You are on page 1of 96

Delft University of Technology

Offshore Wind Plants with VSC-HVDC Connection and their Impact on the Power
System Stability: Modeling and Grid Code Compliance

Ndreko, Mario

Publication date
2016
Document Version
Peer reviewed version

Citation (APA)
Ndreko, M. (2016). Offshore Wind Plants with VSC-HVDC Connection and their Impact on the Power
System Stability: Modeling and Grid Code Compliance. Delft University of Technology.

Important note
To cite this publication, please use the final published version (if applicable).
Please check the document version above.

Copyright
Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent
of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons.
Takedown policy
Please contact us and provide details if you believe this document breaches copyrights.
We will remove access to the work immediately and investigate your claim.

This work is downloaded from Delft University of Technology.


For technical reasons the number of authors shown on this cover page is limited to a maximum of 10.
Offshore Wind Plants with VSC-HVDC
Connection and their Impact on the Power
System Stability: Modeling and Grid Code
Compliance

Author: Mario Ndreko

M.Sc Course EE4545: Electrical Power Systems of the Future

25/4/2016

ESE-IEPG

Delft University of Technology

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
This report is restricted to be used only within TU Delft for the
course EE4545 Power Systems of the Future
Table of Contents
Chapter 1: Introduction to Typical Grid Connection Requirements ......................................... 9
1.1 Requirements for high voltage DC connected wind power plant modules ............... 9
1.1.1 Frequency stability requirements....................................................................... 9
1.2 Requirement for the onshore HVDC converter station ............................................. 9
1.2.1 Requirements for the frequency range and rate of change of frequency
withstand capability ........................................................................................................... 9
1.2.2 Fault ride through (FRT) capability................................................................. 10
1.2.3 Requirements for the reactive current supply during voltage sags .................. 10
1.2.4 Requirements for asymmetrical faulted conditions ......................................... 11
1.2.3 Reactive Power Capability ..................................................................................... 11
Chapter 2 ................................................................................................................................. 13
Wind turbine models for transient stability studies ................................................................. 13
2.1 Introduction ................................................................................................................... 13
2.2 DFIG wind turbine model (Type 3)............................................................................. 14
2.2.1 Generic stability type model ................................................................................... 15
2.3 Full converter direct drive wind turbine model (Type 4) ............................................ 16
2.3.1 Converter interface model ...................................................................................... 18
2.3.2 Converter control module ....................................................................................... 19
2.3.3 Mechanical part model ........................................................................................... 19
2.3.4 Pitch angle control model ....................................................................................... 20
2.3.5 Aerodynamic model ............................................................................................... 21
Chapter 3 ................................................................................................................................. 22
VSC-based HVDC transmission system model for offshore wind power grid connection
studies ...................................................................................................................................... 22
3.1 Introduction ................................................................................................................... 22
3.2 HVDC technology overview for transmission systems ................................................. 22
3.2.1 Classic HVDC technology ..................................................................................... 22
3.2.2 VSC-HVDC technology ......................................................................................... 23
3.2.3 VSC topological considerations ............................................................................. 25
3.2.4 HVDC Submarines cables ...................................................................................... 26
3.2 VSC-HVDC transmission system modeling method .................................................... 27
3.2.1 Model of the HVDC cables in stability studies ...................................................... 28
3.2.2 DC Grid System Power Flow Equations ............................................................. 30
3.2.3 EMT Average-value Model of the onshore converter ............................................ 31

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
3.2.4 RMS value model of the onshore converter ........................................................... 33
3.2.5 EMT Average-value model of the offshore converter station ................................ 34
3.3 LVRT control Logic ................................................................................................ 36
3.4 Methods for tuning VSC controllers ............................................................................. 37
3.4.1 Inner controller tuning ............................................................................................ 37
3.4.2 DC voltage controller tuning .................................................................................. 38
3.5 Control of the VSC-HVDC system in point to point connection .................................. 38
3.6 Control of the VSC-HVDC system in multi-terminal DC grid connection................... 39
3.7 Power based direct voltage droop control ..................................................................... 40
3.8 Assessment of LVRT Control Strategies Using EMT Average Value Models ....... 40
3.8.1 Chopper method assessment with EMT average value model ............................... 42
3.8.2 Artificial offshore grid voltage reduction method ........................................... 44
Chapter 4: Effect of FRT Grid Code Compliance and Reactive Current Injection by VSC-
HVDC on the Transient and Voltage Stability of the Power System ...................................... 49
4.1 VSC-HVDC Dynamic Model............................................................................................ 49
4.2 Selected control parameter sensitivities for evaluation of system impact ......................... 49
4.3 Test system ........................................................................................................................ 50
4.5 Simulation Results ............................................................................................................. 51
4.5.1 Sensitivity of the reactive current boosting gain k ..................................................... 51
Chapter 5: Analysis on Frequency Stability Requirements ..................................................... 55
5.1 Introduction ....................................................................................................................... 55
5.2 Single country HVDC connected wind power plant ......................................................... 56
5.2.1 Single machine equivalent grid model of the AC system .......................................... 56
5.2.2 Hydraulic turbine and governor model for the single machine equivalent model...... 57
5.2.3 Synthetic Inertia emulation controller for offshore wind PPM .................................. 58
5.2.4 Sensitivity of the Kinertia gain of synthetic inertia emulation controller .................. 59
5.2.5 Synthetic Inertia emulation controller using the stored in DC capacitors electro-static
energy .................................................................................................................................. 62
5.2.6 Combined method which uses both wind plant and DC capacitors contribution ....... 64
5.3 Synthetic inertia response by a multi-terminal DC grid .................................................... 65
5.3.1 Effect of Kinertia controller gains .................................................................................. 67
5.4 Impact of reduced system inertia on frequency response of New England test system .... 68
5.5 Participation of MTDC grid in the load-frequency control of power systems –
Demonstration with IEEE 39 bus test system model .............................................................. 71
5.6 Discussion and Conclusions .............................................................................................. 76
References ............................................................................................................................... 77

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Appendix A – Benchmark Power System Model Parameters ................................................. 85
Appendix B - PSS/E type 3 (DFIG) variable speed wind turbine sub-models [Source: PSS/E
User Manual] ........................................................................................................................... 89
Appendix C – Detailed differential equations of the VSC model ........................................... 92

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
List of Acronyms

AC Alternative current
AVR Automatic Voltage Regulator
DC Direct current
DFIG Double Fed Induction Generator
DFAG Double Fed Asynchronous Generator
FCDD Full converter Direct Drive
FRT Fault Ride Through
GSVSC Grid side voltage source converter
HVDC High Voltage Direct Current
IGBT Insulated Gate Bipolar Transistor
MTDC Multi-terminal Direct Current
MPPT Maximum Power Point Tracking
PCC Point of Common Coupling
PI Proportional Integral controller
PLL Phase Lock Loop
PWM Pulse Width Modulation
VSC Voltage Source Converter
OWP Offshore Wind Park
WPVSC Wind Park voltage source converter

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
List of Symbols

I Pcmd Wind turbine active current command (p.u)


I Qcmd Wind turbine reactive current command (p.u)
EQcmd Wind turbine q-axis component of the voltage (p.u)
ref
I Pcmd Wind turbine active current command reference (p.u)
ref
IQcmd Wind turbine reactive current command reference (p.u)

I P max Wind turbine maximum active current limit (p.u)


I Q max Wind turbine maximum reactive current limit (p.u)
Pelec Wind turbine terminal active power (p.u)
Qelec Wind turbine terminal reactive power (p.u)
Pmech Wind turbine mechanical power at the rotor shaft (p.u)
Pwtt Wind turbine terminal active power (p.u)
Qwtt Wind turbine terminal reactive power (p.u)
Paero Wind turbine aerodynamic power given by linear aerodynamic model(p.u)
init
Paero Wind turbine aerodynamic power, initial value (p.u)
Taero Wind turbine aerodynamic torque (p.u)
Kaero Wind turbine, gain for linear aerodynamic model (p.u)
Pref Wind turbine active power reference set-point (p.u)
Tgen Wind turbine generator electrical torque (p.u)
H wtr Wind turbine rotor inertia (s)
H gen Wind turbine generator inertia (s)
Dshaft Wind turbine damping factor for the torsional oscillation
wtt Wind turbine terminal voltage angle (rad)
uwtt Wind turbine terminal voltage (p.u phasor)
ref
u wtt Wind turbine terminal voltage reference (p.u phasor)
iwtt Wind turbine terminal current (p.u phasor)
wtr Wind turbine rotor rotational speed (p.u)
gen Wind turbine generator rotational speed (p.u)
TRV Time constant of the voltage sensor (s)
T fp Time constant of the active power filter (s)
dPmax Maximum active power rate of change (p.u)
dPmin Maximum active power rate of change (p.u)
init Wind turbine speed initial value (p.u)
cmd Pitch angle command given by the pitch angle controller (deg)
 pitch Pitch angle (deg)
Uc Voltage Source Converter terminal voltage before the phase reactor (p.u
phasor)
Us AC voltage at the grid connection point terminal (p.u phasor)

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
I pr Terminal current of the VSC (p.u phasor)
U dc DC voltage at the dc side of the converter station, across the dc capacitor
I dc DC current injection by the VSC
U c ,ref reference voltage of the Voltage Source Converter terminal before the
phase reactor (p.u)
I pr ,ref Reference of the VSC terminal current that flows through the phase
reactor (p.u phasor)
Pdc DC power at the dc terminal of the Voltage Source Converter
Pac Active power at the AC terminal of the Voltage Source Converter (p.u)
Qac Reactive power at the AC terminal of the Voltage Source Converter (p.u)
uc ,abc Instantaneous value of the VSC terminal line voltage behind the phase
reactor
us ,abc Instantaneous value of the grid connection point terminal voltage
uc , Instantaneous value of the VSC terminal voltage in the fictitious two
phase system
us , Instantaneous value of the grid connection point terminal voltage in the
fictitious two phase system
iabc Instantaneous value of the current of phase reactor
ucd d-axis component of the VSC terminal voltage before the phase reactor
(p.u)
u cq q-axis component of the VSC terminal voltage before the phase reactor
(p.u)
id d-axis component of the phase reactor current at the VSC terminal (p.u)
iq q-axis component of the phase reactor current at the VSC terminal (p.u)
idref Reference of the d-axis component of the phase reactor current at the VSC
terminal (p.u)
iqref Reference of the q-axis component of the phase reactor current at the VSC
terminal (p.u)
ucref,q Reference of the d-axis component of the VSC terminal voltage before the
phase reactor (p.u)
ucref,d Reference of the q-axis component of the VSC terminal voltage before the
phase reactor (p.u)
Rdc Resistance of the HVDC cable
Ldc Inductance of the HVDC cable
Cdc Capacitance of the HVDC cable

r Phase reactor resistance


L Phase reactor Inductance
U owp Phasor of the voltage at the offshore wind power park (p.u phasor)
I owp Phasor of the current of the wind power park (p.u phasor)

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Chapter 1: Typical Grid Connection
Requirements

1.1 Requirements for high voltage DC connected wind power plant modules

1.1.1 Frequency stability requirements


Fast active power response requirements refer to the ability of generation and HVDC
transmission units to contribute to the load-frequency control of the power system. The reader
should recall that the frequency in the power system is a continuously controlled variable. It is
primarily affected by the balance between generation and demand [1]. Normally, large
conventional synchronous generation units are equipped with governors are capable to
provide load frequency control.
The installed generation capacity from wind power plants has increased during the last
years. Transmission system operators in order to ensure the power system security of supply
require from large wind plants to provide fast active power response in order to contribute in
the load frequency control. Given that the wind turbines during normal operation does not
have power reserves, they normally are required to response only during over-frequency
events. Hence, AC connected wind plants are required by TSOs to reduce their active power
generation in order to support the frequency stability of the power system. Details about such
requirement can be found analytically in [2].
In the case that VSC-HVDC transmission system is used for the grid connection of the
offshore wind power plants, the onshore power system frequency cannot be detected from the
offshore wind turbines due to the DC circuit decoupling, unless communication link is
utilized. The article 37 of the Network Code (NC) HVDC [3] as it is given by ENTSOE,
stipulates that an HVDC connected wind power plant shall have access to the remote onshore
system frequency where the onshore converter station is connected by means of a fast
communication link. Fiber-optic is the most commonly used technology. It is embedded in the
offshore HVDC cable which connects the onshore and the offshore converters.
1.2 Requirement for the onshore HVDC converter station

1.2.1 Requirements for the frequency range and rate of change of frequency
withstand capability

Any voltage source HVDC converter station, utilized for the grid connection of the
offshore wind power plants shall remain connected to the power system within a pre-specified
frequency ranges by the relevant TSO. This frequency ranges are described in European level
in the network code (NC) HVDC [3]. The range of continuous operation for an HVDC
transmission system lies between 47.0Hz to 52Hz. Furthermore, the HVDC transmission
system shall be capable to withstand frequency rate of change between -2.5Hz/s and 2.5Hz/s.
In addition, the HVDC system and the connected wind power plants shall be capable to
provide a fast active power response (within seconds). The fast active power response of
power electronic interfaced generation and transmission units is commonly referred in the
literature as synthetic inertia or artificial inertia response [4]. Some interesting references for
inertia response from power electronic interfaced generation units can be found in the works
presented in [5], [6], [7]. For the particular case of offshore wind plants with HVDC
transmission, more information is given in references [4], [8], [9].

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
1.2.2 Fault ride through (FRT) capability
Based on NC-HVDC [3] the HVDC converter station shall demonstrate capability remain
connected to the power system during voltage sags down to zero value ensuring stable
operation during the fault and the post-fault period. The relevant transmission system operator
(TSO) is responsible to provide a voltage-against-time profile with pre-defined parameters
that describe the lower limit (or the boundaries) of the line-to-line rms voltage at the
connection point. The latter is given as a function of time in which the converter station shall
stay grid connected. A generic low-voltage-ride-though (LVRT) profile is given in figure 1.

U (p.u)

1 Blocking
allowed
Urec

Ublc
Trip allowed
Uret

Tfault Tclear Tblc Trec


t(s)

Figure 1. Low-voltage–ride-through profile of a HVDC converter station. The diagram expresses the lower limit
of the voltage-against-time profile at the connection points, expressed by the ratio of its actual value and its
nominal value in p.u before, during and after the fault [3].

The variables of this LVRT profile are decided by each TSO. Normally, TSOs based on the
dynamic faulted response (using dynamic simulations) of the voltage at their system, define
the voltage profile for the worst case scenario. The LVRT and FRT envelope has to be
designed such as it is always below the dynamic response of the simulated voltage.
The ability of the wind power plant and their HVDC transmission system to remain
connected while complying with the LVRT curves has been an extensive research topic. The
main problem for the HVDC link used for grid connection of the offshore wind plants is the
over-voltages that appear in the DC link during the fault period.
For the interested reader information about grid code compliance strategies can be found in
[10], [11], [12] and [13]. Beside the grid code compliance strategies, the impact of the FRT
on the power system stability is discussed in [14], [15], [16]. The development of offshore
HVDC grids (termed as multi-terminal DC grid) has raised new challenges about the FRT
compliance of the MTDC grids with FRT and low voltage ride though (LVRT) requirements
[17], [18].

1.2.3 Requirements for the reactive current supply during voltage sags
As discussed in the previous section, under AC system balanced faulted conditions the
onshore HVDC converter station shall remain connected to the system providing a fast acting
(in ms scale) additional reactive short circuit current injection [3]. The later aims to provide a
voltage support to the power system and enhance its fault and post-fault voltage stability. It
will be discussed later in this report that the reactive current injection is beneficial for the
rotor angle stability of interconnected power systems.
The magnitude of the reactive short circuit current should be provided to the grid as a
function of the rms voltage deviation at the grid connection point (see figure 2). In many grid
codes for wind power generation units there is an explicit description of this value [2]. In

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
addition, as described in the NC-HVDC [3], each onshore converter station connected to an
AC power system should be capable of providing at least 2/3 of the rated reactive current for
a given time period decided by the relevant TSO.
The slope of the curve is a variable which the relevant TSO need to specify and it is termed
as ‘k’ gain or “k-factor” as also termed in literature [2]. This is in fact a proportional gain
applied at the outer control loops which relates the reactive current injection to the per unit
voltage deviation [12]. In addition, a voltage set point Uo is needed to cover specific operating
ranges. Finally, if needed a dead band (i.e +/-5% or 10%) shall be selected in the range of
nominal network voltage. Following changes in voltage, the HVDC converter shall be
capable of achieving 90% change in reactive power in a predefined by the relevant TSO rise
time. Normal values of this rise time would be less than 20ms, which are typical time
constants of the power electronic modules.

additional reactive
current
I B / I N
reactive Current gain:

I B / I N
U  U 0  / U N 2  k  10
k with deadband
without deadband
Rise time<20ms

-0.5 -0.25 0 0.2


Voltage Drop
deadband

-x x
U U0  / UN
(overexcited U N : rated voltage
operation) k 2 I N : rated current
U : voltage during fault
-1 I B : reactive current during fault
I B 0 : reactive current before fault
U o : prefault voltage
I B  I B  I B 0
Figure 2. Reactive current boosting slope for voltage support at point of common coupling by HVDC system
during grid faults. This represents a typical curve as would be given at national level grid code requirements.

1.2.4 Requirements for asymmetrical faulted conditions


Given the future large penetration of power electronic based generation and VSC-HVDC
transmission units in the power system, there have been lately proposals to extend the FRT
requirements not only for balanced but also for unbalanced grid faults. In this direction, the
ability to control the negative sequence current of the converter station is considered as a need
in order to assist the protection schemes in the AC transmission lines. It is common practice
today to suppress the negative sequence injection of the converters in order to protect them
from high unbalanced currents. However, this may create malfunctions in the protection
schemes and difficulty to detect line-to ground faults.

1.2.3 Reactive Power Capability


The reactive power capability of the HVDC converter station shall be defined by the
relevant TSO by a given U-Q/Pmax profile. These profiles describe the boundary of reactive
power that each HVDC converter station shall provide at its maximum power capacity. Please
note that these requirements refer to normal operation of the plant and not to faulted response.
Figure 3 introduces the U-Q/Pmax profile as described in NC-HVDC and consists of an outer
fixed envelope and an inner envelope which is to be decided for each synchronous area by the
relevant TSO. The position, size and shape of the inner envelope should be any rectangular

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
but always within the limits of the outer envelope. Finally, the HVDC system shall have the
ability to operate and continuously move to any operating point within the inner envelope in
appropriate time scales to values requested by the relevant TSO.

U [p.u]

Fixed outer envelope


1.118

Voltage
Range
Range of
Q/Pmax

Inner envelope
(different from TSO to TSO)
0.85

Consumption (lead) Production (lag)


-0.5 +0.65 Q/Pmax
0

Figure 3. Boundaries of U-Q/Pmax profile of HVDC system voltage at the grid connection point [3].

