-source converter, weak ac

    systems.

    vi

  • Acknowledgements

    First of all, my deepest gratitude goes to my supervisors, Prof. Hans-Peter Nee and

    Prof. Lennart Harnefors. It is an honor and a pleasure for me to have Prof. Hans-Peter

    Nee as my supervisor. His patience and support helped me to go through the hardest

    moments of the research work. It is also a privilege for me to be a student of Prof. Lennart

    Harnefors. I am grateful for his generosity to share with me his deep understanding on

    scientific work. Without his guidance, this project cannot reach the same level as it is

    today.

    This work has been carried out within Elektra Project 30630 and has been funded by

    Energimyndigheten, ELFORSK, ABB Power Systems, ABB Corporate Research, Ban-

    verket. The financial funding is greatly acknowledged.

    My acknowledgements also go to the members of the steering group: Gunnar As-

    plund (ABB Power Systems), Pablo Rey (ABB Power Systems), Hongbo Jiang (Banver-

    ket), Torbjörn Thiringer (Chalmers University of Technology). During the last two and

    half years, I had many inspiring discussions with the steering group members. Their fruit-

    ful comments and inputs have greatly improved the quality of the research. Especially, I

    would like to thank Gunnar Asplund, who was the chairman of the group before his retire-

    ment from ABB Power Systems. Gunnar Asplund initiated the project and gave valuable

    suggestions at the beginning of the project.

    I would like also to thank my supervisor, Prof. Math Bollen, during my Licentiate

    study at Chalmers. Prof. Math Bollen brought me into the scientific world. I received

    endless support from him during my study at Chalmers and after graduation.

    To my colleagues at ABB, I am grateful for all the supports I have received during

    this period. In particular, I would like to thank Ying-Jiang Häfner, Magnus Öhrström,

    Cuiqing Du, and Rolf Ottersten for interesting discussions as well as many helps with

    thesis writing. Ying-Jiang Häfner carefully reviewed the manuscript of the thesis and

    gave important suggestions. I would like to give a special thank to Pablo Rey, my group

    manager at ABB, for allowing me to be absent from the group for the Ph.D study.

    At KTH, I would like to thank all the colleagues in the Electrical Machines and

    Power Electronics department. In particular, I would like to thank Prof. Chandur Sadaran-

    gani for reviewing the manuscript of the thesis. I am also grateful to Hailian Xie for her

    help with thesis writing, to Peter Lönn for his computer support, to Eva Pettersson and

    Brigitt Högberg for their help with the administrative work.

    vii

  • Many thanks to my parents and my parents-in-law for their love and support. My

    mother-in-law, Prof. Renmu He, is a renowned professor in power systems in China. I

    received many helps from her in my professional life as well as my family life for the past

    years. Her valuable suggestions during her stay in Sweden shed light on my research and

    influenced the content of this thesis. I would like also to thank my sister and nephew for

    their love and encouragement for all the time.

    Last but not least, I would like to thank my beloved wife and daughters. Yibin,

    thank you so much for your endless love, support and understanding. Thank you, Karin

    and Vivianne, for the joys you have brought to my life.

    Lidong Zhang

    Stockholm, Sweden

    April 2010

    viii

  • Contents

    Abstract v

    Acknowledgements vii

    Contents ix

    1 Introduction 1

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

    1.2 Project objectives and outline of the thesis . . . . . . . . . . . . . . . . . 3

    1.3 Scientific contributions of the thesis . . . . . . . . . . . . . . . . . . . . 4

    1.4 List of publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

    2 High-Voltage Direct-Current Transmission 9

    2.1 DC versus AC transmission . . . . . . . . . . . . . . . . . . . . . . . . 9

    2.2 HVDC transmission using line-commutated current-source converters . . 11

    2.3 HVDC transmission using forced-commutated voltage-source converters . 14

    3 Control Methods for VSC-HVDC Systems 21

    3.1 Power-angle control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    3.2 Vector current control . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    3.3 Power-synchronization control . . . . . . . . . . . . . . . . . . . . . . . 30

    3.3.1 Power-synchronization mechanism in ac systems . . . . . . . . . 31

    3.3.2 Power-synchronization control of grid-connected VSCs . . . . . . 32

    3.3.3 Bumpless-transfer and anti-windup schemes . . . . . . . . . . . . 35

    3.3.4 Negative-sequence current control . . . . . . . . . . . . . . . . . 41

    3.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    4 Dynamic Modeling of AC/DC Systems 47

    4.1 Jacobian transfer matrix for ac-system modeling . . . . . . . . . . . . . . 47

    4.1.1 Power-system stability and dynamic modeling . . . . . . . . . . 47

    4.1.2 Feedback-control view of power systems . . . . . . . . . . . . . 50

    4.2 Grid-connected VSCs using power-synchronization control . . . . . . . . 52

    ix

  • Contents

    4.2.1 Impedance-source neglecting the ac capacitor at the filter bus . . . 52

    4.2.2 Impedance-source including the ac capacitor at the filter bus . . . 63

    4.2.3 AC-source feeding from a series-compensated ac line . . . . . . . 70

    4.3 Grid-connected VSCs using vector current control . . . . . . . . . . . . . 74

    4.4 Jacobian transfer matrix for dc-system modeling . . . . . . . . . . . . . . 81

    4.5 Summary of the properties of the Jacobian transfer matrix . . . . . . . . . 87

    4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

    5 Control of VSC-HVDC Links Connected to High-Impedance AC Systems 89

    5.1 General aspects of high-impedance ac systems . . . . . . . . . . . . . . . 89

    5.2 Comparison of power-synchronization control and vector current control . 91

    5.3 Multivariable feedback designs . . . . . . . . . . . . . . . . . . . . . . . 100

    5.3.1 Internal model control . . . . . . . . . . . . . . . . . . . . . . . 102

    5.3.2 H∞ control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1075.3.3 Performance and robustness comparison . . . . . . . . . . . . . . 111

    5.4 Direct-voltage control . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

    5.4.1 Controller design . . . . . . . . . . . . . . . . . . . . . . . . . . 117

    5.4.2 DC-capacitance requirement . . . . . . . . . . . . . . . . . . . . 119

    5.5 Interconnection of two very weak ac systems . . . . . . . . . . . . . . . 122

    5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

    6 Control of VSC-HVDC Links Connected to Low-Inertia AC Systems 131

    6.1 General aspects of low-inertia ac systems . . . . . . . . . . . . . . . . . 131

    6.2 Controller design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

    6.2.1 Frequency droop control . . . . . . . . . . . . . . . . . . . . . . 133

    6.2.2 Alternating-voltage droop control . . . . . . . . . . . . . . . . . 134

    6.3 Dynamic modeling and linear analysis of a typical island system . . . . . 135

    6.3.1 Jacobian transfer matrix . . . . . . . . . . . . . . . . . . . . . . 136

    6.3.2 Integrated linear model . . . . . . . . . . . . . . . . . . . . . . . 147

    6.3.3 Linear analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

    6.4 Simulation studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

    6.5 Jacobian transfer matrix for other input devices . . . . . . . . . . . . . . 155

    6.5.1 Synchronous generator . . . . . . . . . . . . . . . . . . . . . . . 155

    6.5.2 Induction motor . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

    6.6 Subsynchronous characteristics . . . . . . . . . . . . . . . . . . . . . . . 162

    6.6.1 Frequency-scanning method . . . . . . . . . . . . . . . . . . . . 164

    6.6.2 Large ac-system connection . . . . . . . . . . . . . . . . . . . . 165

    6.6.3 Island operation . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

    6.6.4 Summary of the subsynchronous characteristics . . . . . . . . . . 177

    6.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

    x

  • Contents

    7 Conclusions and Future Work 181

    7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    7.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

    References 185

    A Fundamentals of the Phasor and the Space-Vector Theory 197

    A.1 Fundamentals of the phasor theory . . . . . . . . . . . . . . . . . . . . . 197

    A.2 Fundamentals of the space-vector theory . . . . . . . . . . . . . . . . . . 198

    A.3 Implementation of αβ and dq transformations . . . . . . . . . . . . . . . 200

    A.3.1 abc-αβ transformation . . . . . . . . . . . . . . . . . . . . . . . 200

    A.3.2 αβ-dq transformation . . . . . . . . . . . . . . . . . . . . . . . . 201

    B Jacobian Transfer Matrix 203

    B.1 Derivation of the transfer functions in Table 4.1 . . . . . . . . . . . . . . 203

    B.1.1 Transfer function JPθ (s) . . . . . . . . . . . . . . . . . . . . . . 203

    B.1.2 Transfer function JQθ (s) . . . . . . . . . . . . . . . . . . . . . . 205

    B.1.3 Transfer function JUf θ (s) . . . . . . . . . . . . . . . . . . . . . 206

    B.1.4 Transfer function JPV (s) . . . . . . . . . . . . . . . . . . . . . . 207

    B.1.5 Transfer function JQV (s) . . . . . . . . . . . . . . . . . . . . . . 207

    B.1.6 Transfer function JUf V (s) . . . . . . . . . . . . . . . . . . . . . 208

    B.2 Proof of the instability of the dc Jacobian transfer matrix . . . . . . . . . 208

    C Technical Data of the Test System 211

    C.1 The VSC-HVDC link . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211

    C.2 The synchronous generator . . . . . . . . . . . . . . . . . . . . . . . . . 212

    C.3 The induction motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213

    D List of Symbols and Abbreviations 215

    xi

  • Contents

    xii

  • Chapter 1

    Introduction

    This chapter describes the background of the thesis. The aim and the outline, as well as

    the major scientific contributions of the thesis are presented. Finally, a list of publications

    is given.

