100
Steady State Analysis of the Interconnection of Offshore Energy Parks Miguel Jorge da Rocha Barros Marques Dissertação para obtenção do Grau de Mestre em Engenharia Electrotécnica e de Computadores Júri Presidente: Prof. Doutor Paulo José da Costa Branco Orientador: Prof. Doutor Rui Manuel Gameiro de Castro Co-orientador: Profª. Doutora Maria Eduarda de Sampaio Pinto de Almeida Pedro Vogal: Profª. Doutora Sónia Maria Nunes dos Santos Paulo Ferreira Pinto Setembro de 2010

Steady State Analysis of the Interconnection of Offshore ... em ambas as tecnologias. O software PSS/E é usado para simulação de todos os casos. Os resultados obtidos mostram algumas

Embed Size (px)

Citation preview

Steady State Analysis of the Interconnection of Offshore

Energy Parks

Miguel Jorge da Rocha Barros Marques

Dissertação para obtenção do Grau de Mestre em

Engenharia Electrotécnica e de Computadores

Júri

Presidente: Prof. Doutor Paulo José da Costa Branco

Orientador: Prof. Doutor Rui Manuel Gameiro de Castro

Co-orientador: Profª. Doutora Maria Eduarda de Sampaio Pinto de Almeida Pedro

Vogal: Profª. Doutora Sónia Maria Nunes dos Santos Paulo Ferreira Pinto

Setembro de 2010

i

Agradecimentos

Em primeiro lugar quero agradecer ao Professor Rui Castro pela confiança depositada em mim

desde o princípio e pelo apoio prestado em todos os momentos da elaboração desta tese. O meu

obrigado também à Professora Maria Eduarda Pedro, como co-orientadora, pela disponibilidade ao

esclarecimento de dúvidas que surgiram. Uma palavra de agradecimento ainda ao Professor Ferreira

de Jesus pela paciência e disponibilidade quase totais e que em muito contribuíram para a boa

conclusão deste trabalho.

Uma palavra ainda de forte agradecimento aos meus colegas de curso – alguns dos quais se

tornaram verdadeiros amigos – por toda a camaradagem e espírito de entreajuda, essenciais ao

longo de todo o curso.

Uma palavra muito importante de gratidão à minha família. Aos meus pais, não só por terem

viabilizado economicamente os meus estudos mas, sobretudo, pela força e carinho que me deram.

Os seus percursos pessoais e académicos foram sempre um exemplo e uma fonte de inspiração ao

longo do meu trajecto académico. À minha irmã, cujo apoio entusiasmado e companheirismo, ainda

que por vezes à distância, foi também essencial.

Um agradecimento muito especial à minha namorada, por ter estado ao meu lado em todos os

momentos, mesmo nos mais difíceis, e me ter sempre apoiado com ânimo, compreensão e amor,

tornando definitivamente mais fácil o meu percurso.

ii

Resumo

Entre as fontes renováveis de energia, a energia eólica tem ganho relevância, tornando-se uma

alternativa viável às fontes convencionais. As elevadas velocidades do vento e as vastas áreas

disponíveis no mar aumentaram recentemente o interesse global em parques eólicos offshore.

À medida que os parques eólicos offshore se tornam maiores e são instalados mais longe da

costa, a transmissão de energia para a rede em terra torna-se numa característica essencial. As

opções disponíveis são HVAC, HVDC-LCC e HVDC-VSC.

O objectivo deste trabalho é comparar a performance em regime estacionário e transitório de

dois sistemas de transmissão alternativos: HVAC e HVDC-LCC. Alguns esquemas de compensação

do excesso de energia reactiva para o HVAC são também avaliados, tal como as vantagens do

STATCOM em ambas as tecnologias. O software PSS/E é usado para simulação de todos os casos.

Os resultados obtidos mostram algumas diferenças importantes entre as tecnologias. Em regime

permanente, o comportamento do HVDC-LCC e do HVAC com compensação onshore e offshore têm

algumas semelhanças, nomeadamente no factor de potência e nas perdas. Em regime transitório, a

resposta da ligação com HVAC a um defeito na rede é muito diferente da resposta da ligação com

HVDC-LCC. A ligação com HVAC permite o “fault ride through” do parque eólico offshore enquanto a

ligação com HVDC-LCC é bloqueada, interrompendo a transmissão de energia durante o defeito.

Palavras-Chave: Parques Eólicos Offshore, HVAC, HVDC-LCC, STATCOM, PSS/E.

iii

Abstract

Among the renewable energy sources, wind energy has gained relevance, becoming a viable

alternative to conventional sources. High wind speeds and wide available area at sea have recently

increased the global interest on offshore wind farms.

As offshore wind farms become larger and are placed further from the shore, the power

transmission to the onshore grid becomes a key feature. The available options are HVAC, HVDC-LCC

and HVDC-VSC.

The objective of this work is to compare the performance in steady-state and the transient

behaviour of two alternative transmission systems: HVAC and HVDC-LCC. Compensation schemes

for excessive reactive power for HVAC are also evaluated, as well as the benefits of the STATCOM

for both technologies. The PSS/E software is used for simulation of all cases.

The results obtained show some important differences for the technologies. In steady-state, the

behaviour of HVDC-LCC and HVAC with compensation both onshore and offshore hold some

similarities, namely in the power factor and the power losses. In transient regime, the response of the

HVAC link to a fault in the grid is very different to the response of HVDC-LCC link. The HVAC link

allows the “fault ride through” of the offshore wind farm while the HVDC link is blocked, interrupting the

power transmission during the fault.

Keywords: Offshore Wind Farms, HVAC, HVDC-LCC, STATCOM, PSS/E.

iv

List of Figures

Figure 1 – EU power capacity mix: a) 2000 data; b) 2009 data [1]. ........................................................ 1

Figure 2 – Global Cumulative Installed Wind Capacity (1996-2009) [2]. ................................................ 2

Figure 3 – Annual and cumulative installed capacity (in MW) of offshore wind power in Europe (1991 –

2009) [1]................................................................................................................................................... 3

Figure 4 – Europe wind resources over open sea (Offshore Wind Atlas) [6]. ......................................... 8

Figure 5 – Cumulative share of installed capacity by country (end 2009) [1]. ...................................... 10

Figure 6 – Map of operational offshore wind farms in Europe (January 2010) (Adapted from [8]). ...... 11

Figure 7 – Offshore foundations and general characteristics [9]........................................................... 12

Figure 8 – Offshore wind turbine size evolution [9]. .............................................................................. 12

Figure 9 – General layout of an offshore wind farm [13]. ...................................................................... 16

Figure 10 – Layout of an HVAC wind farm[16]. ..................................................................................... 18

Figure 11 – Three-core XLPE submarine cable [17]. ............................................................................ 19

Figure 12 – Equivalent π-model for a submarine cable. ....................................................................... 20

Figure 13 – Maximum transmission lengths with inductive shunt compensation in both ends [17]. ..... 21

Figure 14 – (a) Schematic Diagram of an SVC; (b) an SVC in Radsted, Denmark [18]. ...................... 22

Figure 15 – (a) Schematic Diagram of a STATCOM; (b) a STATCOM installation [18]. ...................... 23

Figure 16 – Loading of an open-end cable due to capacitive charging current for different schemes of

reactive power compensation [19]. ........................................................................................................ 24

Figure 17 – Basic Configuration of a wind farm using an HVDC-LCC [7] Note: F=filter; HFF=High-

frequency filter. ...................................................................................................................................... 26

Figure 18 – Available HVDC Cables: a) OF cable; b) MI cable; c) XLPE cable [22]. ........................... 28

Figure 19 – Mass Impregnated HVDC Cable [22]. ................................................................................ 28

Figure 20 – Six-pulse Converter Bridge [23]. ........................................................................................ 29

Figure 21 – Diagram of a 12-pulse converter [23]. ................................................................................ 31

Figure 22 – HVDC configurations: a) Monopole; b) Bipolar. ................................................................. 32

Figure 23 – Basic Configuration of a wind farm using an HVDC-VSC [11]. .......................................... 34

Figure 24 – Single line diagram of the test grid used (with HVAC transmission, compensations at both

ends). ..................................................................................................................................................... 38

Figure 25 – Layout of the wind turbines in the Lillgrund wind farm [25]. ............................................... 40

Figure 26 – GE 1,5 MW voltage protection characteristics. .................................................................. 43

Figure 27 – PSS/E model of the offshore wind farm. ............................................................................ 44

Figure 28 – HVAC transmission layout in PSS/E. ................................................................................. 45

Figure 29 – HVDC-LCC transmission model in PSS/E. ........................................................................ 47

Figure 30 – STATCOM device in PSS/E. .............................................................................................. 49

Figure 31 – Legend of the values presented in the single-line diagrams. ............................................. 50

Figure 32 – Power Flow result for HVAC with offshore compensation. ................................................ 52

Figure 33 – Power Flow result for HVAC with onshore compensation. ................................................ 53

Figure 34 – Power Flow result for HVAC with offshore and onshore compensation. ........................... 54

v

Figure 35 – Power Flow result for HVAC with STATCOM compensation. ............................................ 55

Figure 36 – Power Flow result for HVDC-LCC transmission. ............................................................... 58

Figure 37 – Power Flow result for HVDC+STATCOM transmission. .................................................... 59

Figure 38 – Single line diagram of the grid used for the dynamic simulations. Note: Bus 3005, where

the fault occurs is marked in the orange rectangle. .............................................................................. 61

Figure 39 – Frequency variation of offshore buses for the Case 1 fault. .............................................. 63

Figure 40 – Voltage variation of offshore buses for the Case 1 fault. ................................................... 63

Figure 41 – Offshore wind turbine speed variation for the Case 1 fault. ............................................... 63

Figure 42 – Frequency variation of onshore buses for the Case 1 fault. .............................................. 63

Figure 43 – Voltage variation of onshore buses for the Case 1 fault. ................................................... 63

Figure 44 – Speed variation of machines on the onshore grid for the Case 1 fault. ............................. 63

Figure 45 – Frequency variation of offshore buses for the Case 2 fault. .............................................. 64

Figure 46 – Voltage variation of offshore buses for the Case 2 fault. ................................................... 64

Figure 47 – Offshore wind turbine speed variation for the Case 2 fault. ............................................... 64

Figure 48 – Frequency variation of onshore buses for the Case 2 fault. .............................................. 64

Figure 49 – Voltage variation of onshore buses for the Case 2 fault. ................................................... 64

Figure 50 – Speed variation of machines on the onshore grid for the Case 2 fault. ............................. 64

Figure 51– Frequency variation of offshore buses for the Case 3 fault. ............................................... 65

Figure 52– Voltage variation of offshore buses for the Case 3 fault. .................................................... 65

Figure 53– Offshore wind turbine speed variation for the Case 3 fault. ................................................ 65

Figure 54– Frequency variation of onshore buses for the Case 3 fault. ............................................... 65

Figure 55– Voltage variation of onshore buses for the Case 3 fault. .................................................... 65

Figure 56– Speed variation of machines on the onshore grid for the Case 3 fault. .............................. 65

Figure 57– Frequency variation of offshore buses for the Case 4 fault. ............................................... 66

Figure 58– Voltage variation of offshore buses for the Case 4 fault. .................................................... 66

Figure 59– Offshore wind turbine speed variation for the Case 4 fault. ................................................ 66

Figure 60– Frequency variation of onshore buses for the Case 4 fault. ............................................... 66

Figure 61– Voltage variation of onshore buses for the Case 4 fault. .................................................... 66

Figure 62– Speed variation of machines on the onshore grid for the Case 4 fault. .............................. 66

Figure 63 – Reactive Power injected/absorbed by the STATCOM on bus 20. ..................................... 66

Figure 64– Frequency variation of offshore buses with HVDC-LCC transmission. .............................. 69

Figure 65– Voltage variation of offshore buses with HVDC-LCC transmission. ................................... 69

Figure 66– Offshore wind turbine speed variation with HVDC-LCC transmission. ............................... 69

Figure 67– Frequency variation of onshore buses with HVDC-LCC transmission. .............................. 69

Figure 68– Voltage variation of onshore buses with HVDC-LCC transmission. ................................... 69

Figure 69– Speed variation of machines on the onshore grid with HVDC-LCC transmission. ............. 69

Figure 70 – Active Power in the converters, for the HVDC-LCC transmission. .................................... 70

Figure 71 – Reactive Power in the converters, for the HVDC-LCC transmission. ................................ 70

Figure 72– Frequency variation of offshore buses with HVDC-LCC+STATCOM transmission............ 71

Figure 73– Voltage variation of offshore buses with HVDC-LCC+STATCOM transmission. ............... 71

vi

Figure 74– Offshore wind turbine speed variation with HVDC-LCC+STATCOM transmission. ........... 71

Figure 75– Frequency variation of onshore buses with HVDC-LCC+STATCOM transmission............ 71

Figure 76– Voltage variation of onshore buses with HVDC-LCC+STATCOM transmission. ............... 71

Figure 77– Speed variation of machines on the onshore grid with HVDC-LCC+STATCOM

transmission. ......................................................................................................................................... 71

Figure 78 –Active Power in the converters, for HVDC-LCC+STATCOM transmission. ....................... 72

Figure 79 – Reactive Power in the converters, for HVDC-LCC+STATCOM transmission. .................. 72

Figure 80 – Reactive Power injected/absorbed by the STATCOM on bus 151. ................................... 72

vii

List of Tables

Table 1: Colour correspondence for each voltage level in Figure 24. ................................................... 39

Table 2: Components of the GE 1,5 MW Wind Turbine Model in PSS/E. ............................................ 42

Table 3: 150 kV cable parameters. ....................................................................................................... 45

Table 4: Onshore and Offshore transformer data [10]. ......................................................................... 46

Table 5: Equations for Rectifier and Inverter Parameters. .................................................................... 47

Table 6: Reactive power compensation alternatives. ............................................................................ 51

Table 7: HVAC selected results for different compensation alternatives. ............................................. 56

Table 8: HVDC-LCC and HVDC-LCC+STATCOM selected results. .................................................... 60

Table 9: HVAC Dynamic Cases Analysed. ........................................................................................... 62

Table 10: Reactance and Susceptance of the lines. ............................................................................. 80

Table 11: 400 kV line calculated parameters. ....................................................................................... 80

Table 12: Two-Winding Transformers Parameters. .............................................................................. 81

Table 13: Load powers. ......................................................................................................................... 81

Table 14: Shunt compensators. ............................................................................................................ 81

Table 15: Wind Turbine aggregate Power Flow Parameters. ............................................................... 82

Table 16: Unit transformer of one single wind turbine. .......................................................................... 82

Table 17: 33 kV Horns Rev cable parameters [10]. .............................................................................. 83

Table 18: 33 kV AC Cable parameters used in the modelled wind farm. ............................................. 83

Table 19: DC Line Power Flow Parameters. ......................................................................................... 83

Table 20: DC Link Converters Power Flow Parameters. ...................................................................... 84

Table 21: “CDC4T” model parameters for the DC Link. ........................................................................ 84

Table 22: STATCOM Power Flow parameters. ..................................................................................... 85

Table 23: “CSTCNT” model parameters for the STATCOM. ................................................................ 86

viii

Abbreviations

A Ampère

A Rotor Swept Area

ABB Swiss-Swedish High Tech Engineering Multinational

AC Alternating Current

B Susceptance

DC Direct Current

EEA European Environment Agency

EU European Union

EWEA European Wind Energy Association

F Farad

FACTS Flexible Alternating Current Transmission System

FP Power Factor

GE General Electric

GW Gigawatt

H Henry

HPFF High Pressure Fluid Filled

HPGF High Pressure Gas Filled

HVAC High Voltage Alternating Current

HVDC High Voltage Direct Current

Hz Hertz

Ic Charging current

Id Output DC current

IEEE Institute of Electrical and Electronics Engineers

IGBT Insulated Gate Bipolar Transistor

kA Kiloampère

km Kilometre

kV Kilovolt

kW Kilowatt

L Inductance

LCC Line Commutated Converters

LPFF Low Pressure Fluid Filled

LPOF Low Pressure Oil Filled

MI Mass Impregnated

MVA Mega Volt Ampère

Mvar Mega Volt Ampère Reactive

MW Megawatt

OF Oil Filled

PCC Point of Common Coupling

PSS/E Power System Simulator for Engineering

p.u. Per unit

Q Reactive Power

R Resistance

Rc Commutating Resistance

ix

rpm Revolutions per minute

Sb Base Apparent Power

SCFF Self Contained Fluid Filled

SCLF Self Contained Liquid Filled

SMES Superconducting Magnetic Energy Storage

STATCOM Static Synchronous Compensator

SVC Static VAr Compensator

TWh Terawatt hours

V Volt

Vb Base Voltage

VLN Line to Neutral rms Voltage

VSC Voltage Source Converters

u Wind Speed

Ud DC output voltage

X Reactance

XLPE Cross Linked Polyethylene

XT Commutating Reactance

Z0 Surface Roughness

Zb Base Impedance

α Firing Angle

β Advance Angle

γ Extinction Angle

ρ Air density

µ Delay Angle

Ω Ohm

ω Angular Speed

x

Table of Contents

1. Introduction ........................................................................................................ 1

1.1. Motivations ............................................................................................................................. 1

1.2. Thesis Objectives and Outlook ............................................................................................ 3

1.3. Thesis Structure .................................................................................................................... 4

2. Wind Energy Offshore ....................................................................................... 6

2.1. Introduction ............................................................................................................................ 6

2.2. The Wind Resource ............................................................................................................... 6

2.2.1. Wind Offshore .................................................................................................................. 7

2.3. Offshore Wind Farms ............................................................................................................ 9