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Chapter 2

Wind turbine models for transient stability studies

2.1 Introduction

This chapter introduces generic dynamic models of variable speed wind turbine generators
which can be used to simulate the dynamic response of offshore wind power plants in large
power system studies [19], [20]. Variable speed wind turbines have been considered the
strongest candidate for the development of offshore wind [21], [22].
A lot of research has been conducted in the past decades for the development of standard
and generic wind turbine models, suitable for bulk power system stability type simulations
[23], [19], [24], [25], [26]. The necessity for such model raise from the fact that detailed
electromagnetic type models (i.e in PSCAD/EMTDC) cannot be used in large system stability
studies mainly for two reasons. First, they are computationally expensive, and second, they
cannot be easily initialized as is the case in stability type studies [1]. In that sense the power
system community has spent efforts in order to simplify detailed EMT type towards stability
type models [23].
The benefit for the development of standard and generic dynamic models is three-fold.
First it makes easier the data and parameter exchange among the different involved parties in
grid connections studies. Second it simplifies the development of wind turbine models in
different commercial software packages (i.e PSS®E, DigSilent/PowerFactory), and third, it
provides a mechanism by which model parameters can be selected in such a way that they
represent the dynamic response of a particular wind power plant [19]. Project partners and
involved manufactures can validate these standard models with their detailed wind power
plant models without confidentiality and intellectual property concerns.
The proposed models use positive sequence RMS value approach, suitable for bulk power
system stability studies, as proposed in the IEC 61400-27. They are capable of reproducing
steady state and dynamic phenomena under different events, such as short circuits (low
voltage ride through) at the AC terminals, loss of generation or load and system separation.
Also the models present different modes of electromechanical oscillations and dynamic
voltage phenomena. The typical simulation time frame of these models is in the range of 10-
30s and could be used for frequency related studies in the range of 0.1 to 10Hz. The
initialization is done in accordance with the grid model, mainly with respect to the load flow
in the benchmark power system.
The main assumptions that this modeling approach applies is that very small time
constants, such as dynamics of the power electronic converters and power electronic
commutation phenomena are neglected. Wind turbines are modeled by fundamental

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
frequency voltages and currents. Finally, the models are modular, consisting of different sub-
models which can be changed or upgraded separately. In the next paragraph, a detailed model
description will be given for type 3 and type 4 wind turbines.

2.2 DFIG wind turbine model (Type 3)

The stator of Double Fed Induction Generator (DFIG) wind turbine, also known as Double
Fed Asynchronous Generator (DFAG), is directly connected to the grid while its rotor is
connected via a back-to-back converter. The main duty of the back-to-back converter is to
adjust the current frequency in the rotor windings. Thus a variable speed operation is achieved
following the Maximum Power Point tracking (MPPT). Figure 4 illustrates the DFIG wind
turbine single line block diagram. This type of wind turbine is referred in the IEC standards as
type 3 generic wind turbine.
Blades
And Rotor
Gear
Box Asynchronous
Generator

= ~
C
~ =
CR

Figure 4. Double fed induction generator (DFIG) wind turbine

For emergency situations, mainly during fault transients, the DFIG wind turbines include a
crowbar (abbreviated as CR in figure 2). The crowbar, is a combination of resistances which
bypass the rotor side converter in order to prevent damages during fault transients, typically
when voltage dips starts and when voltage dips is cleared. However, the ignition of the
crowbar reduces significantly the generator controllability since the rotor of the machine is
short circuited and the machine is acting as an induction motor consuming reactive power.
Such an approach has been abandoned due to the fact that new grid codes demand for
reactive power and voltage support during the fault period. Hence, most of the modern DFIG
wind turbines utilize new control methods for fault ride through which is based on DC
choppers and coordinated control of the grid side and machine side converter. The method
using the DC chopper is making the utilization of the crowbar unnecessary. Thus, the
consumption of reactive power during the AC faults is avoided.
Contrary, reactive current could be dynamically controlled in order to support the terminal
voltage. Such an approach introduces larger flexibility in the grid code compliance from the
DFIG turbines. Grid codes are the set of requirements that the wind turbine has to fulfil in
order to be connected to the system. In the end, due to the assumption that the DC voltage can
be maintained approximately constant during the simulation time, models of the DC link of
the DFIG turbine and crowbar are not considered in this report.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
2.2.1 Generic stability type model
The generic model of the type 3 wind turbine is based on GE’s detailed wind turbine model
as given in [26]. This type 3 generic model is also available in PSS®E, both the source file
and the compiled (.dll) file which can be directly used in dynamic simulations. The model
consists of four sub-models as given in figure 3. Namely, the generator electrical model, the
converter control model, the wind turbine model (two-mass) which also includes a linear
aerodynamic model and the pitch angle control model. The block diagram for each sub-model
of the type 3 wind turbine model is given in Appendix A. More details can be found in [27].
In the stability type simulations model, the dynamics of the electromechanical part of the
generator and generator flux dynamics are eliminated as suggested in [26] and only the
response of the converter to the controller is included. The converter control model consists of
the active power control sub-model and the reactive power control sub-model which gives
voltage and current commands to the generator/converter model.

Vrfq Vterm

Ipcmd
Converter Generator
Control Eq cmd Electrical
Model Model
Pgen, Qgen

P ord ωref ωg Pgen

Pitch
Control Wind
Model ωt Turbine
Model

Pmech
Linear
Aerodynamic
Model
Blade
Pitch
Angle, θ

θ0

Figure 5. Block diagram showing the signal routes of the type3 wind turbine generic model [28]

The reactive power controller can either be operated in constant reactive power control
mode or can utilize an AC voltage controller, emulating the operation of the AVR system.
The active power order is derived from the generator speed and the speed reference which is
obtained by the Maximum Power Point Tracking curve or Power versus Speed curve. A
typical curve as given by GE and applied in this report simulations is given in figure 4. A two
mass model is considered for the representation of the wind turbine mechanical shaft (see in
appendix, figure E5). The aerodynamic power (Pmech) is given by a simplified linear
aerodynamic model which takes as input the pitch angle, as it can be seen in Appendix A.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Optimum active power production vs rotor speed
1
MPPT
0.9

0.8

0.7

Power Set - point (p.u)


0.6

0.5

0.4

0.3

0.2

0.1

0
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3
Rotor Speed (p.u)

Figure 6. Maximum power point tracking curve [26] ( x-axis is rotor speed reference, ωref, and y-axis is the
generated active power, Pgen as it can be found in Appendix E)

2.3 Full converter direct drive wind turbine model (Type 4)

Unlike the DFIG wind turbines, the full converter direct drive (FCDD) variable speed
wind turbine technology applies a back-to-back voltage source converter in order to connect
the generator to the power grid [20]. The generator is often a permanent magnet but an
induction generator could also be applied. Different vendors apply different machine
technologies. The direct drive full converter wind turbine is considered the most reliable
solution due to the elimination of the gearbox, while the permanent magnet is preferable to
the electrically excited option since it allows for black-start capability. The full (back-to-back)
converter of the FCDD decouples the power system frequency from the generator speed,
allowing a wide range of variable speed operation. An advantage is the decoupled operation
of the generator from the grid transients and faults. In that prospect the chopper is used. A
disadvantage is that such wind turbine has zero contribution to the total system inertia. The
total system inertia is defined as the amount of kinetic energy available at rotating machines
and generators. Since there is full decoupling by the AC-DC-AC converter, the kinetic energy
is not available unless adequate control loops are designed to provide it if needed as
frequency response. More information can be found in [24].

synchronous
Generator
uwtt T
= ~ 3
3

CH =
~
Pwtt , Qwtt
Generator Line
Rotor Side Side
Converter Converter

Figure 7. Full converter direct drive wind turbine, CH stands for the DC chopper

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
For the back-to-back converter, different control strategies are implemented depending
mainly on the manufacturer. For example, a common strategy applied in FCDD turbines is
that the Line Side Converter controls the active and reactive power delivered to the grid while
the Generator Side Converter is responsible for variable speed operation and the voltage
control of the DC bus. The IEC 61400-27-1 [20] gives specific recommendations for the
modeling of FCDD turbines referred as type 4. The most important assumption in this
modeling approach for stability type simulations is that the active and reactive power
response is mainly determined by the high bandwidth power electronic modules, which
involve very small time constants which are beyond the scope of dynamic simulations and
can be neglected.
Consequently, only the dynamics of the line side converter controllers are important in
stability type simulations of large scale power systems. Furthermore, the presence of the DC
chopper in the DC bus of the back-to-back converter makes the AC faults and other LVRT-
triggering phenomena invisible to the generator side. GE report [26] suggests that for stability
type simulations, the mechanical part of the model should not be included, because torsional
oscillations are not present. The main reason for this assumption is the activation of the DC
chopper at the DC link of the FCDD turbine during low voltage ride through phenomena. It
practically keeps the DC voltage constant during disturbance. Hence, the generator is
operating in normal steady state condition. The later makes the AC side disturbances invisible
to the synchronous generator. Although, this is valid for some vendors such as GE [26],
Enercon and Vestas [28], there are other manufacturers, like Siemens [28] that do not include
choppers in their turbine and thus the AC side disturbance may trigger torsional modes of
electromechanical oscillations [29]. In order to simulate the dynamic response for these
FCDD turbines, type 4B model is applied. In this approach the wind turbine mechanical part
is modeled by a two-mass model [27] [20]. Figure 8 introduces the block diagram of the type
4B standard model.

ref
uwtt
uwtt

ref
iPcmd
Converter ref
Control iQcmd
Converter
Model Model
Pwtt , Qwtt
init Pwtt
 gen
Pitch
Wind
Control
wtr Turbine
Model
Model
(2 mass)
Linear
Aerodynamic Paero
 pitch Model

Figure 8. Block diagram showing the signal routes of the type 4B wind turbine generic model. The reader is
referred to the list of symbols for clarification of symbols meaning.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
2.3.1 Converter interface model

In stability type simulations, the Thevenin approach is usually followed to represent the
interface between the power electronic interfaced generation units and the power system. In
this approach, the wind turbine is modelled as a current phasor the amplitude and angle of
which is calculated by the control loops. This sub-model as it is presented in figure 9 is the
interface between the power system and the wind turbine model.

iP max
 wtt uwtt  wtt
ref
iPcmd 1
1  sTg iPcmd dq
From Pwtt , Qwtt
Control 0 Current I wtt
Module iQ max Limiter

iQcmd xy
ref
iQcmd 1
1  sTg Reference
frame
iQ min rotation

Figure 9. Block diagram of the type 4 wind turbine converter model. In this model, Tg is the time constant
delay associated with the converter PWM modulation.

As it can be seen, the model is developed in the rotating dq-reference frame. It receives as
inputs the active and reactive current commands in the rotating dq-reference frame which are
termed here as iPcmd and iQcmd. These references are provided from the control module and are
used to calculate the phasor of the terminal current by making use of the dq to xy
transformation.

ix  cos  wtt  sin  wtt  iPcmd 


i     
 y   sin  wtt cos  wtt  iQcmd 

I wtt  ix  ji y

The voltage angle of the wind turbine terminal, which is used for the transformation is
assumed to be instantly available and thus the phase lock loop (PLL) dynamics are not
modeled. The PLL is the control block which is responsible to lock the angle of the voltage at
the grid connection point, θwtt.
Finally, a current limiter is included in the model in order to reduce the active and reactive
current component of the current during AC disturbance to values that respect the overcurrent
capacity of the wind turbine line side converter. The strategy applied to reduce the current
injection to the grid at acceptable levels by the line side converter during a low voltage ride
through determines the active and reactive power response of the FCDD turbine. In addition,
wind power plant modules will be required to inject fault current to the network in support of
voltage dips. Different strategies and sensitivities in the reactive current boosting by the wind
turbine will be discussed during the trade of analysis and is not the objective of this report.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
2.3.2 Converter control module

The block diagram of the control module is illustrated in figure 10. It includes both closed
and open loops for the reactive power control as well as the voltage regulation. It also
includes closed loop control of active power which is determined by the rotor generator speed
in order to model the mechanical oscillations that are transferred to the grid. The initial value
of the rotational speed is determined mainly by the MPPT curve that the FCDD turbine
utilizes (see figure 6). The speed of the generator is calculated from the mechanical model of
the wind turbine which uses a two mass model [20] . The reference value of the active power
is calculated from the power flow. The output of the controllers is sent to the converter model
as real and reactive current commands. The reader should recall that in stability type models
the wind speed is considered to be constant during the simulation and the generated wind
power is assumed to match the power injected to the grid.

uwtt 1
From
Wind turbine 1  sTRV
module
 gen
dPmax x / ref
iPcmd
Pref 1
  
x x 1  sT fp x To converter
/ Module
dPmin

init

qwtt

uwtt 1 0
max
qwtt k pQ
-
k pQ
1  sTRV +
0
  +

1 + ref
iQcmd
q min
kiQ +
wtt
1 k+iQ + To converter
s Module
s
Pwtt 1 x -
 
1  sTpf
x
+

tan  ref
uwtt
Figure 10. Converter control block diagram for type 4 wind turbine stability type model.

2.3.3 Mechanical part model

A two mass model is used to represent the dynamic behavior of the mechanical part of the
wind turbine. This model is capable to produce torsional modes of shaft oscillations. The
initial value of the speed is determined by the MPPT curve. The generator electrical torque,
Tgen, is calculated from the speed of the generator and the generator active power [20]. The
damping of the torsional oscillating mode can be regulated by selection of the damping factor
Dshaft. In steady state the generator speed is equal to the rotor speed. In the case of a
disturbance, such as three phase fault in the wind turbine terminal, the torsional mode will be
excited. These oscillations are observed in the generator and rotor speed. Depending on
generator speed oscillation amplitude it can be transferred to the wind turbine terminal active
power by the model of the active power controller in figure 8 (top).

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
/ T
aero
1 +
Paero x
   wtr
+ 2 H wtr s
From +
Aerodynamic -
module

+
init
D shaft  Pref
-
MPPT
curve

Tgen + +
x 1
Pwtt - 
   gen
2 H gen s +
From
/
Converter
module

Figure 11. Wind turbine two-mass model block diagram.

2.3.4 Pitch angle control model

The pitch angle controller is responsible to regulate the blades pitch angle. The practical
application of the pitch control is that when available wind power is above the equipment
ratings, the blades are pitched to limit the mechanical power delivered to the shaft. This is
achieved by reducing the aerodynamic coefficient, increasing the blades angle. In the case
that the available power is less than the rated, the pitch angle is zero. The latter maximizes the
aerodynamic coefficient and t the aerodynamic power delivered from the wind turbine shaft.
Figure 10 illustrates the model of the pitch controller for type 4B wind turbine. In this model,
the blade position actuators are rate limited (see Rθ variables) and there is a time constant Tθ,
associated with the translation of the blade angle to mechanical output. This time constant
simulates the behavior of the servo-motor drive system of the pitch regulator. The dynamics
of the pitch control are moderately fast, and can have a significant impact on the dynamic
response of the wind turbine model. As given by GE in [26], the speed reference is 1.2 p.u,
for power generation above 0.6 p.u but is reduced to lower level according to the MPPT curve
characteristic. An example of such MPPT curve is illustrated in figure 4. For simplicity, the
same MPPT curve is applied in the simulations reported here for both DFIG an FCDD. In this
curve, it can be seen that at operating power up to 0.6 pu, the wind turbine rotational speed is
set at 1.2 p.u, which is the maximum speed of the MPPT and the speed reference of the pitch
angle control model. In this range of power, a disturbance may accelerate the generator above
the given reference triggering the operation of the pitch controller.

k pp R max
 max
wtr + +  cmd
 
+
 T  pitch
From
- kip + - s
Wind turbine  min
module R min
s
ref
Figure 12. Pitch angle control block diagram

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
2.3.5 Aerodynamic model

The equation that gives the aerodynamic torque for the FCDD wind turbine model is given
by the well-known equation 2.1. If we linearize this equation at an initial operating point we
could obtain the equation for linear model as given in 2.2. In this equation the term P / 
and P / wtr are constant for the initial operating point of the system. P /  is negative
for all the wind speeds while P / wtr is positive for wind speeds below 1.3p.u of rated.
1
Paero   AC p   , Vw3 (2.1)
2

P P
Paero  Paero
init
   init   wtr  init  (2.2)
 init ,init wtr init ,init

Substituting P /  and P / wtr with Kaero1 and Kaero2 we get the final linear model in
figure 11, where the aerodynamic constants can be estimated or given as dynamic parameters
(from measurements done by the manufacturer).

wtr +
 K aero 2
From -
Mechanical Part
Model
init
+
 pitch +
 K aero1
+

+ init
Paero
From -
Pitch control
model
Paero
 pitch
init
To
Wind Turbine
Module
Figure 13. Linearized aerodynamic model block diagram.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Chapter 3

VSC-based HVDC transmission system model for offshore wind


power grid connection studies

3.1 Introduction
Historically, the development of semiconductor technology, in combination with advances
in micro-controllers have led to the high utilization of the voltage source converter (VSC)
technology for industrial applications. Voltage Source Converter (VSC) is a bidirectional
AC/DC converter which uses fully controllable switching devices (with gate turn-off
capability, such as IGBTs). Within the last decades, VSC technology has become applicable
to medium and high voltage transmission systems. VSC-based HVDC transmission systems,
as commonly referred in the literature, is at the moment the state of the art technology for the
transmission of bulk amount of power in large distances. The first generation of VSC-HVDC
utilized two and three level technology. Up-to-date the majority of the manufacturers (ABB,
Siemens and Alstom) provide the state of the art modular multi-level converter (MMC)
HVDC technology. Detailed description about VSC-HVDC technology for addition reading
can be found in [30], [31] .
In the North Sea countries, the integration of large-scale offshore wind power generation
location far from short can be facilitated by the use of VSC-based HVDC transmission
systems [21]. The main reasons for this choice are the technical and economic advantages that
such transmission systems provide [32]. Namely, capability to supply islanded networks with
constant voltage and frequency, capability to operate in weak power systems (with low short
circuit power levels and power system total inertia), capability to independently control active
and reactive power, and black start capability. A good reference for VSC-HVDC technology
can be found in [30].
This chapter discusses the modeling approach followed to simulate VSC-based HVDC
transmission systems, in power system control and stability studies focused on offshore wind
power integration. The focus will be placed to understand what are the most important control
modules which affect the response of the HVDC link, the response of the power system and
the grid code compliance during AC grid faults.

3.2 HVDC technology overview for transmission systems

3.2.1 Classic HVDC technology


The conventional technology for HVDC is based on the line-commutated based
converters (LCC). The LCC-HVDC technology is mainly used today for long distance
bulk point-to-point connection. LCC-HVDC is well established technology with the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
largest project being the Itaipu system in Brazil at 6300MW installed transmission
capacity. Even though there is a great number of applications in different countries LCC
technology has showed significant weakness. The commutation of the converter valves is
closely related to the stiffness of the ac voltage at the grid connection point. For that
reason, the converter does operate properly at connection points with low short circuit
power. The reason is that the LCC cannot create ac voltage itself. It is always necessary
for the operation of the LCC that the system provides the necessary voltage. Another
drawback for the case of LCC-HVDC system is that reactive power compensation is
needed. Significant amount of shunt reactive power compensation and harmonic filters are
required for operation. This makes the substations large, occupying big area making LCC
impractical for very compact sites. Figure 14 provides the components and layout of an
LCC-HVDC substation. In addition, the application of LCC is limited to one direction
current flow through each converter. Hence, in order to reverse the power flow of any
individual terminal the dc voltage polarity must be reversed as well. This is the main
reason why LCC based HVDC technology cannot be used in applications such as multi-
terminal offshore HVDC networks. In a multi-terminal dc network which applies LCC
technology, changing the polarity of the dc voltage for one dc line will change the power
flows in the network.