    1.1 Background

    In 1954, the first commercial high-voltage direct-current (HVDC) link between mainland

    Sweden to Gotland island was commissioned. Since then, the accumulated installed power

    of HVDC transmission worldwide has increased steadily, and recently a dramatic increase

    in volume has been initiated. So far, most of the HVDC systems installed worldwide are

    line-commutated converter (LCC) systems using thyristor valves. However, with grad-

    ually reduced losses and costs, the recently developed voltage-source converter (VSC)

    technology has shown to be more advantageous in many aspects [1–4].

    The conventional line-commutated HVDC technology has an inherent weakness,

    i.e., the commutation of the converter valves is dependent on the stiffness of the alter-

    nating voltage. The converter cannot work properly if the connected ac system is weak.

    Substantial research has been performed in this field [5–9]. The most outstanding con-

    tribution on this subject is [5], which recommends to use short-circuit ratio (SCR) as a

    description of the strength of the ac system relative to the power rating of the HVDC

    link. Both [8] and [9] conclude that, for ac systems with an SCR lower than 1.5, syn-

    chronous condensers have to be installed to increase the short-circuit capacity of the ac

    system. However, synchronous condensers can substantially increase the investment and

    maintenance costs of an HVDC project.

    In contrast to the conventional LCC-HVDC system, the VSC-HVDC system is

    based on self-commutated pulse-width modulation (PWM) technology, i.e., a VSC can

    produce its own voltage waveform independent of the ac system. Thus, a VSC-HVDC

    system has the potential to be connected to very weak ac systems. However, with the

    traditional vector current control the potential of the VSC is not fully utilized [10–12],

    1

  • Chapter 1. Introduction

    e.g., Ref. [11] shows that the maximum power that a VSC-HVDC link using vector cur-

    rent control can transmit to the ac system with SCR = 1.0 is 0.4 p.u. Ref. [12] shows

    that the inner-current controller of vector current control may interact with low-frequency

    resonances that are typically present in weak ac systems. In addition, the phase-locked

    loop (PLL) dynamics of vector current control might also have a negative impact on the

    performance of VSC-HVDC links for weak ac-system connections [10, 11, 13]. The poor

    performance of vector current control for weak-ac-system connections has become an

    obstacle for VSC-HVDC transmission to be applied in more challenging ac-system con-

    ditions.

    The application of high power-electronic devices, such as HVDC systems and

    FACTS devices also imposes new challenges for power-system stability analysis and dy-

    namic modeling. For classical power-system stability analysis, the phasor theory is the

    major mathematical tool. With the phasor approach, the electromagnetic transients of

    the ac network are neglected. This is a practical solution for conventional power sys-

    tems where the electromagnetic transients have negligible effects on the stability issue of

    concern. However, for high power-electronic devices, such a simplification is not accept-

    able. The dynamic frequency range of high power-electronic devices is much higher than

    that of the conventional power-system components. In this frequency, the phasor theory

    cannot properly reflect the dynamic interaction between the ac system and the power-

    electronic devices on the one hand, and between different power-electronic devices on

    the other hand. For example, it has been shown by [14] that the conclusions drawn by

    the phasor-based quasi-static analysis might not always agree with the results obtained by

    time simulations with electromagnetic-transient programs.

    The space-vector theory is based on instantaneous values, and therefore it is able to

    represent the electromagnetic transients of the ac network [15]. Traditionally, the space-

    vector theory is mainly applied for analyzing electrical machines and control of power-

    electronic devices [15–17]. Several methodologies for dynamic modeling of three-phase

    systems based on the space-vector theory have been been proposed. In [18], the complex

    transfer functions are applied for analyzing three-phase ac machines. In [19], a three-

    phase linear current controller is analyzed in the frequency domain based on the space-

    vector approach. In [20], the space-vector theory is applied for modeling of three-phase

    dynamic systems using the transfer matrix concept. For subsynchronous torsional inter-

    action (SSTI) analysis, the ac network is normally required to be modeled by the space-

    vector approach to take into account the electromagnetic transients [21, 22]. In recent

    years, the space-vector theory has also been applied to study the dynamic interactions be-

    tween high power-electronic devices. In [12], the dynamic interaction between an LCC-

    HVDC link and a VSC-STATCOM in the frequency domain is analyzed based on the

    space-vector theory.

    While the space-vector theory has been applied successfully for analyzing high-

    frequency stability phenomena in power systems, the theoretical work to connect the

    2

  • 1.2. Project objectives and outline of the thesis

    high-frequency stability to the classical power-system stability defined by the phasor ap-

    proach is missing in the literature. Such a connection is, however, necessary for HVDC

    systems and FACTS applications. On the one hand, the dynamic frequency range of such

    devices is high. On the other hand, the ratings of those devices are often high enough

    to have a significant impact on most of the classical power-system stability phenomena,

    such as angle stability and voltage stability. In the foreseeable future, the number of such

    devices in power systems is expected to increase considerably. Thus, there is a need for a

    unified modeling approach to address both the high-frequency and low-frequency stability

    phenomena.

    1.2 Project objectives and outline of the thesis

    The objectives of the project are:

    1. Develop a new control method for VSC-HVDC links connected to weak ac sys-

    tems.

    2. Develop a unified approach for dynamic modeling of ac/dc systems.

    3. Investigate various modeling and control issues for VSC-HVDC links connected

    to high-impedance ac systems.

    4. Investigate various modeling and control issues for VSC-HVDC links connected

    to low-inertia ac systems.

    The project is conducted by both theoretical analysis and time simulations. The outline of

    the thesis is:

    Chapter 2 A short introduction of various technologies for HVDC transmission is

    given.

    Chapter 3 Two existing control methods for VSC-HVDC systems, i.e., power-angle

    control and vector current control are described. A novel control method, i.e., power-

    synchronization control, is proposed to solve the problem for VSC-HVDC links

    connected to weak ac systems.

    Chapter 4 A unified dynamic modeling approach, i.e., the Jacobian transfer matrix,

    is proposed for modeling of ac/dc systems. Grid-connected VSCs using power-

    synchronization control and vector current control are modeled by the proposed

    concept.

    Chapter 5 The control issues for VSC-HVDC links connected to high-impedance ac

    systems are investigated. The dynamic performance of a VSC-HVDC link using

    3

  • Chapter 1. Introduction

    power-synchronization control and vector current control are compared. Two mul-

    tivariable feedback-control designs, i.e., internal model control (IMC) and H∞ con-trol are investigated. A direct-voltage controller is proposed. A control structure for

    interconnection of two very weak ac systems is proposed.

    Chapter 6 Power-synchronization control is applied to VSC-HVDC links connected to

    low-inertia ac systems. A frequency droop controller and a voltage droop controller

    are proposed. A linear model of a typical island system is developed for tuning the

    control parameters of the VSC-HVDC link. The subsynchronous characteristics of

    a VSC-HVDC converter are analyzed for both the large ac-system connection and

    island operation.

    Chapter 7 Summarizes the thesis and provides suggestions for future work.

    1.3 Scientific contributions of the thesis

    The main contributions of the thesis are:

    • A novel control method for grid-connected VSCs, i.e., power-synchronization con-trol, is proposed. The VSC using power-synchronization control basically resem-

    bles the dynamic behavior of a synchronous machine. A group of additional con-

    trol functions, such as high-pass current control, current limitation function, anti-

    windup schemes, etc. are proposed to deal with various practical issues during op-

    eration.

    • A novel modeling concept, i.e., the Jacobian transfer matrix, is proposed as a uni-fied dynamic modeling technique for ac/dc systems. With the proposed concept, a

    synchronous power system is viewed upon as a multivariable feedback control sys-

    tem. The proposed concept is intended to be a unified framework for analyzing both

    the low-frequency and high-frequency stability phenomena in power systems.

    • The theoretical connections between the stability defined by the Jacobian trans-fer matrix concept and the classical power-system stability defined by the phasor

    approach are analyzed. It is discovered that the transmission zeros of the Jacobian

    transfer matrix have a close relationship with angle and voltage stability in power

    systems.

    • A similar modeling concept, i.e., the dc Jacobian matrix, is proposed for modelingof dc systems. By using a π-link dc model, it is mathematically proven that the

    dc system (constructed by VSCs) is an inherently unstable process, where the dc

    resistance gives a destabilizing effect.

    4

  • 1.3. Scientific contributions of the thesis

    • Grid-connected VSCs using power-synchronization control and vector current con-trol are modeled by the Jacobian transfer matrix concept. The transfer functions are

    validated with frequency-scanning results from PSCAD/EMTDC.

    • The Jacobian transfer matrix concept is also applied for modeling of two conven-tional power components, i.e., the synchronous generator and the induction motor.