2.3.1. Current Status ................................................................................................................ 10

2.3.2. Technologies and Future Trends ................................................................................... 11

2.4. Grid Integration .................................................................................................................... 12

2.4.1. Transmission Technologies ........................................................................................... 13

2.4.2. Grid Connection Requirements ..................................................................................... 14

3. Transmission Technologies for Offshore Wind Farms ................................. 16

3.1. Introduction .......................................................................................................................... 16

3.2. HVAC Transmission ............................................................................................................ 17

3.2.1. General Aspects ............................................................................................................ 17

3.2.2. Topology and Main Components................................................................................... 17

3.2.3. AC Submarine Cables ................................................................................................... 18

3.2.4. Reactive Power Compensation ..................................................................................... 21

3.2.4.1. FACTS Devices ..................................................................................................... 22

3.2.4.2. Distribution of the Compensation .......................................................................... 23

3.3. HVDC-LCC Transmission ................................................................................................... 25

3.3.1. General Aspects ............................................................................................................ 25

3.3.2. Topology and Main Components................................................................................... 25

3.3.3. DC Submarine Cables ................................................................................................... 27

3.3.4. Converter Technology ................................................................................................... 29

3.3.4.1. 6-pulse Bridge Converter ....................................................................................... 29

3.3.4.2. 12-Pulse Bridge Converter .................................................................................... 31

3.3.4.3. HVDC Schemes ..................................................................................................... 31

3.3.5. HVDC-LCC + STATCOM .............................................................................................. 32

3.4. HVDC-VSC Transmission ................................................................................................... 34

xi

4. PSS/E Modelling ............................................................................................... 35

4.1. Introduction .......................................................................................................................... 35

4.2. Grid Structure ...................................................................................................................... 35

4.2.1. Lines .............................................................................................................................. 36

4.2.1.1. 400 kV Lines .......................................................................................................... 36

4.2.2. Transformers ................................................................................................................. 37

4.2.3. Loads and Shunt Compensators ................................................................................... 37

4.2.4. Diagram of the Grid ....................................................................................................... 38

4.3. Offshore Wind Farm Model ................................................................................................. 40

4.3.1. Wind Turbines ................................................................................................................ 41

4.3.1.1. Steady-State Model ............................................................................................... 41

4.3.1.2. Transient Model ..................................................................................................... 41

4.3.1.3. Fault Ride Through Capability ............................................................................... 42

4.3.2. Step-Up Transformers ................................................................................................... 43

4.3.3. 33 kV Cables ................................................................................................................. 43

4.3.4. Layout of the offshore wind farm ................................................................................... 43

4.4. HVAC Transmission ............................................................................................................ 45

4.4.1. AC Submarine Power Cable .......................................................................................... 45

4.4.2. Onshore and Offshore Transformers............................................................................. 46

4.5. HVDC-LCC Transmission ................................................................................................... 47

4.5.1. Steady-State Model ....................................................................................................... 47

4.5.2. Transient Model ............................................................................................................. 48

4.6. STATCOM ............................................................................................................................. 48

4.6.1. Steady-State Model ....................................................................................................... 49

4.6.2. Transient Model ............................................................................................................. 49

5. Power Flow Results ......................................................................................... 50

5.1. Introduction .......................................................................................................................... 50

5.2. HVAC Power Flow ............................................................................................................... 50

5.2.1. Offshore Compensation ................................................................................................. 52

5.2.2. Onshore Compensation ................................................................................................. 53

5.2.3. Offshore and Onshore Compensation ........................................................................... 54

5.2.4. STATCOM Compensation ............................................................................................. 55

5.2.5. Discussion ..................................................................................................................... 56

5.3. HVDC Power Flow ............................................................................................................... 58

5.3.1. HVDC Configuration ...................................................................................................... 58

5.3.2. HVDC+STATCOM Configuration .................................................................................. 59

5.3.3. Discussion ..................................................................................................................... 60

xii

6. Dynamic Results .............................................................................................. 61

6.1. Introduction .......................................................................................................................... 61

6.2. HVAC Dynamic Behaviour .................................................................................................. 62

6.2.1. Case 1 results ................................................................................................................ 63

6.2.2. Case 2 results ................................................................................................................ 64

6.2.3. Case 3 results ................................................................................................................ 65

6.2.4. Case 4 results ................................................................................................................ 66

6.2.5. Discussion ..................................................................................................................... 67

6.3. HVDC Dynamic Behaviour .................................................................................................. 69

6.3.1. HVDC Configuration ...................................................................................................... 69

6.3.2. HVDC + STATCOM Configuration ................................................................................ 71

6.3.3. Discussion ..................................................................................................................... 73

7. Conclusions ..................................................................................................... 75

7.1. Final Remarks ...................................................................................................................... 75

7.2. Future Work.......................................................................................................................... 76

References .............................................................................................................. 78

Appendix A – Grid Parameters .............................................................................. 80

Appendix B – GE 1,5 MW Wind turbine aggregate “dyr” record file .................. 87

1

1. Introduction

This chapter presents an overview of some general aspects regarding wind energy and, more

specifically, offshore wind energy. The motivation that led to the execution of this thesis and the main

objectives of the work are also presented.

1.1. Motivations

For a long time, conventional energy sources such as coal, oil, natural gas, hydro power and

nuclear power were used as main fuels for the production of electricity. The energy crisis that affected

the world in the 1970s and 1980s led to major changes in the hitherto energy paradigm. Reducing the

world’s dependency of fossil fuels became a priority, as the finiteness of these resources was found to

be truly concerning. The need to guarantee diversity and reliability of energy sources, together with

the fast growing concern for environmental issues also contributed to a greater interest in alternative

energy sources.

The need for a long-lasting solution to fulfil the growing worldwide energy demand, a solution that

is both environmentally friendly and economically safe was found in renewable energies. In fact,

renewable energy sources such as wind, solar photovoltaic and hydro power, to name a few, have

suffered significant increases in recent years. The installed power over the last ten years has

increased in such way that renewables account for more than 50% of new installations in the EU,

cementing a rising trend initiated over a decade ago [1]. Such data can be verified in Figure 1, where

the share of total installed power in the EU in 2000 and in 2009 is presented.

Figure 1 – EU power capacity mix: a) 2000 data; b) 2009 data [1].

In Figure 1, wind power clearly stands out as the renewable source that experienced the largest

increase in installed power: wind power’s share of total installed capacity has increased from 2% in

2000 to 9% in 2009. These facts reflect a global trend on new power installations, as 2009 was the

2

second year running where more wind power was installed than any other generating technology in

the EU [1]. Wind energy is considered one of the most promising renewable energy sources,

benefiting from a mature technology, developed especially in Europe and in the USA. It is now a totally

established technology that has increased remarkably in recent years. Wind power is, in fact, the

fastest growing renewable energy: since 1996 the installed wind capacity increased from 6100 MW to

158505 MW as it is shown in Figure 2.

Figure 2 – Global Cumulative Installed Wind Capacity (1996-2009) [2].

The fast development of wind energy throughout the world has opened up new frontiers in wind

energy generation in the form of offshore wind farms. Placing wind turbines at sea became a

possibility, as many of the windiest locations onshore are already occupied. Higher wind speeds at

marine locations and wider installation areas, together with the possibility of moving wind turbines

away from population – to which wind turbines may cause some discomfort – were the main reasons

driving the investment on offshore wind energy.

The Vindeby wind farm in the Baltic Sea, off the coast of Denmark, was the first offshore wind

farm in the world. It was built in 1992 and it consists of eleven 450 kW wind turbines. Since then, and

in particular in the last 7/8 years, offshore wind power has experienced an enormous growth, with the

installation of several new offshore wind farms. A graph representing the development of the offshore

wind power installed capacity from 1991 to 2009 can be seen in Figure 3, in which the referred growth

can be observed.

3

Figure 3 – Annual and cumulative installed capacity (in MW) of offshore wind power in Europe (1991 – 2009) [1].

The current trends for offshore wind farms indicate new offshore wind farms will be higher rated

(as a result of larger wind turbines and a higher number of wind turbines) and placed further from the

sea. The power transmission of large amounts of power over long distances, from the wind farm to the

onshore grid, makes the transmission system a key feature of the offshore wind farm installation.

Currently available technologies for the transmission system to shore are high voltage alternating

current (HVAC) and high voltage direct current (HVDC). For HVDC connections, there are two

technical options: line commutated converter based HVDC (known as HVDC-LCC) and voltage source

converter based HVDC (HVDC-VSC). The HVDC alternatives differ in the technology used for the

power converters.

The technology used for the interconnection of offshore wind farms influences in a very significant

manner the whole operation of the system. As so, the main characteristics of operation for each

transmission system have to be assessed and thoroughly studied, including steady-state analysis and

dynamic behaviour analysis of the link (when disturbances occur), making this an important field of

investigation in the near future.

1.2. Thesis Objectives and Outlook

The main objective of this thesis is to compare the performance in steady-state and the transient

behaviour of two alternative transmission systems for offshore wind farms: HVAC technology and

HVDC-LCC technology.

In steady-state, the power flow analysis of a grid with an integrated offshore wind farm is

performed, for both transmission systems. Special attention is paid to both the power factor and the

reactive power flow in the point of connection of the wind farm to the onshore grid. For the HVAC

technology, compensation schemes for the reactive power generated by the submarine cable are also

4

evaluated. Compensation with shunt reactors placed onshore, offshore and at both ends of the cable

and also the use of a STATCOM are subject to analysis based on simulations. For HVDC-LCC

technology, the configuration with a STATCOM as a way of compensating reactive power

consumption of the converters is also analysed.

The transient behaviour of the wind farm, and particularly of the connection (AC or DC Link), is

investigated. The dynamic behaviour of the system, as a response to onshore grid faults, is tested

through simulation as different faults are applied to the grid. The influence of the STATCOM is also

assessed.

All the analysis are done considering simulation results obtained after careful design,

dimensioning and testing of the offshore grid performed in PSS/E software.

1.3. Thesis Structure

The content of this thesis is divided in seven chapters.

The present chapter, chapter one, is a general introduction to the dissertation, providing

information on the current status of offshore wind energy and its integration among the global energy

panorama. Motivations, objectives and structure of the work developed in the course of the thesis are

also outlined.

Chapter two presents an overview of the most important issues regarding offshore wind farms.

The main characteristics of offshore wind farms, main advantages and constraints associated with

these wind farms, current status and grid integration issues are also approached.

Chapter three provides a closer look on the transmission system used for the connection of

offshore wind farms with the onshore grid. The chapter presents a detailed description of the three

available options: High Voltage Alternating Current Transmission (HVAC); High Voltage Direct Current

based on Line Commutated Converter Transmission (HVDC-LCC) and High Voltage Direct Current

based on Voltage-Source Converter Transmission (HVDC-VSC). Main features, operating issues and

limitations of each technology are discussed, with special focus on the HVAC and the HVDC-LCC

alternatives.

Chapter four focuses on the modelling carried out in the PSS/E software of the offshore wind farm

studied. The models used and the parameters selected for each model, for both the AC and the DC

transmission, are described with sufficient detail. Differences in modelling for steady-state analysis

and for transient behaviour analysis of the grid are also summarized.

Chapter five presents the results of the power flow simulation of the offshore wind farm. For

HVAC transmission, some alternatives for reactive power compensation are reviewed. For HVDC

transmission, the possible benefits of a STATCOM are also studied. A comparative analysis of the

transmission alternatives in steady-state is also performed.

5

Chapter six presents the results obtained for the dynamic simulation of the wind farm. Dynamic

behaviour of the HVAC or HVDC-LCC transmissions to different grid faults on the onshore grid is

analysed. The influence of the STATCOM on the dynamic behaviour of the link is studied.

In chapter seven, the main conclusions of the work are presented, together with a reference to

future studies that may offer important future contributions to the theme of this thesis.

6

2. Wind Energy Offshore

This chapter will present an overview of some general aspects of wind farms offshore: reasons

and motivations for building offshore, their advantages and disadvantages, grid integration and

transmissions options to shore.

2.1. Introduction

The increasing requirement to produce electrical power from renewable and clean energy

sources and, particularly, the major investment in wind energy made over the last 15 years, led to a

growing interest in harnessing the wind power that lies off the coast of many countries. Offshore wind

power became, therefore, the next evolution in wind power technology. A number of reasons gave rise

to this bulk investment. The offshore wind characteristics lead to a larger produced energy but the

difficult conditions at sea, the integration of the wind farm and other technical requirements have to be

seriously considered. It is however relevant to notice that after years at the starting block, the offshore

wind energy is taking off and becoming a viable alternative for energy production.

2.2. The Wind Resource

Winds are caused by pressure differences across the earth’s surface due to differential solar

heating of the earth’s atmosphere. Therefore, wind energy is strongly influenced by solar energy: the

amount of solar radiation absorbed at the earth’s surface is greater at the equator then at the poles,

resulting in faster heating of the air in the equator [3]. This warm air, heated in the equatorian regions,

rises to high altitudes and then flows toward the poles. At about 30ºN and 30ºS, the air begins to cool

and sink and so a return flow of air takes place in the lower layers of the atmosphere. Zones of high

pressure are created by descending air, as zones of low pressure are formed where air is ascending.

This horizontal gradient drives the flow of air from high to low pressure. The greater the pressure

gradient, the greater is the force on the air and the higher is the wind speed [4].

Stronger, constant and more persistent winds occur at about 10 km above the earth’s surface.

However, as placing turbines at such high altitudes is unfeasible, the area of interest is limited to only

a few tens of meters high [3].

In addition to this global phenomenon resulting in wind, there are also local effects to be

considered. The nature of the terrain has an important effect. Wind turbines utilise wind energy from a

lower area known as boundary layer. In this region, wind speed is retarded by frictional forces on the

earth’s surface, which means that wind speed increases with height (this gradient is known as wind

shear) [4].

7

2.2.1. Wind Offshore

As stated previously, the nature of the surface influences the wind profile at a given region.

Offshore, the smoothness of the surface of the ocean results in a low surface roughness (a parameter

referred to as Z0). The surface roughness length is dependent on the sea state, increasing with the

local wave conditions which are, in turn, influenced by wind itself. As a result of a low surface

roughness there is also low turbulence and wind shear on a marine environment. Offshore locations

are in fact windier and the winds are more persistent than those at an onshore location. These effects

increase with distance from shore in the downwind direction, making it more appealing to locate

offshore wind farms further from land [5].

Data collected from offshore meteorological stations has been compiled to create the wind

offshore atlas of Europe, which is presented in Figure 4. The mean wind speed at each location is

given for five different heights: 10, 25, 50, 100 and 200 m from sea level. The highest values of wind

speed are for the 200 m height, making it the most economically interesting solution for placing

offshore wind turbines.

The geographical distribution of wind speed offshore, given by Figure 4, shows that, apart from a

small region in southern France, the North Sea is the windiest location in Europe. In fact, for this

reason (and also for the fact that water depth in the North Sea increases slowly with distance from

shore) most of the wind farms in Europe are located in the North Sea, as part of solid investments by

English and Danish governments in offshore farms.

8

Figure 4 – Europe wind resources over open sea (Offshore Wind Atlas) [6].

As the power that can be extracted by a wind turbine is proportional to the cube of the wind speed

(as given by Equation (1)), the extracted power is highly dependent on the wind speed [3].

=1

2 (1)

Therefore, and given the fact that wind speed offshore is potentially higher than onshore, a 10%

increase in wind speed results in an approximate 30% increase in power production, making wind

power offshore a very interesting option.

According to the European Environment Agency (EEA) the technical potential of offshore wind by

2020 is around 25000TWh, six to seven times greater than the projected electricity demand [1]. These

figures suggest that a significant wind resource lies off the coast of many countries. The data

presented refers to European estimations but some studies done elsewhere point out similar offshore

potential in other regions of the world.

9

2.3. Offshore Wind Farms

Following the rising interest in the wind resource offshore, a way of harnessing the wind was

considered and steps were taken towards creating platforms located offshore, in order to gain

experience and test problems and potential benefits of this technology. The first offshore wind turbine

was then installed in Sweden in 1991. The following year, the Vindeby wind farm, the first of its kind,

was installed in shallow water (2-5 m deep) off the coast of Denmark. It consists of eleven 450 kW

machines, located about 3 km from the shore [5].

Small wind farms, consisting of few wind turbines (less than 10) with low ratings (around 500 kW)

were erected in other European countries but it wasn’t until the 2000s that serious investment was put

into the design and construction of offshore wind farms. Denmark has led the way, building the Horns

Rev wind farm in 2002, the largest at the time of construction and a reference for following projects.

Many new offshore wind farms have followed, and the remarkable growth in new installations

over the last 7/8 years (as can be seen in Figure 3), denotes that, in fact, offshore wind energy has

gain relevance as an alternative to consider for energy production.

The fact that offshore wind farms are located at sea offers several advantages but also poses

major challenges for potential investors. Some relevant constraints associated with the offshore wind

farms that need to be considered are the following:

• Installation and maintenance of a large structure located far from shore is a complex

procedure, given the distance and the need for special ships;

• Accessibility of turbines may prove difficult, as harsh weather conditions and extreme

waves are likely to exist at sea;

• As a consequence, materials used for wind turbines and additional equipment need to be

resistant to physical impact as well as to salt corrosion (when the wind farm is placed in

salt water);

• Overall, at present time, the construction and maintenance of offshore wind energy is

more expensive than on shore, given the higher costs of foundation, installation and

electrical connection of wind farms [4].

However, as the number and size of offshore wind farms increase, technology is likely to advance

to deal with these problems and reduce transportation and installation costs. The advantages will

outweigh the disadvantages making offshore wind farms a reliable investment [4].