Figure 14. LCC-HVDC Converter [33]

3.2.2 VSC-HVDC technology

The solution to the above problem is given by the utilization of the more recent
version of HVDC technology known as the voltage source converter based high voltage
direct current technology (VSC-HVDC). The development of VSC technology is based
both on the improved performance and the increasing rating of the insulated gate bipolar
transistors (IGBT). Also important role plays the controllability that IGBTs illustrate by
means of capability to turn on and off and thus becoming a self-commutated rather than
line commutated converter. Furthermore, sinusoidal pulse width modulation technique
(SPWM) gives flexibility and improves the performance of operation generating less
harmonic distortion.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
With the use of the VSC-HVDC technology, there is no need to change the DC
voltage polarity in order to change the power flows, as it is the case of LCC technology.
The last characteristic is very attractive for implementation of the VSC-HVDC at
offshore HVDC networks. Furthermore, there is no need for reactive power
compensation and no need to install large filters to suppress harmonic distortion. As a
result, the converter station of VSC-HVDC is more compact in comparison to the LCC
technology with financial benefits on the construction of compact and flexible offshore
sub-stations. Last but not least, with VSC-HVDC it is possible to undergo stage
development of meshed ac-dc networks, with fast and cost efficient planning. Figure 15
provides the main layout of the VSC-HVDC substation.

Figure 15. VSC-HVDC converter substation [33].

The most important technical advantage of VSC-HVDC compared to the LCC is the
ability to produce its own voltage waveform independently from the ac network. VSC
can produce the required sinusoidal ac voltage by proper switching of the IGBTs, even
connected to a passive network. Furthermore, by utilization of suitable power
synchronization control, it is possible to connect VSC-HVDC systems to high
impedance weak ac systems, where the short-circuit level is low or even zero. Thus,
VSC-HVDC technology has only limited requirements on the short circuit power of the
system it connects to. On the contrary, it contributes limited short-circuit capacity to the
ac system during the AC faults. Moreover, it can provide black start and system
recovery capability in case of ac faults without losing its controllability. In addition, by
implementing proper control scheme, such as vector control along with PWM, the VSC
can adjust and control both active and reactive power independently. The latter is
important for dynamic voltage support at the point of common coupling with the ac
network. Also the ability to adjust active power gives inherent capability of frequency
control in the synchronous area it operates.
In conclusion, the VSC-HVDC transmission system is increasingly getting more
competitive in the field of transmission system technology and its applications are
extended from point-to-point connection of asynchronous systems to more complex
multi-terminal configurations, for the integration of large scale offshore wind power to
the onshore power system network.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
3.2.3 VSC topological considerations

A classic VSC converter consists of three legs as illustrated figure 16. Each leg is
identical to a single phase converter. There are separate controls signals for each leg. The
control signals are sinusoidal waveforms with 50Hz frequency, 120o shifted, as it can be
seen in Σφάλμα! Το αρχείο προέλευσης της αναφοράς δεν βρέθηκε.. The sinusoidal
signals are compared with a triangular high frequency carrier signal in order to create the
pulses that switch on and off the IGBTs. The higher the frequency of the carrier signal, the
less the harmonic distortion that appear. A typical value of the carrier frequency is in the
region of 2-4kH. The amplitude of the sinusoidal signal, also known as modulation index,
ma should be in region of (0<ma<1) for linear operation.

For each of the three legs, the ac voltage fundamental frequency component with respect
to point N from figure 2.5 can be given as:

Vcontrol , peak Vd V
 vAN 1, peak   ma d (2.1)
Vtri , peak 2 2

Thus the fundamental component of the output voltage depends on the dc voltage at the
input and the modulation index ma as long as the modulation is in the linear region (ma<1).
From the above it is possible to calculate the fundamental component of the line-to-line
RMS voltage such as:

3 3 Vcontrol , peak
 vLL 1, peak   v 1   v 1   v 1, peak  Vd  0.612maVd (2.2)
2 2 2 Vtri , peak

Figure 16. classical VSC technology [34].

It is important to notice that for stability studies and dynamic simulations in general, only
the fundamental frequency of the voltage is of main importance. Hence, the converter can
be represented as controlled in amplitude and phase voltage source with 50Hz.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Figure 17 PWM of VSC converter station [34]

3.2.4 HVDC Submarines cables


HVDC submarines cables have been used since 1950s for interconnection of
asynchronous systems via classic HVDC technology. Up to date there is a rapid market
development especially due to the interest from offshore wind industry. HVDC cables do
not “suffer” from the charging current limitations of ac cables. Transmission distances for
HVDC cables are considered high. In general, their operation voltages are higher than the
ac equivalent cables. The highest installed HVDC submarine cable up to date is operating
at ±500kV and has a conductor cross section 3000mm2 capable of carrying 1400MW per
cable (Kii channel project crossing Japan).

Each submarine cable consists of the conductor (or core). The main materials used for the
conductor is stranded copper or aluminium (for lighter applications). HVDC cables are
generally single core, but it can also be double cores allowing one bipole circuit to be laid
as a single cable (application in the NorNed cable). Surrounding the core is the insulation
material. The thickness of the insulation material depends on the operational voltage. A
metallic sheath and a plastic outer layer protect the cable and prevent moisture ingress.
Two types are widely used: the cross linked polyethylene (XLPE) and mass impregnated
paper (MI), both illustrated figure 18.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Figure 18. Marine cable technology [32]

3.2 VSC-HVDC transmission system modeling method


The time-averaged VSC modeling approach is commonly used to represent VSC-based
HVDC transmission systems in bulk power system control and stability studies [30]. Let
assume the time average approach for two-level HVDC converter. As it can be seen in figure
17, the voltage of the VSC after the phase reactor is a sinusoidal voltage with higher order
harmonics. Assuming the presence of only the fundamental voltage waveform, the EMT
average value model of the VSC in the AC side is a controllable three phase AC voltage
source, while at the DC side, it is a DC current injection in parallel with the model of the
HVDC cables (which will be discussed below). Figure 19 introduces the EMT time-averaged
model of the VSC as used for two and three level VSC.
As it can be seen, uc ,abc are the three-phase controllable instantaneous values of the
converter voltage behind the phase reactor. These three-phase voltages are controllable, in
amplitude and angle, by the modulation indexes, ma , mb , mc . Moreover, ia , ib , ic are the line
instantaneous currents in each phase of the phase rector, and us ,a , u s ,b , u s ,c are the line
voltages at the grid connection point. The DC side of the converter is represented as a DC
current injection, behind a DC capacitor. This DC current injection is calculated from the
power balance between the AC and DC side (conservation of the power) [36].
VSC
uc , a r L ia
+
uc , b r L ib
U dc us ,a
Cdc I dc uc , c r L ic
-
u s ,b

mc
U dc us ,c
2
mb
ma U dc
U dc 2
2

Figure 19. Time-averaged modeling approach for the VSC-HVDC system.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
A VSC-HVDC system for offshore wind power integration includes three sub-models,
namely: the VSC converters, the HVDC transmission cables and the Offshore Wind Park
(OWP) model. Figure 20 provides the model architecture and the way that signals are
communicated per module. Please note that the model is a mathematical description of the
real system and not the exact copy of it.
In this diagram, I dcj , U dcj represent the DC current injection and the voltage at the offshore
(j=1) and offshore terminal (j=2) of the HVDC link respectively. In addition, I pr is the phasor
of the current flowing through the phase reactor and U s is the phasor of the grid voltage at
the grid connection point. Finally, in the offshore wind park side, I owp is the phasor of the
current of the offshore wind park and U owp is the phasor of the voltage at the offshore wind
power plant terminal.

AC
1 Network
I owp I dc I 2
dc
I pr

OWP VSC HVDC cables VSC I=YV


State Space Model
U owp 1
U dc Us
U dc2

Figure 20. Block diagram showing signal routes and the architecture for the VSC-HVDC transmission system
model. [14]

As it can be seen from figure 20, the AC network side or grid side VSC (GSVSC) model
receives as input the DC terminal voltage and the phasor of the AC terminal voltage. The
output signals are the current injections in the DC side. At the AC side, the output signal is
the phase reactor current phasor. The calculation of the DC output is performed, as already
mentioned, by the conservation of power between the AC and the DC side of the VSC.
However, in the OWP side, the VSC provides the AC voltage reference to the offshore wind
power plant. Thus, the wind park VSC (WPVSC referred in this report) is always providing a
constant voltage reference in amplitude and frequency to the offshore wind power plant
modules. Within the following paragraphs a detailed description will be given for the models
of the VSC, HVDC cables and controllers which can be used for stability type studies.

3.2.1 Model of the HVDC cables in stability studies

The applied approach to represent the HVDC cables in power system stability and control
type studies is using lumped π-equivalent model given by state space differential equations in
time domain. It should be noted that here the DC capacitors of the VSC station are added to
the DC capacitor of the HVDC cable model. If at a DC node there is no converter connected
(i.e it is junction node), only the DC cable capacitance is used, as shown in figure 21.
Providing the state space model of each cable, it is possible to extend it to any DC grid
topology by making use of the incidence matrix, which analytically describes the DC network
topology.
A graphical overview of the MTDC grid transmission state space model is shown in figure.
22. In the proposed state space generic form, the state space vector x and the DC current in-
feed vector u is defined in (1) and (2) respectively. In this model, the information of the DC
grid topology is included in the matrix A according to (3)-(7). The input of the model is the
DC current in-feed vector u, while the output is the state space vector x. Vector u, includes

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
the DC current in-feed as computed from the converters model by conservation of power at
each node.
Matrix A and B are constant, defined at the initialization stage of the dynamic model,
according to (3)- (11). Hence, for each DC cable section there are three states: two referring
to the DC voltage at the sending and receiving end, and one referring to the current flowing
through each cable section. Normally, HVDC links are bipolar, however assuming that the
AC system remains balanced during disturbed conditions in the AC side, the DC circuit if
bipolar can be represented by a mono-polar circuit in power system stability studies with half
the total current.
Sending Receiving
End End

I dc(1) Ldc Rdc I dc(2)


VSC VSC
I cable
U dc(1) Cdc Cdc U dc(2)

Cdccable
Cdc  Cdc
VSC

2

Figure 21. π-equivalent model of a DC cable section.

I dc( n )

U dc( n ) U dc(1) I dc(1)

dx
 Ax  Bu
U dc(3) dt

I dc(3) U dc(2)
I dc(2)

Figure 22. State space model to represent a MTDC grid dynamic behaviour [33].
T
x  U dc  U dc  U dc       
n 1 n2 nm
1 2 n
I cable I cable I cable (1)
 1x ( m n )
T
    
u   I dc
1 2 n
I dc I dc (2)
  (1xn )

 a11
nn a12
n m
 (3)
A   21 22 
 mn
a a mm  ( n  m )( n  m )

a11nn  0nn (4)


T

  1 1 1    (5)
nm  M mn * diag  
a12    

  Cdc1 Cdc 2 
Cdcn  nn 

 1 1 1 
mn  diag 
a21   M mn (6)
 Ldc1 Ldc 2 Ldcm  mm

 Rdc1 Rdc 2 Rdcm 


mm  diag 
a22  (7)
 Ldc1 Ldc 2 Ldcm  mm

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
For a MTDC grid with n nodes and m branches, as in (5)-(7), Rdci, Ldci, with i=...m, is the
resistance and inductance of each DC cable section while Cdci is the sum of the converters and
cable capacitance at each DC node. The incidence matrix, M for a DC grid with n number of
nodes and m number of branches is a mxn matrix defined as:
 1 , if current flows from i to j
 (8)
M mn  i, j    1, if current flows from j to i
0, if there is no branch between i and j

The B matrix is written as:


 b11
nn
 (9)
B   21 
b mn  ( n  m ) xn

Where the sub-matrices are defined as:


 1 1 1  (10)
b11nn  diag  ... 
 dc1
C Cdc 2 Cdcn  nn

mn   0mn
b21 (11)

More detailed modelling of the cable (with more sections) is not considered necessary, at
least for frequency stability type simulations of bulk power systems. However, this would not
be the case for DC side disturbances, such as DC faults where a more detailed modelling of
the DC grid cables is necessary.

3.2.2 DC Grid System Power Flow Equations


An important detail for performing dynamic simulations is the initialization of the states
related to the DC grid model, for any given operating point which is of interest. Hence, a DC
grid power flow is necessary in order to define the steady state voltages and currents, and
properly initialize the MTDC grid model. In this paper, a simple Newton-Raphson power
flow script is developed in Matlab (for initializing the user defined Simulink model) and in
Python for initializing the PSS®E model. In this power flow the VSC stations are considered
lossless. For a DC grid with n number of nodes and m number of branches, first the
admittance matrix of the DC grid is computed for any given topology by making use of the
incidence matrix given in (8):
1 1
Ydcnn  MT diag ( ... )M
Rdc1 Rdcm (12)

Furthermore, the vector which contains all the DC node voltages is defined such as:
T
U dc  U dc(1) U dc(2) ... U dc( n 1) U dcslack  (13)
nx1

Given (12) and (13) the mismatch equations can be calculated as shown in (14):
T
 P1 P N 1 P slack 
F(U dc )  Ydcnn U dc   dc(1) ... dc(1) dc( n )  0 (14)
U dc U dc U dc  nx1

Considering the slack as being the last DC node (which means that for the slack bus the DC
voltage is known, while DC power injection is un-known) we define the state variable vector
such as.
T
X  U dc(1) ... U dc( n1) Pdcslack 
(15)
The reader should notice that more than one slack node could be used. Using a Newton-
Raphson iterative method the mismatch equation can be solved by:

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
U kdc1  U dc
k
 J 1F  U dc 
(16)
Where, J is the Jacobian matrix of F with respect to X. Hence, for any DC grid topology,
using the results of the DC grid power flow it is possible to initialize the states of the DC grid
state space model and the states related to the controllers of the VSC model (i.e inner and
outer). Finally, the reader should notice that due to the focus of this paper study, the operating
points of the converters are chosen in such a way that the DC voltage maximum values are
not violated.

3.2.3 EMT Average-value Model of the onshore converter


The onshore VSC module for HVDC application consists of the power electronic
interface, the modulator which generates the PWM pulses, the inner current controller which
is a current feedback loop and gives the voltage reference set points at the VSC, and finally,
the outer controllers.
Figure 23 illustrates the control subsystems that constitute the VSC station for HVDC
transmission system applications. In this picture iabc, is the instantaneous current flowing
through the phase reactor while Udc is the DC voltage of the converter. In the DC side, the
system consists of a DC chopper. It is mainly a controlled (in term of dissipated power)
resistor which ensures the protection of the HVDC system from over-voltages in the DC link
when faults are present in the AC terminals. Please note that in the case of an HVDC station
for offshore wind application, the power flow is from the DC to the AC side of the grid side
VSC. Hence, when balanced three phase faults occur in the AC side, the AC power drops to
zero. As a result, the DC link voltage rises since the offshore wind plant continuous to inject
active power during the fault period to the HVDC link. In order to prevent such high direct
voltage levels, the chopper is triggered and it ensures normal operation value voltage levels,
typically 1.1p.u. The application of the DC chopper is the common LVRT and FRT strategy
applied by manufacturers up to date for the grid code compliance of the HVDC link during
AC grid faults.
Figure 24 introduces the outer controllers, which control the DC voltage, the active and the
reactive power of the HVDC station onshore. The control of DC voltage will be discussed in a
separate paragraph for the MTDC grid case. For the case of reactive power control, the
converter could be either in AC voltage control mode, emulating the behaviour of the AVR in
synchronous generators or in reactive power or power factor control mode which supplies to
the system constant reactive power irrespective of voltage variations.
Although the use of detailed EMT models (where the IGBTs are represented) in grid code
compliance and stability studies would lead to more accurate results, the very small time
constants associated with the PWM modulator and the power electronic switching modules
increases significantly the simulation cost (time for simulation) of the model. Hence, time
average EMT models are used and presented here within the scope of this course. Certain
assumptions are made in the representation of the VSC station for stability type simulations.
First of all, all currents and voltages are represented by positive sequence values. Three phase
voltages are assumed symmetrical. Harmonics and commutation phenomena related to the
switching of the IGBTs are out of the scope of dynamic simulations and thus are completely
neglected. All variables in the AC side refer to phasors while in the DC side to absolute
values.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
DC Chopper VSC
+𝑈𝑑𝑐
uc ,abc us ,abc
2 𝐶𝑑𝑐
iabc
2
~
r 3
𝐶𝑑𝑐 Rch L
−𝑈𝑑𝑐
2 =
2
PWM pulses

modulator

Converter voltage setpoints


Inner
Controller

Phase reactor current set points

DC AC
Voltage Voltage
Controller Controller
Measurements Active Power Reactive Power Measurements
Controller Controller

Outer
Controllers

Setpoints

Figure 23. The single line diagram of a bipolar VSC-based HVDC station.
support, (d) Reactive power controller, (e) constant active power controller [33].

A bipolar VSC-HVDC link is assumed here with identical inner and outer current control
loops per each pole-converter. A typical voltage oriented vector control is used (using a SRF-
PLL with adequate bandwidth) at the onshore VSCs as shown in figure 25. More details can
be found in Appendix C. With reference to the outer control loops, PI-based control strategies
are applied. The active current (id) loop is either in direct voltage control mode or in constant
active power control mode. The reactive current control loop (iq) provides voltage control
during normal conditions and reactive short circuit current injection during AC system
failures, as given in typical grid code requirements. Figure 25 presents the block diagram of
the average value EMT type model.

id 0 iq 0
DC Voltage Droop Regulator Q Controller

U ref idmax Qref iqmax


dc + k p _ dc
+ ref
i +
kp_Q 
ki _ Q +
iqref
- - d - s -
idmin iqmin
U dc (a)
Q
(b)
id 0 iq 0
Active Power Controller Uac Controller
iqmax
Ps ,ref + kp_P 
ki _ P +
idmax
i ref U ref
s
+
kp_Q 
ki _ Q
+
iqref
s - d s
- -
idmin iqmin
Ps Us
(c) (d)

Figure 24. Outer controllers block diagram (a) DC link voltage droop controller, (b) DC link voltage
constant control (c) AC voltage controller for AC terminal voltage [32].

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
GSVSC
U dc  uc , a
+
I dc  r L ia
U dc 
Cdc
Rch
uc , b r L ib us ,a
Cdc
- uc , c r L ic
U ref
U dc  I dc  u s ,b
dc
us ,abc
Voltage us ,c
Outer
Us limiter
mabc
U dc   U c ,max
3 2
Controllers PLL
U sref abc us ,abc
dq
iqref Current
idref abc dq
limiter Inner + 
Controller
+ +
𝐾𝑖 +
𝑘𝑝 + +
+
-
𝑠
- ucd,ref usq usd
id ωL

iq iabc
ωL

+
-
𝐾𝑖
- ucq,ref abc dq
𝑘𝑝 + +
𝑠

id iq
Figure 25. Time-average model of the grid side voltage source converter (GSVSC) with its control
modules [18].