    It is discovered that the transmission zeros of the Jacobian transfer matrix are useful

    for interpreting some classical concepts, such as the synchronizing torque for the

    synchronous generator and the pull-out slip for the induction motor, from a feed-

    back control point of view. The transfer functions of the Jacobian transfer matrices

    are also validated with frequency-scanning results from PSCAD/EMTDC.

    • The performance of power-synchronization control and vector current control arecompared for VSC-HVDC links connected to weak ac systems, where it is con-

    cluded that power-synchronization control is more suitable for weak-ac-system con-

    nections.

    • Two multivariable feedback-control design methods, i.e., IMC and H∞ control areinvestigated for VSC-HVDC links connected to high-impedance ac systems. The

    performance and robustness of various control designs are compared and discussed.

    • A two-degree-of-freedom direct-voltage controller for VSC-HVDC system is pro-posed where a prefilter is applied to remove the overshoot of the direct voltage. A

    notch filter is proposed to reduce the dc-resonance peak.

    • The requirement of dc capacitance for VSC-HVDC links connected to weak acsystems is derived.

    • A control structure for VSC-HVDC links interconnecting two very weak ac sys-tems is proposed. The linear model is validated with time simulations from PSCAD/-

    EMTDC for each major design step.

    • A frequency droop controller and an alternating-voltage droop controller are pro-posed for VSC-HVDC links connected to low-inertia systems.

    • A complete linear model is developed for a typical island system which includesa synchronous generator, an induction motor, a VSC-HVDC link and some RLC

    loads. The root-locus technique is applied to tune the control parameters of the

    VSC-HVDC link.

    • Simulation studies are performed to demonstrate the flexibility of power-synchroni-zation control for various operation modes related to island operation.

    • The subsynchronous characteristics of a VSC-HVDC converter using power-synch-ronization control are analyzed using the frequency-scanning method.

    5

  • Chapter 1. Introduction

    1.4 List of publications

    - The doctoral thesis has resulted in the following publications:

    I L. Zhang and H.-P. Nee, “Multivariable feedback design of VSC-HVDC connected

    to weak ac systems”, in PowerTech 2009, Bucharest, Romania, 2009.

    II L. Zhang, L. Harnefors and H.-P. Nee, “Power-synchronization control of grid-

    connected voltage-source converters”, IEEE Trans. Power Systems, vol. 25, no. 2,

    pp. 809-820, May 2010.

    III L. Zhang, L. Harnefors and H.-P. Nee, “Modeling and control of VSC-HVDC

    links connected to island systems” accepted for publication at IEEE Power and

    Energy Society General Meeting, 2010, Minneapolis, USA.

    IV L. Zhang, L. Harnefors and H.-P. Nee, “Interconnection of two very weak ac

    systems by VSC-HVDC links using power-synchronization control”, accepted for

    publication in IEEE Trans. Power Systems.

    V L. Zhang, H.-P. Nee and L. Harnefors, “Analysis of stability limitations of a

    VSC-HVDC link using power-synchronization control”, submitted to IEEE Trans.

    Power Systems.

    - The author has co-authored the following publication during the course of the Ph.D

    study:

    VI L. Harnefors, L. Zhang and M. Bongiorno, “Frequency-domain passivity-based

    current controller design”, IET Power Electron., vol. 1, no. 4, pp. 455-465, 2008.

    - During the course of the licentiate study, the author has authored and co-authored the

    following publications:

    VII L. Zhang and M. H. J. Bollen, “A method for characterizing unbalanced voltage

    dips with symmetrical components”, IEEE Power Engineering Letter, pp. 50-52,

    July 1998.

    VIII L. Zhang and M. H. J. Bollen, “A method for characterization of three-phase un-

    balanced dips from recorded voltage waveshapes”, in International Telecommuni-

    cation Energy Conference, Copenhagen, Danmark, 1999.

    IX L. Zhang and M. H. J. Bollen, “Characteristics of voltage dips in power systems”,

    IEEE Trans. Power Delivery, vol. 15, no. 2, pp. 827-832, April 2000.

    X L. Zhang, “Three-phase unbalance of voltage dips”, Licentiate thesis, Techni-

    cal Report no. 322L, ISBN 91-7197-855-0, Chalmers University of Technology,

    Göteborg, Sweden, 1999.

    6

  • 1.4. List of publications

    XI M. H. J. Bollen and L. Zhang, “Analysis of voltage tolerance of ac adjustable-speed

    drives for three-phase balanced and unbalanced sags”, IEEE Trans. Ind. Applicat.,

    vol. 36, no. 3, pp. 904-910, May/June 2000.

    XII M. H. J. Bollen, J. Svensson, and L. Zhang, “Testing of grid-connected power con-

    verters for the effects of short circuits in the grid”, in European Power Electronics

    Conference, Lausanne, Switzerland, 1999.

    - In addition, during the working time in ABB, the author has authored and co-authored

    the following publications that are relevant to the subjects of the doctoral and licentiate

    studies:

    XIII L. Zhang and L. Döfnäs, “A novel method to mitigate commutation failures in

    HVDC systems”, in International Conference on Power System Technology, Kun-

    ming, China, 2002.

    XIV L. Zhang, L. Harnefors and P. Rey, “Power system reliability and transfer capa-

    bility improvement by VSC-HVDC (HVDC Light)”, in Cigre Regional meeting,

    Tallin, Estonia, 2007.

    XV M. H. J. Bollen and L. Zhang, “Different methods for classification of three-phase

    unbalanced voltage dips due to faults”, Electric Power Systems Research, vol. 66,

    no. 1, pp. 59-69, July 2003.

    7

  • Chapter 1. Introduction

    8

  • Chapter 2

    High-Voltage Direct-Current

    Transmission

    This chapter presents general aspects of HVDC transmission. Two major HVDC tech-

    nologies, i.e., HVDC transmission using line-commutated current-source converters and

    HVDC transmission using forced-commutated voltage-source converters are described.

    2.1 DC versus AC transmission

    The history of electric power systems began with direct-current (dc) transmission. In

    1882, Thomas Edison built the first power system with dc transmission with a low voltage

    level. However, dc transmission was quickly replaced by three-phase alternating current

    (ac) transmission because of several advantages of the latter. The most prominent advan-

    tage of ac transmission is that power can be transformed to different voltage levels. By

    using transformers, long-distance power transmission becomes possible. In addition, cir-

    cuit breakers for alternating current can take advantage of the natural current zeros that

    occur twice per cycle, and ac motors are cheaper and more robust than dc motors.

    In spite of the principal use of ac transmission in power systems, the interests on dc

    transmission still remain [23]. In 1954, the first commercial HVDC link between main-

    land Sweden to Gotland island was commissioned. Since then, the accumulated installed

    power of HVDC transmission systems worldwide has increased steadily, and recently a

    dramatic increase in volume has been initiated. Given the extra costs and losses related

    to the converter stations, HVDC transmission is justified by some particular conditions

    where the dc technology is the most feasible or may be the only solution:

    • Power transmission via cables. Due to their physical structures, cables have muchhigher capacitance than overhead lines. The capacitive current in cables created by

    the alternating voltage makes ac power transmission over long distance impossible.

    Even for a moderate length (50km), the losses created by the capacitive current can

    9

  • Chapter 2. High-Voltage Direct-Current Transmission

    be so high that reactive compensating equipment has to be installed in the middle

    of the cable [24]. However, installation of reactive compensating equipment is ex-

    pensive and not always practical, e.g., with submarine cable transmission under sea.

    On the other hand, if the power is transmitted by direct currents, there will be no

    losses related to capacitive currents. Therefore, for long-distance submarine cable

    transmission, HVDC transmission is the only feasible technical solution.

    • Bulk-power transmission over long-distance. Interestingly, given the fact that acwon the “battle of currents” due to its possibility to transmit power over long dis-

    tance [25], HVDC transmission wins the battle back after a century. To transmit

    the same amount of power, dc transmission needs fewer power lines than ac trans-

    mission. Accordingly, the costs and losses of the converter stations get balanced by

    savings on the overhead lines where the break-even distance is around 400 km to

    700 km depending on the land conditions and project specifications [26]. Besides,

    dc transmission does not have the stability limitation related to ac transmission over

    long distance.

    • Unsynchronized ac-system connection. AC transmission is only possible if thetwo interconnected ac systems have the same nominal frequency and operate syn-

    chronously, but dc transmission does not have such requirements. Many back-to-

    back HVDC links have been built for such purposes.

    Besides the above essential arguments, there are additional benefits by having embedded

    HVDC links in ac systems:

    • Power-system stability improvement. One of the major features of the HVDC tech-nology is its capability to manipulate large amount of power in a very short time,

    which can often be utilized to improve the stability of the ac system. One example is

    the improvement of transient stability by running up or running back the dc power

    for emergency power supports [4, 27]. Another example is that HVDC system can

    be used to damp low-frequency oscillations in ac systems by having an auxiliary

    damping controller [28].