Advantages of offshore wind power are as follows:

• Wind speeds are higher offshore than onshore, which means that a large potential for

power production lies offshore. In fact, the annual average full-load hours offshore is

between 3500-4000 hours, while for onshore this figure is between 1500 and 3000 (the

average being around 2000 hours) [7];

• Winds at marine locations are more persistent and less turbulent than those on land;

10

• Due to the proliferation of wind parks onshore, some countries are running out of suitable

onshore locations, as the windier sites are already occupied;

• A wider area for installation of wind turbines is also likely to be available at sea,

compared to land;

• Wind turbines that might be intrusive and cause visual and noise impact on land might be

acceptable if sited away from shore, far from population.

2.3.1. Current Status

As of January 2010 all offshore wind farms were located in Europe. According to EWEA data [1],

there are 38 operational offshore wind farms in Europe, representing a total capacity of 2055,9 MW of

installed capacity. Denmark has been a pioneer and a worldwide reference as far as offshore wind

farms are concerned. For the last few years, however, the UK became the number 1 country in

offshore installed capacity, with a total of 882,8 MW, representing almost 43% of the European total.

Complete division of installed capacity by country is presented in Figure 5.

Figure 5 – Cumulative share of installed capacity by country (end 2009) [1].

As stated previously, the North Sea is the windiest region in Europe. The relatively low water

depth (less than 20 m) of the sea has also contributed to the widespread installation of offshore wind

farms in the region. Figure 6 shows precisely that, as can be seen that most of the operational

offshore wind farms are at the North Sea.

11

Figure 6 – Map of operational offshore wind farms in Europe (January 2010) (Adapted from [8]).

Offshore wind energy is a growing market, and a significant number of offshore wind farms are

either under construction or planned for the next 10 years. Around 40 GW of installed capacity are

expected to exist by 2020 [1].

2.3.2. Technologies and Future Trends

The trends and perspective for future offshore wind farms are likely to demand new technology,

creating new challenges for the industry. As the technology develops and experience is gained, future

projects will be higher rated and placed in deeper waters, further from shore. Looking at the wind

farms proposed by developers, the industry will go beyond the so-called 20:20 envelope (20m water

depth, 20 km from shore) in which current wind farms are included.

Some experience gained from wind farm projects in land is useful but offshore wind farms differ

significantly from the onshore ones.

Submarine support structures are a very important part of the offshore project, as they represent

a significant proportion of offshore development costs. The choice of structure is determined

essentially by water depth and consistency of seabed. The currently available options for foundation

are represented in Figure 7, the more commonly used being the monopole and gravity foundations.

For deepwater application some prototypes of floating foundations are under development, with the

first experience being the Floating Hywind in Norway, at a depth of 220 m. These floating devices will

be important as the wind farms move further from shore.

12

Figure 7 – Offshore foundations and general characteristics [9].

Another issue of concern regarding the technology used in offshore wind farms is the wind

turbines used. The wind profile offshore (less turbulent and with higher average and extreme wind

speeds than onshore) along with challenging access conditions demand special requirements for the

design and manufacture of wind turbines. These include robustness, corrosion protection, high

reliability and low maintenance requirements.

Recent trends for wind turbines indicate that future offshore wind turbines will be larger machines,

with higher rotor diameter and increasingly higher capacity, as indicated in Figure 8. Planned wind

farms are expected to include turbines with rated power as high as 5 MW.

Figure 8 – Offshore wind turbine size evolution [9].

2.4. Grid Integration

Integrating an offshore wind farm in an electrical network poses a significant challenge to the grid.

The impact varies with the strength of the grid and the size of the wind farm. As the offshore wind

capacity grows, grid integration issues may arise, as increasingly large amounts of electricity are fed

into networks, either in distribution or transmission systems, at points not specifically designed for

13

such infeeds [4]. As such, the transmission technology – AC or DC – used to link the wind farm to the

onshore grid is of key relevance to meet the grid requirements of the network.

An overview of the available technologies is presented as well as a description of the main

considerations of connecting an offshore wind to the grid.

2.4.1. Transmission Technologies

The installation of offshore wind farms has given rise to new challenges regarding electrical

connections, both for the internal electric system of the wind farm and for its connection to the main

grid. Existing wind farms onshore are connected via AC cables or overhead lines to the national grids.

An AC network within the farm collects the power production of each wind turbine. The voltage level of

generation is typically around 700V, which is then stepped up to a medium voltage level (typically 33

kV) by a transformer installed in the nacelle or the tower base [4].

The same scheme was applied to the first wind farms installed offshore. However, as wind farms

tend to be larger, with longer distances between turbines and longer distance to shore, alternative

arrangements were studied. Special focus is, then, on the transmission from the wind farm to shore.

Currently available technologies for transmission system to shore are high voltage alternating current

(HVAC) and high voltage direct current (HVDC). For HVDC connections, there are two technical

options: line commutated converter based HVDC (known as HVDC-LCC) and voltage source

converter based HVDC (HVDC-VSC).

The AC connection has been used in all offshore wind farms. With the increase in the size of wind

farms, the transmission to shore is performed at higher voltage levels than the 33 kV used in

collection, usual values being 132 kV (used in the Nysted wind farm, in Denmark) or 150 kV (as is the

case of the Horn Rev wind farm). The HVAC solution is indeed the most straightforward technical

approach as it features some interesting advantages such as ease of interconnection, installation and

maintenance; operational reliability and cost effectiveness. Some disadvantages include the limits for

transmitted power, associated to the charging current on AC cables, which will limit the maximum

distance of transmission. Some scheme of reactive power compensation must then be used in AC

cable systems.

HVDC solutions might prove a reliable option for large wind farms further from shore as this

technology has some very promising characteristics. One of the most important is the ability to

transmit large amounts of power over long distances with lower losses than HVAC. Since the

transmission is done using DC cables, the transmission distance is not affected by cable charging

current. Main disadvantages include the complexity and associated cost of installation and

maintenance of the HVDC system, which may well be overcome in the future, due to economies of

scale.

In the line commutated based HVDC the conversion is performed with thyristor converters. The

HVDC-LCC is a totally established technology, as the first HVDC link dates back to 1954 in Gotland,

14

Sweden. Such proven track record has become one of the most important advantages of this option,

in addition to the features referred in the previous paragraph.

The HVDC-VSC transmission is a relatively recent technology that uses insulated gate bipolar

transistors (IGBT) converters for the inversion/rectifier operations. This technology was made possible

by advances in high-power electronics and is used as a viable alternative to thyristor based LCCs

used in “Classic HVDC”. Unlike line-commutated converters, VSCs do not need an AC source

commutation voltage; the independent control of both active and reactive power and the reduced

injection of harmonics also represent important advantages of HVDC-VSC over HVDC-LCC [4].

More detailed description of the working principles of each of the referred technologies – with

special emphasis on HVAC and HVDC-LCC – is presented in Chapter 3.

2.4.2. Grid Connection Requirements

Until some years ago wind farms were allowed, or even required, to disconnect from the grid

during a disturbance in the grid [10]. This has changed significantly, specially due to the proliferation of

large amounts of installed wind power capacity and, predictably in a near future, installed capacity

offshore. The disconnection of a large offshore wind farm would result in a significant loss of

generation that could cause some stability problems to the network. Transmission system operators

require nowadays for offshore wind farms to stay connected under certain disturbances in the grid.

These requirements are known as the fault ride through capability of the wind farm and are generally

regulated in grid codes. As established in most grid codes, only under certain circumstances shall

wind farms be disconnected from the grid following a grid fault, remaining otherwise connected in

order to assist in the stabilization of the grid frequency or the voltage during fault, providing voltage

back-up.

Apart from the fault ride through capability, other technical requirements must be fulfilled by the

wind farm, since the increasing size of offshore wind farms means that the rating of such installations

will be comparable to that of traditional generating plants on the grid. These requirements include [11]

[12]:

• Control of active and reactive power (operation under a specified range for power factor);

• Frequency range (with time durations for extreme conditions, permissible reduction at

frequency extremes, if any);

• Contribution to network stability;

• AC voltage control capability.

As the proliferation of offshore wind power increases, wind farms will be bound to meet these

demands, which may prove difficult depending, to great extent, on the transmission system used

between the wind farm and shore. The charging currents affecting AC cables represent a limitation for

15

the HVAC option and so some form of compensating the surplus reactive power generated by the

cable may prove necessary to met grid requirements. As far as HVDC is concerned, capacitive

currents are nonexistent but other issues are of concern, such as the behaviour of the DC link

following an onshore grid fault and the power factor at the connection point in HVDC-LCC, as thyristor

based converters absorb reactive power. FACTS devices can be used to reduce or eliminate these

problems, therefore enhancing the performance of the HVDC system.

Further aspects of the issues concerning connection, both in HVAC and in HVDC-LCC will be

discussed in Chapter 3.

16

3. Transmission Technologies for Offshore Wind Farms

This chapter provides an overview of the available transmission systems for offshore wind farms.

Several aspects of the connection in either AC or DC are discussed.

3.1. Introduction

One of the most important aspects of the design of offshore wind farms is the connection to the

electrical grid located in land. The electrical energy generated in the offshore wind farm requires one

or more submarine cables to transmit the power to the onshore utility grid that services the end-users.

Depending on the wind farm size and distance from shore, the connection poses a challenge, not

only to the wind farm developer, but also to the onshore network itself, as some grid code

requirements must be met.

There are currently three different transmission technologies available:

• HVAC – High Voltage Alternate Current;

• HVDC-LCC – High Voltage Direct Current with Line Commutated Converters;

• HVDC-VSC – High Voltage Direct Current with Voltage Source Converters;

These technologies present significant differences, on the operation mode and the components

used. A detailed explanation of the options is presented throughout this text, in 3.2, 3.3 and 3.4

Special focus is given to the HVAC and HVDC-LCC technologies.

The general layout of an offshore wind farm is represented in Figure 9. This scheme can refer to

both the HVAC and the HVDC transmission. The components present at the “collecting point” and at

the “wind farm grid interface” vary according to the technology used to connect the generation to the

PCC (Point of Common Coupling) with the grid.

Figure 9 – General layout of an offshore wind farm [13].

17

As of 2010 all operating offshore wind farms are connected via AC cables. However, as wind

farms become larger and are placed further from the coast, the use of HVDC solutions will become

technically and economically interesting alternatives.

3.2. HVAC Transmission

3.2.1. General Aspects

Connecting the wind farm to the grid by an AC cable is the most straightforward technical

solution, as both the power generated by the wind farm and the onshore transmission grid are AC. In

fact, all the operational offshore wind farms use HVAC for the connection link to shore. The HVAC

transmission offers some advantages over the DC solutions such as [11] [14]:

• Proven and low-cost technology;

• Easy to integrate in existing power systems;

• Low losses over small distances.

On the other hand there are some constraints of the HVAC system that limit significantly the use

of this technology, namely [11] [14]:

• There is an excessive amount of reactive power produced in the AC submarines cables;

• Increase in the cable length means increase in its capacitance which results in a reactive

power increase, resulting in a transmission distance limit for AC systems;

• Necessary use of compensation systems (shunt reactors, STATCOMS, SVC, etc) at the

ends of the cable;

• Load losses are significantly higher for longer distances;

• For large wind farms several cables may be necessary, increasing line losses.

3.2.2. Topology and Main Components

A transmission system based on HVAC technology includes the following main components [13]

[15]:

• AC based collector system within the wind farm;

• Three core XLPE HVAC submarine transmission cable(s);

• Offshore transformer(s);

• Reactive power compensation (onshore and/or offshore);

18

• Onshore transformer;

Figure 10 illustrates the layout of an HVAC based offshore wind farm, with the components

mentioned above.

Figure 10 – Layout of an HVAC wind farm[16].

The voltage used in the collecting system in the wind farm is usually medium voltage, in the range

of 33-36 kV. For relatively small wind farms, with short distances to shore, the offshore transformer

may not be necessary, so the link is done at the same voltage used in the wind farm.

3.2.3. AC Submarine Cables

The submarine cable used in the transmission system is one of the most important components

of the system, as the cost of the cable represents a large fraction of the total cost of the offshore wind

farm investment. The laying of the cable is a complex process performed with special laying vessels,

so as to guarantee the full operation of the system. The potentially severe conditions at sea and other

hazards, such as fishing or dropped objects, also demand special attention on the mechanical

resistance of the cable. Technically, different cable types have different characteristics that influence

the transmission and can even limit the maximum distance of transmission. A submarine cable has

typically the following structure [14]:

• Conductor core, typically copper or aluminium;

• Electrical Insulation, either solid dielectric (XLPE cables) or oil impregnated paper (OIP);

• Shielding, a conductor layer of paper or polymer;

• Sheathing, metallic layer used to ground the cable and also protect it from water;

• Armour, an outer metallic armature, to increase mechanical resistance;

• Optic Fibre, for communications;

The link can be achieved with either three single-core cables or one three-core cable (more than

one may be used, depending on the transmitted power). The latter is the most common type, as it

19

represents lower cable and installation costs and lower electromagnetic fields and induced current

loss, compared with separate cables.

The main types of AC cables differ in the electrical insulation used in the cable. Historically,

cables used lapped paper insulation impregnated with insulating oil. These cables are:

• High-pressure pipe-type, either fluid-filled (HPFF) or gas filled (HPGF);

• Low-pressure oil-filled (LPOF), also referred as self-contained liquid-filled (SCLF) and

Low-pressure fluid-filled (LPFF);

The need of a pumping system in both ends of the lines to circulate the fluid and the

environmental risk of pipe rupture, resulting in oil spill are serious disadvantages of these cables. As

so, these cables are only considered for very short distances.

Cross-linked polyethylene cable (XLPE) cables are currently the most cost-effective and modern

alternative. The insulation is made of solid dielectric (also called extruded dielectric or polymeric

insulated) and it presents several advantages over the fluid-filled options: higher mechanical

resistance, lower weight and, consequently, easier installation process [14]. Furthermore, XLPE

cables have lower capacitance and lower dielectric losses than the fluid-filled cables.

One three-core XLPE cable is shown in Figure 11, where the structure described above can be

seen.

Figure 11 – Three-core XLPE submarine cable [17].

The major problem concerning the connection of wind farms with AC submarine cables is the fact

that cables generate significant amounts of reactive power. This reactive power is produced by the

high shunt capacitance of cables (significantly higher than overhead lines). In the AC system, the

cable must carry the load current and the reactive current generated by the cable capacitance, which

reduces the power rating of the cable.

Considering the cable is modelled by an equivalent π-model, the submarine cable can be

represented by the scheme shown in Figure 12. The charging current at each end are represented by

and .

20

Figure 12 – Equivalent π-model for a submarine cable.

Neglecting the voltage drop across the resistance and the reactance of the cable, one can

consider that = = . As so, the total charging current affecting the cable is the sum of and

(as represented in Figure 12) and is represented by , which is given by Equation (2):

= ∙ ∙ (2)

In the equation,

• is the angular frequency (rad/s),

• C is the capacitance (F)

• V is the terminal voltage (V).

From Equation (2) it is clear that, as the cable system voltage is increased to minimize losses, the

charging currents also increase, worsening the situation.

The reactive power generated by the shunt capacitance is then given by Equation (3):

= − ∙ ∙ (3)

Note that the negative signal means the capacitance supplies reactive power to the grid. From

Equation (3) it can be seen that the reactive power generated by the cable is a function of the

capacitance and the voltage, being particularly dependent on the voltage (since it is squared).

Since the cable capacitance is distributed along the entire length of the cable, the longer the

cable the higher the capacitance and the resulting generated reactive power. As a consequence, the

load carrying capability of the cable is reduced, which means that a cable can only transmit a certain

amount of power for a given distance. This results in an obvious limitation for the length of AC links.

The maximum rating for a three-core submarine cable with a voltage rating of 150-170 kV is limited to

about 200MW, with compensation at both ends, and a maximum cable length of 200 km [11].

A comparison of the transmission capacity of different cables operated at certain voltage levels

(132, 150, 230 and 400 kV) is presented in Figure 13. The maximum distances for an AC submarine

cable system varies according to the rated voltage for the link and the maximum transmitted power

decreases with the length of the cable.

21

Figure 13 – Maximum transmission lengths with inductive shunt compensation in both ends [17].

The transmitting distance represents an obvious limitation for the HVAC system using a

submarine cable. Compensating the reactive power produced by the cable can be a solution since it

increases significantly the maximum distance of transmission. Reactive power compensation options

are discussed in 3.3.4.

The capacitance of AC cables may also lead to other undesirable effects, like overvoltages, high

harmonic currents and resonances between the cable’s capacitance and the reactance of the

generators. Real power losses within the cable also limit the distance for HVAC cable transmission.

These losses in the submarine cables derive from dielectric losses (relatively small) and ohmic losses

in the conductors (the higher component of losses), the metallic shield and the steel wire armour [14].

3.2.4. Reactive Power Compensation

The solution for the large amounts of reactive power at the cable is to compensate the reactive

power produced by absorbing reactive power, thus reducing the additional losses and increasing the

maximum transmitting distance. The compensation is usually realized by fixed or electronically

controlled shunt reactors. The fixed shunt reactor is the simplest device but the progress in FACTS

(Flexible AC transmission system) devices, such as SVC (Static VAr compensator) or STATCOM

(Static Synchronous Compensator), considerably extends the reactive power and voltage control

possibilities offered by the switched shunt reactors.