3.2.4 RMS value model of the onshore converter

As already discussed in [35] certain simplifications are made for the model of the onshore
VSC. These assumptions are based on the fact that time constants associated with the inner
controller are very small in comparison to AC system and outer controllers. In addition,
including these time constants in stability type software such as PSS®E complicates both the
modelling and the computation time, let alone it involves risks of numerical instability.
Figure 26 presents the block diagram of the onshore VSC-HVDC system. As it can be
seen, the outputs of outer controllers are fed to the current limiters and the phase reactor
current is then calculated. The current limiter could be operated in either equal priority for the
active and reactive current or in pure active or pure reactive current priority [40].
Another issue that arises when it comes to building RMS value models for stability type
simulations is the Phase Lock Loop (PLL). The three-phase PLL is measuring instantaneous
three phase voltages and by making use of Park transformations, the angle of the voltage is
calculated. This angle is used to align the dq reference frame with the phasor of the system
voltage. In general, the PLL is a very fast tracking device, which does not introduce delays. It
is widely accepted that the PLL can track changes in system frequency accurately. In stability
type simulations an approach that is often followed is to model the PLL by a time delay
between the measured angle at the point of common coupling and the angle passed on to the
dq transformation [36]. In this work, following the approach adapted in IEC 61400 for full
converter interfaced wind turbines, we decided to completely neglect the dynamics associated
with the PLL.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
I pr I source DC cable model
VSC
+

jX pr Cdc U dc
U s e j -

Case where inner


Ps
I dc  controller and phase
U dc reactor are included
Us
I source  I pr 
jX pr
I pr

I pr
xy
I pr  ix  ji y dq

iy ix iq
xy
id
dq Inner usd
 ix  cos   sin    id  Controller

i      iq id and phase

 y   sin  cos   i q  Current


Reactor model
usq
Limiter
ref ref ref iq* id*
Us i
q id U dc Current
Limiter
U sref Outer U dc
Controllers
Q In Fig. 2 Ps iqref idref
Qref Ps ,ref
(a) (b)

Figure 26. PSS®E model of the onshore VSC-HVDC terminal

3.2.5 EMT Average-value model of the offshore converter station

The offshore VSC-HVDC converter station is responsible for controlling the AC voltage
and frequency of the offshore island. The station operates as a power sink for the offshore
wind generated power. During normal operation the offshore converter builds up the AC
voltage and frequency for the offshore collector system. By applying a control strategy which
keeps the AC terminal voltage amplitude and frequency constant, it allows the wind turbine to
be grid connected at the collector island as it is the case for AC connection. This power is
then transmitted by the HVDC link to the onshore power system. The control strategy
followed for the offshore converter is shown in figure 27. The controller is designed on a dq-
rotating reference frame. A PI regulator is used to control the output of the converter’s AC
voltage. This is succeeded by keeping the d-axis voltage to 1p.u. The q-axis reference is set to
be zero, providing 50 Hz rotating voltage reference to the collector system. The 50Hz
frequency of the offshore collector system is made by giving to the PLL a three phase,
symmetrical 50Hz frequency.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
VSC
+
Cdc L
~ R
U dc
=
iabc 3

uc ,abc us ,abc
-

PWM pulses

modulator
mc ,abc
 sref
abc PLL
dq
mcref,abc
ucd,ref Oscillator

ucq,ref  0 𝑘𝑝 +
𝐾𝑖
𝑠
f=50Hz

- +
Us Us ref

Figure 27. Offshore VSC-HVDC station subsystem and control.

mabc
U dc 3
2  sref
abc
dq e jx
ucd,ref
u q
c , ref 0 1
𝐾𝑖 𝑠
𝑘𝑝 +
𝑠
1.15
+ u - + s  2 f s
K LVRT Us Us ref
- 0.0
1.15

1.4 U ref
1.0 LVRT
/
*

U dc*  U dc 
Figure 28. Offshore EMT average value VSC model and the relevant LVRT method. Only one LVRT
strategy is used at a time for this paper analysis but both are presented here for completeness [18].

As it can be seen in figure 28, the q-axis voltage component is set to be zero, while the d-
axis voltage component is controlled using a PI-regulator. The PI-regulator controls the
amplitude of the offshore collector grid voltage to the reference value. An oscillator is used to
maintain the frequency of the offshore collector grid at the 50Hz reference. The time average
EMT type model of the offshore HVDC station is show in figure 23.
Beside the normal operating control blocks, the offshore HVDC station is equipped with
LVRT control blocks. As it has been discussed in the previous paragraph, during faults in the
onshore AC grid, over-voltages are observed in the HVDC link. The cause of the over-voltage
the mismatch between the power flow from the offshore to the onshore HVDC station. These
over-voltages can be mitigated by making use of the DC chopper as shown in figure 20.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
An alternative to the chopper based FRT strategy, is the reduction of the AC offshore grid
voltage. Once, a fault is applied onshore and the direct voltage rises above the normal
operating point threshold, the FRT control module of the offshore converter station reduces
the offshore AC island voltage reference. Hence, the power infeed from the offshore HVDC
station to the HVDC link is reduced, reducing the power mismatch between the onshore and
the offshore converters. This strategy is known as artificial AC voltage drop control scheme
for FRT grid code compliance of HVDC links. Simulation results will be presented in
following paragraphs. More information can be found in [18].

3.3 LVRT control Logic

During AC system fault conditions, the onshore converter should provide additional
reactive current injection to the AC system as given by the relevant grid codes. In order to
ensure that the total current will be always below the converters overcurrent capacity, an
appropriate control logic should be introduced which prioritizes active or reactive current
components of the converter fault current. Normally, reactive current priority is given during
severe voltage sags in order to utilize the capacity as reactive current. In that case the active
current is reduced accordingly (or even limited to zero) as a result of the injected reactive
current.
Finally, the different operating states of the converter station (i.e normal operating state,
faulted condition and the post fault period) are coordinated using the state machine of figure
29. The converter shifts from the normal operation, to faulted state and post-fault state
depending on the voltage levels as measured by the abc/dq Park transformation and the use of
the PLL. The coordination of the state machine with respect to the other control module is
presented in figure 24.
usd  0.5
usd  0.5
&&
S0: Normal time>Tr+Tfault S3:
Operation
Ramping_Up
With R (p.u/s)

usd  0.5
&& usd  0.5
FLAG_FRT = 0 &&
S2:
FLAG_FRT = 1
FRT
MODE

Figure 29. State machine for the LVRT logic which allows the implementation of the ramp-up
function during post-fault period (Tr is the ramping time calculated based on the ramping rate R, while
Tfault is the time that the fault is applied). [18]

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
U dc ,thr U dc

us ,abc Engage Chopper D Chopper


Chopper Control uc ,abc
PLL

dqàabc
usd FRT Logic
Start/Stop
Ramping
Freeze ucd,ref ucq,ref
Direct
Voltage Voltage
Ramp Up
Control Loop function Limiter

U dc DC
id** idref
ucd ucq
Current Inner Current
Voltage
Droop
Limitation Control Loop

U dc ,ref Contol iq**


AC Voltage
iqref
Control Loop

Pdc Pdcref
Figure 30. Coordination of the control modules via the state machine for the LVRT logic which allows
the implementation of the ramp-up function during post-fault period (Tr is the ramping time calculated
based on the ramping rate R, while T fault is the time that the fault is applied).

Please note that the d-axis voltage component at the PCC point is equal to the amplitude of
the PCC voltage normalized in per unit. Once, the fault is cleared and the voltage rises to
normal values, the additional reactive current injection of the converter stops and the
converter returns to its pre-fault active current value. However, in order to protect the power
electronics components from the fast transients during the post-fault period, the converter’s
active current is ramped-up with the so called ramp-up rate given in p.u/s. Typical grid codes
indicate the ramp-up rate of the converter station in p.u/s. This ramp-up rate during the post
fault period should be activated by an appropriate LVRT control logic.

3.4 Methods for tuning VSC controllers


The selection of the control parameters for the inner and the outer controllers can be
quantified. The method is based on the frequency bandwidth that the VSC is required to
operate. The method of internal model control (IMC) is used for tuning the inner controller.

3.4.1 Inner controller tuning

Based on the IMC method, the design parameter for the selection of the inner controller is
the rise time Tr of the VSC. This corresponds to the time needed from 10% to 90% of the
steady state value when the system is subjected to step response. Given the rise time, the
closed looped bandwidth α (rad/s) can be calculated as:
ln 9
 (17)
Tr

With the given calculated bandwidth, the proportional and the integral gain of the inner
controller can be found as a function of the phase reactor parameters.

kp   L (18)

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
ki   r (19)

Thus, for any given VSC, with known phase reactor parameters, it is possible to calculate the
proportional and integral gains of the inner controller. Of course this is only an estimation and
always sensitivity analysis to these parameters needs to be performed.

3.4.2 DC voltage controller tuning


The same method can be used for the selection of the DC voltage controller. The equations
are given as:

k p   d Cdc (20)

ki   d2Cdc (21)

Where,  d is the design criterion, given in rad/s. Similar to the inner controllers, a
modification of the calculated parameters may be necessary in order to achieve better
response in terms of overshoot and oscillatory behaviour.

3.5 Control of the VSC-HVDC system in point to point connection

For a point-to-point VSC-HVDC link used for the grid connection of offshore wind power
plants, a master-slave control strategy is applied. The grid side VSC (GSVSC) is responsible
to control the HVDC link voltage (slave). The wind plant side VSC station (termed as
WPVSC) is injecting its DC current in the HVDC link irrelevant of DC voltage variations
(master). The main control target of the offshore converter station is to maintain constant the
offshore AC terminal voltage and frequency at the offshore island grid.
The reader should recall that the voltage at the DC terminals of VSC-HVDC link is
controlled by the active power at the AC terminal of the GSVSC. Note that the balance
between the WPVSC and GSVSC is affecting the direct voltage variations. This active power
injection by the GSVSC to the onshore power system is controlled by controlling the active
(d-axis here) component of the current.
There are two main control strategies commonly applied in the control of the DC link
voltage, at the GSVSC terminal. Mainly, either a constant voltage control method or a droop
method (see figure 24 for the control implementation). In the constant voltage control method,
the DC voltage is controlled at its rated value by making use of a PI controller. In the direct
voltage droop control strategy, the voltage is allowed to vary proportionally to a given Udc-Pdc
characteristic. Figure 31 gives the characteristics of both types of control schemes. The slope
of the Udc-Pdc characteristic is determined by the droop value which is the inverse of the
proportional gain controller.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
U dc U dc
Inverter Rectifier
Inverter Rectifier

Pdc Pdc

Constant Voltage Droop Control


Control

Figure 31. Control strategies applied for dc voltage control of the HVDC link by the GSVSC. Positive
Pdc means rectification while negative inversion.

3.6 Control of the VSC-HVDC system in multi-terminal DC grid connection

The control of DC voltage is very important in the operation of future offshore HVDC
grids. The DC voltage is related to the power balance in the MTDC grid [42]. The MTDC, in
contrast to AC transmission systems, is an active system, which involves continuous control
actions in order to operate properly. One of the most important control actions is the control
of the HVDC voltage. This is achieved as result of balancing the power generated by the DC-
connected plants offshore and the power fed by the onshore converters into the AC system(s)
[38]. The control of voltage in a MTDC grid is assigned to the onshore VSC-HVDC stations.
At least one converter station onshore needs to be in voltage control mode, assigned with the
duty to balance the HVDC grid and in such way maintain the DC voltage within stiff
operational limits. This converter is often named a “slack” converter. For redundancy, more
than one converter would be assigned with the duty to balance the DC grid, depending on
both the grid size and converters capacity [39].
Three are the main control strategies for controlling the voltage in a MTDC grid: the
constant power control, the constant voltage control and the power based droop control. In
[30], [39], a detailed description has been given of these MTDC grid voltage control
strategies for the interested reader.

U dc U dc
Inverter Rectifier Inverter Rectifier

Pdc Pdc

Constant Power Constant Power


Control Mode Control Mode

(a) (b)

Figure 32. Constant active power control mode of the onshore VSC-HVDC station: (a) in inversion mode (b) in
rectification mode.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
3.7 Power based direct voltage droop control

As it is presented in figure 24, the direct voltage droop controller is operating on the active
current control loop. It is actually a proportional controller applied on the active current of the
onshore converter station. An alternative approach is the power based direct voltage droop
control scheme as it shown in figure 33. It operates on the active power controller of the
onshore converter station. Hence, the direct voltage variation affects the converter active
power reference as it can be seen in figure 33. The power based direct voltage droop control
is the commonly used control strategy for multi-terminal HVDC grids. Beside the control of
direct voltage, the VSC stations constituting part of the HVDC grid will be required to
participate in the share of load frequency control. Especially when the HVDC grid is
connecting two AC power systems. One such example will be the connection of the British
power system to the Dutch or German power system with VSC-based HVDC grids [14]. In
that frame, the same approach as in figure 33 can be used in order to provide active power
response or primary frequency. In this case, a frequency droop controller (red dashed block)
ensures primary load frequency response.
Frequency droop
cotroller

Δf 1
Rf
Pf _ droop
Active current
DC Voltage Droop Controller
Pdc _ droop max component of
+ i
d the VSC-HVDC
+ - ki _ P
U dc 0 k p _ dc kp_P 
- + s min
i
idref
d

U dc Ps ,0 +
-

Ps

Figure 33. Frequency droop controller implemented as additional loop on the power based DC voltage
droop controller.

3.8 Assessment of LVRT Control Strategies Using EMT Average Value Models

So far, we have discussed the modeling approach followed for grid connection studies
related to offshore wind plants in VSC-HVDC connection. This paragraph aims to provide
simulation examples on the dynamic response of a VSC-HVDC link during balanced three
phase faults at the onshore HVDC converter station terminal. The example will discuss the
main issues associated with the grid code compliance of offshore HVDC links and their
connected wind plants.
Two grid code compliance control strategies will be presented and discussed in this section.
The goal is twofold: first to demonstrate the capability of the proposed methods to ensure
FRT compliance according to the FRT grid codes. Please recall that the main problem for the
FRT compliance of the VSC-HVDC link for offshore wind connection is the increased over-
voltages in the DC side. Second, the paragraph tries to demonstrate important control
parameters sensitivities and trade-offs which affect the combined HVDC system and the
connected offshore wind power plants.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
The HVDC link in this case study is rated here at +/-320kV. It connects two offshore wind
power plants as it is shown in figure 34. The link is assumed symmetrical bipolar with
grounded return. The wind power plant for this analysis is assumed to be of full-converter
(type 4A) modelled with average value EMT models. Please note that the same control
schemes are used as presented above for the type 4 wind plant. The offshore collector grid
island consists of a step-up transformer, 150kV HVAC offshore cables, and 33kV/150kV
transformers which make the interface with the 33kV feeder. The topology and ratings given
in this example are typical as used in the offshore wind projects in Germany.
TYPE4A
WT

dc
5km HVAC
ac
Cable
33kV
TYPE4A
WT
3 S k''
2 3
110kV 303 X /R
WPVSC GSVSC
Station1 Station 2
dc 150kV PCC
ac PCC
33kV 3
3km HVAC
Cable

Figure 34. Test system for the EMT average-value model developed in Matlab/Simulink. Both onshore
and offshore converters are modelled as shown in figure 3 & 4.

The first FRT compliance method uses the classical chopper based solution [40] where an
IGBT-based controllable resistance is installed at the DC terminals of the onshore converter
station, dissipating the excess of the DC link power during AC voltage sags. This method
despite of increasing the investment cost, has been widely adapted and is the current LVRT
control strategy followed on real project up to date in Germany (Borwin and Helwin wind
plants connection). The main benefit of this approach is the decoupled dynamics between the
offshore AC collector grid and the onshore power system voltage during onshore faults.
The second solution for grid code compliance avoid the use of chopper. It uses a
coordinated and communication-free control strategy. The coordination is between the
onshore and the offshore converter station which enables the power reduction of wind park
side VSC station during onshore AC fault conditions [11]. In this method, we are artificially
transferring the AC voltage drop from the onshore converter terminals to the offshore
converters terminals by means of control. The latter, ensures lower overvoltage in the HVDC
link during the fault period. In this way power balance is brought between onshore and
offshore converter terminals. Figure 28 presents the LVRT strategy controller followed for
AC voltage reduction. Please note that no communication is needed in that case and the only
measured needed is the direct voltage variations.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
3.8.1 Chopper method assessment with EMT average value model

A self-cleared 150ms three phase fault is applied at the onshore connection point of the
onshore converter VSC station at 1 sec (bus 303), as it can be seen in figure 34. Figure 35
presents the simulation result for particular case study. During the fault period, as the grid
code requires, the onshore converter station provides an additional reactive current injection
(see d-axis current) at the grid connection point using a k gain equal 2. The duration and the
voltage drop reflects the strict TenneT TSO requirement for offshore wind power plants [2].
This means that for a voltage sag of 1.0p.u at the grid connection point, 2 p.u reactive current
is injected by the converter station, which of course is far beyond the capacity of the
converter, which is here 1.1p.u. As a result, the reactive current will be limited to the
maximum capacity current of the converter while the active current will be reduced to zero.
The latter is performed by the current limiting control block. During the post-fault period, the
active current returns to its post fault value with a pre-given active current ramping rate which
is 2 (p.u/s) in this case study.
325
Udc [kV]

320
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
1
iabc [p.u]

-1
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
1
id [p.u]

0.5
0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
0
iq [p.u]

-0.5
-1
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
1
uabc [p.u]

-1
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
400
MW

200

0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
time [s]
Figure 35. LVRT response of the onshore converter station using chopper based strategy for a k gain
equal 2 and reactive current priority as current limitation strategy.

This ramping up of the active current is a variable that each HVDC owner and system
operator may decide based on HVDC system needs. In order to demonstrate the effect of the
ramping rate of the converter station, the HVDC system dynamic response for 4 selected
ramping rate values has been shown in figure 36. From the results, it can be concluded that
the slower is the ramping up of the onshore converter station active power, the slower will be
the direct voltage recovery to the normal pre-fault value. However, this will not impact the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
reactive current injection of the converter station as soon as the direct voltage is kept at stiff
limits. It is important at this point to mention that during the fault and post fault period, the
chopper is responsible for controlling the direct voltage while the direct voltage control loop
is blocked. The reader should observe the dissipated power by the chopper for different values
of the ramping rate. The slower is the ramping rate, the bigger is the burden placed to the
chopper mechanism. Finally, the ramping up of the converter may impact the rotor angle
response of the AC system depending on the SCR and the installed generation capacity in the
system as shown in [16].
Onshore converter DC terminal votlage
326
R= 1
324
Udc [kV]

R= 3
322 R= 5
R= 9
320

1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3


time [s]
active current component
1
id [p.u]

0.5

0
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]

reactive current component


0
iq [p.u]

-0.5

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Chopper dissipated Power
400
MW

200

0
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Figure 36. Influence of the selected ramping-up rate on the FRT response of the VSC-HVDC link.

In addition, during the period that the chopper is active, the state of the direct voltage
controller of the converter station is set to “freeze” in order to avoid negative interactions
during the post-fault period. It is only after a delay of 0.15s (time chose by trial and error)
after the ramping is completed that the direct voltage control is restored to the converter
station. Once the direct voltage control is restored, the direct voltage will return to its pre-fault
value followed by the active current spike as it can be observed in figure 36. This is actually
an instant discharge of the converter station DC caps.
In general, it can be concluded at this point that the dynamic response of the VSC-HVDC
link and the wind power plants is robust for the case of chopper based LVRT. The offshore
AC grid transients are decoupled from the onshore AC grid voltage. This is demonstrated here
both by EMT switch models as well by average value models. The chopper solution has been
tested for different operating point of the HVDC link and has demonstrated adequate
response.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
3.8.2 Artificial offshore grid voltage reduction method
The alternative and cost-effective method to the chopper based LVRT strategy is the
implementation of the direct voltage based artificial voltage reduction at the offshore collector
grid, applied by the offshore converter station [11]. The simulation results for this case are
presented in figure 37. The controller is continuously in operation and uses as input the ratio
of the actual measured to the rated direct voltage. Once the ratio is above the chosen threshold
value, the AC voltage reference is reduced proportionally to the measured direct voltage. In
this way a direct coupling is achieved between the direct voltage in the HVDC link and the
offshore AC collector grid voltage reference. Please note that this coupling is active only
above the threshold value. Initially it is assumed here that offshore wind power plants are not
injecting reactive current to the offshore AC grid when the voltage is artificially reduced. This
is equivalent to the case where the k gain of the wind turbines is equal to zero.