    • Firewall function. Large interconnected ac systems have many well-known advan-tages, e.g., the possibility to use larger and more economical power plants, reduction

    of reserve capacity in the systems, utilization of the most efficient energy resources,

    as well as achieving an increase in system reliability [25,29]. However, larger inter-

    connected ac systems also increase the system complexity from the operation point

    of view. One of the consequence of such complexity is the large blackouts in Amer-

    ica and Europe [30]. In this aspect, HVDC links have the “firewall” function in

    preventing cascaded ac-system outages spreading from one system to another [31].

    10

  • 2.2. HVDC transmission using line-commutated current-source converters

    -

    -

    -

    +

    +

    +

    aU

    N

    bU

    cU

    L

    L

    L

    aI

    bI

    cI

    Ldci

    dcu

    V1

    V6

    V5

    V2V4

    V3

    Fig. 2.1 Graetz bridge for LCC-HVDC system.

    2.2 HVDC transmission using line-commutated current-

    source converters

    The converter technology used for HVDC transmission in the early days was based on

    mercury valves. The major problem with mercury-arc technology was ark-back fault

    which destroyed the rectifying property of the converter valve and consequently triggered

    other problems [23]. In the late 1960s, the thyristor valve technology was developed that

    overcame the problems of mercury-arc technology. Converters based on either mercury

    valves or thyristor valves are called line-commutated converters (LCCs), or current-source

    converters (CSCs). The basic module of an LCC is the three-phase full-wave bridge cir-

    cuit shown in Fig. 2.1. This topology is known as the Graetz bridge. Although there are

    several alternative configurations possible, the Graetz bridge has been universally used

    for LCC-HVDC converters as it provides better utilization of the converter transformer

    and a lower voltage across the valve when not conducting [32].

    The Graetz bridge can be used for transmitting power in two directions, i.e., the

    rectifier mode and the inverter mode. This is achieved by applying different firing angles

    on the valves. If the firing angle is lower than 90◦, the direct current is flowing from the

    positive terminal of the dc circuit, thus the power is following from the ac side to the dc

    side; If the firing angle is higher than 90◦, the direct voltage changes polarity, thus the

    direct current is flowing from the negative terminal of the dc circuit. The power is then

    flowing from the dc side to the ac side. An HVDC link is essentially constructed by two

    Graetz bridges, which are interconnected on the dc sides. The interconnection could be

    an overhead line, a cable, or a back-to-back connection.

    The application of LCC-HVDC technology has been very successful and the instal-

    lations of LCC-HVDC links are expected to grow at least in the near future. However, the

    LCC technology suffers from several inherent weaknesses.

    11

  • Chapter 2. High-Voltage Direct-Current Transmission

    0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04

    -1

    0

    1

    ud

    c (

    p.u

    .)

    0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04-2

    0

    2

    I v (

    p.u

    .)

    0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04

    -1

    0

    1

    Uv (

    p.u

    .)

    time (sec)

    V5 V1 V3 V1 V3

    V4 V6 V2 V4 V6 V2

    Uab Ubc Uca Ucb Uac Ubc Uca Ucb Uba Uac Uab Uba

    Uab Uac Uac Uab Uab

    Fig. 2.2 Commutation failures of an LCC-HVDC inverter. Upper plot: direct voltage. Middle plot:

    valve currents. Lower plot: valve voltage.

    One problem is that the LCC always consumes reactive power, either in recti-

    fier mode or in inverter mode. Depending on the firing angles, the reactive power con-

    sumption of an LCC-HVDC converter station is approximately 50 − 60% of the ac-tive power. The reactive-power consumption requires compensation by connecting large

    ac filters/capacitors at the converter stations. For a common LCC-HVDC link, the fil-

    ters/capacitors not only increase the costs, but also occupy large amounts of space of

    the converter stations. Besides, large filters/capacitors also contribute to the temporary

    overvoltage (TOV) and low-order harmonic resonance problems of the HVDC link when

    connects to a weak ac system [5].

    Another well-known problem of the LCC-HVDC system is the occurrence of com-

    mutation failures at the inverter station typically caused by disturbances in the ac system.

    Either depressed voltage magnitude or phase-angle shift of the alternating voltage may

    reduce the extinction volt-time area of the inverter valve [33, 34]. If the extinction angle

    of the inverter valve is smaller than 5 − 6◦, the previously conducted valve will regaincurrent, which will end up with a commutation failure. Fig. 2.2 shows plots of a typical

    commutation failure. A disturbance on Uab appears during the commutation between V1

    and V3. Because V1 does not get the reverse voltage that is needed to switch off the cur-

    rent, V1 continues to be conducted and the valve current of V3 goes down to zero again.

    When the next commutation occurs between V2 and V4, V1 and V4 conduct at the same

    time. From Fig. 2.2 it can be observed that the commutation failure, in fact, creates a short

    circuit on the dc side, which essentially temporarily stops the power transmission.

    Commutation failures are common phenomena of LCC-HVDC systems. A single

    12

  • 2.2. HVDC transmission using line-commutated current-source converters

    commutation failure generally does no harm to either the converter valves or the ac sys-

    tem. However, a number of repeated commutation failures may force the HVDC link to

    trip [35].

    While the above two problems can be mitigated by relatively easy measures, the

    third problem is more fundamental, which can become a limiting factor for LCC-HVDC

    applications. For LCCs, the successful commutation of the alternating current from one

    valve to the next relies on the stiffness of the alternating voltage, i.e., the network strength

    of the ac system. If the ac system has low short-circuit capacity relative to the power

    rating of the HVDC link, i.e., low SCR, more problematic interactions between the ac

    and the dc systems are expected. Besides, the SCR of the ac system also imposes an

    upper limitation on the HVDC power transmission, which is often described by the well-

    known maximum power curve (MPC) [5]. As mentioned before, an LCC-HVDC link

    normally requires reactive-power compensation by connecting larger ac filters/capacitors

    at the converter stations. These ac filters/capacitors create additional problems in weak ac

    systems as described below.

    One such problem is the aforementioned TOV issue. In case of a sudden change in

    the active power, or the blocking of converter, the large filters/capacitors at the converter

    station together with the high inductance of the ac system cause a temporary overvolt-

    age before the protection system disconnects the filters/capacitors. The magnitude of the

    overvoltage is directly related to the strength of the ac system. Ref. [5] gives the following

    estimation of the fundamental components of TOV (TOVfc) regarding the SCR of the ac

    system:

    • SCR > 3: TOVfc lower than 1.25 p.u.

    • 2 < SCR < 3: TOVfc higher than 1.25 p.u. but lower than 1.4 p.u.

    • SCR < 2: TOVfc higher than 1.4 p.u.

    TOVs can influence the design and costs of the dc stations. TOVs can also lead to satura-

    tion of the converter transformer or transformers close to the dc station.

    Another problem with weak-ac-system connections is the low-order harmonic res-

    onance. The high inductance of the ac system and the large filters/capacitors of the HVDC

    link create a resonance with low frequency fres, which can be estimated approximately by

    fres ≈ f1√

    2 · SCR (2.1)

    where f1 is the nominal frequency of the ac system. That is to say, the resonance frequency

    tends to be lower for weak ac systems. Generally speaking, the lower the resonance fre-

    quency, the greater the risk for harmful interaction with the converter control system.

    An improved topology of the LCC-HVDC system to overcome part of the above

    mentioned problems is the capacitor-commutated converter (CCC)-HVDC technology,

    13

  • Chapter 2. High-Voltage Direct-Current Transmission

    dci

    dcu

    -

    -

    -

    +

    +

    +

    aU

    N

    bU

    cU

    L

    L

    L

    aI

    bI

    cI

    Fig. 2.3 Two-level voltage-source converter.

    where ac capacitors are inserted in series between the valves and converter transform-

    ers [36]. The series-connected capacitors not only supply the reactive power consumed

    by the valves, it also improve the dynamic performance of the HVDC system. However,

    the major drawback of the CCC concept is that the series capacitors increase the insula-

    tion costs of the valves. Thus, the CCC-HVDC technology has been so far only applied

    to back-to-back HVDC links, where the voltage level of the valves is much lower.

    2.3 HVDC transmission using forced-commutated voltage-

    source converters

    Voltage-source converters are a new converter technology for HVDC transmission [2].

    The first commercial VSC-HVDC (HVDC Light) link with a rating of 50 MW was com-

    missioned in 1999 in Gotland island of Sweden, close to the world’s first LCC-HVDC

    link.

    Voltage-source converters (VSCs) utilize self-commutating switches, e.g., gate turn

    off thyristors (GTOs) or insulated-gate bipolar transistors (IGBTs), which can be turned

    on or off freely. This is in contrast to the LCC where the thyristor valve can only be

    turned off by reversed line voltages. Therefore, a VSC can produce its own sinusoidal

    voltage waveform using pulse-width modulation (PWM) technology independent of the

    ac system.

    Many different topologies have been proposed for VSCs. However, for HVDC ap-

    plications, they have been so far limited to three major types: two-level converter, three-

    level converter, and modular multilevel converter (M2C) [37–39].