Shunt reactors are the most commonly used devices as they represent simple and robust

solutions with low installation costs. They also have the advantage of requiring no transformer for the

connection, thus having no additional power losses. One of the disadvantages is that the reactors are

22

designed for a single operational mode, usually to compensate the cable at full load. Another

disadvantage is the fact that the reactive power absorbed by the shunt reactor is proportional to the

square of the terminal voltage, as demonstrated by Equation (4), where L stands for the inductance of

the reactor.

=

(4)

3.2.4.1. FACTS Devices

The SVC and the STATCOM are part of the FACTS device family, used for voltage regulation

and power system stabilization, based on power electronics. These devices are capable of both

generating and absorbing reactive power. The flexibility of use is, therefore, the main advantage of

these equipments, since they allow the continuous variable reactive power absorption (or supply). The

reactive power is not proportional to the voltage at the bus, also another advantage of these

equipments. The FACTS devices can also contribute in the improvement of the voltage stability and

the recovery from network faults.

The similarity of the SVC and STATCOM devices led to them being sometimes referred generally

as “Static VAr Compensators”. These are, however, different equipments. The SVC (“SVC Classic” for

some manufacturers) is based on conventional capacitor banks together with parallel thyristor

controlled inductive branches. These inductive branches can either be TCR (Thyristor Controlled

Reactor), used for linear injection of reactive power or TSC (Thyristor Switched Capacitor), used for

stepwise injection of reactive power. A SVC device is represented in Figure 14, where a linear

diagram and an SVC installation (in an offshore wind farm) are represented.

Figure 14 – (a) Schematic Diagram of an SVC; (b) an SVC in Radsted, Denmark [18].

The STATCOM device uses a power electronic voltage source (VSC). The converter uses

semiconductors with turn-off capability, such as Insulated Gate Bipolar Transistors (IGBTs). The

benefits of the STATCOM (commercially known as “SVC Light” by ABB or “SVC Plus” by Siemens),

compared with the SVC, are the fact that the capacitor banks used are smaller and also there is no

23

need for big air-cored inductors. Further advantages of the STATCOM are also found in the dynamic

behaviour (such as faster transient response). A simplified schematic diagram of a STATCOM and an

example of an installation is shown in Figure 15.

Figure 15 – (a) Schematic Diagram of a STATCOM; (b) a STATCOM installation [18].

3.2.4.2. Distribution of the Compensation

The distribution of the reactive power compensation devices is also an important matter. Figure

16 presents several options for compensation: on one end only (offshore, in the case), on both ends or

distributed along the cable. The figure also compares the loading of an open-end cable. It is obvious

that the worst case is with compensation installed at one end only. The best solution is to install

distributed compensation along the cable, since the charging current generated would flow towards

the reactor, not increasing the rating of the cable so significantly. It is however important to notice that

the placement of reactive power compensators in the middle of the cable is likely to not be possible,

as the installation of these interstitial reactive power compensators would present an added

technological and economic challenge.

24

Figure 16 – Loading of an open-end cable due to capacitive charging current for different schemes of reactive power compensation [19].

25

3.3. HVDC-LCC Transmission

3.3.1. General Aspects

Current trends in offshore wind farms indicate future installations will be rated at hundreds of MW

and further from shore (as far as 200/300 km). These transmission requirements will not be feasible

using AC submarine cable transmission, given the limitations mentioned in 3.2. Connecting offshore

wind farms through a DC link has then been considered since there is a significant amount of

experience in transmitting large amounts of power over long distances through HVDC links.

There are two different HVDC transmission technologies: the line commutated converter HVDC

(HVDC-LCC, also referred as “Conventional” or “Classic” HVDC) using thyristors in the converters and

the voltage source converter HVDC (HVDC-VSC) which uses IGBTs.

The main advantage of HVDC-LCC over the HVDC-VSC is its proven track record, since there is

an accumulated experience of decades for this technology. The first commercial HVDC-LCC

connection was installed in 1954 and since then many other conventional HVDC links were installed

all over the world.

Some of the most important advantages that HVDC-LCC transmission offers over AC are [20]

[21]:

• Asynchronous connection, since sending and receiving end frequencies can differ;

• Transmission distance using DC is not limited by cable charging current;

• Low cable power losses;

• Higher power transmission capability per cable;

• Power flow is fully defined and controlled;

• HVDC does not transfer short circuit current.

Some of the constraints of HVDC-LCC transmission are the following [20] [21]:

• No experience in connecting offshore wind farms;

• Production of harmonics in the converter, making the use of filters necessary;

• The converters at each end consume reactive power.

3.3.2. Topology and Main Components

The HVDC-LCC transmission system for an offshore wind farm can be represented by Figure 17.

26

Figure 17 – Basic Configuration of a wind farm using an HVDC-LCC [7] Note: F=filter; HFF=High-frequency filter.

An HVDC-LCC transmission system consists of the following main components (represented in

Figure 17) [11]:

• AC based collector system within the wind farm;

• Offshore three-phase two-winding converter transformers;

• Auxiliary power set (STATCOM or Diesel generator);

• AC and DC Filters;

• Smoothing Reactor;

• DC submarine cables;

• Onshore converter station (with transformer and LCC converter) and filters.

The AC collector system is typically rated at 33 kV, as in the HVAC wind farms. This medium

voltage is stepped-up to the necessary transmission line voltage in the offshore transformers. Usually

the transformers are star-star and delta-star connected to power the 12-pulse converter, as this

configuration cancels several harmonics. The design of the transformers is critical as their insulation

must withstand the AC component of the voltage and the DC component from the thyristor valves.

Tappings are also included for proper system control.

The LCC power converters (onshore and offshore) are the most important elements in the

system, as they perform the AC/DC conversion offshore and the DC/AC conversion onshore. The LCC

converters are based in thyristor valves, capable of standing 8 kV and DC currents up to 4 kA. With

these characteristics it is possible to convert up to 1000 MW for land connections and 500 MW for

submarine transmissions [14]. In the LCC converter the current is always lagging the voltage due to

the control angle of the thyristors; hence these converters consume reactive power. For this reason,

reactive power compensation is necessary at both ends to provide reactive power to the system.

27

Capacitor banks or STATCOM devices are considered for this effect. Detailed explanation of the

functioning of the thyristor converter is provided in 3.3.4.

A line-commutated converter requires an AC voltage source for its commutation, so an auxiliary

power set is required at the offshore station to provide the necessary commutation voltage for the

HVDC-LCC connection, when there is little or no wind. The auxiliary power set is also used to supply

power to other devices in the wind farm, when the wind farm is disconnected from the grid.

The AC filters mentioned are used to absorb harmonic currents generated by the HVDC

converters, thus reducing the impact of the harmonics on the AC system. These filters also supply

reactive power to the converter station. The DC filters are used to avoid the generation of circulating

AC currents in the cable.

The smoothing reactors are large inductances connected in series with each pole of the DC link.

They are used to prevent current interruption at minimum load, limit the DC fault current, reduce

voltage and current harmonics and also prevent resonance in the DC circuit.

The DC submarine cable is approached in detail in section 3.3.3.

3.3.3. DC Submarine Cables

The submarine cable is one of the most important components of the HVDC-LCC system, since it

is the link for power transmission from the offshore wind farm.

The elements comprising a DC cable are the same as the ones in an AC cables. As so, like the

HVAC cables, the HVDC cable main technologies available differ in the electrical insulation used:

• Oil Filled (OF) cables;

• Mass Impregnated (MI) cables;

• Cross-linked Polyethylene (XLPE) cables.

These cables are represented in Figure 18.

28

Figure 18 – Available HVDC Cables: a) OF cable; b) MI cable; c) XLPE cable [22].

In Oil Filled cables (also known as SCFF – Self Contained Fluid Filled) the insulation consists of

paper impregnated with low-viscosity oil. The core of the cable is covered by a hollow shaft where oil

is circulated by pumps at both ends of the line. Oil Filled cables can be used to DC voltages up to 600

kV and great sea depths. The need for pumping at each end of the system may prove difficult, and

may limit significantly the maximum length of the cable to less than 50 km [14]. The danger of oil spill

and the need for cable protection are also obvious disadvantages of the Oil Filled cable technology.

Mass Impregnated cables have similar construction but the paper insulation is impregnated in

resin and high viscosity oil and no oil circulation system is required. It is the most used cable

technology in existing HVDC systems and so the track record and high reliability are some of the

advantages of the MI cable.

As can be seen in Figure 19, the MI cable consists of different layers that include, in addition to

the conductor and the paper insulation, layers used for mechanical protection of the copper conductor:

the polyethylene and lead sheaths, jackets and steel armour.

Figure 19 – Mass Impregnated HVDC Cable [22].

29

The MI cable technology is available for voltages up to 500 kV and transmission power of up to

800 MW. The capacity of the cable is only limited by the conductor temperature of about 50ºC.

In XLPE cables the insulating material is made of solid dielectric, also known as extruded

dielectric. It is a relatively new technology, developed to overcome some of the limitations of the

previously referred technologies. XLPE cables have all the advantages of the MI cables and

additionally can carry nominal current with a cable temperature of 90ºC. Increased bending capability,

higher mechanical resistance and lower weight are also advantages of the XLPE cables, since the

installation process is easier than for other cables.

3.3.4. Converter Technology

HVDC-LCC employs line commutated, current source converters, with thyristor valves. The use

of the thyristor grants this HVDC technology a limited operating flexibility, since the lack of turn-off

controllability of the conventional thyristor results in poor power factors and considerable waveform

distortion. Thus, some reactive power compensation and filtering must be provided.

3.3.4.1. 6-pulse Bridge Converter

The basic building block used for HVDC conversion is the three-phase, full-wave bridge referred

to as a 6-pulse or Graetz bridge, as represented in Figure 20. The term 6-pulse is due to six

commutations or switching operations per period resulting in a characteristic harmonic ripple of 6

times the fundamental frequency in the dc output voltage. Each 6-pulse bridge is comprised of 6

controlled switching elements or thyristor valves. Two thyristors are simultaneously conducting at each

instant.

Figure 20 – Six-pulse Converter Bridge [23].

The following analysis is done for the rectifier mode of operation, but similar considerations can

be made about the inverter mode. In order to conduct, the thyristor valve must be positively biased

30

(with a positive voltage between the anode and the cathode) and also a pulse has to be applied to the

gate. Therefore, the current will only flow through the valve once the firing pulse has been applied.

This enables the possibility of controlling the instant when current begins to flow through a valve, or to

commutate from one valve to another, by postponing the firing pulse. The angle between the instant at

which the thyristor valve becomes forward biased and the start of commutation (given by the firing

pulse) is referred to as the firing delay angle, α. The method of delaying the firing instant allows the

change in the average value of the DC voltage (in Figure 20) and, consequently, of the transmitted

power. As so, the average output voltage of the converter, working as a rectifier, is given by equation

(5) [24].

=3√6

∙ ∙ cos − ! ∙ (5)

Where:

• is the DC output voltage;

• is the line-to-neutral rms commutating voltage referred to the secondary of the

converter transformer;

• is the firing angle;

• ! is the commutating resistance;

• is the output DC current.

Note that the equation includes the commutation resistance that accounts for the internal voltage

drop in the converter due to commutation. The commutation delay accounts for the fact that the

commutating operation is not instantaneous, and so there is a small interval where three valves are

conducting at a time. The angle associated with this delay in commutation is called commutation

angle, µ. The value of µ is generally in the range of 15º to 25º [24].

The power factor of the converter can be proved (in [24]) to be given by equation (6):

" =3

∙ cos (6)

Therefore, it is of great interest to use a reduced value of firing delay angle, in order to have a

high power factor. The value of α is usually between 10º and 20º [24]. As the firing delay angle

increases, the converter absorbs more reactive power from the AC grid.

The inverter performs the DC/AC operation and is located at the onshore converter station.

The analysis carried out for the rectification process is equally applicable to the inverter operation,

although the inverter equations are expressed in terms of the advance angle β (β =180º-α) and the

extinction angle γ (γ= β - α). The advance angle defines the moment at which the valves of the inverter

31

start conducting and the extinction angle refers to the interval between the end of commutation and

the next zero crossing of the commutating voltage.

3.3.4.2. 12-Pulse Bridge Converter

The series-connected double bridge converter has become the standard configuration in HVDC

transmission. This converter is represented in Figure 21.

Figure 21 – Diagram of a 12-pulse converter [23].

The 12-pulse converter is a series-connection of two 6-pulse converter bridges and it requires two

3-phase transformers, one with a star-star and the other with star-delta (i.e. with 30º phase shift

between them).

This arrangement has some advantages when compared to the 6-pulse bridge converter, namely

the fact that the converter injects harmonic currents of orders n=12k±1 into the AC system, and so no

filtering of lower order harmonics, like the 5th and 7th is necessary (like in the 6-pulse converter) [24].

The alternate voltage supplied by the AC network is also more sinusoidal, meaning it has lower

harmonic content, requiring less filtering. The voltage on the DC side repeats itself 12 times during the

cycle, making it closer to the ideal continuous voltage. In the 6-pulse converter the voltage repeats

itself six times during a period at AC grid frequency.

3.3.4.3. HVDC Schemes

There are two basic configurations used in HVDC systems: the monopole and the bipolar

configurations, shown in Figure 22.

32

Figure 22 – HVDC configurations: a) Monopole; b) Bipolar.

In the monopole configuration one single DC cable is used. The return path is assured by sea

electrodes. Alternatively, due to high earth resistivity, metallic structures in the proximity or

environmental constraints, a metallic conductor may be used, either in the form of a low voltage AC or

DC cable or integrated in the cable itself.

In the bipolar scheme, two DC cables with opposite polarities are used and so the return path is

guaranteed. This configuration is the most common in HVDC transmission, since it offers some

advantages over the monopole option, namely higher transmission capacity and higher availability,

since during maintenance or outages of one pole, it is still possible to transmit part of the power. More

than 50% of the transmission capacity can be used, limited only by the overload capacity of the pole

[23].

3.3.5. HVDC-LCC + STATCOM

As referred previously, in a LCC transmission systems both converters absorb reactive power, as

in the converters the current always lags the line voltage due to the requirements for a positive

commutation voltage at the firing of the thyristors. Thus, this type of converter needs reactive power

for operation, which has to be provided by reactive power devices connected to the AC network. As

these converters depend on the line voltage for commutation, a minor disturbance in the AC voltage

might also result in commutation failures in the converters.

Capacitor banks or synchronous compensators may be used for reactive compensation at both

ends of the DC link. A more advanced solution consists of a STATCOM. The STATCOM provides the

33

necessary commutation voltage to the HVDC converter, continuous AC voltage control, fast reactive

power compensation to the network under transient conditions and removal of possible non-

characteristic harmonic interactions [18]. The STATCOM can also provide limited active power support

to the network during transient conditions, such as active power changes of the wind farm output.

However, the ability to provide active power support depends on the energy storage on the DC side,

which in the case of a conventional DC capacitor will be very limited. Larger energy storage could be

provided by means of batteries or SMES (Superconductive Magnetic Energy Storage).

The HVDC-LCC transmission system may benefit significantly from the STATCOM technology

since it can, in steady state, provide the reactive power consumed by the converter. Offshore, the

STATCOM can also work as an auxiliary power set. In AC grid fault conditions, the STATCOM can

provide reactive power to the network, thus contributing to a less severe voltage dip off the onshore

bus [12]. The working principles of a STATCOM device are mentioned in section 3.2.4.1.

34

3.4. HVDC-VSC Transmission

HVDC with voltage source converters is a fairly recent technology (first installed in 1999) and it is

known commercially as “HVDC Light”, for ABB and “HVDC Plus”, for Siemens. The basis of the

HVDC-VSC technology is the self commutated converters, with IGBT (Insulated Gate Bipolar

Transistor) valves. Figure 23 depicts a general layout of an HVDC-VSC system applied to an offshore

wind farm.

Figure 23 – Basic Configuration of a wind farm using an HVDC-VSC [11].

VSC converters are self-commutating, not requiring an external voltage source for its operation.

Also, the reactive power flow can be independently controlled at each AC network and the reactive

power control is independent of active power control. These features make VSC transmission an

interesting option for connection of offshore wind farms. In addition to the referred HVDC advantages,

the VSC technology also offers the following main benefits [11] [14]:

• Total control of active and reactive power;

• Minimum risk of commutation failures;

• Smaller size of converters and filters than HVDC-LCC;

• Possibility of the converters starting with a dead grid, not needing any start-up

mechanism (“Black-start” capability).

However, there are some constraints to the use of VSC technology. The main constraint is the

considerably lower experience in HVDC transmission, compared with the LCC option. HVDC-LCC has

more than 30 years of service experience, with high availability rates.

Also, VSC transmission has higher system power losses, when compared with an LCC system.

Typically, the power loss at full load for the converters is about 4-6% for VSC and 2-3% for LCC

transmission [11]. The IGBT valves are also much more expensive than the power thyristors used in

LCC technology.

Apart from these disadvantages, some of which easily surpassed once VSC links become

widespread and further experience is gained, the voltage source converter technology is a promising

option for HVDC in years to come.

35

4. PSS/E Modelling

This chapter describes the models used in the PSS/E software for the simulation of the

alternatives for grid connection of an offshore wind farm. The simulations performed in this software

are also described.

4.1. Introduction

In order to simulate the behaviour of an offshore wind farm and of the required transmission

system, appropriate models of the wind farm, the transmission system (either HVAC or HVDC) and

the electrical grid have to be constructed. Only then is it possible to analyse the steady-state and the

transient behaviour of the power system.

The software used to model the system and perform the simulations in the work of this thesis is

the PSS/E software. PSS/E stands for Power System Simulator/Engineering and it is a software tool

provided by Siemens Power Technologies International (PTI). It is used by most utilities in the world to

perform power system simulations, as it allows the performance of power flow analysis, dynamic

simulations and stability studies, among other features. PSS/E is composed of a comprehensive set of

programs for studies of power system transmission network performance in both steady-state and

dynamic conditions, obvious reasons for its widespread use by transmission and distribution systems

operators.