Figure 37. LVRT of the VSC-HVDC link with artificial voltage reduction method at the offshore AC
grid.

Observing the simulation results of figure 37, it can be seen that the offshore collector
grid voltage has been reduced to zero during the fault period in order to keep the direct
voltage at controllable levels. Hence in this way the onshore terminal fault is “artificially”
shifted to the offshore terminal by the dedicated control loop applied at the offshore converter
station. In this way the offshore converter station is not absorbing any power from the wind
turbines during the fault period, keeping the direct voltage at controllable levels. The
generated by the wind turbines power which is not flowing in HVDC the link is dissipated in
this way by the individual choppers of each wind turbine DC link. In practice, with this

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
approach, instead of having a central HVDC chopper, we use the individual per each wind
turbine choppers, which are by default installed at each wind turbine. The latter reduces the
grid code compliance cost as it avoids the investment of the HVDC chopper equipment and
reduces significantly the size of the onshore substation.
Observing the dynamic behavior of the onshore converter station, it can be seen that there
are no significant differences on the active and reactive current response between the two
different LVRT schemes. This holds true as the direct voltage is kept at controllable normal
levels (0.95-1.1p.u) and the modulation index is not affected. The main difference observed
between the two strategies is the direct voltage dynamic response, which in the case of the
chopper method is more stable. Looking at the response of the offshore AC grid voltage, it
can be commented that a sustained voltage drop is observed as a result of the ramping rate at
the onshore converter station terminal.
In order to assess the impact of the onshore converter ramping rate on the combined wind
park and VSC-HVDC system dynamic response, the HVDC system response with three
different ramping rates have been plotted in figure 38. From the simulation results, it can be
concluded that a higher (in pu/s) active current post-fault ramping rate, will reduce the period
of the voltage drop at the offshore converter station. The same applies also for the case of the
direct voltage variation as it can be observed in figure 38. This is due to the smaller period
that the direct voltage stays above the relevant threshold.

Figure 38. Impact of the ramping rate on the artificial voltage reduction LVRT method.

In general, it can be concluded that for the artificial voltage reduction method, the impact
of the ramping rate of the onshore converter during post the fault period should be carefully
assessed. Furthermore, it is important to ensure that the LVRT curve of the offshore wind
turbines will not be met, avoiding unnecessary tripping of the wind turbines. Finally, the
control loops for creating the artificial voltage reduction shall be carefully designed and co-

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
optimized with respect to the AC offshore grid resonance frequency in order to avoid the
trigger of offshore harmonic instability while complying with the grid code. Hence, it become
clear that the application of advanced LVRT schemes utilizing the high control capability of
the converters is possible but it may be performed under very careful analytical studies.

3.8.3 Case when the offshore wind turbines comply with the Grid Code given at the
offshore AC collector grid

In the previous case study, it has been assumed that the offshore wind turbines are not
injecting reactive current during the artificial (by the offshore converter station) voltage sag at
the offshore AC grid. This choice was made intentionally with purpose to illustrate the effect
of the voltage reduction method itself. However, this is not in line with what common grid
codes specify, where wind turbines connected to the offshore AC grid shall provide additional
reactive current injection during voltage drops at their AC terminals. This requirement has
been mainly made available in order to assist the offshore AC collector grid protection
schemes to detect and isolate symmetrical permanent faults. The reader should note that at the
isolated AC offshore grid the only controlled reactive current sources are the wind turbines
and the offshore converter station.
As an alternative approach, figure 39-40 demonstrates the case where the offshore wind
turbines are injecting reactive current at the offshore collector grid during the artificial voltage
drop. Furthermore, it is assumed here that the wind turbines in this case study are ramping up
their active power with a ramping rate of 3 (p.u/s) during the post fault period. Please note
that the ramping up is performed only if the active current is reduced to zero as a result of
reactive current injection. Recall that the injection of reactive current reduces the active since
the fault current capacity of the wind turbines is only 1.1 p.u of the rated current. The latter is
performed automatically by means of dedicated control schemes.
The same ramping rate is also assumed for the case of the onshore converter station. From
the simulation results, it can be observed that through this voltage reduction mechanism and
the injection of reactive current by the wind turbines, a better fault ride through (FRT)
response can be achieved. This is because of the fact that the active current injection of the
wind turbines is reduced to zero as a result of reactive current support. Hence, the offshore
converter station would create a smaller voltage drop (0.6p.u) at the offshore AC collector
system for the same onshore grid fault. So it is clear that for this particular LVRT strategy,
requiring the wind turbines to inject reactive current is beneficial because it will limit the
artificial voltage drop created offshore ensuring the grid code compliance.
Important however, is to mention that there is a need for coordinated operation between
the FRT strategy (voltage reduction method here) and the protection schemes applied at the
offshore AC grid. Since distance protection cannot be easily applied due to the small network
impedances, the protection scheme shall be capable to distinguish between cases of a
permanent fault at the offshore AC collector grid from a case where artificial voltage
reduction is created due to the fault at the onshore terminal. This can be ensured by means of
smart protection schemes and the utilization of communication systems between the wind
turbines and the offshore converter.
Finally, the post-fault ramping rate of the wind turbines and the post-fault ramping rate of
the onshore converter station shall be coordinated. When the ramping rate of the wind
turbines is faster than the ramping rate of the onshore converter, the active power of the wind
turbines will be recovered in faster rate than what will be injected to the onshore grid
terminals. Hence, there is risk of over-voltage in the HVDC link, which may re-trigger the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
LVRT scheme. When the opposite occurs, the onshore converter may inject more power to
the grid than the wind turbines are feeding creating the risk of direct voltage collapse in the
link. The reader should note that during ramping up, the converter’s direct voltage control
loop is blocked. Hence, it is crucial to ensure the smooth coordination between the two
ramping rates. In general, it is advised that the two ramping rates are equal or at least the
ramping rate of the wind turbines is higher. Finally, figure 39 and 40 shows the simulation
results with the onshore converter metrics for the case where the ramping rate of active
current are equal.
In conclusion, it can be mentioned that when a cost effective LVRT strategy has to be
applied, the coordinated artificial voltage reduction method could be used under careful
considerations and studies related to the offshore collector grid. Coordination is needed in
order to avoid tripping of the units. In addition, the control loops added shall ensure that they
do not trigger the resonance frequency and harmonics at the offshore collector grid hence
thorough application of the method is needed.
WT terinal voltage at 33kV
u abc offshore[p.u]

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
WT instantaneous current
1
iabc wt [p.u]

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
WT active curretn
id WT [p.u]

1
0.5
0
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]

WT reactive current
0
iq WT[p.u]

-0.5

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
1
uabc [p.u]

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
WPVSC voltage
1
Uac[p.u]

0.5

0
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Figure 39. Offshore collector grid response with the wind turbines injecting reactive current during the artificial by
the offshore converter station voltage drop. This response is triggered from a balanced three phase fault at the
onshore grid. From top to bottom, the wind turbine 33Kv voltage, the wind turbine current, the active and reactive
current (dq-frame), the onshore VSC-HVDC station voltage and the offshore VSC station controlled AC voltage
drop.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
DC terminal voltage
330

Udc [kV]
325

320
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
AC terminal current
1
iabc [p.u]
0

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Active current component of the GSVSC
1
id [p.u]

0.5

0
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Reactive current component of the GSVSC
0
iq [p.u]

-0.5

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]
Voltage at the onshore AC terminal
1
uabc [p.u]

-1
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
time [s]

Figure 40. Onshore HVDC converter station metrics. From top to bottom, the onshore DC terminal
voltage, the instantaneous p.u current of the onshore VSC station, the active (d-axis) and reactive (q-
axis) current and the AC terminal voltage.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Chapter 4: Effect of FRT Grid Code Compliance and
Reactive Current Injection by VSC-HVDC on the
Transient and Voltage Stability of the Power System

This chapter demonstrates how far and large offshore wind power plants and their VSC-
HVDC system delivery can contribute in improving power system voltage and rotor angle
stability under faulted conditions. The report performs a parametric sensitivity analysis on the
additional reactive current injection gain k, the choice of the dead-band and the current
limitation strategy followed by the onshore VSC-HVDC. The analysis is applied on a multi-
machine power system such as the IEEE 39-bus test system. The paper focuses on the
onshore HVDC converter station as the offshore converter station and the wind power plants
operate fully decoupled from the onshore power system. Thus, even if the HVDC
transmission system is part of offshore wind power plant delivery, the onshore HVDC
converter station is playing the most important role in the compliance with grid connection
requirements imposed at the onshore point of common coupling (PCC). Please see the chapter
3 for details.

4.1 VSC-HVDC Dynamic Model

A positive sequence, quasi-steady state modelling approach has been chosen to represent
the dynamic behaviour of the VSC-HVDC station in stability type studies [14]. The
associated with the inner current controller time constants are considered very small with
respect to much larger outer controller and AC system dynamics time constants. Thus,
without loss of important information, the fast dynamics of the inner current controller could
be neglected at least in stability time studies with large power system. As a result, the outer
controller is the most important control module which determines the behaviour of the VSC-
HVDC system. With regard to the Phase Lock Loop (PLL), it has been assumed that the PLL
can perfectly track the voltage angle at the point of common coupling (PCC), and thus the
voltage angle at the PPC point is directly used into the dq to xy transformation [41].
The DC side of the converter is modelled as a DC current source, while the DC cables by a
lumped π section DC circuit. State space equations in time domain have been used to model
the dynamic behaviour of the DC cables [41]. The necessity to represent DC cables with
dynamic models appears as a need to model the dynamic response of the DC voltage,
especially under faulted conditions. The latter influences the response of DC voltage
controller, hence the active current injection of the VSC-HVDC station through the DC
voltage control loop [41]. Finally, it is worth mentioning that the operation of the VSC-
HVDC link is limited by mainly two factors: the AC terminal current (especially under
faulted condition) and by the DC voltage levels [41].

4.2 Selected control parameter sensitivities for evaluation of system impact

The VSC-HVDC onshore converter, at the presence of a symmetrical three-phase fault in


the system, should provide additional reactive short circuit current in order to support the
voltage at the point of common coupling. The value of this additional short circuit current

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
injection is primarily determined by the AC voltage controller (implemented as PI controller),
which controls the reactive current component of the VSC-HVDC, and by the DC voltage
controller which controls the active current component.
An important sensitivity which influences the short circuit current provided by the VSC-
HVDC, and thus the transient stability and voltage response under faulted conditions is the
selection of the gain k (defined as the slope of curve in figure 1). The value of gain k
influences the reactive current injection of the converter station during faulted conditions.
Normal values of the gain k (given in p.u) are between 2 and 10. The reader should note that
the AC voltage controller in this paper analysis is assumed to be in continuous operation.
However, the short circuit current capability of the converter station is limited due to the
ability of the IGBTs to handle large short circuit currents. A typical value of the over-current
capability is 110-115% of the rated value. In the same direction, another important sensitivity
is the current limitation strategy followed by the onshore VSC-HVDC converter to restrain
the short circuit current at the maximum over-current capability. It will be shown that the
method prioritising active or reactive current component, in order to keep the short circuit
current contribution of the onshore VSC-HVDC converter to the maximum acceptable value
influences the voltage and transient stability of the system. Finally, a third sensitivity is the
choice of the dead-band. It is defined as the margin of the voltage drop which the AC voltage
controller is insensitive (see figure 1). Previous studies performed for DFIG based AC
connected offshore wind power plants [42] have recommended that the removal of the dead-
band would be favourable for the overall transient stability of the system. Such sensitivity will
be also addressed in this paper for VSC-HVDC systems.

4.3 Test system


A modified version of the original New England test system is used in this paper to
demonstrate sensitivities and trade-offs for VSC-Based HVDC systems (see figure 6). Two
VSC-HVDC links are connected at buses 2 and 9 respectively, with identical control
parameters and strategies using step up transformers with short-circuit voltage uk equal to
10%. Both onshore HVDC converters are rated at 450 MVA while the maximum delivered
active power is 400 MW. The HVDC system in both cases is connecting two offshore wind
power plants located in distance of 250 km from shore. The HVDC link is considered as a
mono-polar rated at 300kV. Furthermore, two onshore wind power plants have been added at
selected buses, 21 and 25 respectively. Both onshore and offshore wind power plants are
modelled following IEC-61400 standard models. All synchronous generators have been
modelled by IEEE standard models, with IEEE SEXS type excitation system and TGOV1
type governor with standard parameters.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
28

37 29
26
41
25 38

24
27
30
16
2 35
18 17
HVDC link
21 22
dc ac
15
ac dc 3
301 40
400 MW 1 14 23
4 19
36
5 20 33
12
39 6 34
7
11 13
8 31

9 convenional generation = 5202.4 MW


10

302
onshore wind generation = 1600 MW
32
dc ac offshore wind generation = 800 MW
ac HVDC link dc
Total load = 7467.2 MW
400 MW

Figure 41. Single-line diagram for the point-to-point HVDC connection of offshore wind power plants to IEEE
39-bus test system.

4.5 Simulation Results

4.5.1 Sensitivity of the reactive current boosting gain k


Figure 42, introduces the voltage profiles at buses 301 and 302 (see figure 41) for a 150 ms
symmetrical three phase fault at bus 39. In this set of simulations, no dead-band is used at the
reactive current boosting slope. It can be seen that a relevantly high gain (k=6) leads to
improved voltage drop at the grid connection point, during the fault period. This can be
explained by observation of the reactive current component of each converter station, hence
the reactive power delivered to the system. However, the over-current capability of the VSC-
HVDC station is limited and the additional reactive current adds on top of the steady state
pre-fault AC current injection of the converter station (which is 0.8 in both cases). Hence, the
additional reactive short circuit current component may cause the converter to reach its
maximum over-current capability, which in this case study is selected to be 1.15p.u. Once the
maximum over-current capability of the converter is reached, the current limitation controller
will limit the additional reactive short circuit current.
From the results of figure 42, it can be seen that the increase in the gain k, enhances the
voltage profile at PCC during the fault period contributing a positive effect on the power
system voltage stability. However, once the converter’s over-current capability is reached, the
current limitation strategy followed to restrain the value of the fault current contribution will
influence or sometimes even counter-act this positive effect. This is strongly dependent on the
current limitation strategy. Hence, the active or the reactive current is restricted. One proposal
would be to oversize the converter in order to leave more available space to the additional
short circuit current. This under circumstances could be a solution, but it may lead to non-
desired results as discussed by authors in [12] depending on the R/X ratio at PCC.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Converter Upcc NE301 Converter Upcc NE302

1 1
K=1
K=6 0.4

p.u

p.u
0.5 1 0.5 0.3
0.2
0.8
1 1.1 1.2
0 1 1.1 1.2 0
1 1.5 2 2.5 1 1.5 2 2.5
time[s] time[s]

Converter Reactive Power NE301 Converter Reactive Power NE302


1.2 1.2
0.9 0.9
0.6 0.6
p.u

p.u
0.3 0.3
0 0
-0.3 -0.3
1 1.5 2 2.5 1 1.5 2 2.5
time[s] time[s]
id iq for VSC 301 id iq for VSC 302
1 1
0.5 Active Current of VSC 0.5 Active Current of VSC
p.u

p.u
0 0
-0.5 -0.5
Reactive Current of VSC
-1 Reactive Current of VSC
-1
1 1.5 2 2.5 1 1.5 2 2.5
time[s] time[s]
I terminal VSC 301 I terminal for VSC 302
1.2 1.2

1 1
p.u
p.u

0.8 0.8

0.6 0.6

1 1.5 2 2.5 1 1.5 2 2.5


time[s] time[s]

Figure 42. Simulation result for the IEEE 39-bus test system which shows the sensitivity of the k-factor on the
onshore AC terminal of the VSC-HVDC station (No dead-band with equal-priority current limitation).

Figure 43 demonstrates the effect of the gain k on the DC voltage response and the
dissipated power by the DC choppers of the VSC-HVDC link. As it is observed, in the case
with high gain k, due to smaller voltage drop, less power is dissipated in the chopper in both
HVDC links. This reduces the stress on the DC chopper, by means of requiring less power
needed to be dissipated during the fault. Furthermore, three different values of the gain k are
selected in order to demonstrate the effect of the power system response to these parameters.
Figure 44, demonstrates the voltage drop at selected buses. From the results, it can be seen
that an increase in the gain k leads to lower voltage drop, at least for this case study
disturbance.
In addition, the critical clearing time (CCT) for two selected generators is demonstrated in
figure 45. As it can be observed the short circuit current contribution of the VSC-HVDC
system is also beneficial for the power system transient stability. The reader should notice that
for each generator the CCT is defined for a symmetrical fault at the low voltage side terminal
of the step up transformer. Thus, the improvement on the CCT of each generator occurs due
to improved post-fault voltage response and due to smaller rotor angle displacement between
the online generators with respect to the faulted generator. Finally, from these results it is
clearly observed the improvement in the power system dynamic response from the case where
the VSC-HVDC link does not participate in providing additional reactive short circuit current
(k=0).

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Udc 301 Udc 302
1.2 1.2
Chopper triggered
above 1.1 p.u k=1
1.15 DC voltage 1.15 k=6

1.1

p.u
1.1

p.u
1.05 1.05

1 1
1 1.2 1.4 1.6 1.8 2 1 1.2 1.4 1.6 1.8 2
time[s] time[s]

P chopper at 301 in p.u of 500MW P chopper at 302 in p.u of 500MW


1 1
Fault 0.3
0.8 occurs 0.8 0.8
here 0.2
0.6 0.6 0.7
0.1
p.u

p.u
0 0.6
0.4 0.4
0.5
1 1.1 1.2 1 1.05
0.2 0.2

0 0
1 1.2 1.4 1.6 1.8 2 1 1.2 1.4 1.6 1.8 2
time[s] time[s]

Figure 43. DC Voltage and dissipated power at the choppers of the HVDC links for the two k-factor
gains applied at the AC voltage controllers.

Voltage levels for fault 150ms at bus 39


1
0.9
0.8
0.7
0.6
U [p.u]

0.5 k= 2
0.4 k=4
0.3
0.2 k=6
0.1
0 k=0
30 2 6 5 7 9 1 38 35 33
NE system bus number

Figure 44. System minimum voltage level during 150ms fault at bus 39, for four selected k-factor
gains (No dead-band, equal priority).

Sensitivity of gain k for VSC-HVDC on the CCTs of online


generators
350
286 298
300 263
250 225
195 210
CCT[ms]

180 k =2
200
140
150 k=4
100 k=6
50
k=0
0
G30 G31
Selected Generator

Figure 45. CCTs (in ms) for faults applied at each generator terminal as a function of gain k of the
VSC-HVDC system (No dead-band, equal priority).