    Fig.2.3 shows a two-level grid-connected VSC. The two-level bridge is the simplest

    topology that can be used in order to build up a three-phase forced-commutated VSC

    bridge. The bridge consists of six valves and each valve consists of a switching device

    14

  • 2.3. HVDC transmission using forced-commutated voltage-source converters

    dci

    dcu

    -

    -

    -

    +

    +

    +

    aU

    N

    bU

    cU

    L

    L

    L

    aI

    bI

    cI

    Fig. 2.4 Three-level neutral-point-clamped voltage-source converter.

    and an anti-parallel diode. For an HVDC link, two VSCs are interconnected on the dc

    side. For high-voltage applications, series connection of switching devices is necessary.

    The operation principle of the two-level bridge is simple. Each phase of the VSC can be

    connected either to the positive dc terminal, or the negative dc terminal. By adjusting the

    width of pulses, the reference voltage can be reproduced, as shown in the upper plot of

    Fig. 2.6. After filtering by phase reactors and shunt filters, this series of voltage pulses

    resembles the voltage waveform of the reference voltage.

    The three-level VSC shown in Fig. 2.4 is also called neutral-point-clamped (NPC)

    converter. The key components that distinguish this topology from the two-level converter

    are the two clamping diodes in each phase. These two diodes clamp the switch voltage to

    half of the dc voltage. Thus, each phase of the VSC can switch to three different voltage

    levels, i.e., the positive dc terminal, the negative dc terminal and the mid-point. Con-

    sequently, voltage pulses produced by a three-level VSC match closer to the reference

    voltage. Therefore, the three-level NPC converter has less harmonic content as shown in

    the middle plot of Fig. 2.6. Additionally, the three-level NPC converter has lower switch-

    ing losses. Compared to two-level VSCs, three-level NPC VSCs require more diodes for

    neutral-point clamping. However, the total number of switching components does not

    necessarily have to be higher. The reason for this is that, for HVDC applications, a valve

    consists of many series-connected switches. In the two-level case a valve has to with-

    stand twice as high voltage than in the three-level case. Accordingly, the total number of

    15

  • Chapter 2. High-Voltage Direct-Current Transmission

    A

    B

    A

    B

    A

    BA

    B

    A

    B...

    -

    +aU

    A

    B

    A

    B

    A

    B

    ...

    dcu

    (a)

    (b)

    Positive arm

    Negative arm

    V1

    V2

    Fig. 2.5 Modular multilevel voltage-source converter. (a) One M2C module. (b) One phase topol-

    ogy.

    switches is approximately equal.

    The NPC concept can be extended to higher number of voltage levels, which can

    result in further improved harmonic reduction and lower switching losses [40]. However,

    for high-voltage converter applications, the neutral-clamped diodes complicate the insu-

    lation and cooling design of the converter valve. Therefore, NPC concepts with a number

    of voltage levels higher than three has never been considered for HVDC applications [37].

    The recently proposed modular multilevel converter (M2C) concept [39, 41–43]

    attracts significant interests for high-voltage converter applications. Fig. 2.5 shows the

    M2C topology for one phase. Compared to the above two topologies, one major feature

    of the M2C is that no common capacitor is connected at the dc side. Instead, the dc

    capacitors are distributed into each module, while the converter is built up by cascade-

    connected modules.

    Fig. 2.5(a) shows an M2C module. Each M2C module consists of two valves which

    can be switched in three different ways:

    • V2 is turned on and V1 is turned off, the capacitor is inserted into the circuit fromA to B. The module contributes with voltage to the phase voltage. The capacitor is

    charged if the current is from A to B, and discharged otherwise.

    • V1 is turned on and V2 is turned off, the capacitor is by passed. The module doesnot contribute with voltage to the phase voltage.

    • Both V1 and V2 are turned off, the module is blocked.

    16

  • 2.3. HVDC transmission using forced-commutated voltage-source converters

    0 0.01 0.02 0.03 0.04 0.05 0.06

    -100

    0

    100

    UL

    1 (

    kV

    )

    0 0.01 0.02 0.03 0.04 0.05 0.06

    -100

    0

    100

    UL

    1 (

    kV

    )

    0 0.01 0.02 0.03 0.04 0.05 0.06

    -100

    0

    100

    time (sec)

    UL

    1 (

    kV

    )

    Fig. 2.6 Pulse-width modulation for different converter topologies. Upper plot: two-level con-

    verter. Middle plot: three-level converter. Lower plot: M2C with five modules .

    The M2C concept is especially attractive for high-voltage applications, since the con-

    verter can be easily scaled up by inserting additional modules in each arm. If consider-

    able amounts of modules are cascaded (approximately 100 modules would be common

    for HVDC applications), each module theoretically only needs to switch on and off once

    per period, which greatly reduces the switching losses of the valves. However, prelimi-

    nary investigation indicates that slightly higher switching frequencies are necessary. The

    lower plot of Fig. 2.6 shows the voltage waveform produced by a five-module (five for

    each arm) M2C. With only five modules, the waveform already resembles much better

    the sinusoidal voltage reference than the other two topologies. With M2C, the harmonic

    content of the voltage produced by the VSC is so low that additional filtering equipment

    is almost unnecessary.

    An additional benefit of the M2C is that the control system has an extra freedom in

    dealing with faults at the dc side. The dc capacitors are not necessarily discharged during

    faults. Thus, the fault recovery can be faster [39].

    Compared to the other two topologies, the major drawback of the M2C topology is

    that the required switching components are doubled since only one of the valves of each

    module contributes to the phase voltage when the module is inserted in. In addition, the

    design and control of the M2C are generally more complex at least than the two-level con-

    verter. However, since the switching frequency of the M2C can be kept very low switches

    with higher blocking voltages may be used, which in turn limits the increase in number

    of switches. On the other hand, the reduction of switching losses and savings on filtering

    equipment of the M2C may eventually justify its application for HVDC transmission.

    17

  • Chapter 2. High-Voltage Direct-Current Transmission

    5.0

    0.1

    5.0 0.1

    Reactive power

    Over-voltage

    limitation

    Under-voltage

    limitation

    Converter current

    limitation

    Active power

    Fig. 2.7 PQ diagram for a typical VSC-HVDC converter.

    No matter what converter topology is used, the VSC can always be treated as an

    ideal voltage source where the control system has the freedom to specify the magnitude,

    phase, and frequency of the produced sinusoidal voltage waveform. However, for control

    design and stability analysis, it is important to take into account the limitation of the

    converter in terms of active and reactive power transfer capability.

    One such limit is the converter-current limitation, which is imposed by the current

    carrying capability of the VSC valves. Since both the active power and the reactive power

    contribute to the current flowing through the valves, this limitation is manifested as a

    circle in a PQ diagram. Accordingly, if the converter is intended to support the ac system

    with reactive-power supply/consumption, the maximum active power has to be limited to

    make sure that the valve current is within the limit.

    Another limitation which determines the reactive-power capability of the VSC is

    the over/under voltage magnitude of the VSC (modulation index limitation). The over-

    voltage limitation is imposed by the direct-voltage level of the VSC. The under-voltage

    limit, however, is limited by the main-circuit design and the active-power transfer capa-

    bility, which requires a minimum voltage magnitude to transmit the active power. In this

    respect, the tap-changer of the converter transformer can play an important role to extend

    the reactive-power limitation of the VSC. This could be an argument to have converter

    transformers in VSC-HVDC systems. Fig. 2.7 shows the PQ diagram with the above

    mentioned limitations for a typical VSC-HVDC converter [44].

    VSC-HVDC technology overcomes most of the weaknesses of the LCC-HVDC

    technology. In addition, it supports the ac system with reactive-power supply/consumption.

    18

  • 2.3. HVDC transmission using forced-commutated voltage-source converters

    Similar to an LCC-HVDC system, a VSC-HVDC system can quickly run up or run back

    the active power for ac system emergency-power support, but it can also instantly reverse

    the active power [4].

    Since the direct voltage of a VSC-HVDC system varies in a much smaller range

    than that of a LCC-HVDC system, extruded cables can be used for VSC-HVDC systems.

    The extruded cable reduces the cable cost and the construction cost. The latter makes

    long-distance land-cable transmission possible [2].

    Besides the above features, the most essential one is that a VSC-HVDC system

    has an unlimited connection capability with ac systems, i.e., with properly designed con-

    trol systems, VSC-HVDC system has the potential to be connected to any kind of ac

    system with any number of links. This outstanding property will eventually bring the dc-

    transmission technology to ever broader application fields!

    19

  • Chapter 2. High-Voltage Direct-Current Transmission

    20

  • Chapter 3

    Control Methods for VSC-HVDC

    Systems

    This chapter describes various control methods used for VSC-HVDC systems. In Sec-

    tion 3.1 and Section 3.2 two existing control methods, i.e., power-angle control and

    vector current control are described. A novel control method, i.e., the so-called power-

    synchronization control, is introduced in Section 3.3. The major results of this chapter are

    summarized in Section 3.4. Some results of this chapter are included in [45].

    3.1 Power-angle control

    Power-angle control is also called voltage-angle control. It is perhaps the most straightfor-

    ward controller for grid-connected VSCs [46–48]. The principle of power-angle control

    is based on the following well-known equations

    P =U1U2 sin θ

    X

    Q =U21 − U1U2 cos θ

    X(3.1)

    where P and Q are the active and reactive powers between two electrical nodes in ac

    systems with voltage magnitudes U1 and U2. The quantities θ and X are the phase-angle

    difference and line reactance between the two nodes. From (3.1) it follows that the active

    power is mainly related to the phase angle θ, while the reactive power is more related to

    the voltage-magnitude difference. These mathematical relationships are the foundation of

    power-angle control, i.e., the active power is controlled by the phase angle of the VSC

    voltage, while the reactive power or filter-bus voltage is controlled by the magnitude of

    the VSC voltage.