In the particular case of this thesis, the PSS/E version 30 was used for the features it offers for

network simulation: the extensive library of power system components, including wind turbine, DC

links and FACTS are extremely useful for the simulations performed in the scope of the thesis. The

wide range of possibilities available in PSS/E modelling increases significantly the complexity of the

simulation but also add to the interest of this software for network modelling.

4.2. Grid Structure

The electrical grid that will be used for steady-state and transient analysis – and in which the

offshore wind farm will be integrated – is the “savnw” network, provided by PSS/E as an example of a

relatively large grid, as it has 23 buses, 6 generators and 7 loads. In order to properly use this grid, a

few changes were made, namely to the operating frequency and voltage levels. These changes were

made taking into account values used in the European electrical grid.

Additional changes to the original “savnw” network were performed since the offshore wind farm

added (rated at 110MW) will be placed in a bus where the existing generation is of 750 MW.

Therefore, adjustments in active and reactive power of the network are required, which can be made

by adapting the power of the loads and the shunt compensators.

As so, the changes made to the original network can be summarised:

36

• The frequency was changed to 50 Hz (the frequency in use in most parts of the world,

including Portugal) since the frequency of original example grid is 60 Hz (typical of

American grids);

• The voltage levels were changed from 500 kV, 230 kV and 22 kV to, respectively, 400

kV, 220 kV and 33 kV. These voltage levels are common in most European Grids,

including Portugal;

• The 110 MW offshore wind farm was added, replacing an existing power plant of 750

MW;

• The inductive shunt compensators were removed;

• The parameters of the 400 kV lines were adapted;

In order to accommodate for the changes reported above, some network parameters have to be

adapted, which are described in the next sections. Most of the parameters of the grid are presented in

detail in Appendix A.

4.2.1. Lines

As a result of changing the grid frequency to 50 Hz, the reactance and the susceptance of the

branches have to be converted, since these parameters depend on the grid frequency. This

conversion is done using Equations (7) and (8).

#$%&' = #(%&' ×$%&'

(%&' (7)

*$%&' = *(%&' ×$%&'

(%&'

(8)

The line parameters (except those of the 400 kV lines) used in the test grid are presented in

Table 10, in Appendix A.

4.2.1.1. 400 kV Lines

The parameters of the 400 kV lines were changed to values of 400 kV lines existing in the

Portuguese grid. This decision was made upon studying the values of the existing 400 kV line

parameters and verifying these resemble the ones of a power cable, since significant amounts of

reactive power were being produced by the line. This issue does not occur in the 220 kV and 33 kV

lines.

Considering the base power +, = 100 / and the base voltage , = 400 1 , the base

impedance is 2, =34

5

64= 1600 Ω.

37

The line parameters: resistance, reactance and susceptance in p.u. can, respectively, be

calculated with Equations (9), (10) and (11).

!8.:. =!;< × =

2, (9)

#8.:. =#;< × =

2, (10)

*8.:. = (*;< × =) × 2, (11)

Where:

• @AB is the line resistance in Ω/1D;

• EAB is the line reactance in Ω/1D;

• FAB is the line susceptance in S/1D;

• H is the line length in 1D;

• IJ is the base impedance in Ω;

The values for the 400 kV line parameters calculated through the equations above are shown in

Table 11, in Appendix A.

4.2.2. Transformers

For the two-winding transformers in the “savnw” grid, the transformer reactance is frequency

dependent, so their values also have to be adapted to the 50 Hz frequency, according to Equation (7).

Tap changing was used in the network transformers, changing the voltage magnitude of the

transformers, and, thus regulating properly the reactive power flow.

The values of the results of the calculations for the transformers in the grid are shown in Table

12, in Appendix A.

4.2.3. Loads and Shunt Compensators

As mentioned previously, adding the offshore wind farm rated at 110 MW to bus 102 of the grid in

the place of the existing 750 MW conventional power plant leads to an imbalance of active and

reactive powers. This imbalance can be compensated by reducing the consumed power in the loads.

As so, the values of the load powers of the grid are presented in Table 13, in Appendix A.

The shunt compensators are also regulated in order to maintain the voltage in the buses at

values within reasonable values (between 0,95 p.u. and 1,05 p.u.). Table 14 (in Appendix A) outlines

the values of the shunt compensators of the grid.

38

4.2.4. Diagram of the Grid

The resulting grid, after the changes performed, is presented in the single-line diagram shown in

Figure 24, in which the offshore wind farm is already included. Bus 102, referred above, is marked in

the figure. The legend of the colours used in the figure for each voltage value is presented in Table 1.

Figure 24 – Single line diagram of the test grid used (with HVAC transmission, compensations at both ends).

39

Table 1: Colour correspondence for each voltage level in Figure 24.

Colour Rated Voltage (kV)

Red 400

Black 220

Purple 150

Dark Green 33

Blue 20

Orange 13,8

Bordeaux Red 0,7

40

4.3. Offshore Wind Farm Model

The model for the offshore wind farm is based in an existing installation: the Lillgrund Wind Farm.

This wind farm is located off the coast of Sweden, at a distance of 9 km from the Point of Connection

in the onshore grid and the transmission is achieved by a combination of an AC sea cable (7 km long)

and an AC land cable (2 km long). With 48 wind turbines, rated at 2,3 MW each, the total capacity of

the wind farm is 110 MW. Figure 25 depicts the layout of the wind turbines in the Lillgrund wind farm,

used as a reference.

Figure 25 – Layout of the wind turbines in the Lillgrund wind farm [25].

As can be seen in Figure 25, the internal grid of the Lillgrund wind farm consists of 33 kV sea

cables divided in five feeders and each of these feeders connects 9 or 10 wind turbines to the offshore

substation. The total is 48 wind turbines rated at 2,3 MW.

The modelled offshore wind farm is based on this layout. The turbines are joined in aggregates

and then connected to the offshore substation bus via 33 kV AC cables.

41

4.3.1. Wind Turbines

The wind turbines used are the GE 1,5 MW model, available in the PSS/E Wind package. This

model is of a DFIG (Doubly Fed Induction Generator) wind turbine developed by General Electric and

released for PSS/E simulation and testing.

In order to match the 110 MW of the reference wind farm, a total of 74 1,5 MW wind turbines

were used in the model. These are joined in five aggregates: two aggregates of 24 MW (16 wind

turbines for each aggregate) and three of 21 MW (14 wind turbines per aggregate). This actually adds

up to 111 MW, which is the value assumed hereby for the offshore wind farm power.

For PSS/E simulation of the wind farm, two distinct models are designed for the wind turbines: the

steady-state model (which allows the power flow simulations) and the transient model (used for

dynamic simulations).

4.3.1.1. Steady-State Model

The load flow provides initial conditions for dynamic simulations. In the Load Flow parameters of

the PSS/E, the wind turbines were modelled as five conventional generators, rated at 24 MW and 21

MW. The values specified on the existing generator record are outlined in Table 15, in Appendix A.

Note that the values in Table 15 were calculated considering the values in the Individual WTG Power

Flow Data of the PSS/E Wind User Guide [26] and multiplying by the number of lumped elements, as

recommended by the PSS/E guide.

Since the model used is of a DFIG wind turbine, both the active and reactive power can be

controlled. As so, for the wind turbine aggregates, a fixed power factor of 0,9 (injecting reactive power

in the grid) was chosen, resulting in a value for Qgen, which is 11,62 Mvar for the 24 MW aggregates

and 10,16 Mvar for the 21 MW aggregates.

The values calculated by this manner were used in the HVDC connection alternative. However,

for the HVAC transmission system, the wind turbines were regulated for a unit power factor, which

means no reactive power is generated by the wind turbines. The reason is that the AC cable already

produces a significant amount of reactive power, so an additional quantity of reactive power generated

by wind farm would deteriorate the system behaviour.

4.3.1.2. Transient Model

The transient model of the GE 1,5 MW wind turbine comprises nine models for each of the five

aggregates. These models are an accurate and complete representation of the behaviour of the GE

Wind turbine, including, for example, the simulation of wind gusts and ramps. Table 2 indicates the

models used and their description.

42

Table 2: Components of the GE 1,5 MW Wind Turbine Model in PSS/E.

Model Code Description

GEWTG1 GE Wind Turbine Generator/Converter

GEWTE1 GE Wind Turbine Electrical Control

GEWTT Two Mass Shaft

WGUSTC Wind Gust and Ramp

GEWTA GE Wind Turbine Aerodynamics

GEWTP GE Pitch Control

GEWTPT Plotting Output Variables as VARs

VTGTPA Under Voltage/Over Voltage Generator Bus Disconnection Relay

FRQTPA Under Frequency/Over Frequency Generator Bus Disconnection Relay

Values for the parameters of these models were based on typical values given in the PSS/E Wind

manual example [26]. The number of aggregated wind turbines was changed in the models to 14 for

the 21 MW aggregates and 16 for the 24 MW aggregates.

An example of the “.dyr” files used for one of the aggregates, where the parameters of each

models are written, is presented in Appendix B.

4.3.1.3. Fault Ride Through Capability

The generator model includes the Under/Over voltage protective functions (modelled by

VTGTPA). These functions are defined such that wind plants must not trip for events that are less

severe than the defined thresholds and time durations. So, the wind turbines can “ride through” a fault

that causes a voltage dip on the terminal voltage of the generator, as long as the voltage dip and

respective duration respect the curves in Figure 26.

43

Figure 26 – GE 1,5 MW voltage protection characteristics.

4.3.2. Step-Up Transformers

The 0,69/33 kV transformers that adapt the voltage at the generation buses (690 V) to the voltage

of the internal grid of the wind farm (33 kV) have the parameters presented in Table 16, in Appendix A.

As each aggregate contains 14 or 16 wind turbines, the rated power of the transformers is,

respectively, 24,5 MVA and 28 MVA.

4.3.3. 33 kV Cables

The cables used to link the five buses where the wind turbines are connected to the offshore bus

are 33 kV AC cables, XLPE insulated. The cable parameters are presented in Table 17, in Appendix

A.

Considering the base power +, = 100 / and the base voltage , = 33 1 , the base impedance

yields: 2, =34

5

64= 10,89 Ω. Therefore, using Equations (9), (10) and (11), the values in Table 17 – and

bearing in mind that * = × – the parameters for the medium voltage cables are presented in Table

18, in Appendix A. The lengths of these cables were attributed considering typical distances in

existing offshore wind farms, such as the Lillgrund wind farm.

4.3.4. Layout of the offshore wind farm

Figure 27 illustrates the offshore wind farm designed in PSS/E, where the wind turbines, the step-

up transformers and the 33 kV sea cables mentioned above are represented.

44

Figure 27 – PSS/E model of the offshore wind farm.

45

4.4. HVAC Transmission

The HVAC transmission is the simplest alternative for the transmission of the electrical power of

the offshore wind farm to the onshore grid. The main components are the submarine cable and the

two transformers: onshore and offshore. Figure 28 illustrates the HVAC transmission scheme, with the

AC cable and both transformers. Note that the 102 bus in Figure 28 corresponds to the 102 bus in

Figure 27. For proper compensation of the reactive power generated in the cable, shunt reactors (or a

STATCOM device) were applied to one or both ends of the cable. They are not, however, represented

and described in detail in this chapter, as they vary according to the case simulated.

Figure 28 – HVAC transmission layout in PSS/E.

4.4.1. AC Submarine Power Cable

The submarine power cable chosen is a 100 km 150 kV XLPE cable. 150 kV is a typical option for

offshore wind farms, used in, for example, the Horns Rev wind farm, in Denmark. Hence, using the

parameters of the submarine cable in Horns Rev, we have the following set of values: ! =

0,039 Ω 1D⁄ ; # = 0,12 Ω 1D⁄ ; = 0,19 P"/1D [10].

Considering the base power +, = 100 / and the base voltage , = 150 1 , the base

impedance is 2, =34

5

64= 225 Ω. So, using Equations (9), (10) and (11) the 150 kV cable parameters are

depicted in Table 3.

Table 3: 150 kV cable parameters.

Cable @AB

[S/AB] EAB

[S/AB] UAB

[VW/AB] Length

[AB] @X.Y. EX.Y. FX.Y. From

Bus To

Bus

20 21 0,039 0,12 0,19 100 0,0173 0,053 1,34235

46

4.4.2. Onshore and Offshore Transformers

The offshore transformer adapts the voltage from the 33 kV of the internal grid to the 150 kV of

the submarine cable. As for the onshore transformer, it is used to increase the voltage from the 150 kV

of the cable to the 400 kV of the onshore grid. Table 4 shows the offshore and onshore transformers

data.

Table 4: Onshore and Offshore transformer data [10].

Buses

Location

Rated Voltage [kV/kV]

Rated Voltage

Impedance [%]

Rating [MVA]

E [X. Y. ] Z[[ \]^ F_`aH

From Bus

To Bus

20 102 Offshore 33/150 13,8 160 0,08625

21 151 Onshore 150/400 15 200 0,075

47

4.5. HVDC-LCC Transmission

The Line Commutated Converter based HVDC technology requires a DC cable and two

converters, one onshore and another offshore. In PSS/E there are appropriate models for the

simulation of a transmission with HVDC-LCC for both steady-state and transient situations.

4.5.1. Steady-State Model

The steady-state model enables the power flow analysis of the DC Link and establishes the initial

values for dynamic simulations.

In PSS/E, the data requirements for the “Two-terminal DC Line” (as mentioned in [27]) fall into

three groups: control parameters/set points, converter transformers and DC line characteristics. In the

Load Flow Module of PSS/E 30, the two tabs refer to “DC Line” and the respective “Converters”:

Rectifier and Inverter.

Graphically, the DC line is represented by Figure 29, where the converter is represented by a

diode-like symbol. Note that the rectifier and inverter models include the converter transformer,

despite these transformers not being visible.

Figure 29 – HVDC-LCC transmission model in PSS/E.

The following values were assumed:

bcdefg= 110 /h

bcdefg= 150 1

= 18°

j = 18°

P = 20°

The current in the DC line is given by: bcdefg=

klmnofg

3lmnofg= 733 A.

The model parameters were calculated considering [27] and [24]. Some of the equations used for

the rectifier and the inverter are expressed in Table 5.

Table 5: Equations for Rectifier and Inverter Parameters.

Parameter Rectifier Inverter

Converter Power Factor cos(r) =cos( ) + cos ( + P)

2 cos(r) =

cos(j) + cos (j + P)

2

48

Transformer Nominal Power +tuvd =

bcdefg

cos(r)

AC Converter Voltage wxdyzg=

×

3√2

Converter Transformer Ratio

Du = xdyzg

|uczg

Dc = xdyzg

|uczg

Commutating Reactance #t[Ω] = #t[p. u. ] × 2,, with 2, =3gnzg

5

6n

Commutating Resistance ! =3

× #t

Using values calculated with the equations in Table 5, assuming some values recommended in

[27], the parameters used in the load flow model are reported in Table 19 and Table 20, in Appendix

A.

4.5.2. Transient Model

In order to simulate the dynamic behaviour of the HVDC link, the transient model used is the

“CDC4T”. The parameters of this model were calculated using the equations and the recommended

values in the PSS/E Manuals [28] and [29].

The parameters used in the “CDC4T” model are reported in Table 21, in Appendix A.

4.6. STATCOM

PSS/E provides appropriate models to simulate FACTS devices. A variety of such devices is

available for power flow and dynamic simulations, divided in either shunt devices (such is the case of

STATCOM), series devices and combined devices (with both series and shunt elements).

The STATCOM, as described in detail in Chapter 3 is a FACTS device, used in the network to

absorb or provide reactive power, thus providing voltage support to the network, in either steady-state

or transient situation.

In the work developed in this thesis, the STATCOM is used in the HVDC network, to inject

reactive power in the onshore bus, reducing the voltage dip originated by the fault in the AC grid. In

HVAC transmission, the STATCOM also provides this voltage support in case of a fault and,

additionally in steady-state, it is used for reactive power compensation of the AC cable.

ABB’s “PCS 6000 STATCOM double-outdoor” is used as a reference for the modelling of the

STATCOM device. It has a capacity of 2x32 Mvar.

49

4.6.1. Steady-State Model

For power flow analysis, the STATCOM is modelled as a FACTS device with the parameters

adjusted as to simulate the behaviour of this device. In both cases of the use of the STATCOM (in

HVAC and HVDC transmission) it was modelled with the same parameters.

Figure 30 is the graphic representation of a STATCOM in PSS/E.

Figure 30 – STATCOM device in PSS/E.

The parameters presented in Table 22 (Appendix A) are the most important values used in power

flow analysis of the STATCOM and they are obtained using PSS/E Manual [27] for FACTS devices.

4.6.2. Transient Model

The dynamic behaviour of the STATCOM is simulated using the “CSTCNT” model. The

parameters of this model were calculated using the equations and the recommended values in the

PSS/E Manuals [28] and [29].

The values of the “CSTCNT” model are presented in Table 23, in Appendix A.

50

5. Power Flow Results

Chapter 5 presents the results of the steady-state simulations performed on the grid with the

offshore wind farm, as described in Chapter 4. HVAC and HVDC-LCC transmission are analysed.

Discussion of the results is also performed.

5.1. Introduction

The power flow analysis carried out comprehends numerical calculations of active and reactive

power flows and node voltages. PSS/E software is used, in the scope of this thesis, for the power flow

analysis of transmission power alternatives from offshore wind farms. Special attention is given to the

Point of Common Coupling (hereby designated as PCC), i.e., the point of connection of the wind farm

with the remaining grid. The voltage at the PCC, the active power injected and the reactive power

injected/absorbed in the PCC is analysed as a part of the power flow study. Power losses for each

transmission system are also assessed.