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Finally, the influence of the dead-band on the voltage response at the grid connection point
is demonstrated in figure 46. A 15% dead-band is assumed for this case study. The value of
gain k is assumed to be 6 while the disturbance is the same with the previous set of
simulations. Observing the post-fault response of the voltage at the point of common
coupling, it can be seen that in the case without dead-band a better system voltage recovery is
achieved. This improvement in the system voltage response is the result of continuous
additional reactive current injection to the system by the VSC-HVDC system. Finally, the
reader should notice that a 500ms continuation of the AC voltage control (as suggested in
some grid codes) after return in the dead-band is not considered in this paper.

Converter Upcc NE301 Converter Upcc NE302


1.3
1.2 1.2
1.1 1.1
1 1
0.9 0.9
0.8 0.8
p.u

p.u
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4 15% Dead-band
0.3 0.3 NO Dead band
0.2 0.2
1 1.5 2 2.5 3 1 1.5 2 2.5 3
time[s] time[s]

Id iq for VSC 301 Id iq for VSC 302


1.2
1.1
1 Active current (id)
0.6 Active current (id) 0.5
0.1
0
p.u

p.u

-0.4
-0.5
Reactive curretn (iq) Reactive curretn (iq)
-0.9
-1

-1.4
1 1.5 2 2.5 3 1 1.5 2 2.5 3
time[s] time[s]

Figure 46. Effect of dead-band on VSC-HVDC and PCC voltage response (K=6, dead-band x=15%
and equal priority current limitation strategy).

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Chapter 5: Analysis on Frequency Stability
Requirements

5.1 Introduction
At present, there is an ongoing trend for increased penetration of power electronic based
generation in power systems (mainly large PV and full converter based wind power plants)
[21] [22]. This large in-feed of power electronic based generation, in combination with
replacement of large conventional power plant units (coal and nuclear) will eventually lead to
the decrease of power system inertia [6], [7], [9]. Reduced system inertia will cause faster and
larger frequency deviations in the presence of severe events (i.e tripping of significant amount
of generation following a system disturbance).
The North American Electrical Reliability Corporation (NERC) has already reported the
continuous declining of the frequency response due to inverter interfaced generation in the
last decade [7]. Also, the European Wind Energy Association (EWEA) suggests the
establishment of systematic system planning to integrate large wind power plants, and
suggests the understanding of system interactions [22].
The effects of the reduction in system inertia could be further increased when HVDC
transmission system technology is used for the grid connection of large offshore wind power
plants, located far from shore [4]. Given these future plans, system inertia is gaining more
attention within studies related to the planning of future power systems [43].
Furthermore, the effect of increased penetration of DC connected PPMs is not only
associated with the frequency response of the AC system but moreover with the transient
stability of the online generators [42]. Reduced system inertia will reduce the critical clearing
time of the online generators. For the above reasons, TSO’s in order to ensure system security
of supply have introduced non-exhaustive frequency stability related grid connection
requirements for AC connected and now also for DC connected Power Plant Modules (PPMs)
[3].
This chapter discusses the compliance of DC connected offshore wind PPMs with
frequency stability requirements imposed by TSOs, with reference the latest draft of ENTSO-
E. Both point-point connection to single country and MTDC connection between two
asynchronous power systems will be assessed in this chapter. The aim is to illustrate
sensitivities which exist with regard to frequency related requirements, indicate possible
limitations from the offshore wind PPM and from VSC-HVDC system point of view. The
chapter is organised as follow: first the single country DC connected case will be discussed
demonstrating simulation results with a user-defined test system in Matlab/Simulink. The
model allows for the exploitation of all possible control strategies for implementation of both
synthetic inertia and primary frequency response while assessing all possible sensitivities for
grid code compliance. Additionally, the MTDC case will be stressed out in order to indicate
possible interactions which exist between AC systems, DC systems and offshore wind PPMs.
Finally, for the MTDC case, time domain simulations will be performed by making use of
IEEE 39-bus known as New England multi-machine test system in PSS/E.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
5.2 Single country HVDC connected wind power plant

A minimum working case study, to demonstrate possible trade-offs for frequency related
grid connection requirements of a VSC-HVDC link connecting offshore wind PPM is
introduced in figure 47. The time-averaged rms value modelling approach is used for the
VSC-HVDC station [44] [30]. A user-defined dynamic model of the VSC-HVDC station as
shown in figure 48 is developed in Matlab/Simulink [45]. The HVDC link is considered as a
symmetrical bipolar while the AC system is assumed balanced (both in steady state and under
transients). Hence, under this assumption we can assume an equivalent mono-polar
representation of the HVDC link, while maintaining conservation of energy between AC and
DC side. The HVDC cables are modelled by state space differential equations, as presented in
the previous chapter.

1 Frequency
response
Td s  1 Communication 400MW
link Offshore Wind
Df’
AC Power Plant
f
Equivalent system

M=5s
Sg=500MVA
2 1 dc
dc
ac ac
Cdc Cdc

Frequency
f
response
HVDC cable
L12=300km

Figure 47. A simple test system of single country DC connected PPM implemented in
Matlab/Simulink

The offshore wind PPM is modelled as an aggregate full converter wind turbine,
following IEC-61400 standards [20]. The model of wind turbine consists of four sub-models,
the converter control model, the converter interface, the pitch control model, the two-mass
mechanical part model and the aerodynamic model. One major difference with the suggested
by IEC-61400 model [20] is that in the present report case study the aerodynamic part is
modelled by non-linear aerodynamic model instead of linear model. This is considered more
accurate for frequency related studies due to the fact that functions such as synthetic inertia
response will create deviation of the wind turbine states from steady state values which
cannot be captured accurately from the linear model.

5.2.1 Single machine equivalent grid model of the AC system


In power system studies related to load-frequency control, the interest is mainly related to
the collective behaviour of the generators which contribute to the load frequency control
(mainly large generator equipped with governors and speed regulators). Power system
oscillations or fast transient dynamics (higher than 2Hz) associated with the power system
components are neglected and not modelled in such studies. Thus, the common approach is to
represent the system by an equivalent large generator. This equivalent generator has an
equivalent inertia constant Meq equal to the inertia constant of all the generating units in the
power system and is driven by the combined mechanical outputs of the individual turbines. In
addition, the effect of the system load in frequency changes is modelled by a single damping

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
constant D. Figure 48 introduces such an approach. One should note that this is the first but
simple enough approach to study and design load frequency related control schemes and is
widely used and introduced in chapter 9 of [1] [6].
All metrics are represented in per unit on system base while frequency in per unit is equal
to generator speed deviation. An increase in the power system load by ΔPload will cause a
change in frequency change and thus via the governor a change in generators mechanical
power output ΔPm which is considered equal in absolute value with electrical power delivered
to the system. In addition, the effect of the VSC-HVDC links in this model is modeled by an
additional term ΔPvsc which is the additional change or the deviation of steady state value of
the power injected to the system by the DC connected PPM equipped with frequency
sensitive control loops. This control loop for the DC connected PPM will be introduced in
following paragraphs.

Conventional Units
ΔPvsc
ΔPm1 +
+ 1 Δf
ΔPm2
ΔPm- M eq s  D
ΔPmn
ΔPload

Figure 48. Single Machine equivalent grid model for frequency stability related studies [1]

5.2.2 Hydraulic turbine and governor model for the single machine equivalent model

In this single machine representation of the power system frequency response, two types of
turbines could be used either steam or hydro. Both demonstrate different time domain
response for the same load change. For this case study, it was decided to use the hydro turbine
model in order to demonstrate sensitivities and trade-offs for the participation of DC
connected offshore wind PPMs in synthetic inertia response. This choice is justified by the
fact that hydro turbines are more sluggish than gas turbines and it will be easier to
demonstrate the influence of synthetic inertia response provided by DC connected PPMs.
Hydraulic turbines are of two types: impulse turbine (Pelton wheel) and reaction turbines
(Francis propeller). The classic transfer function (which describes the turbine output change
as a function of change in gate opening) of the ideal lossless hydraulic turbine is given as:

Pm 1  Tw s
 (13)
g 1  Tw
2s

Figure 49 shows the transfer function of the hydraulic turbine used and the model of the
governor used in this report analysis. More details about this model and parameters used can
be found in chapter 9 of [1]. In this model, we have included the active power response of the
VSC-HVDC station, represented as ΔPvsc.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
ΔPvsc
Hydro
Model +
Δf_ref=0+ 1 1  sTR 1 Δf
1  sTw +

- 1  sTG R
1  s T TR T
1 s w ΔPm M eq s  D
Rp 2
-

Governor model of
Hydro ΔPload

1
Rp

Figure 49. Equivalent single machine linear model for load-frequency analysis as given in [1].

5.2.3 Synthetic Inertia emulation controller for offshore wind PPM

Offshore wind plant with HVDC connection operate fully decoupled with respect to the
onshore power system frequency. Hence a communication link is needed in order to provide
fast frequency response, as shown in figure 47 [4]. This link is used in order to allow the
offshore wind plants to have access to the onshore power system frequency. In this report we
are going to discuss the method which utilizes communication link in order to provide
synthetic inertia response from offshore wind power plants. Furthermore, beside the
communication loop, an additional controller is needed at the control modules of offshore
wind power plant [4] [48]. This additional control loop takes as input the measured AC power
system frequency at the onshore connection point of VSC-HVDC station (which is send
offshore with communication link) and has as output the change in active power reference of
the PPM, as demonstrated in figure 50 [9].
uwtt 1
From
Wind turbine 1  sTRV
module
 gen
dPmax x / ref
iPcmd
Pref + 1
  
x x 1  sT fp x To converter
-
/ Module
dPmin

init

d
f K inertia
dt

Synthetic Inertia
Controller

Figure 50. Synthetic inertia emulator controller implemented at the active power controller of the
offshore wind PPM.

The output of the synthetic inertia controller is fed as additional power reference on the
model of the PPM. Once the frequency is in steady state value, the output of the synthetic
inertia controller is zero. During the presence of frequency deviation (let say by tripping of a
unit) an additional power reference term is added to the previous steady state reference. In

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
such a way, the PPM changes its active current component and thus the injected to the
collector system active power. This additional active power provided to the system is
provided from the kinetic energy stored in the rotating shaft of the wind turbines. The kinetic
energy of the rotating mass of wind turbines is a function of blades mass and rotational speed,
which is wind speed depended [46]. Thus, the operating-point in terms of generated power
plays a significant role in the de-loading capability of the wind turbines (by means of
supplying additional active power). Finally, the reader should note that the PPM model is
assumed to operate under maximum power point tracking delivering the maximum power
which could be captured from wind. No reserves of wind generation which could be released
for frequency response will be addressed in this report. The focus will be placed only on
assessing limitations with the method of using the kinetic energy of the wind turbines.
Special care need to be given in the amount of power withdrawn from the kinetic energy in
order to prevent the turbines to stall. For given power system frequency deviation, it becomes
clear that the additional active power provided is a function of K inertia control gain. The
sensitivity of such control constant on the system and PPM response will be demonstrated by
time domain simulation in the following paragraph. Finally, the VSC-HVDC link will
transport this additional active power to the system. It will be shown that the small time
constants associated with the controllers and VSC-HVDC link are very fast with respect to
system dynamics and thus will not add significant delay in the combined PPM and HVDC
link response to the power system frequency deviations.
Different sensitivities will be exposed in order to decide possible trade-offs for grid code
compliance of the DC connected PPM. Namely, the parametric sensitivity of the inertia
emulator controller. In addition, a communication-less scheme will be discussed which uses
the stored in over-sized DC capacitors of the converter station energy, as first given in [46]
[4].

5.2.4 Sensitivity of the Kinertia gain of synthetic inertia emulation controller

Figure 51 demonstrates by means of simulations the effect of the synthetic inertia


controller (Kinertia) on the frequency response of single machine test system. From these
simulation results, the frequency response is improved between the case of no participation
(k=0) and with participation of the offshore wind PPM on the inertia response. Both the rate
of change of frequency (ROCOF) and the minimum frequency deviation (NADIR) are
improved. By increasing the inertia controller gain between 0 and 40, has a positive effect on
the power system frequency response. For a value of the inertia controller 60, we observe a
reduction in NADIR and shift in the time instant which NADIR occurs. This is mainly related
to the fact that more power is withdrawn from the kinetic energy of the wind turbines, which
decreases its rotational speed. The reduction per each utilized gain of the synthetic inertia
controller is presented in figure 44 along with the delivered additional power from the
offshore wind PPMs. It is important for the reader to notice that the additional power injection
from the wind power is followed by a drop in the active power below its pre-disturbance
value. This is due to the reduction in the speed of the wind turbines. The higher is the gain the
bigger the contribution of the wind turbines. This translates into larger speed reduction
following the first seconds of the response. The effect of this reduction in active power
response after the first seconds can be also observed in slower frequency recovery of the
system. Such response is known and observed in large systems, as shown in [7]. The reader
should note that the inertia controller gain is strongly related to the energy provided by the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
PPMs [8] [48]. Hence associated to a synthetic inertia constant H, in seconds which can be
assigned to the PPMs.
Sensitivity of K inertia gain on system frequensy response Sensitivity of K inertia on Pelec of PPMs

50 0.49
0.48

Pelec PPM (p.u)


49.8 Unstable!!! ROCOF 0.47
f (Hz)

49.65 0.46
49.6 0.45
K inertia = 0
49.6 0.44 K inertia = 20
49.4 0.43 K inertia = 40
49.55 K inertia = 60
3.8 4 4.2 4.4 0.42
K inertia = 80
49.2 0.41
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time (s) time (s)

Rotor speed of the PPM model Deviation of Power from steady state
0.15
1

0.8 0.1
speed (p.u)

Change in hydro
0.6
p.u
0.05 Change in load
0.4 Change in active power of offshore wind PPM
0
0.2

0 -0.05
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time (s) time (s)

Figure 51. Sensitivity of inertia emulator controller gain on system and offshore wind PPM response
[25]

From this discussion, it can be understood that there is a stability limit or let say a limit of
maximum contribution from offshore wind PPMs in the fast inertia response. In this case
study, for inertia controller gain value of 80, the aggregate wind turbine model will stall and
the system frequency will completely collapse (for this test system). The reduction in speed
and the final stall of the machine is demonstrated in the rotor speed and active power response
of the wind turbines. This effect is well known in literature and is the strongest limitation of
the offshore wind PPMs in contribution of synthetic inertia response [46] [4] [8] [49].
Furthermore, the additional power which the wind turbines can deliver is operating point
dependent as already mentioned. Operating points at high wind speeds, where the rotational
speed is higher can release more kinetic energy than low wind speeds operating points. GE
has performed laboratory tests and has demonstrated for their wind turbines the operating
point dependence to its synthetic inertia controller [26] [25].
As a result, functions such as inertia response provided by offshore wind PPMs need
meticulous design of controls schemes which need to be tuned online for each wind and
rotational speed. In other words, the amount of synthetic inertia which can be provided to the
system is changing during a day and system operators should be aware of such changes in
order to take actions keeping higher spinning reserves. Hence knowing the wind speed per
offshore wind PPM it is possible to determine analytically the available power which could be
delivered as synthetic inertia response. Such methods will not be discussed as they are out of
the scope of this report.
Another more robust method would be to operate the wind turbines under non-optimum
power point tracking which will allow some kind of reserve which can be released for
frequency response functions [48]. Such, approach however places the burden to the project
developers as it has financial impact. The wind plant project developer need to harvest as

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
much power as possible from wind and triggers discussions about possible markets of
ancillary services.
Finally, it should be commented that inertia response is a non-mandatory requirement for
the grid connection of offshore wind PPMs which leaves the flexibility to the relevant TSO to
decide based on the system specific needs. Normal practice is that TSO requires for such
control functions to be installed at the hardware, without in many cases up to date, asking for
activation.
Sensitivity of Kinertia gain on system frequensy response

49.52

49.5
X: 6.658
Y: 49.47 X: 9.366
49.48
Y: 49.47
f (Hz)

49.46 X: 6.462
Y: 49.44

49.44
K inertia = 0
49.42 K inertia = 20
X: 6.448 K inertia = 40
49.4 Y: 49.39 K inertia = 60
K inertia = 80
49.38
4 5 6 7 8 9 10 11 12 13
time (s)

Figure 52. Zoomed frequency response of the test system

It was mentioned above that the presence of the VSC-HVDC link does not add any
significant delay in the inertia response provided to the onshore power system by offshore
wind PPMs. The time constants associated with the control modules of the VSC-HVDC
system are infinitely fast in comparison to slow time constants of system frequency response
[4] [47]. This is demonstrated in figure 53, where the power delivered by the offshore wind
PPM is plotted with the power at the receiving end of the VSC-HVDC system.
Active Power at the PPM and
at the Grid-Side VSC-HVDC station
240
Received power onshore
Power at PPMs offshore
235

230
MW

225

220

215
0 10 20 30 40 50 60 70
time (s)

Figure 53. Power at remote offshore and onshore converter station for the participation of
PPM in frequency response

The most important control module in the case of the VSC-HVDC link is the DC voltage
controller at the onshore converter station. Once the power delivered by the offshore wind
power plant is increased/decreased, as a result of inertia emulator controller, the DC voltage
will vary accordingly and the onshore converter will modify the active power injection to the
system. In such a way, the same variation of power from offshore wind PPMs is send to the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
grid connection point without any significant delay. At this point, the reader should note that
the DC voltage variation in the DC link is dependent on the active power generated by the
offshore wind PPM. Figure 54 demonstrates the sensitivity of synthetic inertia controller on
the DC link voltage response. The reader should note that in the case of stalling of the
machines, the DC voltage of the link will collapse, leading to the disconnection of the HVDC
link.
Sending DC Terminal
253

252.5
Udc (kV)

252 K inertia = 20
K inertia = 40
251.5 K inertia = 60
K inertia = 60

251
0 10 20 30 40 50 60 70
time (s)
Receiving DC Terminal
250.5

250
Udc (kV)

K inertia = 20
K inertia = 40
249.5 K inertia = 60
K inertia = 80

249
0 10 20 30 40 50 60 70
time (s)

Figure 54. Variation of sending and receiving DC terminal voltage for the participation of the
DC connected PPM in the frequency response.

As a result, both the DC voltage variations and the possibility of DC voltage collapse
should carefully be assessed in the design process of such control schemes in order not to
jeopardise system normal operation or not interact with over-voltage or under-voltage control
schemes of the VSC-HVDC link. The latter argument is also vital when we talk about MTDC
connections where the DC side voltage variations influence the operation of remote and in
many case asynchronously connected power systems [32]. Such a case study will be
discussed in following paragraphs.