    Fig. 3.1 shows the main-circuit and control block diagram of a VSC-HVDC con-

    verter using power-angle control. Lc is the inductance of the phase reactor, and Ln is the

    21

  • Chapter 3. Control Methods for VSC-HVDC Systems

    nL cL

    -+refPP

    QP,

    E

    VSC

    ref

    av

    ref

    bvref

    cv

    +

    -fu

    fC

    APC

    PLL

    RPC/AVC-+refQQ V∆

    ci

    v

    +

    -

    Voltage

    reference

    control

    -+

    fU

    refU

    Fig. 3.1 Main-circuit and control block diagram of a VSC-HVDC converter using power-angle

    control.

    inductance of the ac system. Cf is the ac capacitor connected at the filter bus. The bold

    letter symbols E, uf , and v represent the voltage vectors of the ac source, the filter bus,

    and the VSC respectively. P and Q are the active power and reactive power from the VSC

    to the ac system. The quantity ic is the current vector of the phase reactor.

    To produce three-phase alternating voltages, the VSC needs three variables: magni-

    tude, phase angle and frequency. With power-angle control, these three variables are given

    by three different controllers, i.e., the reactive-power controller (RPC) or the alternating-

    voltage controller (AVC), the active-power controller (APC), and the phase-locked loop

    (PLL). The above controllers are briefly described in below:

    • Reactive-power controller. The reactive power to/from the VSC is controlled bythe magnitude of the VSC voltage. A proportional-integral (PI) controller can be

    applied, e.g.,

    ∆V =

    (KQp +

    KQis

    )[Qref −Q] . (3.2)

    where the output ∆V gives the change in magnitude of the VSC reference voltage.

    • Alternating-voltage controller. Alternatively, the VSC-HVDC converter controlsthe filter-bus voltage instead of the reactive power. The output of the AVC is the

    same as that of the RPC. A PI controller can be applied, e.g.,

    ∆V =

    (KUp +

    KUis

    )[Uref − Uf ]. (3.3)

    • Active-power controller. The active power to/from the VSC is controlled by thephase angle of the VSC voltage. A proportional-integral (PI) controller can be ap-

    22

  • 3.1. Power-angle control

    plied, i.e.,

    θv =

    (KPp +

    KPis

    )[Pref − P ]. (3.4)

    where the output θv gives the change in phase angle of the VSC reference voltage.

    • Phase-locked loop. The function of the PLL is to synchronize the VSC to the acsystem. Below a description of a PLL design suitable for power-angle control is

    given.

    If ω1 is the angular frequency of the ac system, and ω is the angular frequency of the

    VSC, a PLL controller has the objective to follow the phase angle of the filter-bus

    voltage by minimizing

    e = (ω1 − ω)t. (3.5)

    If the error e in (3.5) is sufficiently small, (3.5) can be approximated by

    e ≈ sin(ω1t− ωt)= sinω1t cosωt− cosω1t sinωt (3.6)

    where sinω1t and cosω1t can be obtained by αβ transformation of the filter-bus

    voltage as shown below.

    The phase quantities of the filter-bus voltage can be defined as

    ufa = Uf0 cos(ω1t)

    ufb = Uf0 cos(ω1t− 120◦)ufc = Uf0 cos(ω1t− 240◦). (3.7)

    The corresponding real and imaginary parts of the vector uf in the stationary αβ

    reference frame (see Appendix A) can be written as

    ufα = Uf0 cosω1t, ufβ = Uf0 sinω1t. (3.8)

    Substituting (3.8) into (3.6), yields

    e ≈ ufβUf0

    cosωt− ufαUf0

    sinωt. (3.9)

    A PI controller can be used to minimize the error e, i.e.,

    θPLL =

    (KPLLp +

    KPLLis

    )e. (3.10)

    Fig. 3.2 shows the control block diagram of the PLL. The angle change θPLL is

    added to a reference frequency signal ωreft.

    23

  • Chapter 3. Control Methods for VSC-HVDC Systems

    abc

    αβ

    trefω

    fau

    fbu

    fcu

    COS

    SIN

    0

    1

    fU

    0

    1

    fU

    s

    KK ip

    PLLPLL +

    αfu

    βfu

    +

    - PLLθ

    +

    +

    Fig. 3.2 PLL for power-angle control.

    • Voltage-reference control. By having ∆V , θv and ωt, the three-phase referencevoltages of the VSC can be formulated as:

    vrefa = (V0 + ∆V ) cos(ωt+ θv)

    vrefb = (V0 + ∆V ) cos(ωt+ θv − 120◦)vrefc = (V0 + ∆V ) cos(ωt+ θv − 240◦) (3.11)

    where V0 is a nominal voltage reference, e.g., V0 = 1.0 p.u.

    As shown in this section, the design and implementation of power-angle control

    is simple and straightforward. However, power-angle control practically has never been

    applied to any real VSC-HVDC system, since it suffers from two fundamental problems:

    1. The control system has no general means to damp the various resonances in the

    ac system. Therefore, the bandwidth of the controller is very much limited by the

    resonances in the ac system, especially the one at the grid frequency [48].

    Fig. 3.3 shows a plot of active-power and reactive-power step responses with power-

    angle control. The resonance at the grid frequency can be easily observed. Although

    the resonance can be damped out by applying notch filters in the active-power and

    reactive-power controllers, or canceled by some model-based control designs [49],

    the effects of such measures are doubtful since the ac system is a highly uncertain

    process where not all of the resonance frequencies are known.

    2. The control system does not have the capability to limit the valve current of the

    converter. This is a serious problem, as the converters of a VSC-HVDC link usually

    do not have over-current capability. It is very important for the control system to

    limit the valve current to prevent the converters from being blocked (tripped) at

    disturbances.

    24

  • 3.2. Vector current control

    0 0.5 1 1.5 2-0.05

    0

    0.05

    0.1

    0.15

    time (sec)

    Pre

    f, P

    (p

    .u.)

    Pref

    P

    0 0.5 1 1.5 2-0.05

    0

    0.05

    0.1

    0.15

    time (sec)

    Qre

    f, Q

    (p

    .u.)

    Qref

    Q

    Fig. 3.3 Step response of active power (upper plot) and reactive power (lower plot) with power-

    angle control. Observe the resonance at the grid frequency.

    3.2 Vector current control

    Vector current control of VSCs has initially been applied to variable-speed drives, where

    the VSC is connected to an ac motor [50,51]. By utilizing the dq decoupling technique, the

    control system is able to control the torque and flux independently. Once field orientation

    is obtained the ac motor can be controlled using principles quite similar to those of dc-

    motor control. Vector current control is generally considered as a substantial step in ac-

    motor control [52].

    The application of vector current control on grid-connected VSCs is often consid-

    ered as a dual problem of drive control [17,53]. The basic principle is to control the active

    power and the reactive power independently through an inner-current control loop [54,55].

    As shown in Fig. 3.4, the essence of vector current control is that the control system

    creates a converter dq frame, where a PLL is applied to make sure that the d-axis of the

    converter dq frame is always aligned with the filter-bus voltage in order to synchronize

    the VSC to the ac system. A simple implementation of a current controller is achieved by

    the proportional-type control law

    vcref = αcLc (iref − icc) + jω1Lcicc +HLP(s)ucf (3.12)

    where αc is the desired closed-loop bandwidth of the inner-current controller, iref is the

    converter current reference, and vcref is the voltage reference of the VSC. The superscript

    c denotes the converter dq frame. The term jω1Lcicc is used to remove the so-called cross-

    coupling. The function HLP(s) is a low-pass filter to improve the disturbance rejection

    25

  • Chapter 3. Control Methods for VSC-HVDC Systems

    Grid-q

    Grid-d

    fu

    v

    E

    VSC-d

    VSC-q

    Fig. 3.4 Converter dq frame of vector current control.

    capability of the current controller. HLP(s) has the following expression

    HLP(s) =αf

    s + αf(3.13)

    where αf is typically chosen with a bandwidth (40 − 100 rad/s) [13].By applying the control law in (3.12), vector current control is claimed to be able to

    control the active power and the reactive power independently by the d and q components

    of the current reference iref . However, this is a conditional correct conclusion, which is

    only valid if the filter-bus voltage is sufficiently stiff such that its dynamics are negligible.

    A brief analysis is given below.