The power flow results are presented by the single-line diagrams of the network (from the PSS/E

load flow software). The active and reactive flows at each end of the branches and the voltage

magnitude and angle at each bus are depicted in each figure of this chapter. Figure 31 depicts how

each of the values from the power flow calculations are represented in the power flow result figures.

Figure 31 – Legend of the values presented in the single-line diagrams.

5.2. HVAC Power Flow

For the HVAC transmission system, as presented in Figure 28, the power flow analysis focuses

on the reactive power flow in the submarine 100 km cable. Therefore, with the objective of

compensating the reactive power generated in the cable, some reactive power compensation options

are studied. The alternatives considered are reviewed in Table 6.

51

Table 6: Reactive power compensation alternatives.

Designation Onshore (Mvar) Offshore(Mvar)

Offshore shunt reactor 0 60

Onshore shunt reactor 80 0

Onshore and offshore shunt reactors 40 40

Offshore STATCOM 0 64

The main criteria for the choice of the value of the shunt reactors (defined in terms of reactive

power injected, in Mvar) is the power factor at the PCC, which is chosen to be approximately of 0,9, a

typical value in grid integration of wind farms [11]. As so, the shunt reactor chosen for offshore

compensation only absorbs 60 Mvar, as the 0,9 power factor at the PCC is guaranteed with this value.

An estimate of the reactive power produced by the 150 kV AC cable (data on Table 3) can be

made taking into account Equation (3). The approximate amount of reactive power produced by the

100 km cable is given by (12):

= × × × = 2 × 50 × 0,19 × 10( × 100 × (150 × 10) = 134,3 / (12)

So, for the HVAC transmission, approximately 60 % of this value of reactive power is

compensated by the schemes presented in Table 6.

Note that, for HVAC power flow, the wind turbine generators supply no reactive power, since

there is already an excess of reactive power, as a consequence of the shunt capacitance of the AC

cable. Therefore, the capability of the DFIG machines of providing voltage support to the grid, by

supplying reactive power, is not considered for the present study.

For the first case analysed (offshore compensation), the presented grid includes buses 3004, 201

and 211, closest to the wind farm. For the following power flow results, however, the results relative to

these buses are omitted as they are very similar to the results for the offshore compensation only

transmission.

52

5.2.1. Offshore Compensation

In Figure 32 the results of the grid connected by an AC cable with compensation offshore are

shown. The 60 Mvar shunt reactor (marked in the figure with the orange rectangle) absorbs 64,6 Mvar

of reactive power, given the dependency of the reactive power of the voltage at the bus (as expressed

by equation (4)).

All the bus voltages near the AC link (buses 102, 20, 21 and 151) are within acceptable values

(i.e. below 1,05 p.u.) and the power factor at PCC is approximately 0,88.

Figure 32 – Power Flow result for HVAC with offshore compensation.

53

5.2.2. Onshore Compensation

The most important remark regarding Figure 33 is the fact that using onshore compensation only

leads to overvoltage on the offshore bus. Other values for the onshore reactor where experimented

but led to negligible improvements on the magnitude of the voltage of bus 20. The power factor at the

PCC is, however, acceptable, since the shunt reactor (marked in orange rectangle on the figure)

absorbs a large portion of the reactive power produced at the cable: 83,4 of the 128,3 Mvar produced

by the cable.

Figure 33 – Power Flow result for HVAC with onshore compensation.

54

5.2.3. Offshore and Onshore Compensation

Figure 34 shows the power flow result for a scheme of compensating reactive power at both ends

of the cable by two 40 Mvar reactors. The compensators are marked in orange on the figure. No

overvoltages occur in any bus, onshore or offshore.

Figure 34 – Power Flow result for HVAC with offshore and onshore compensation.

55

5.2.4. STATCOM Compensation

Figure 35 presents the results for the transmission using an offshore STATCOM. The results are

very similar to the use of offshore compensation with a shunt compensator, since the amount of

reactive power consumed by the STATCOM and the reactor are identical. The STATCOM is shown

inside the orange rectangle on the figure.

Figure 35 – Power Flow result for HVAC with STATCOM compensation.

56

5.2.5. Discussion

The compensation alternatives studied bare some similarities but also important differences

between them.

All the schemes were designed with the purpose of obtaining a power factor of approximately 0,9

(injecting reactive power in the grid) at the PCC. Issues like reactive power circulation, bus voltages,

power rating of the cable and power losses are analysed.

Table 7 shows the results obtained for the power factor at the PCC, the voltages at each end of

the cable and the active and reactive power at each end of the cable for the four configurations

studied. The values are taken from the power results presented above.

Table 7: HVAC selected results for different compensation alternatives.

Power Factor at the PCC

V onshore bus (p.u.)

V offshore bus (p.u.)

P at the offshore end (MW)

P at the onshore

end (MW)

Q at the offshore

end(Mvar)

Q at the onshore

end(Mvar)

Offshore 0,88 1,023 1,038 110,8 108,8 76,2 60,3

Onshore 0,93 1,021 1,070 110,8 108,3 10,8 128,3

Offshore and Onshore

0,94 1,020 1,046 110,8 108,8 55,1 82,1

STATCOM offshore

0,88 1,023 1,037 110,8 108,8 77,9 58,4

Compensating onshore only is the option that leads to less favourable results. The excess of

reactive power at the onshore bus leads to an excessive voltage magnitude at bus 20, as seen in

Figure 33. The influence of bus voltage amplitude on the transmission of reactive power is thus

demonstrated by the results. Another important issue to consider is the power rating of the AC cable.

In the onshore compensation scheme the power transmitted by the cable reaches its peak near the

onshore bus. This may lead to the overloading of the cable, a direct consequence of the significantly

high amount of reactive power transmitted through the cable.

The other compensation alternatives present more satisfactory results. As far as the power rating

of the cable is concerned, compensating offshore only leads to a more evenly distributed loading of

the cable (as presented in Figure 16) as similar amounts of reactive power flow to each end of the

cable. The bus voltages at each end of the cable (bus 20 offshore and bus 21 onshore) are below 1,05

p.u. (as can be seen in Table 7) so no overvoltage exists, since the excess of reactive power at the

offshore bus no longer exists. Therefore, compensation offshore is proven to be necessary, in order to

maintain the stability of the bus voltages. It is, however, important to notice that the installation of

compensators offshore is likely to be more expensive and technically challenging than the placement

of such devices at the onshore substation.

The STATCOM compensation presents very similar results to the compensation offshore using a

shunt reactor. This option was introduced since the STATCOM device may offer benefits for the

system in its dynamic behaviour, as approached in Chapter 6.

57

The power losses of the system are the same for all compensation schemes, since the devices

used for compensation are considered as being lossless. The active power generated by the wind

farm is 111 MW and the active power at the PCC is 108,8 MW. Therefore, the losses for the HVAC

transmission can be calculated by:

= − kww

=

111 − 108,8

111= 1,98%

In Table 7, it is obvious that for any of the cases studied the power factor achieved is close to the

0,9 power factor, set as an objective of the compensation. Without compensation of any sort, the

reactive power injected in the grid would be very high, leading to a much deteriorated power factor at

the PCC. The bus voltages would also be above 1,05 p.u., due to the excess of reactive power.

The solution considered to be the most effective and cost-efficient, given the smaller size of the

compensators is the offshore and onshore compensation. This configuration is considered for the

dynamic simulations in Chapter 6.

58

5.3. HVDC Power Flow

The HVDC power flow analysis is carried out for two configurations: one without STATCOM and

another with a STATCOM at the onshore bus (bus 151). Note that for the studies with HVDC

transmission the wind turbines are working with a 0,9 power factor (injecting reactive power in the

grid). The objective is once again to guarantee an approximate 0,9 power factor at the PCC.

The active and reactive power flows and the voltages at the buses at each end of the link are

compared for these two HVDC transmission options and also taking into account the HVAC

transmission power flow results.

5.3.1. HVDC Configuration

Figure 36 depicts the single-line diagram with the power flow results for the HVDC transmission.

Figure 36 – Power Flow result for HVDC-LCC transmission.

59

5.3.2. HVDC+STATCOM Configuration

Figure 37 shows the power flow results for the HVDC+STATCOM alternative. The STATCOM can

be seen at bus 151, injecting reactive power in steady-state. The introduction of the STATCOM in the

system influences the power flows at the nearby lines.

Figure 37 – Power Flow result for HVDC+STATCOM transmission.

60

5.3.3. Discussion

Comparing with the HVAC transmission, the HVDC configuration holds no reactive power issues,

since these issues yield from the shunt capacitance of the AC cable, inexistent in HVDC.

Table 8 presents some results for the two HVDC configurations studied, including power factor at

the PCC, voltage magnitudes on both ends of the transmission scheme and active and reactive

powers circulating on both ends of the submarine cable.

Table 8: HVDC-LCC and HVDC-LCC+STATCOM selected results.

Power Factor at the PCC

V onshore

bus (p.u.)

V offshore

bus (p.u.)

P at the offshore end (MW)

P at the onshore

end (MW)

Q at the offshore

end(Mvar)

Q at the onshore

end(Mvar)

HVDC-LCC 0,89 1,012 0,978 110 109 44 54

HVDC-LCC+STATCOM 0,89 1,016 0,978 110 109 44 55

From Table 8 it can be noted that the two configurations bare similarities. Offshore, as the wind

farm is the same, the voltage magnitude and the active and reactive powers transmitted in the cable

are equal. Onshore, the influence of the STATCOM – although slight – can be noted, as the voltage is

higher, due to the injection of reactive power by the STATCOM.

As expected, the HVDC-LCC converters consume reactive power, which is a consequence of

having lagging current. Therefore, as can be seen in Table 8, the rectifier (located offshore) consumes

44 Mvar and the inverter (onshore) consumes 54 Mvar and 55 Mvar, for the HVDC and for the HVDC-

STATCOM transmission, respectively. As so, the power factor at the PCC is, for both cases, 0,89

(consuming reactive power).

The use of a STATCOM is justified in steady-state for the injection of reactive power in bus 151,

in order to compensate the reactive power absorbed by the inverter. So, in the power flow analysis,

the STATCOM is working as a capacitor bank, injecting 65 Mvar – its maximum value for the bus

voltage at bus 151. In this case, the use of the STATCOM changes the reactive power that flows to

and from the lines connected to bus 151, because of the behaviour of the LCC converter (consuming

reactive power). For the purpose of this study, in steady-state, capacitor banks could have also been

used but the STATCOM is chosen as to also provide support to the grid in transient situations.

The power losses in the HVDC transmission are very similar to the ones in HVAC (around 2%),

as the generated power is the same (111 MW) and the power at the PCC is similar (109 MW). So the

power losses are given by:

=111 − 109

111= 1,8%

In HVDC, the losses are originated in the DC line (characterized by a resistance). The line

reactance and capacitance concepts do not exist and so neither does reactive power in the link.

61

6. Dynamic Results

Chapter 6 presents the results of the dynamic simulations performed on the grid under analysis.

The response of the HVAC and HVDC-LCC transmission systems to an AC grid fault is analysed and

results are discussed.

6.1. Introduction

The analysis of the dynamic behaviour of the offshore wind farm comprehends the response of

the wind farm to voltage and frequency disturbances in the grid.

The study of this work is focused on the response of both the offshore wind farm and the onshore

grid to a three-phase fault in an onshore bus, bus 3005, as marked in Figure 38.

Figure 38 – Single line diagram of the grid used for the dynamic simulations. Note: Bus 3005, where the fault occurs is marked in the orange rectangle.

62

The behaviour of some parameters of the grid, namely the voltage and the frequency in each bus,

during and immediately following the disturbance are analysed for the two transmission systems under

analysis: HVAC and HVDC-LCC.

The parameters of the grid that are presented in the results, in 6.2 and 6.3, are the frequency and

voltage variation in the offshore and onshore buses and the speed variation of the offshore and

onshore machines. For the HVDC transmission the active and reactive power in the rectifier and

inverter are also presented.

The analysis is carried out for the two transmission alternatives and for different fault types, with

the objective of assessing the fault ride through capability of the offshore wind farm, i.e. the

requirement for the wind farm to stay connected to the grid during the disturbance, thus contributing to

the reestablishment of the normal operation. The fault ride through capability of the wind turbines in

the offshore wind farm is guaranteed by the under/over voltage disconnection relays of the wind

turbine generators. These devices allow the operation of the wind turbines even when the terminal

voltage decreases. This capacity of “riding through” a fault is limited to defined voltage dips and fault

durations. So, and according to the implemented fault ride through characteristic (as defined in Figure

26), the wind turbine will only trip if the fault that occurs across the terminals of the machine are

outside the defined limits.

6.2. HVAC Dynamic Behaviour

In order to evaluate the behaviour of the offshore wind farm connected to the onshore by an

HVAC link, different dynamic cases are analysed, as presented in Table 9.

Table 9: HVAC Dynamic Cases Analysed.

Case Fault Duration

(ms) Fault Severity

Reactive Compensation Scheme

Case 1 100 Moderate Shunt reactors at both ends

Case 2 300 Moderate Shunt reactors at both ends

Case 3 100 Severe Shunt reactors at both ends

Case 4 100 Severe STATCOM offshore

Cases 1 and 2 are two different short-circuits with similar severity but different clearance times,

simulating two different types of protections: fast and slow operating. Case 3 represents a more

severe short-circuit, with a higher short-circuit power. All these three cases are studied with the

compensation at both ends of the cable (as proposed in 5.2.3). Case 4 has the same conditions as

Case 3, except for the fact that the compensation is assured by a STATCOM placed offshore and

possible benefits of this device are studied.

The results for the outlined cases 1 to 4 are presented in 6.2.1 to 6.2.4, in the proper order. In

6.2.5 the most significant aspects regarding the results obtained are discussed.

63

6.2.1. Case 1 results

Figure 39 – Frequency variation of offshore buses for the Case

1 fault.

Figure 40 – Voltage variation of offshore buses for the Case 1

fault.

Figure 41 – Offshore wind turbine speed variation for the Case

1 fault.

Figure 42 – Frequency variation of onshore buses for the Case

1 fault.

Figure 43 – Voltage variation of onshore buses for the Case 1

fault.

Figure 44 – Speed variation of machines on the onshore grid

for the Case 1 fault.

0 0.5 1 1.5 2 2.5 3 3.5 449.98

50

50.02

50.04

50.06

50.08

50.1

50.12

50.14

Time [s]

Fre

qu

en

cy

[H

z]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 20

Bus 102

0 0.5 1 1.5 2 2.5 3 3.5 40.85

0.9

0.95

1

1.05

1.1

Time [s]

Vo

lta

ge

Ab

so

lute

Va

lue

[p

.u.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

Bus 20

0 0.5 1 1.5 2 2.5 3 3.5 4

2999.7

2999.8

2999.9

3000

3000.1

3000.2

3000.3

3000.4

3000.5

Time [s]

Ma

ch

ine

Sp

ee

d [

rpm

]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

0 0.5 1 1.5 2 2,5 3 3.5 4 4.5 549.98

50

50.02

50.04

50.06

50.08

50.1

50.12

50.14

Time [s]

Fre

qu

en

cy [

Hz]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 0.5 1 1.5 2 2.5 3 3.5 40.8

0.85

0.9

0.95

1

1.05

1.1

1.15

Time [s]

Vo

ltag

e A

bs

olu

te V

alu

e [

p.u

.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 52999

3000

3001

3002

3003

3004

3005

3006

3007

3008

3009

Time [s]

Mac

hin

e S

pe

ed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

64

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 52995

3000

3005

3010

3015

3020

Time [s]

Mach

ine S

peed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

6.2.2. Case 2 results

Figure 45 – Frequency variation of offshore buses for the Case

2 fault.

Figure 46 – Voltage variation of offshore buses for the Case 2

fault.

Figure 47 – Offshore wind turbine speed variation for the Case

2 fault.

Figure 48 – Frequency variation of onshore buses for the Case

2 fault.

Figure 49 – Voltage variation of onshore buses for the Case 2

fault.

Figure 50 – Speed variation of machines on the onshore grid for the Case 2 fault.

0 0.5 1 1.5 2 2.5 3 3.5 449.95

50

50.05

50.1

50.15

50.2

50.25

50.3

50.35

Time [s]

Fre

qu

en

cy

[H

z]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 20

Bus 102

0 0.5 1 1.5 2 2.5 3 3.5 40.85

0.9

0.95

1

1.05

1.1

1.15

1.2

1.25

Time [s]

Vo

ltag

e A

bso

lute

Valu

e [

p.u

.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

Bus 20

0 0.5 1 1.5 2 2.5 3 3.5 4

2999.7

2999.8

2999.9

3000

3000.1

3000.2

3000.3

3000.4

3000.5

Time [s]

Mac

hin

e S

pee

d [

rpm

]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 549.95

50

50.05

50.1

50.15

50.2

50.25

50.3

50.35

Time [s]

Fre

qu

en

cy

[H

z]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 0.5 1 1.5 2 2.5 3 3.5 40.8

0.85

0.9

0.95

1

1.05

1.1

1.15

1.2

Time [s]

Vo

lta

ge

Ab

so

lute

Va

lue

[p

.u.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

65

6.2.3. Case 3 results

Figure 51– Frequency variation of offshore buses for the Case 3

fault.

Figure 52– Voltage variation of offshore buses for the Case

3 fault.

Figure 53– Offshore wind turbine speed variation for the Case 3

fault.