5.2.5 Synthetic Inertia emulation controller using the stored in DC capacitors electro-
static energy

Next to the method which utilizes the stored kinetic energy in the wind turbine shaft, a
second approach has been proposed by the authors in [8]. This method utilizes the stored
electro-static energy in the DC link capacitors of the VSC-HVDC station. For such a case
study, the DC capacitors have been significantly oversized. The offshore wind PPMs does not
actively participate in such a scheme, and continue to inject the same amount of active power
to the HVDC link. Thus, the inertia emulator controller for such a case study is placed as
additional control loop on the onshore VSC-HVDC station, and communication link is not
required.
Figure 55 introduces the control loop for synthetic inertia response provided by the DC
capacitors of the VSC-HVDC onshore station. As it can be seen, for the case of synthetic
inertia emulator applied at the VSC-HVDC converter, the same method which utilizes the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
derivative of the frequency is used. The frequency is directly measured at the grid connection
point at the onshore converter station.
Synthetic Inertia
Controller

f s K inertia
0.1s  1

Active current
component of
idmax
the VSC-HVDC
U dc 0 + -
kp
- id + i min idref
d
DC Voltage Droop
U dc Controller i d0
(b)

Figure 55. Synthetic inertia control loop implemented as additional control loop on the DC voltage
controller of VSC-HVDC converter station

An important limitation in this method is that the large DC capacitors need to be added to
the DC circuit. Authors in [8] suggested values which are between 2-8mF that are
significantly higher with what is commonly used in the industrial applications [50]. Figure 56,
demonstrates the effect of different capacitances on the system frequency response of the
simple test system, when such synthetic inertia method is used. From the simulation results,
we may see that the ROCOF is improved in the first seconds after the disturbance is applied.
However, the discharge of the DC capacitors will be followed by a DC voltage drop. This DC
voltage drop will cause the DC voltage controller of the receiving end converter to reduce the
active power injection reducing significantly the injected to the AC system power. Thus, for
the given test system and case study, even though the ROCOF is improved, the NADIR is
deteriorated to values much lower than the case without additional synthetic inertia control
loop. So at least for this given case study, and for the selected method & control parameters,
such a synthetic inertia method proves to be non-robust and non-reliable due to the impact of
DC voltage on the dynamic response of the receiving-end converter. Higher values of the DC
capacitor may reduce the DC voltage drop and improve system response, but still the active
power reduction is significant. A more detailed and meticulous sensitivity analysis will
demonstrate all possible limitations of such a method.
As it has been discussed in the method which uses kinetic energy stored in the wind
turbines shaft, the sensitivity of the inertia emulator gain is important. Figure 57,
demonstrates that higher gains will normally lead to more active power injected to the system
which during the first seconds will improve the ROCOF but eventually will bring NADIR to
lower values. This reduction in active power is due to the DC voltage drop, as it can be seen
from the simulation results.
In general, the conclusion which can be drawn based on this small test system and the
simulation results of this report is that the method which utilized DC caps energy will provide
an improvement in the ROCOF during the first seconds of the disturbance but will eventually
lead to higher values of NADIR.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
frequency response Deaviation of GSVSC from Steady State
0.08
50 Cdc=2mF
0.06 Cdc=4mF
Cdc=6mF
49.8 0.04
ROCOF Cdc=8mF
50 0.02 No Synthetic Inertia

f (Hz)

p.u
49.6
49.9 0
0.07
49.8 -0.02
49.4 0.065

-0.04 0.06

1.5 2 2.5 3 3.5 1 1.2 1.4


49.2 -0.06
0 10 20 30 0 5 10 15 20 25 30
time (s) time (s)

Udc - sending DC terminal Udc - Receiving DC terminal


260 256

254

255 252

250
kV

kV
250 248

246

245 244

242

240 240
0 5 10 15 20 25 30 0 5 10 15 20 25 30
time (s) time (s)

Figure 56. Sensitivity of the DC capacitance on synthetic inertia provided by VSC-HVDC system.

frequency response VSC-HVDC active power deviation


50.2 0.08
0.06
50 0.06 0.04
49.8 0.04 0.02
f (Hz)

0
49.6 ROCOF 0.02
p.u

1 1.2 1.4
50
49.4 0

49.2 49.9
-0.02

49 1.2 1.4 1.6 1.8 2 2.2 -0.04


0 10 20 30 40 0 10 20 30 40
time (s) time (s)

DC terminal votlage at sending end DC terminal votlage at receiving end


260 255

255
250
kV

kV

250

k inertia =20 245


245 k inertia =50
k inertia =100

240 240
0 10 20 30 40 0 10 20 30 40
time (s) time (s)

Figure 57. Sensitivity of control synthetic emulator controller gain on the system response.

5.2.6 Combined method which uses both wind plant and DC capacitors contribution

A new proposal which could be made in this report is combining both methods using a
communication link. In this approach for the same disturbance both controllers will
contribute. As it can be seen in figure 58, such a control scheme provides better results in
terms of frequency response for our test system. Both the ROCOF and the NADIR are
improved from the case without frequency response.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Combine method of synthetic inertia
50.1

50

49.9 ROCOF
50
49.8
49.95

f (Hz)
49.7

f (Hz)
49.9
49.6
49.85
49.5
1 1.5 2
49.4 K inertia = 0
K inertia = 40
49.3
DC cap method (4mF) + PPM inertia
49.2
0 10 20 30 40 50 60 70
time (s)

Figure 58. Proposed combined method of inertia response by both PPMs and DC caps

Of course, in order to draw very strong conclusions about this method a meticulous and
detailed sensitivity analysis need to be performed. For now, due to the limited space of this
report we will conclude that such a method is promising and overcomes the limitations
discussed.

5.3 Synthetic inertia response by a multi-terminal DC grid

In the previous paragraph the synthetic inertia response provided by VSC-Based HVDC
connected offshore wind PPMs has been discussed. The same control schemes will be
addressed for the case where the PPMs are connected in MTDC configuration which extends
between two asynchronous power systems. Authors in [50] have suggested a communication-
free scheme for synthetic inertia response provided by offshore wind PPMs connected in
MTDC grids, extending the proposals in [4] [47].
For single country (point-to-point) DC connected PPMs, it was shown in this report that
there are two main sources of energy used to provide synthetic inertia response, either kinetic
withdrawn from the shaft of the offshore wind turbines or electrostatic, stored in the oversized
DC capacitors of the VSCs. Differences and limitations of these two methods have been
stressed out in previous paragraph. However, for the MTDC grid connection, used for a
connection of two asynchronous power systems and offshore wind PPM, the required energy
could be additionally (except from wind PPM) withdrawn from the remotely MTDC
connected power system. The advantage of this approach is that the required additional power
is available anytime, as soon as the remote converter is connected to a relevantly strong
network. With such a proposed scheme the power injections from the AC power systems into
the DC grid are coordinated in such a way that the frequency deviations of all the AC areas
stay close to each other despite power imbalances.
Furthermore, in such a method there is no need for communication link between the
onshore converters, which provide synthetic inertia response. The variation of the DC voltage,
as a result of synthetic inertia response provided by the onshore converters will adjust
accordingly the power delivered or withdrawn from the remote system, based on DC voltage
droop characteristics that the remote system onshore converter applies [43] [39] [51].

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
U dc
Rectifier Inverter
B
U dc B
Droop line equation
1
U dc 0 A U dc   id  id 0   U dc 0
kp
1
slope 
kp
0
id 0 idB id
(a)

DC Voltage Droop Active current


Controller component of
idmax
the VSC-HVDC
U dc 0 + -
kp
- id + i min idref
d

U dc id 0
(b)

Figure 59. (a) DC voltage droop line (b) controller for implementation.

The reader should recall that DC voltage control in MTDC offshore grids is performed by
application of droop controllers [44] and is strongly related to power balance between
offshore and onshore converters [51] [39]. Each onshore converter is controlling locally its
active power injection to the AC system, following a droop line characteristic between DC
voltage and active power injection to the onshore terminal. In such a way it contributes to the
balancing of the MTDC grid, hence in control of DC voltage control. Figure 59,
demonstrates such a droop line characteristic, and the controller for its implementation. Note
that in this figure 59, the droop controller is applied at the active current component of the
converters.
Thus, for the MTDC case study demonstrated in this report the focus will be placed only on
synthetic inertia response provided by the onshore VSC-HVDC stations, while the PPMs are
not participating at all. In future work we will deal and compare the methods which use
additional power from the offshore wind PPM as recommended in [50]. Figure 60
demonstrates a minimum MTDC grid extending between two AC power systems. The DC
grid connects a total amount of 400MW wind generation to both power systems. Both AC
power systems are modelled following the single machine equivalent approach. It is assumed
for the following simulations that power is flowing from offshore wind PPMs to onshore
converter stations.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
400MW
Offshore Wind
Power Plant
HVDC cable
L12=L13=300
km AC 1
slack
Equivalent system
3 ac Meq=7s
400MW
ac dc
Df Sg=500MVA
dc 1
Cdc
AC 2
Cdc
Equivalent system
200MW

Sg=500MVA dc 2
Meq=5s ac Cdc

Figure 60. Simple multi-terminal DC grid coupling two AC power systems

5.3.1 Effect of Kinertia controller gains

Let us assume that there is an increase of 0.1p.u load in system AC2, same as in the point-
point case study. Normally, if the VSC-HVDC station 2 is not equipped with additional
synthetic inertia control loop, it will not respond in the frequency change. With the synthetic
inertia controller in operation, the onshore converter station will contribute in improving the
frequency response. The question which then arises is how different the system response
could be and what the interactions are between the different converters in the MTDC grid,
when one is equipped with synthetic inertia controller. The reader at this point should notice
that the offshore converter station continues to inject constant wind generated active power
during the disturbance and is not participating in the load-frequency response.
As it can be seen in figure 61, with the synthetic inertia control loop in operation, for a 0.1
increase in load, the system frequency response is improved in terms of ROCOF. Increasing
the synthetic inertia controller gain, larger active power contribution from the HVDC
converter station is achieved with further reduction in the ROCOF. However, same as in the
case of point-point DC connected PPM the NADIR is reduced when higher synthetic inertia
controller gains are applied. The main reason for that is the reduction in the DC voltage at all
the DC nodes, as a result of power extraction from the DC grid, followed by discharge of DC
capacitors. Once the DC voltage drops at the DC terminal, the droop controller will lower the
injection of active power and in such a way counter-act the synthetic inertia controller.
In addition, as demonstrated in figure 61, the DC voltage drop is experienced
instantaneously at all DC nodes. The remotely DC connected onshore converter station, is
operating also in DC voltage control mode following a given droop line characteristic. Once
the DC voltage drops, the droop DC voltage controller will reduce the active current
component of the VSC (d-axis) following the DC voltage variation. Hence, the remotely
MTDC connected onshore converter station will disturb the second remotely connected
system. Figure 61 demonstrates the frequency response of AC1 system as a result of synthetic
inertia response in systems AC2.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
AC 2 system - frequency response AC 1 system - frequency response
50
50.5

49.8

f [Hz]

f [Hz]
50
49.6

49.4 49.5

0 10 20 30 40 0 10 20 30 40
time (s) time (s)
Deviation of active power Deviation of active power
0.15 0.05

0.1
DP (p.u)

DP (p.u)
Hydro
0.05 0

VSC-HVDC station
0

-0.05 -0.05
0 10 20 30 40 0 10 20 30 40
time (s) time [s]

DC terminal 2 volltage DC terminal 3 volltage


321 321
320 320
319
Udc [kV]

Udc [kV]

319
318 320 320
318
317
315 317
316 315
1 1.5 2 2.5 1 1.5 2
315 316
0 10 20 30 40 0 10 20 30 40
time (s) time (s)
No inertia emulator
K inertia = 20
K inertia = 50

Figure 61. sensitivity of the synthetic inertia emulator controller on the DC grid and power system
response

Based on these set of simulation results with this small test system, it becomes clear that
the MTDC connection of offshore wind PPMs, could be beneficial for the system where
synthetic inertia response in provided. However, due to the coupling of the two systems, a
disturbance is added to the remotely connected AC system. Of course the magnitude of the
disturbance is very much related to the stiffness, to equivalent inertia constant of this system
and to the short circuit ratio of the converter station. Finally, the reader should notice that the
DC capacitors are not oversized in this case study.

5.4 Impact of reduced system inertia on frequency response of New England test system

In previous paragraphs, methods and control strategies for the emulation of synthetic
inertia response by DC connected offshore wind PPMs have been demonstrated. The single
machine representation has been used for demonstrating controls functionality [6] [1]. This
paragraph will focus more on the impact of large scale penetration of DC connected offshore
wind generation on the frequency response of the IEEE 39 bus system. The latter will provide
better insight about the response of real multi machines systems. Figure 63 introduces the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
topology of the MTDC test system. The IEEE 39-bus system is coupled via the MTDC grid to
Great Britain benchmark model.

30
400 MW

247.4 kV 2
ac
dc
301
600 MW ac 3 1
dc NE system
1
252.3 kV
1002 39
1001 1008
8
9
dc
250 kV 302
ac 247.4 kV ac
303 5
252.3 kV dc
4
1003 1004 380 MW
2 ac 400 MW
dc
600 MW
1005
Meshed DC grid (250kV)
1007 L13, L15, L25, L24, L34=100km
1006 L56, L46, L63 =100km

Figure 62. A meshed MTDC grid connecting

Figure 63 demonstrates the three scenarios for the IEEE 39-bus (NE system) test system,
as first used in [33]. For these three snapshots in figure 63, the inertia constant H in seconds,
is provided in figure 64 per each area of IEEE 39-bus system. Important is to observe that
loads are represented as a constant impedance and does not affect the results. In these set of
simulations, only the conventional units are participating in load frequency control or else
primary frequency response. The control parameters of the conventional generator governors
can be found in [33]. None of the two onshore wind PPMs and HVDC converters of IEEE 39-
bus system are equipped, for these results with any frequency response control loops.

Penetration per given scenario


100
90 VSC-Based
Offshore Wind
80
Generation
70
60 Conventional
50 Generation
[%]

40
30
Onshore Wind
20 Generation
10
0
Scenarios (A: left, B: middle, C: right)

Figure 63. Selected scenarios for IEEE 39 bus test system [33]

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Inertial Constant per Area

20 18.9

15.2
15 13.3

Base case

H [s]
10
case b
5 3.2 2.7
2.4 3 3 case c
2.2

0
1 2 3
Area

Figure 64. Reduction in equivalent inertia constant per area [33]

The impact of such a decrease of system inertia constant on the frequency response is
shown in figure 65 for a trip of the onshore wind generation unit at bus 21. In both cases, a
same disturbance is added. System frequency response is compared between scenario A (the
base case) and scenario B which is the case where DC connected offshore wind is replacing a
conventional unit. Comparing the two cases, it can be seen that both ROCOF and NADID are
lower in the case of scenario B which is justified by the reduction in total system inertia
constant.

Frequency response of New England (IEEE 39-bus System)


50.1
49.8
0.125Hz/sec

50 49.75 0.2Hz/sec

49.7 ROCOF
49.9
2.2 2.4 2.6 2.8
f (Hz)

49.8

49.7

49.6
Base Case- Scenario A
Scenario B
49.5
0 5 10 15 20
time[s]
Figure 65 frequency response of IEEE 39-bus system for two scenarios presented in figure 64

In general, it can be concluded that the reduction of equivalent system inertia constant H is
followed by faster and larger frequency deviations. In addition, control actions need to be
taken in order to compensate this response either by re-tuning (change in droop
characteristics) of the governors of the conventional units (frequency droop) or introducing
frequency response control loops for participation of VSC stations in load-frequency
response.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
5.5 Participation of MTDC grid in the load-frequency control of power systems –
Demonstration with IEEE 39 bus test system model

Such a MTDC grid, as shown in figure 62, could be used to provide load–frequency control
functionality. As discussed, the advantage of the VSC to control active power could be used
in order to stabilize [52] the system and/or provide load-frequency control [43]. The case of
synthetic inertia response has been discussed and is not going to be stressed out in this
paragraph. The focus will be placed only on primary frequency response and interaction
between two MTDC connected systems.
Primary frequency response is the response of the synchronous generator governors in the
presence of a frequency deviation. VSC-HVDC converters operating in MTDC configurations
could also participate in such control schemes, especially when coupling two asynchronous
power systems. In addition, this could be combined with the participation of offshore wind
PPMs if reserves are applied. However, for now only the frequency response provided by
VSC stations operating in MTDC grids will be addressed.
The frequency droop controller used in these simulation results operates on the active
current component, as shown in figure 66 [9]. A dead-band is not used for simplicity. Having
as input the frequency measures at the grid connection point, a droop controller with a
predefined droop constant is implemented which controls the active current reference set
point, hence the active power delivered to the system.

Frequency droop
cotroller

1
Δf
Rf

Active current
max component of
i
d the VSC-HVDC
U dc 0 + -
kp
- id + i min idref
d
DC Voltage Droop
U dc Controller i d0

Figure 66 Frequency droop controller implemented as additional loop on the DC voltage droop controller

Applying the control loop of figure 66 on the DC voltage droop controller of VSC 301 and
VSC 302, we observe that the frequency response of the IEEE 39-bus system is improved
(see figure 61). Both NADIR and ROCOF metrics are better in the latter case and closer to
the base case. In addition, figure 68 demonstrates the change of active power delivered by the
three onshore converters in the MTDC grid. As it can be seen, VSC 301 and VSC 302
increase their active power injection to the IEEE 39-bus system participating in the load-
frequency response.
On the remotely MTDC converter 303, the active power is reduced (see figure 60). This
happens as a result of the DC voltage controller reaction of VSC 303 which in order to control
the DC voltage at normal operating levels will reduce the delivered active power, thus
bringing power balance between onshore and offshore converters. Hence, given that the
offshore wind generation, thus the injected to the MTDC grid power by the offshore VSCs
does not change, the increase in active power of VSC 301 & 302 is followed by decrease in
VSC 303. Simulation results of DC voltage will be illustrated in next paragraph where the

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
sensitivity of the DC voltage droop constant to system frequency response will be observed.
Finally, the reader should notice the different droop constant (Rf) selection for VSC 301 and
VSC 302 (which is 0.03 and 0.05 respectively). This choice is made in order to demonstrate
the impact of frequency-droop constant to the active power response of the converters station.
Figure 69 demonstrates the sensitivity of frequency droop gain on active power response of
the converters.

Frequency response of New England (IEEE 39-bus System)


50.1
49.8
0.125Hz/sec
50 49.75 0.2Hz/sec
49.7 ROCOF
49.9
2.2 2.4 2.6 2.8
f (Hz)

49.8

49.7

49.6 Base Case- Scenario A


Scenario B
With freq. response by HVDC station
49.5
0 5 10 15 20
time[s]

Figure 67. Frequency response of the New England test system in loss of conventional
generation.

Active Power Deviation from steady state for VSC-HVDC stations in MTDC
0.3
DP VSC-HVDC 301
0.2 DP VSC-HVDC 302
DP VSC-HVDC 303

0.1

0
Both Converters of NE Ssystem will
change the active power injeciton
DP [p.u]

-0.1

-0.2

-0.3 The remote converter which performs


DC voltage contro
will reduce injection of active power
-0.4

-0.5
0 5 10 15 20
time[s]

Figure 68. The active power response of VSC stations participating in load-frequency response.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
It is worth mentioning that the active power delivered by the onshore VSC station should
always respect the maximum capacity of the converter. This means that in situations where
the converter is over-loaded, the participation in systems load-frequency response may lead to
the activation of current limitation strategies, which depending on priority given may cancel
load-frequency response, due to reduction in active current component of the VSC station.
Finally, the oscillations observed in the active power response are power system oscillations
which are triggered in the AC system by the tripping of the generation unit. These oscillations
are present in the frequency signal measured at the grid connection point of VSC 301 & 302.
Hence, they are transferred via the frequency droop to the active power set point change of
the VSC stations.
301

0.3
0.25
Dp [p.u]

0.2
0.15
0.1 Rf=0.05
0.05 Rf=0.04
0
Rf=0.03
-0.05
0 2 4 6 8 10 12 14 16 18 20
time[s]
302

0.3
0.25
DP [p.u]

0.2
0.15
0.1
0.05
0
-0.05
0 2 4 6 8 10 12 14 16 18 20
303
0.05
0
-0.1
-0.2
DP [p.u]

-0.3
-0.4
-0.5
-0.6
-0.7
0 2 4 6 8 10 12 14 16 18 20

Figure 69. Sensitivity of frequency-droop constant Rf on active power reponse of the


converers.