    In a stationary frame, the dynamic equation of the phase reactor of the VSC can be

    described by Kirchhoff’s voltage law as

    Lcdis

    dt= vs − usf (3.14)

    where the superscript s denotes the stationary reference frame. If it is assumed that

    ω = ω1 (3.15)

    at all times, i.e., the angular frequency ω of the converter dq frame equals the angular

    frequency ω1 of the grid, the following relations are established

    usf = ucfe

    jω1t, isc = icce

    jω1t, vs = vcejω1t. (3.16)

    Substituting (3.16) into (3.14) yields the dynamic equation in the converter dq frame

    Lcdiccdt

    = vc − ucf − jω1Lcicc. (3.17)

    If the switching-time delay is neglected and it is assumed that |vcref | does not exceed themaximum voltage modulus, then vc = vcref . Substituting (3.12) into (3.17) yields

    Lcdiccdt

    = αcLc(iref − icc) −s

    s + αfucf . (3.18)

    26

  • 3.2. Vector current control

    By writing (3.18) in component form and applying Laplace transform (s = d/dt), yields

    iccd =αc

    s+ αcirefd −

    s

    Lc(s+ αc)(s+ αf )ucfd

    iccq =αc

    s+ αcirefq −

    s

    Lc(s+ αc)(s+ αf )ucfq. (3.19)

    The above design approach for the inner-current controller is often referred to as internal-

    model control (IMC) design [51], since the bandwidth of the inner-current control is

    explicitly specified in the control parameters. Another common design approach is the

    deadbeat-current control design [56, 57], which can only be realized by digital imple-

    mentations. Generally speaking, if the bandwidth αc of IMC is chosen sufficiently high,

    IMC and deadbeat-current control give similar results. For either of the control design,

    the control bandwidth is basically limited by the switching frequency of the PWM and

    the sampling period of the computer. Moreover, both methods rely on a good knowledge

    of the value of Lc.

    The following analysis will establish the relations between the active/reactive power

    and the current references of the inner-current control. Assuming per unit quantities, the

    instantaneous active power and reactive power from the VSC to the filter bus are given by

    P = Re{ucf(i

    cc)

    ∗}, Q = Im

    {ucf (i

    cc)

    ∗}. (3.20)

    Linearizing (3.20) yields the following expressions

    ∆P =

    [iccd0

    iccq0

    ]T [∆ucfd

    ∆ucfq

    ]+

    [ucfd0

    ucfq0

    ]T [∆iccd

    ∆iccq

    ]

    ∆Q =

    [icd0

    −icq0

    ]T [∆ucfq

    ∆ucfd

    ]+

    [ucfd0

    ucfq0

    ]T [ −∆iccq∆iccd

    ](3.21)

    where the subscript 0 denotes the operating-point value. In the converter dq frame, in the

    steady state, the q component of the filter-bus voltage equals zero and the d component

    equals the voltage magnitude, i.e.,

    ucfd0 = Uf0, ucfq0 = 0. (3.22)

    If the dynamics of the filter-bus voltage are neglected, it follows that

    ∆ucfd = ∆ucfq = 0. (3.23)

    By substituting (3.22) and (3.23) into (3.21), the expressions of ∆P and ∆Q can be

    simplified as

    ∆P = Uf0∆iccd, ∆Q = Uf0∆i

    ccq. (3.24)

    27

  • Chapter 3. Control Methods for VSC-HVDC Systems

    Linearizing (3.19) and further substituting it into (3.24) [with the condition in (3.23)]

    yields the following relationship in transfer matrix form

    [∆P

    ∆Q

    ]=

    [Uf0

    αcs+αc

    0

    0 Uf0αc

    s+αc

    ]

    ︸ ︷︷ ︸J(s)

    [∆irefd

    ∆irefq

    ]. (3.25)

    The transfer matrix J(s) is called Jacobian transfer matrix in this thesis, which is a gen-

    eral concept for dynamic modeling of ac/dc systems that is to be introduced in Chapter 4.

    Eq. (3.25) shows that the Jacobian transfer matrix J(s) is diagonal, i.e., ∆P is

    only related to ∆irefd while ∆Q is only related to ∆irefq , and no cross-coupling between

    the two loops exists. However, (3.25) is derived based on the assumptions of (3.15) and

    (3.23). Both of the assumptions are related to the stiffness of the filter-bus voltage. If the

    ac system is strong enough, i.e., Ln ≪ Lc, the dynamics of the filter-bus voltage can beneglected. However, if the ac system is weak, the assumptions in (3.15) and (3.23) no

    longer hold. Therefore, the weaker the ac system, the higher the off-diagonal elements in

    J(s), i.e., the more interactions between the active-power and the reactive-power control.

    Consequently, to analyze the stability of vector current control for VSC-HVDC

    links connected to weak ac systems, the dynamics of the filter-bus voltage have to be

    considered. That is, the Jacobian transfer matrix J(s) in (3.25) should take into account

    the grid inductance Ln and the dynamics of the PLL. Such a model will be developed in

    Chapter 4. An in-depth analysis of the difficulty with vector current control for weak-ac-

    system connections will be given in Chapter 5.

    Fig. 3.5 shows the main-circuit and control block diagram of a VSC-HVDC con-

    verter using vector current control. The active-power controller and the reactive-power or

    the alternating-voltage controller of vector current control can be designed in a similar

    way as power-angle control but with irefd and irefq as outputs. However, the PLL can be de-

    signed in a more concise way by utilizing the concept of the converter dq frame , i.e., a PI

    controller is applied to minimize the q component of the filter-bus voltage in the converter

    dq frame

    θPLL =

    (KPLLp +

    KPLLis

    )Im{ucf}. (3.26)

    In this way, the VSC is synchronized to the ac system. Fig. 3.6 shows the control block

    diagram of the PLL for vector current control. With vector current control, the voltage

    reference of the VSC is formulated by vrefd , vrefq and ωt. This is essentially the same as

    power-angle control where the reference of the VSC voltage is formulated by the magni-

    tude, the phase angle and the frequency. In the former case the rectangular form is used

    while the latter uses polar form. The mathematical expressions of the dq-αβ and αβ−abcblocks in Fig. 3.5 are given in Appendix A.3.

    For vector current control, given sufficiently high bandwidth, the dq components of

    the converter current always follow the corresponding current references. Consequently,

    28

  • 3.2. Vector current control

    nL cL

    -+refPP

    QP,

    E

    +

    -fu

    fC

    APC

    PLL

    RPC/AVC-+refQQ

    ci

    v

    +

    -

    Inner-

    current

    controller

    dqαβ abc

    αβ

    fu ci-

    +refU

    fU

    ref

    αv

    ref

    βv

    ref

    dv

    ref

    qvref

    di

    ref

    qi

    VSC

    ref

    av

    ref

    bvref

    cv

    Fig. 3.5 Main-circuit and control block diagram of a VSC-HVDC converter using vector current

    control.

    ++

    s

    KK ip

    PLLPLL +

    c

    fu}{Im cfu

    PLLθ tω

    trefω

    Fig. 3.6 PLL for vector current control.

    by limiting the modulus of the current references, the valve current of the converter is

    limited. The simulation results in Fig. 3.7 show the fault ride-through capability of a VSC-

    HVDC link using vector current control. A three-phase ac-system fault with 0.2 s duration

    is applied at 0.1 s close to the filter bus. The modulus of the current reference [|irefc | =√(irefd )

    2 + (irefq )2] reaches the current limit Imax immediately after the fault occurrence.

    The control system automatically limits the converter current. After the fault is detected,

    the control system reduces the fault current to half of the maximum load current (or any

    other desired values to minimize the short-circuit current contribution to the ac system)

    except a very short current spike at the fault inception. In VSC applications, regardless

    of the control principle, the converter always tries to protect itself from excessive over

    currents. This fast protection is often implemented as a low-level hardware system, and

    its objective is to protect the converter in cases where the higher levels of control fail.

    Since the current spike in Fig. 3.7 is so short ( < 1.6 pu in magnitude and < 5 ms in

    duration ), it neither does any harm to the converter valve, nor contributes much to the

    short-circuit current to the ac system.

    Besides the fault-current limitation capability, the current control also has a damp-

    ing effect on resonances in the ac system. Therefore, vector current control overcomes

    the two fundamental problems of power-angle control. In practice, vector current control

    29

  • Chapter 3. Control Methods for VSC-HVDC Systems

    0 0.2 0.4 0.6 0.8 10

    0.5

    1

    P (

    p.u

    .)

    0 0.2 0.4 0.6 0.8 1

    0

    0.5

    1

    i re

    f

    d, i r

    ef

    q (

    p.u

    .)

    0 0.2 0.4 0.6 0.8 10

    1

    2

    time (sec)

    | i c

    | (p

    .u.)

    i ref

    d

    i ref

    q

    Fig. 3.7 Fault ride-through capability of vector current control. Upper plot: active power from the

    VSC. Middle plot: dq components of the current reference. Lower plot: modulus of the

    converter current.

    has been successfully applied to a number of commercial VSC-HVDC links. However, a

    major drawback of vector current control is its poor performance for VSC-HVDC links

    connected to weak ac systems, which becomes an obstacle for VSC-HVDC transmission

    to be applied in more challenging ac-system conditions.

    To overcome the problem of vector current control with weak-ac-system connec-

    tions, a novel control method, i.e., power-synchronization control, is proposed in the next

    section. In some sense, power-synchronization control might be viewed upon as a combi-

    nation of power-angle control and vector current control.