Figure 54– Frequency variation of onshore buses for the

Case 3 fault.

Figure 55– Voltage variation of onshore buses for the Case 3

fault.

Figure 56– Speed variation of machines on the onshore grid

for the Case 3 fault.

0 0,5 1 1,5 2 2,5 3 3,5 4 4,5 549.8

49.9

50

50.1

50.2

50.3

50.4

50.5

50.6

Time [s]

Fre

qu

en

cy [

Hz]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 20

Bus 102

0 0,5 1 1,5 2 2,5 3 3,5 4 4,5 50.4

0.5

0.6

0.7

0.8

0.9

1

1.1

Time [s]

Vo

lta

ge A

bs

olu

te V

alu

e [

p.u

.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

Bus 20

0 1 2 3 4 5 6 7 8 9 102970

2975

2980

2985

2990

2995

3000

3005

3010

3015

3020

Time [s]

Ma

ch

ine

Sp

ee

d [

rpm

]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

0 0,5 1 1,5 2 2,5 3 3,5 4 4,5 5 5,5 6

50

50.1

50.2

50.3

50.4

50.5

50.6

50.7

Time [s]

Fre

qu

en

cy

[H

z]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

Time [s]

Vo

ltag

e A

bso

lute

Valu

e [

p.u

.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 62995

3000

3005

3010

3015

3020

3025

3030

3035

Time [s]

Mach

ine S

peed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

66

6.2.4. Case 4 results

Figure 57– Frequency variation of offshore buses for the Case 4

fault.

Figure 58– Voltage variation of offshore buses for the Case 4

fault.

Figure 59– Offshore wind turbine speed variation for the Case 4

fault.

Figure 60– Frequency variation of onshore buses for the Case 4

fault.

Figure 61– Voltage variation of onshore buses for the Case 4

fault.

Figure 62– Speed variation of machines on the onshore grid for

the Case 4 fault.

Figure 63 – Reactive Power injected/absorbed by the STATCOM on bus 20.

0 1 2 3 4 5

49.95

50

50.05

50.1

50.15

50.2

50.25

50.3

50.35

50.4

Time [s]

Fre

qu

en

cy

[H

z]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 20

Bus 102

0 1 2 3 4 50.5

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

Time [s]

Vo

lta

ge

Ab

so

lute

Va

lue

[p

.u.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

Bus 20

0 1 2 3 4 5 6 7 8 9 102980

2985

2990

2995

3000

3005

3010

3015

Time [s]

Ma

ch

ine

Sp

ee

d [

rpm

]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

0 1 2 3 4 5

50

50.1

50.2

50.3

50.4

50.5

50.6

50.7

49.9

Time [s]

Fre

qu

en

cy

[H

z]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 1 2 3 4 50.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

1.2

Time [s]

Vo

ltag

e A

bso

lute

Valu

e [

p.u

.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

Bus 21

0 1 2 3 4 5 62995

3000

3005

3010

3015

3020

3025

3030

3035

Time [s]

Mach

ine S

peed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

0 1 2 3 4 5-50

-25

0

25

50

75

100

Time [s]

Reacti

ve P

ow

er

[MV

Ar]

67

6.2.5. Discussion

Observing the results, some similarities are encountered between the figures obtained for the

different analysed cases. An important conclusion that can be drawn is that the behaviour of the

offshore wind farm is largely governed by the behaviour of the wind turbines, and, more specifically,

the DFIG machines modelled.

When the fault occurs, the electromagnetic torque drops, caused by the terminal voltage drop of

the generators. This causes the machine to accelerate, a behaviour noticed in all cases under

analysis. When the fault is cleared (either 100ms or 300ms after the application), the machine speed

decreases until it reaches a stable value close to the initial. For case 3 and 4 (Figure 53 and Figure

59), the rotor speed variation does not reach the initial value within the 10 seconds of the simulation,

unlike cases 1 and 2 (Figure 41 and Figure 47). This is caused by the aggravated severity of the short

circuit applied in these cases, which also causes the machine to accelerate to a much higher speed

than the speed in cases 1 and 2. The speed variation of the machines in the onshore grid is very

similar for all four cases: the machines accelerate during the fault and then decelerate following the

disturbance.

The electrical frequency in the offshore buses also increases, followed by the stabilization to the

initial 50 Hz. Similar results are found for onshore buses, for all the cases studied.

The voltage dips in offshore and onshore buses present some differences in the cases. Cases 1

and 2 present voltage dips less profound than the ones in Cases 3 and 4, due to the difference in the

short-circuit severity. In Cases 1 and 2 the voltage at the offshore buses reach approximately 0,9 p.u.

(Figure 40 and Figure 46). In Case 3, the voltage at the offshore wind farm buses reach about 0,6 p.u.

In Case 4 the voltage decreases to approximately 0,65 p.u.

The influence of the STATCOM installed offshore can be seen in case 4. The figures resemble

the ones in case 3, since the fault applied is the same. However, the STATCOM provides the voltage

support it was supposed to. The voltage dip in the offshore buses (in Figure 58) is less severe

(approximately 0,65 p.u.) than the voltage dip in case 3 (Figure 52) (0,6 p.u.). The wind turbine speed

increases slightly less than in the case 3. The effect of the STATCOM in the onshore buses (Figure

60, Figure 61 and Figure 62) is not significant, derived from the fact that the STATCOM is placed

offshore. Figure 63 shows the reactive power injected/consumed by the STATCOM, which explains

the reason for the improved behaviour of the offshore buses as a result of the action of the

STATCOM. This device is consuming reactive power before the fault (working as a reactor). During

the fault, the STATCOM injects reactive power, thus contributing to the rise of voltage in all offshore

buses, making the dips less profound. Following the clearance of the fault, the STATCOM slowly

tends to consume reactive power, as in steady-state.

In all four cases, the recovery of the voltage magnitudes, the machine frequencies and the

machine speeds indicate that stability is reached in all four cases, regardless of the differences in the

behaviour of the cases. This means that the faults to which the grid is subjected allow the fault ride

through of the wind farm, as these faults are not severe enough to allow the disconnection of the

68

machines by the under/over voltage relays. If more severe faults and/or with a larger fault duration

(outside the characteristic of the voltage relays, defined in Figure 26) were to be applied to the grid,

the disconnection of the machines would occur, thus having the wind farm disconnected from the grid.

This case was not, however, studied in this work.

69

6.3. HVDC Dynamic Behaviour

For HVDC transmission between the offshore wind farm and the grid the study carried out is

similar to the one for HVAC transmission. In this case, the fault applied is a 100 ms moderate fault

(similar to the one applied in Case 1, for HVAC transmission). Figure 70 and Figure 71 are present

only for HVDC transmission as they show the active and reactive power in the converters.

6.3.1. HVDC Configuration

In this section the results for the HVDC+STATCOM are presented. Note the additional figure,

Figure 80, which represents the reactive power injected or absorbed by the STATCOM.

Figure 64– Frequency variation of offshore buses with HVDC-LCC

transmission.

Figure 65– Voltage variation of offshore buses with HVDC-LCC

transmission.

Figure 66– Offshore wind turbine speed variation with HVDC-LCC

transmission.

Figure 67– Frequency variation of onshore buses with HVDC-

LCC transmission.

Figure 68– Voltage variation of onshore buses with HVDC-LCC

transmission.

Figure 69– Speed variation of machines on the onshore grid with

HVDC-LCC transmission.

0 0.1 0.2 0.3 0.4 0.50

50

100

150

200

Time [s]

Fre

quen

cy [

Hz]

Bus 90001

Bus 90002

Bus 90003Bus 90004Bus 90005

Bus 102

0 0.5 1 1.5 2 2.5 30

0.2

0.4

0.6

0.8

1

1.2

1.4

Time [s]

Vo

ltag

e A

bso

lute

Valu

e [

p.u

.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.53000

3000.05

3000.1

3000.15

3000.2

3000.25

Time [s]

Mac

hine

Spe

ed [

rpm

]

Bus 90001

Bus 90002Bus 90003Bus 90004

Bus 90005

0 1 2 3 4 5 6 7 8 9 1049.8

49.85

49.9

49.95

50

50.05

50.1

50.15

50.2

Time [s]

Fre

qu

en

cy [

Hz]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

0 0.5 1 1.5 2 2.5 30.8

0.85

0.9

0.95

1

1.05

1.1

1.15

Time [s]

Voltage A

bsolu

te V

alu

e [p.u

.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

0 1 2 3 4 5 6 7 8 9 102985

2990

2995

3000

3005

3010

Time [s]

Mach

ine S

peed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

70

Figure 70 – Active Power in the converters, for the HVDC-LCC

transmission.

Figure 71 – Reactive Power in the converters, for the HVDC-LCC

transmission.

0 0.5 1 1.5 2 2.5 3-150

-100

-50

0

50

100

150

Time [s]

Ac

tiv

e P

ow

er

[MW

]

Rectifier

Inverter

0 0.5 1 1.5 2 2.5 30

10

20

30

40

50

60

70

80

Time [s]

Re

ac

tiv

e P

ow

er

[Mv

ar]

Rectifier

Inverter

71

6.3.2. HVDC + STATCOM Configuration

Figure 72– Frequency variation of offshore buses with HVDC-

LCC+STATCOM transmission.

Figure 73– Voltage variation of offshore buses with HVDC-

LCC+STATCOM transmission.

Figure 74– Offshore wind turbine speed variation with HVDC-

LCC+STATCOM transmission.

Figure 75– Frequency variation of onshore buses with HVDC-

LCC+STATCOM transmission.

Figure 76– Voltage variation of onshore buses with HVDC-

LCC+STATCOM transmission.

Figure 77– Speed variation of machines on the onshore grid with

HVDC-LCC+STATCOM transmission.

0 0.1 0.2 0.3 0.4 0.50

50

100

150

200

Time [s]

Fre

quen

cy [

Hz]

Bus 90001

Bus 90002

Bus 90003Bus 90004

Bus 90005

Bus 102

0 0.5 1 1.5 2 2.5 30

0.2

0.4

0.6

0.8

1

1.2

1.4

Time [s]

Vo

lta

ge

Ab

so

lute

Valu

e [

p.u

.]

Bus 90001

Bus 90002

Bus 90003

Bus 90004

Bus 90005

Bus 102

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.53000

3000.05

3000.1

3000.15

3000.2

3000.25

Time [s]

Mac

hine

Spe

ed [

rpm

]

Bus 90001

Bus 90002Bus 90003Bus 90004Bus 90005

0 1 2 3 4 5 6 7 8 9 1049.8

49.85

49.9

49.95

50

50.05

50.1

50.15

50.2

Time [s]

Fre

qu

en

cy [

Hz]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

0 0.5 1 1.5 2 2.5 30.8

0.85

0.9

0.95

1

1.05

1.1

1.15

1.2

1.25

Time [s]

Voltage A

bsolu

te V

alu

e [p.u

.]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

Bus 151

0 1 2 3 4 5 6 7 8 9 102985

2990

2995

3000

3005

3010

Time [s]

Mach

ine S

peed

[rp

m]

Bus 101

Bus 211

Bus 3011

Bus 3018

Bus 206

72

Figure 78 –Active Power in the converters, for HVDC-

LCC+STATCOM transmission.

Figure 79 – Reactive Power in the converters, for HVDC-

LCC+STATCOM transmission.

Figure 80 – Reactive Power injected/absorbed by the STATCOM on bus 151.

0 0.5 1 1.5 2 2.5 3-150

-100

-50

0

50

100

150

Time [s]

Acti

ve P

ow

er

[MW

]

Rectifier

Inverter

0 0.5 1 1.5 2 2.5 30

10

20

30

40

50

60

70

80

Time [s]

Reacti

ve P

ow

er

[Mvar]

Rectifier

Inverter

0 1 2 3 4 5 6-80

-70

-60

-50

-40

-30

-20

-10

0

Time [s]

Re

acti

ve P

ow

er

[MV

Ar]

73

6.3.3. Discussion

From the results for HVDC and HVDC+STATCOM transmission options, the most relevant

conclusion is that the offshore wind farm connected in such ways does not possess fault ride through

capability, i.e., the wind farm does not remain connected to the grid during the disturbance. Rather,

the voltage drop causes commutation failure in the inverter, which means that the current does not

transfer from one semiconductor switch to another. The consequence of the commutation failure is the

bypassing of the inverter by short-circuiting its input and opening its output. The bypassing occurs at

time 0,502 s – Figure 70 and Figure 78 depict this behaviour (similar in both transmission systems).

During the disturbance, with the inverter bypassed, the rectifier continues to circulate a lower level of

direct current through the shorted inverter, resulting in a lower active power drawn by the converters

as a consequence of the voltage dip (represented in Figure 65 and Figure 73). Active and reactive

power at the rectifier and the inverter continue to flow during the fault until the AC voltage at the

rectifier finally goes to zero, at which time the DC link is blocked, at 0,602 s. The active and reactive

power circulated through the converters become equal to zero.

As the frequency of the offshore buses increases very significantly (as noted in Figure 64 and

Figure 72) the over frequency generator bus disconnection relays try to disconnect the wind

generators. The under-voltage relays also step in, due to the voltage dip in generator buses. Finally,

the offshore machines trip, the voltage at the offshore buses drops to zero (Figure 65 and Figure 73).

After the disconnection of the machines, the frequency values are not relevant, so the time interval

shown in Figure 64 and Figure 72 is only of approximately 0,5 seconds. At this point, all the buses

located offshore are isolated from the onshore grid, as they are disconnected.

From the grid point of view, the wind turbine speed (Figure 66 and Figure 74) is not relevant after

the fault so the time interval does not include the time period following the clearance of the fault, since

the machines are disconnected.

From the point of view of the onshore grid, the disconnection of the offshore wind farm is a loss of

a 110 MW generator. As so, the frequency of the onshore buses increases but it does not reach the

value it had before the disturbance, 50 Hz. This is caused by the fact that the governors of the

onshore generators have first order control and so the reference level of power is the same before and

after the fault, even though the amount of active power has changed. This means that the frequency

establishes in a lower value, approximately 49,85 Hz (Figure 67 and Figure 75). Similar reasoning is

applied to the speed of the onshore machines (Figure 69 and Figure 77). The voltage drops at the

onshore buses are, much like all the other results here analysed, identical for both transmission

alternatives. The voltage drops, on bus 151, to about 0,8 p.u. (Figure 68 and Figure 76).

All the results are analysed together for the HVDC and the HVDC+STATCOM transmissions as

the differences between these two are negligible. The conclusion is that the use of the STATCOM

does not improve the performance of the transmission system. The objective was that this device

would inject reactive power in the onshore bus (bus 151) during and immediately after the grid

disturbance and thus contributing to a less severe voltage dip in the bus at which it is connected.

74

However, as can be verified in Figure 80, the STATCOM is already injecting the maximum reactive

power it can supply before the fault and so its action is very limited, as it does not have the ability of

injecting more than 70 Mvar. As such, the benefits of the STATCOM to the dynamic behaviour of an

HVDC-LCC are not significant enough to consider the HVDC+STATCOM a viable alternative to the

HVDC transmission. The utilization of a larger STATCOM could, in theory, solve this limitation.

The disconnection of the wind farm during the fault becomes the only viable solution due to the

operating problems of the LCC converters. Limited improvements are available: increasing the

minimum extinction angle of the inverter is possible although it represents an increased reactive power

consumption of the converter.

It is noticeable that the behaviour of the wind farm is largely influenced by the behaviour of the

HVDC-LCC transmission technology, and, it this particular case, the limitation it has when dealing with

commutation failures.

The reconnection of this offshore wind farm, following the grid fault, is not simulated in this work,

but it is likely that this reconnection leads to severe transient phenomena.

75

7. Conclusions

In this chapter, the main conclusions of the work are presented, together with a reference to

future studies that may offer important future contributions to the theme of this thesis.

7.1. Final Remarks

The development of global wind energy over the last 15 years has been remarkable. In particular,

Europe has invested strongly in new installations. In the last two years (2008 and 2009) more wind

power was installed than any other generating technology in the EU. The significantly growing wind

energy has given rise to the creation of offshore wind farms. In just 10 years the installed capacity of

offshore wind energy has reached 2 GW and the previsions indicate this energy source will experience

a breathtaking increase, with forecasted scenarios of 40 GW of installed capacity by 2020.

The higher wind speed offshore, the available area of installation and the possibility of moving

away from population possibly disturbing wind turbines have been the main driven forces behind the

recent developments on offshore wind energy. As trends indicate that future wind farms will be larger

and further from shore, the transmission technology from the offshore wind farm to the onshore grid

has gained relevance and HVAC and HVDC solutions are under analysis.

The objective of this thesis was to compare two transmission technologies, HVAC and HVDC-

LCC, by carrying out steady-state and dynamic analysis of the offshore wind farm and the onshore

grid for both transmission alternatives. The offshore wind farm was modelled in PSS/E software as a

110 MW wind power plant placed 100 km from the shore.

The power flow analysis showed that the steady-state performance of HVAC and HVDC is

satisfactory, even though different operation issues arise for each technology.

For HVAC the main issue is the large amount of reactive power produced at the submarine cable.

As such, some compensation schemes were studied, in order to absorb the surplus reactive power.

The application of a compensator device of some sort (a reactor or a STATCOM) located offshore was

found to be necessary. The scenario of onshore compensation only (no offshore device) led to an

unacceptable voltage rise on the offshore bus voltage, as a result of the excessive amount of reactive

power. The performance of the remaining compensation schemes was similar: the compensation at

both ends of the cable, the compensation offshore only (by a shunt reactor) and the STATCOM

offshore compensation all lead to a high power factor at the point of connection of the wind farm to the

remaining grid (around 0,9, injecting reactive power in the grid). The application of a STATCOM

offshore leads to almost the same results as the shunt reactor offshore. The offshore and onshore

compensation alternative led to the most favourable results: lowest reactive power injected in the

onshore grid and an even distribution of the cable loading (low power rating of the cable). The need of

smaller shunt reactors (40 Mvar) are also an advantage of this alternative.

76

The HVDC link allows the transmission of the power generated at the wind farm with low power

losses (around 2%) and no reactive power problems are originated by this technology. However, both

line-commutated converters (rectifier and inverter) always consume reactive power, which can be

compensated. The use of a STATCOM (as part of the HVDC+STATCOM alternative considered)

provides that compensation, supplying reactive power to the onshore grid.

The dynamic behaviour of the offshore wind farm and the onshore grid was studied for the

response of this power system to an AC grid fault. In this case, the results differ very much according

to the transmission system.

The offshore wind farm connected to the grid by HVAC presents favourable results, as the wind

farm is able to stay connected to the grid during the disturbance. This fact is valid for all three bus

faults applied (differing in duration and severity) which means this wind farm is able to “ride through”

the fault. The performance is improved by the application of the STATCOM, as this device injects

reactive power during the fault, reducing the voltage dips and also reducing the frequency rise and

offshore wind turbines speed increase.

The HVDC transmission, however, gives totally different results. The AC onshore grid fault

causes commutating failure of the inverter, which is bypassed instantly after the disturbance. The

tripping of the machines leads to the blocking of the DC link. The offshore buses are disconnected

from the grid and the power generation is lost. This means that the converter technology governs the

behaviour of the wind farm, since the limitations of the LCC converters cause the wind farm to be

disconnected. The disconnection during the fault and the subsequent reconnection after the fault is the

solution for operating one such system, although the reconnection is not analysed in the course of the

thesis. The STATCOM does not improve the performance of the transmission, since it is already, in

steady-state, supplying reactive power to the grid, not having, therefore, the capability of injecting

further amounts of reactive power (since it is already injecting initially the maximum reactive power it

can).

In conclusion, one can state that if HVAC is used to implement the grid connection, the response

of the wind farm to disturbances is determined by the wind turbines, since the connections themselves

are passive elements. With the DC link, the wind turbines are electrically decoupled from the analysed

system. As so, the response of the wind farm is governed by the technology in the LCC converter.

7.2. Future Work

The modelling of the wind farm may be the starting point to future simulations, with the possibility

of exploring other options, resolving some of the limits found on the course of this thesis work.

The limited action of the STATCOM that has been encountered may be overcome by using larger

devices, with larger capability of reactive power injection. Two STATCOM devices working together

can also be considered, if limited space is not a problem.

77

For HVDC transmission, in particular, some auxiliary power devices may be used to absorb the

surplus active power when an AC grid fault occurs, and by doing so, reducing the frequency instability

of the offshore grid. Large batteries and more technologically advanced solutions, like

Superconducting Energy Storage (SMES) systems can be considered in the future.

Applying this study to different scenarios can also be advantageous to the knowledge of these

systems. Working with a larger wind farm, installed further from the shore can be considered. The

power losses are likely to be higher and, in AC, the longer cable will need to have larger compensation

schemes. For increased distances, the use of HVDC is likely to be more favourable, since the HVAC

will reach its limit.

Finally, one of the most important advances in HVDC transmission, the HVDC-VSC (voltage

source converters) may well prove to be a valid alternative for the connection of offshore wind farms.

This technology, owing to the turn-off capability of the IGBTs of the converters, allows total control of

active and reactive power, which could represent fault ride through capability for offshore wind farms

connected with HVDC-VSC.

78

References

[1]. EWEA - European Wind Energy Association. www.ewea.org. [Online] [Cited: 26 April 2010.]

[2]. GWEC - Global Wind Energy Council. www.gwec.net. [Online] [Cited: 25 June 2010.]

[3]. Castro, Rui. Folhas da Cadeira de Energias Renováveis e Produção Descentralizada. 2008.

[4]. Freris, Leon. Renewable Energy in Power Systems. s.l. : Wiley, 2008.

[5]. Manwell, James F., McGowan, Jon G. and Rogers, Anthony L. Wind Energy Explained:Theory,

Design and Application. s.l. : Wiley, 2002.

[6]. Wind Atlases of the World. www.windatlas.dk. [Online] [Cited: 24 April 2010.]

[7]. Danish Energy Authority. www.ens.dk. [Online] [Cited: 26 April 2010.]

[8]. Delivering offshore wind power in Europe - EWEA Report. www.ewea.org. [Online] [Cited: 29 April

2010.]

[9]. E-ON Offshore Factbook. www.eon.com. [Online] [Cited: 29 April 2010.]

[10]. Enquist, Johan. Fault Ride-Through of Offshore Wind Parks - Master of Science Thesis. s.l. :

Chalmers University of Technology, 2007.

[11]. Ackermann, Thomas. Wind Power in Power Systems. s.l. : Wiley, 2005.

[12]. Xu, Lie and Andersen, Bjarne R. Grid Connection of Large Offshore Wind Farms Using HVDC.

s.l. : Wiley Interscience, 2005.

[13]. Lazaridis, Lazaros P. Economic Comparison of HVAC and HVDC Solutions for Large Offshore

Wind Farms under Special Considerations of Reliability - Master's Thesis. s.l. : Royal Institute of

Technology, 2005.

[14]. Alegría, Iñigo Martinez de, et al. Transmission alternatives for offshore electrical power. s.l. :

Elsevier, 2008.

[15]. Negra, N. Barberis, Todorovic, J. and Ackermann, T. Loss Evaluation of HVAC and HVDC

transmission solutions for large offshore wind farms. s.l. : Elsevier, 2006.

[16]. Zubiaga, M., et al. Evaluation and selection of AC transmission lay-outs for large offshore wind

farms. s.l. : IEEE, 2009.

[17]. Nexans. www.nexans.com. [Online] [Cited: 7 May 2010.]

[18]. Siemens Energy. www.energy.siemens.com. [Online] [Cited: 18 May 2010.]

79

[19]. Wiechowski, W. and Eriksen, P. Selected Studies on Offshore Wind Farm Cable Connections -

Challenges and Experience of the Danish TSO. s.l. : IEEE, 2008.

[20]. Wright, Sally D., et al. Transmission Options for Offshore Wind Farms in the United States. s.l. :

AWEA, 2002.

[21]. Kirby, N. M., et al. HVDC Transmission for Large Offshore Wind Farms. s.l. : IEEE, 2002.

[22]. Prysmian Cables. www.prysmian.com. [Online] [Cited: 25 May 2010.]

[23]. High Voltage Direct Current Transmission - Proven Technology for Power Exchange.

www.energy.siemens.com. [Online] [Cited: 15 March 2010.]

[24]. Sucena Paiva, José Pedro. Redes de Energia Eléctrica - Uma Análise Sistémica. s.l. : IST

Press, 2005.

[25]. Technical Description Lillgrund Wind Power Plant. s.l. : Vattenfall, 2008.

[26]. PSS/E Wind Modelling Package for 1.5/3.6/2.5 MW Wind Turbines - User Guide. 2009.

[27]. PSS/E 30 Users Manual. 2004.

[28]. PSS/E 30 Program Operation Manual Volume II. 2004.

[29]. PSS/E 30 Program Application Guide Volume II. 2004.

80

Appendix A – Grid Parameters

This appendix presents the parameters used in the test grid, represented in Figure 24.

A.1. Onshore Grid Parameters

A.1.1. Line Parameters

Table 10: Reactance and Susceptance of the lines.

Line E[

[X. Y. ]

F[

[X. Y. ] From

Bus

To

Bus

153 154 0,0375 0,08333

153 154 0,045 0,125

153 3006 0,01 0,025

154 203 0,0333 0,08333

154 205 0,00278 0,075

154 3008 0,01833 0,25

203 205 0,0375 0,06667

203 205 0,0375 0,06667

3001 3003 0,00667 0

3002 3004 0,045 0,075

3003 3005 0,045 0,075

3003 3005 0,045 0,075

3005 3006 0,025 0,05833

3005 3007 0,02083 0,05

3005 3008 0,04167 0,1

3007 3008 0,02083 0,05

A.1.2. 400 kV Line Parameters

Table 11: 400 kV line calculated parameters.

Line H

[AB]

@AB

[S/AB]

EAB

[S/AB]

FAB

[S/AB] @X.Y. EX.Y. FX.Y. From

Bus

To

Bus

151 152 100 0,0292154 0,3215977 3,514E-06 0,00183 0,0201 0,5623

151 152 100 0,0292154 0,3215977 3,514E-06 0,00183 0,0201 0,5623

151 201 50 0,0292154 0,3215977 3,514E-06 0,00091 0,01005 0,28115

152 202 80 0,0292154 0,3215977 3,514E-06 0,00146 0,01608 0,44984

81

152 3004 70 0,0292154 0,3215977 3,514E-06 0,00128 0,01407 0,39361

201 202 50 0,0292154 0,3215977 3,514E-06 0,00091 0,01005 0,28115

201 204 40 0,0292154 0,3215977 3,514E-06 0,00073 0,00804 0,22492

A.1.3. Transformer Parameters

Table 12: Two-Winding Transformers Parameters.

Buses E[ [X. Y. ] From

Bus To

Bus 101 151 0,01133

152 153 0,00417

201 211 0,01771

202 203 0,01354

204 205 0,0125

205 206 0,01111

3001 3002 0,0125

3001 3011 0,00833

3004 3005 0,01354

3008 3018 0,07083

A.1.4. Loads

Table 13: Load powers.

Bus _H []

_H [\_]

153 200 100

154 500 450

154 400 350

203 200 100

205 1000 700

3005 100 50

3007 200 75

3008 100 50

A.1.5. Shunt Compensators

Table 14: Shunt compensators.

Bus F`Y []

151 0

153 150

82

154 300

201 300

203 100

205 300

A.2. Offshore Wind Farm Parameters

A.2.1. Wind Turbines

Table 15: Wind Turbine aggregate Power Flow Parameters.

24 MW Aggregates

21 MW Aggregates

Pgen [MW] 24 21

Pmax [MW] 24 21

Pmin [MW] 1,1 1

Qgen [Mvar] 11,62 10,16

Qmax [Mvar] 11,62 10,16

Qmin [Mvar] 0 0

Mbase [MVA] 26,72 23,38

Xsource [p.u.] 0,8 0,8

A.2.2. Step-up transformers

Table 16: Unit transformer of one single wind turbine.

Transformer Parameters Value

Unit Rating [MVA] 1,75

Unit Rated Voltage [kV/kV] 0,69/33

Unit Impedance [%] 5,75

Unit X/R 7,5

83

A.2.3. 33 kV Cables

Table 17: 33 kV Horns Rev cable parameters [10].

Cable Parameters Value

Resistance [Ω/km] 0,042

Reactance [Ω/km] 0,11

Capacitance [µF/km] 0,28

Table 18: 33 kV AC Cable parameters used in the modelled wind farm.

Cable Length

[AB] @X.Y. EX.Y. FX.Y. From

Bus To

Bus

1 102 1 0,003857 0,010101 0,000957

2 102 0,8 0,003085 0,008081 0,000766

3 102 0,6 0,002314 0,006061 0,000574

4 102 0,4 0,001543 0,00404 0,000383

5 102 0,2 0,000771 0,00202 0,000191

A.3. HVDC-LCC Modelling Parameters

A.3.1. Steady-State DC Line Parameters

Table 19: DC Line Power Flow Parameters.

Parameter Value

Line Number 1

Control Mode Power

Rdc [Ω] 2,0

Rcmp [Ω] 0,0

Delti [p.u.] 0,0

Setval [MW] 110

Vschedule [kV] 150

Dcvmin [kV] 0,0

Vcmode [kV] 145

Metered Rect

CCC Itmax 20

Ccc Accel 1,0

84

A.3.2. Steady-State Converter Parameters

Table 20: DC Link Converters Power Flow Parameters.

Parameter Rectifier Inverter

Bus Number 102 151

Max Firing Angle [º] 18 19

Min Firing Angle [º] 8 18

Bridges in Series 1 1

Primary Base [kV] 33 400

Commutating Resistance [Ω] 0 0

Commutating Reactance [Ω] 14,22 13,20

Measuring Bus 0 0

Trans Ratio [pu] 3,36 0,28

Tap Setting [pu] 0,91875 0,91875

Max Tap Setting [pu] 1,5 1,5

Min Tap Setting [pu] 0,51 0,51

Tap Set [pu] 0,00625 0,00625

Reactance of CCC [Ω] 0 0

AC Tx From Bus 0 0

AC Tx To Bus 0 0

AC Tx Id 1 1

A.3.3. Transient Model Parameters

Table 21: “CDC4T” model parameters for the DC Link.

Parameter Value

ALFDY [º] 8

GAMDY [º] 18

TVDC [s] 0,05

TIDC [s] 0,05

VBLOCK [pu] 0,6

VUNBL [pu] 0,65

TBLOCK [s] 0,1

VBYPAS [kV] 75

VUNBY [pu] 0,8

TBYPAS [s] 0,1

RSVOLT [kV] 145

RSCUR [A] 73,33

VRAMP [pu/s] 5

CRAMP [pu/s] 5

85

C0 [A] 0

V1 [kV] 100

C1 [A] 586,67

V2 [kV] 147

C2 [A] 733,33

V3 [kV] 150

C3 [A] 733,33

TCMODE [s] 0,1

A.4. STATCOM Parameters

A.4.1. Steady-State Parameters

Table 22: STATCOM Power Flow parameters.

Parameter Rectifier

Device Number 1

Terminal Bus 0

Control Mode Normal

P Setpoint [MW] 0,00

Q Setpoint [MVAR] 0,00

V send setpoint 1,02

Shunt Max [MVA] 64

RMPCT [%] 100

Bridge Max 0

V term max [pu] 1,1

V term min [pu] 0,9

V series max [pu] 1

I Series Max [MVA] 64

Dummy Series X [pu] 0,05

V Series Reference Sending end Voltage

86

A.4.1. Transient State Parameters

Table 23: “CSTCNT” model parameters for the STATCOM.

Parameter Value

T1 [s] 0,1

T2 [s] 0,1

T3 [s] 0,1

T4 [s] 0,1

K 25,116

Droop [pu] 0,03

Vmax 999

Vmin -999

Icmax [pu] 1,25

ILmax [pu] 1,25

Vcutout [pu] 0,2

Elimit [pu] 1,2

Xt [pu] 0,1

Acc 0,5

STBASE [MVAr] 64

87

Appendix B – GE 1,5 MW Wind turbine aggregate “dyr”

record file

The following data was used for the dynamic simulation of one aggregate of GE 1,5 MW wind

turbines. Similar records were written for the other aggregates.

90001 'USRMDL' 1 'GEWTG1' 1 1 2 11 3 5 0 14 1.5000 0.8000 0.50000 0.90000 1.1100 1.2000 2.0000 0.40000 0.80000 5.0000 0.20000E-01/ 90001 'USRMDL' 1 'GEWTE1' 4 0 10 62 18 7 90001 0 0 1 0 0 1 0 0 0 0.15000 18.000 5.0000 0.0000 0.0000 0.50000E-01 3.0000 0.60000 1.1200 0.40000E-01 0.43660 -0.43600 1.1000 0.20000E-01 0.45000 -0.45000 5.0000 0.10000 0.90000 1.1000 40.000 0.50000 1.4500 0.50000E-01 0.50000E-01 1.0000 0.15000 0.96000 0.99600 1.0040 1.0400 1.0000 0.95000 0.95000 0.40000 1.0000 0.20000 1.0000 0.25000 -1.0000 11.000 25.000 3.0000 -0.90000 8.0000 0.25000 10.000 1.0000 1.7000 1.1100 1.2500 5.0000 0.0000 0.0000 10.000 0.25000E-02 1.0000 5.5000 0.10000 -1.0000 0.10000 0.0000 / 90001 'USRMDL' 1 'GEWTT' 5 0 1 5 4 3 0 5.2900 0.0000 0.0000 1.8800 2.3000 / 0 'USRMDL' 0 'WGUSTC' 8 0 3 6 0 4 90001 '1 ' 0 9999.0 5.0000 30.000 9999.0 9999.0 30.000 / 0 'USRMDL' 0 'GEWTA' 8 0 3 9 1 4 90001 '1 ' 0 20.000 0.0000 27.000 -4.0000 0.0000 1.2250 35.250 72.000 1000.0 / 0 'USRMDL' 0 'GEWTP' 8 0 3 10 3 3 90001 '1 ' 0 0.30000 150.00 25.000 3.0000 30.000 -4.0000 27.000 -10.000 10.000 1.0000 / 0 'USRMDL' 0 'GEWTPT' 8 0 2 0 0 17 90001 '1 ' / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.1 5.0 0.2 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.2 5.0 0.733 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.3 5.0 1.267 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.4 5.0 1.8 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.5 5.0 2.333 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.6 5.0 2.867 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.7 5.0 3.4 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.8 5.0 3.933 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.9 5.0 4.467 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.0 1.1 0.1 0.08 / 0 'USRMDL' 0 'VTGTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 0.0 1.15 0.0 0.08 / 0 'USRMDL' 0 'FRQTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 47.0 55.0 0.02 0.08 / 0 'USRMDL' 0 'FRQTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 47.5 55.0 10.0 0.08 / 0 'USRMDL' 0 'FRQTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 45.0 51.25 30.0 0.08 / 0 'USRMDL' 0 'FRQTPA' 0 2 6 4 0 1 90001 90001 '1' 0 0 0 45.0 52.0 0.02 0.08 /