    3.3 Power-synchronization control

    With the two control methods described in the previous sections, a PLL is used to synchro-

    nize the VSC with the ac system. This has since long been believed to be a pre-condition

    for any grid-connected VSC. However, there is, in fact, an internal synchronization mech-

    anism in ac systems that the VSC can utilize to synchronize with the ac system. In this

    section, a control method based on such type of synchronization is proposed. The major

    goal of the proposed control method is to overcome the problem of vector current control

    with weak-ac-system connections.

    30

  • 3.3. Power-synchronization control

    1SM 2SMX

    222

    2 mem TT

    dt

    dJ −=

    ω11

    11 em

    m TTdt

    dJ −=

    ω

    Fig. 3.8 Synchronization mechanism between SMs in an ac system.

    3.3.1 Power-synchronization mechanism in ac systems

    In this sub-section, the power-synchronization mechanism between synchronous machines

    (SMs) in ac systems is described. The mechanism is illustrated by a simple system con-

    sisting of two interconnected SMs as shown in Fig. 3.8. SM1 operates as a generator and

    SM2 operates as a motor. The reactance X is the sum of the reactances of the SMs and

    that of the line interconnecting the two SMs. All resistances and other damping effects

    are disregarded.

    Initially, it is assumed that the two SMs operate at steady state, as described by the

    phasor diagram in Fig. 3.9(a). The phasorsE1 andE2 represent the line-to-line equivalents

    of the inner emfs of the two SMs respectively. These emfs are assumed to be constant at

    all times (even during transients). The electric power transmitted from SM1 to SM2 is

    given by

    P =E1E2 sin θ

    X(3.27)

    where θ is the electrical angle separating the two emfs E1 and E2. The mechanical torque

    Tm1 of SM1 is now increased by a certain amount for a short duration and then brought

    back to its initial value. As a consequence of the temporary increase of Tm1, the me-

    chanical angle of the rotor of SM1 advances, as predicted by the generator-mode swing

    equation

    J1dωm1dt

    = Tm1 − Te1 (3.28)

    where J1 is the total inertia of the shaft-system of SM1 , ωm1 is the rotor speed, and Te1 is

    the electromagnetic torque of of SM1. Since the emf of a synchronous machine is tightly

    connected to the rotor position, the advance of the mechanical angle of the rotor of SM1inevitably causes an advance of the phase of the emf of SM1, as indicated by the phasor

    E′

    1 in Fig. 3.9(b). The initial position of E1 is shown as a dashed line in Fig. 3.9(b). Due

    to the phase advancement indicated by E′

    1, the phase difference between the emfs of the

    two SMs is increased. According to (3.28), this translates into an increase of the electric

    power transmitted from SM1 to SM2 . This increase in power is equivalent to an increase

    in the electromagnetic torque Te2 of SM2 . Assuming that SM2 has a constant load torque

    Tm2, the rotor of SM2 starts to accelerate as dictated by

    J2dωm2dt

    = Te2 − Tm2 (3.29)

    31

  • Chapter 3. Control Methods for VSC-HVDC Systems

    '1E1E 2E

    θ2E

    (a) (b)

    Fig. 3.9 Phasor diagrams describing power synchronization.

    where J2 is the total inertia of the shaft-system of SM2 , and ωm2 is the mechanical angular

    velocity of SM2 . As the rotor of SM2 starts to accelerate, the same thing occurs with the

    phase of E2, as indicated by the arc-shaped arrow in Fig. 3.9(b). The acceleration of the

    phasor E2 results in a reduction of the phase difference between the emfs of the two

    SMs. After a transient, which in reality involves a certain amount of damping, the phase

    difference between the emfs of the two SMs is brought back to its initial value (as the

    transmitted electric power), and the system is again at steady state.

    The synchronization mechanism described above is known to all power system spe-

    cialists, i.e., the synchronization process is achieved by means of a transient power trans-

    fer. The same kind of synchronization also appears in large systems of interconnected

    synchronous machines.

    Due to the fact that the synchronous machines can maintain operation in various ac-

    network conditions while the vector-current-controlled VSCs are prone to fail, it makes

    sense to suggest a control method based on a synchronization process where the electric

    power is the communicating medium. In the next sub-section, a controller based on power

    synchronization is proposed.

    3.3.2 Power-synchronization control of grid-connected VSCs

    From the discussion in the preceding sub-section it is known that the SMs in an ac system

    maintain synchronism by means of power synchronization, i.e., a transient power trans-

    fer. This power transfer involves a current which is determined by the interconnecting

    network. Generally, this current is unknown. If power synchronization should be used to

    control a VSC, therefore, it cannot be combined with a vector current controller, which

    requires a known current reference. As will be shown below, the active power output from

    the VSC is instead controlled directly by the power-synchronization loop and the reactive

    power (or alternating voltage) is controlled by the magnitude of the VSC voltage. Conse-

    quently, an inner current loop is not necessary from a power and voltage control point of

    view.

    However, as it was mentioned in Section 3.2, besides power and voltage control,

    the current controller is also important in

    1. Providing damping effects to poorly-damped resonances in ac systems.

    2. Limiting the valve current of the converter during severe ac-system faults.

    32

  • 3.3. Power-synchronization control

    To make the essence of power-synchronization control easier to be captured, the ini-

    tial proposal of the control law only considers the damping issue, while the fault-current-

    limitation issue is proposed as a modification of the initial design.

    Accordingly, an initial control design based on the power-synchronization law is

    proposed as

    • Power-synchronization loop (PSL). The control law is given by

    θv =kps

    (Pref − P ). (3.30)

    where θv supplies the synchronization input to the VSC, i.e., ωt = ωreft + θv. The

    power-synchronization loop is essentially an emulation of the swing equation, how-

    ever, not an exact copy. Since the mechanical angular velocity ωm is the derivative

    of the angular position, (3.28) represents a double integration when going from

    torque (or electric power) to angular position. This double integration, inherently,

    yields a poor phase margin even with considerable damping. Therefore, the pro-

    posed power-synchronization law in (3.30) employs only a single integration.

    • Alternating-voltage control (AVC). The control law is given by

    ∆V =kus

    (Uref − Uf ). (3.31)

    where ∆V gives the change in magnitude of the VSC reference voltage. The AVC

    can also be viewed as an emulation of the exciter control of a synchronous ma-

    chine. A normal exciter control of a synchronous machine is of proportional type.

    However, it is found to be more beneficial to have integral process for the VSC to

    suppress high-frequency disturbances. If there are other voltage-controlling devices

    connected close to the filter bus, a load compensation should be applied to avoid

    voltage hunting. This issue will be discussed in Chapter 6.

    • Reactive-power control (RPC). When operating against a weak ac system, theVSC-HVDC converter should preferably be operated in AVC mode to give the ac

    system the best possible voltage support. In case reactive-power control is required,

    the output of this controller should be added to the alternating-voltage reference,

    and the added amount should be limited. The PI-type controller proposed in (3.2)

    can be used for the RPC but with voltage reference change ∆Uref as output.

    • Voltage-vector control law. The control law of the voltage vector of the VSC isproposed as

    vcref = (V0 + ∆V ) −HHP (s) icc (3.32)where V0 is the nominal value, e.g., V0 = 1.0 p.u., and ∆V is given by the AVC.

    HHP (s) is a high-pass filter for damping purpose, which is expressed by

    HHP (s) =kvs

    s+ αv(3.33)

    33

  • Chapter 3. Control Methods for VSC-HVDC Systems

    where αv should be chosen low enough to cover all the possible resonances in the

    ac system. Typically αv should be chosen between 30 rad/s and 50 rad/s to also

    cover the subsynchronous resonance in ac systems. The gain kv determines the level

    of the damping effect with typical values between 0.2 p.u. and 0.6 p.u. The effects

    of HHP (s) for resonance damping will be further analyzed in Chapter 4.

    In the following, a current limitation scheme is proposed as a modification of the

    initial design. The principle is to have the control system in current-limitation mode au-

    tomatically once the converter current is above the limit.

    In case that the converter current is above the maximum limit Imax, the desired

    control law of the VSC is the inner-current control law of vector current control in (3.12).

    However, instead of giving a constant current reference to (3.12), the value of iref in (3.12)

    is given by

    iref =1

    αcLc

    [(V0 + ∆V ) −HHP (s) icc −HLP(s)ucf − jω1Lcicc

    ]+ icc. (3.34)

    The current reference in (3.34) is designed in such a way that the control law in (3.12)

    becomes (3.32) in normal operation. This can be easily verified by substituting (3.34)

    into (3.12). However, the current reference iref in (3.34) gives an indication of the actual

    converter current. During ac-system faults, current limitation is automatically achieved by

    limiting the modulus of iref to the maximum current limit Imax. A brief analysis of this is

    given below.

    The dynamics of the converter current in the converter dq frame can be described

    by

    Lcdiccdt

    = vc − ucf − jω1Lcicc. (3.35)

    Assuming vc = vcref , substituting (3.12) into (3.35) yields,

    Lcdic

    dt= αcLc (iref − icc) −

    s

    s+ αfucf . (3.36)

    By setting the time derivative and the Laplace operator s to zero, it is found that

    iref = icc. (3.37)

    That is, the current reference is identical to the actual converter current in the steady state.

    In other words, by limiting the modulus of the current reference, the converter current is

    limited.

    Fig. 3.10 shows the overview of power-synchronization cont