The coupling between DC voltage and active power response of the VSC in MTDC
configuration is well understood and has been discussed in the case of synthetic inertia
response. The same principles apply also for the case of primary frequency response with the
only difference being the fact that participation of the VSC stations in load-frequency control
actions will change the DC terminal voltage levels when steady state is achieved. Figure 70
demonstrates the sensitivity of the DC voltage response on different values of frequency
droop constant Rf. For example, between 0.03 and 0.05, the first case involves higher
participation from the converters, which means both larger voltage drop and steady state DC
voltage deviation. Finally, such a relation between the frequency and DC voltage variation
could be quantified with simple formulas, but is beyond the scope of the present report.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
DC grid terminal voltage 201 DC grid terminal voltage 202
1.01 1.01

Udc [p.u

Udc [p.u
1.005 1.005

1 1

0.995 0.995
0 5 10 15 20 0 5 10 15 20
time[s] time[s]
DC grid terminal voltage 301 DC grid terminal voltage 302
1 1
Udc [p.u

Udc [p.u
0.99 0.99

0.98 0.98

0 5 10 15 20 0 5 10 15 20
time[s]
DC grid terminal voltage 303 time[s]
1.015

1.01 Rf=0.05
Udc [p.u

1.005 Rf=0.04

1 Rf=0.03
0.995
0 5 10 15 20
time[s]

Figure 70. Senstivity of droop-frequency constant on DC voltage response.

Finally, the sensitivity of the frequency-droop constant Rf on the system frequency


response of the IEEE-39 bus system and the disturbance added to the second system (Great
Britain Benchmark), is demonstrated in figure 71. From these simulation results we observe
that for the case of 0.03 droop, due to higher participation, there is slightly higher NADIR.
Important however, is to focus more on the impact on the second system, the remotely MTDC
connected Great Britain benchmark. From these results, the lower is the frequency-droop gain
Rf, the bigger the disturbance as a result of larger power withdrawn from the VSC 303. The
reader should also notice that for this case study, offshore wind generated power is imported
in both cases from two power systems, thus the type of disturbance in Great Britain system
has the form of frequency drop. In any case, this impact should be considered and assessed in
the planning and expansions studies of power systems with MTDC connections.
Another aspect which may create problems in the case where VSC-HVDC stations are
participating in fast frequency response functions is the overvoltage which may be triggered
as a result of fast change in active current injection by the converter station. In equations (5)
& (12) it has been discussed the relation between the AC connection point voltage change as
a function of active and reactive current component of the converter station. The sensitivity of
voltage variation at the grid connection point to the active current injection from the converter
station is very much depended on the R/X ratio. Hence, in systems with relevantly high R/X
ratio the participation of the VSC station in synthetic inertia or in primary response may lead
to over-voltages. The latter should be taken into account when such control schemes are
applied.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
IEEE New England frequency
50
49.62
Rf=0.05
Rf=0.04
49.6
Rf=0.03

Hz
49.8 3.5 4 4.5 5 5.5

49.795

49.79
49.6 19.4 19.45 19.5 19.55
0 5 10 15 20
time[s]
UK system frequency
50

49.99
Hz

49.98

49.97
0 5 10 15 20

Figure 71. Sensitivity of frequency droop gain Rf on (up) IEEE-39 bus system (down) Great Britain
benchmark system

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
5.6 Discussion

With reference to the frequency stability requirements, it has been demonstrated that the
large penetration of DC connected FLOW power plant modules will lead to decrease in
system inertia with higher risks in frequency stability of the power system. The contribution
of point-point DC connected power plants in the system frequency response, by additional
control loops, has been also demonstrated. This additional control loops implemented at
offshore power plant modules have shown that synthetic inertia response could be provided
by utilization of the wind turbines kinetic energy. The necessary condition for this scheme is
the operation of the wind turbine above the minimum speed. A communication link between
the onshore connection point and the offshore wind power plant is necessary for such scheme
(although different proposals in literature have been made for communication-less schemes).
The reader should note that such a synthetic inertia response is operating point depended
and special care is needed in the design of synthetic inertia emulators control schemes. It is
important to mention that if the emulated inertial response of wind turbines is appropriately
tuned, it will lead to improved system response. However, it was shown that attempts to make
inertia response as large as possible, will result in deeper frequency drops (NADIR) or even
stall of the wind turbines. A second approach for synthetic inertia response which utilizes the
energy stored in the DC capacitors is discussed as well. Simulation results and sensitivity
analysis have shown that the ROCOF is improved but the maximum frequency deviation
(NADIR) of the system is deteriorated. Finally, a combined method is proposed which
exhibits improved performance however more extensive sensitivities are necessary to draw
strong conclusions.
The general conclusion from this analysis (based on the given test system and models) is
that the single country DC connected PPMs are capable of contributing synthetic inertia
response to the onshore power system. However, this contribution is operating point
dependent. It is well known that variable speed wind turbines can only contribute an emulated
inertia response if operating above their minimum speed. The latter, raise questions about the
degree that onshore power systems and mainly TSOs would rely on synthetic inertia response
provided from wind generation. If significant uncertainty exists over the stored inertial
energy available at time of low synchronous inertia on the system, there is a risk of frequency
excursions greater than normal. In order to determine the aggregate inertial response
capability of offshore wind power plants, TSOs would need to be able to predict how many of
the turbines are operating above the minimum speed and thus are capable of providing inertia
response. Online monitoring systems which will have as input wind speed at given sites could
help for the calculation of available inertia from offshore wind PPMs.
Finally, the multi-terminal DC connection of power plants, extended between two remote
asynchronous power systems demonstrate benefits in terms of frequency response functions
(both inertia and primary) due to the available energy which could be exchanged between
these systems for frequency support. A fully operational MTDC grid with offshore wind
plants can be considered as a large (virtual) power plant capable of proving functions such as
synthetic inertia and primary load-frequency control. However, in the case of MTDC
connection, interactions between the HVDC converter station (mainly DC voltage controllers)
need to be understood and addressed before such schemes are applied. Such interactions are
very important not only for the case of frequency response but also for understanding the
operation of combined AC/DC systems both under normal and faulted conditions

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
References

[1] P. Kundur, Power System Stability and Control, McGraw-Hill Inc..

[2] T. T. Gmbh, "Requirements for Offshore Grid Connections," October 2010.

[3] ENTSO-E, "ENTSO-E Draft Network Code on High Voltage Direct Curretn Connections and
DC Connected Power Park Modules," Brussels, November 2013.

[4] Haileselassie, T.M.; Torres-Olguin, R.E.; Vrana, T.K.; Uhlen, K.; Undeland, T., "Main grid
frequency support strategy for VSC-HVDC connected wind farms with variable speed wind
turbines," PowerTech, 2011 IEEE Trondheim , vol., no., pp.1,6, 19-23 June 2011.

[5] Z. S. Zhang, Y. Z. Sun, J. Lin, G. J. Li, "Coordinated frequency regulation by doubly fed
induction generator-based wind power plants".IET Renewable Power Generation, 2012,
vol.6,Iss.1, pp.38-47.

[6] Yunfei Mu, Jianzhong Wu, Ekanayake. J, Jenkins. N, Hongjie Jia, "Primary Frequency Response
From Electric Vehicles in the Great Britain Power System," Smart Grid, IEEE Transactions on ,
vol.4, no.2, pp.1142,1150, June 201, 2013.

[7] N. W. Miller, M. Shao, S. Venkataraman, C. Loutan, M. Rothleder, "Frequency response of


California and WECC under high wind and solar conditions," in Power and Energy Society
General Meeting, 2012 IEEE , vol., no., pp.1,8, 22-26 July 2012.

[8] Jiebei Zhu, Booth, C.D., Adam, G.P., Roscoe, A.J., Bright, C.G., "Inertia Emulation Control
Strategy for VSC-HVDC Transmission Systems".Power Systems, IEEE Transactions on , vol.28,
no.2, pp.1277,1287, May 2013.

[9] Lingling Fan; Zhixin Miao; Osborn, D., "Wind Farms With HVDC Delivery in Load Frequency
Control".Power Systems, IEEE Transactions on , vol.24, no.4, pp.1894,1895, Nov. 2009.

[10] Arjen A. van der Meer, Ralph L. Hendriks and Wil L. Kling, "A survey of fast power reduction
methods for VSC connected wind power plants consisting of different turbines types," EPE-WE
Seminar , 2009.

[11] Christian Feltes, Holge Wrede, Friedrich Koch and Istvan Erlich, "Fault Ride-Through of DFIG-
Based Wind farms connected to the Grid through VSC-Based HVDC link," Power Systems,
IEEE Transactions on V.24, Issue3, p.p 1537-1546 , 2009.

[12] Erlich, I.; Shewarega, F.; Engelhardt, S.; Kretschmann, J.; Fortmann, J.; Koch, F, "Effect of wind
turbine output current during faults on grid voltage and the transient stability of wind parks,"
Power & Energy Society General Meeting, 2009. PES '09. IEEE , vol., no., pp.1,8, 26-30 July
2009.

[13] S. K. Chaudhary, R.Teodorescu, P. Rodriguez and P.C. Kjar, "Chopper Controlled Resistors in
VSC-HVDC Transmission for WPP with Full-scale converters," in Sustainable Alternative
Energy (SAE), 2009 IEEE PES/IAS Conference, pp.1-8, 28-30 Sept. 2009.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
[14] M. Ndreko, A. A. van der Meer, J. Bos, M. Gibescu, K. Jansen and M.A.A.M van der Meijden,
“Transient Stability Analysis of an Onshore Power System With Multi-Terminal Offshore VSC-
HVDC: A Case study for The Netherlands”.IEEE PES General Meeting, July 2013, Vancouver
BC.

[15] Ndreko, Mario, Marjan Popov, Jens C. Boemer, and Mart van der Meijden. , Sensitivity analysis
on short-circuit current contribution from VSC-HVDC systems connecting far and large offshore
wind power plants, In Innovative Smart Grid Technologies Conference Europe (ISGT-Europe),
2014 IEEE PES, pp. 1-6. IEEE, 2014., 2014.

[16] Arjen van der Meer, Mario Ndreko, Madeleine Gibescu, and Mart van der Meijden., "The Effect
of FRT Behavior of VSC-HVDC Connected Offshore Wind Power Plants on AC/DC System
Dynamics," IEEE Transactions on Power Delivery, vol. 99, no. 99, p. 99, 2015.

[17] Silva, B.; Moreira, C.L.; Leite, H.; Lopes, J.A.P., "Control Strategies for AC Fault Ride Through
in Multiterminal HVDC Grids".Power Delivery, IEEE Transactions on , vol.29, no.1,
pp.395,405, Feb. 2014.

[18] M. Ndreko, A. M. Bucurenciu, M. Popov and M.A.M.M van der Meijden, "On Grid Code
Compliance of Offshore MTDC Grids: Modeling and Analysis," in IEEE PowerTech,
Eindhoven, Netherlands, 2015.

[19] Ellis, A.; Kazachkov, Y.; Muljadi, E.; Pourbeik, P.; Sanchez-Gasca, J. J., "Description and
technical specifications for generic WTG models — A status report," in Power Systems
Conference and Exposition (PSCE), 2011 IEEE/PES , 20-23 March 2011.

[20] "IEC 61400-27-1: Wind Turbines - Part 27-1: Electrical Simulation models for wind power
generation," IEC, vol. Committee draft (CD), no. Ed.1, 04 2012.

[21] EWEA, "Delivering Offshore wind Power in Europe," June 2012. [Online]. Available:
http://www.ewea.org/index.php?id=1976.

[22] EWEA, "The European Offshore wind Industry key 2011 trends and statistics," June 2012.
[Online]. Available: http://www.ewea.org/index.php?id=1976.

[23] Asmine, M.; Brochu, J.; Fortmann, J.; Gagnon, R.; Kazachkov, Y.; Langlois, C.-E.; Larose, C.;
Muljadi, E.; MacDowell, J.; Pourbeik, P.; Seman, S.A.; Wiens, K., "Model Validation for Wind
Turbine Generator Models," Power Systems, IEEE Transactions, pp. vol.26, no.3, pp.1769,1782,
Aug. 2011.

[24] Ekanayake, J. and Jenkins, N., "Comparison of the respons of doubly fed and fixed speed
induction generators wind turbines to changes in network frequency".Energy Conversion, IEEE
Transactions on , vol.19, no.4, pp. 800- 802, Dec. 2004.

[25] Kara Clark, Nikolas W. Miler & Juan J. Sanchez Gasca, "Modeling of GW wind turbines for grid
studies," GE, 2008.

[26] Kara Clark, Nicholas W. Miller and Juan J. Sanchez-Gasca, "Modeling of GE Wind Turbine-
Generators for Grid studies," NY, USA, 2010.

[27] Siemens PTI, PSSE Operational Manual.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
[28] Pouyan Pourbeik, "Proposed Changes in WECC WT4 Generic Model for Type 4 Wind Turbine
Generators," Knoxville, USA, 2011.

[29] Nikos Flouretzou, Vassilios Angelidis and Georgios D. Demetriadis, "VSC-Based HVDC Power
Transmission Systems: An Overview," IEEE Transactions on Power Electronics, Vol.24, NO.3,
March , 2009.

[30] J. Beerten, S. Cole and R. Belmans, “Modeling of Multi-terminal VSC HVDC Systems with
Distributed DC Voltage Control,” IEEE Transactions on Power System,s vol.PP, no.99, pp.1,9,
0, 2012.

[31] M. Ndreko, "Offshore Wind Power Connected to the Dutch Transmission System by VSC-
HVDC Networks: Modeling and Stability Analysis," 2012.

[32] M.Ndreko, A. A. van der Meer, M. Gibescu and M.A.M.M van deer Meijden, "Impact of DC
Voltage Control Parameters on AC/DC System Dynamics Under Faulted Conditions," in IEEE
PES GM 2014 (submitted), 2014.

[33] Mario Ndreko & Madeleine Gibescu, "Multi-terminal DC connected Offshore Wind Power Plant
Modules," FLOW project report, Deliverable 2B, 2013.

[34] Mario Ndreko, Madeleine Gibescu, "Offshore wind power plant - Single country connection,"
Delft University of Technology, 2013.

[35] C. Ismunandur, "Control of Multiterminal VSC-HVDC for Wind Power Intergration," in MSc
thesis TU delft.

[36] Stijn Cole, Jef Berteen and Ronnie Belmans, “Generalized Dynamic VSC MTDC Model for
Power System Stability Studies,” Electric Power system Research - Science Direct, 2011.

[37] Eduardo Prieto-Araujo, Fernando D, Bianchi, Adria Junyent-Ferre and Oriol Gomis-Bellmunt,
"Methodology for Droop Control Dynamic Analysis of Multi-terminal VSC-HVDC Grids for
Offshore Wind Farms," IEEE Transactions on Power delivery, Vol.26, NO 4, October, 2011.

[38] Lie Xu, Liangzhong Yao and Masoud Basargan, “DC grid Management of a Multi-terminal
HVDC Transmission System for Large Offshore Wind Farms,” in Sustainable Power Generation
and Supply, 2009. SUPERGEN '09. International Conference on .

[39] Pinto, R.T.; Rodrigues, S.F.; Bauer, P.; Pierik, J., "Comparison of direct voltage control methods
of multi-terminal DC (MTDC) networks through modular dynamic models," in Power
Electronics and Applications (EPE 2011), Proceedings of the 2011-14th European Conference
on , 2011.

[40] S. K. Chaudhary, R.Teodorescu, P. Rodriguez and P.C. Kjar, "Chopper Controlled Resistors in
VSC-HVDC Transmission for WPP with Full-scale converters," in Sustainable Alternative
Energy (SAE), 2009 IEEE PES/IAS Conference, pp.1-8, 28-30 Sept. 2009.

[41] Cole, S.; Beerten, J.; Belmans, R, "Generalized Dynamic VSC MTDC Model for Power System
Stability Studies".Power Systems, IEEE Transactions on , vol.25, no.3, pp.1655,1662, Aug. 2010.

[42] F. Shewarega, I. Erlich, J. L. Rueda, "Impact of large offshore wind farms on power system
transient stability," in Power Systems Conference and Exposition, 2009. PSCE '09. IEEE/PES ,

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
vol., no., pp.1,8, 15-18 March 2009.

[43] N. R. Chaudhuri, R. Majumder, B. Chaudhuri, "System Frequency Support Through Multi-


Terminal DC (MTDC) Grids".Power Systems, IEEE Transactions on , vol.28, no.1, pp.347,356,
Feb. 2013.

[44] D. V. H. a. R. B. Jef Berteen, "VSC MTDC Systems with a Distributed DC Voltage Conrol - A
power Flow Approach," in PowerTech, 2011 IEEE Trondheim , vol., no., pp.1-6, 19-23 June
2011.

[45] Mario Ndreko, Madeleine Gibescu, "Offshore wind power plant - Single country connection,"
FLOW project report, Deliverable 2A, Delft University of Technology, 2013.

[46] Y. Pipelzadeh, B. Chaudhuri, T. C. Green, "Inertial response from remote offshore wind farms
connected through VSC-HVDC links: A Communication-less scheme," in PES GM 2012 IEEE ,
vol., no., pp.1,6, 22-26 July 2012.

[47] Silva, B.; Moreira, C.L.; Seca, L.; Phulpin, Y.; Peas Lopes, J.A., "Provision of Inertial and
Primary Frequency Control Services Using Offshore Multiterminal HVDC
Networks".Sustainable Energy, IEEE Transactions on , vol.3, no.4, pp.800,808, Oct. 2012.

[48] I. Erlich, M. Wilch, "Primary frequency control by wind turbines," in ower and Energy Society
General Meeting, 2010 IEEE , vol., no., pp.1,8, 25-29 July 2010.

[49] Yi Wang, Xiaorong Zhu and Heming Li, "Contribution of VSC-HVDC Connected Wind Farms
to Grid Frequency Regualtion and Power Damping".

[50] ABB, April 2012. [Online]. Available: http://www.abb.com/industries/us/9AAC30300394.aspx.

[51] C.D. Barker and R.Whitehouse, "Autonomous Converter control in a Multi-terminal HVDC
System," in AC and DC Power Transmission, 2010. ACDC. 9th IET International Conference on
, vol., no., pp.1-5, 19-21 Oct. 2010.

[52] M.Ndreko, A.van der Meer, M. Gibescu, B. Rawn and M.A.M.M van der Meijden, "Damping
Power System Oscillations by VSC-Based HVDC Networks: A North Sea Grid Case Study," in
Wind Energy Integration in Large Power Systems workshop, London, Oct. 2013, 2013.

[53] ENTSOe, "Offshore Transmission Technology," 2011. [Online]. Available:


https://www.entsoe.eu/resources/publications/.

[54] Anderson, Paul M. and Fouad, A. A, Power System Control and Stability, Canada: IEEE press
power engineering series, 2002.

[55] n ppel, T.; Nielsen, J.N.; Jensen, K.H.; Dixon, A.; Ostergaard, J.; , "Small-signal stability of
wind power system with full-load converter interfaced wind turbines," Renewable Power
Generation, IET , pp. vol.6, no.2, pp.79-91, March 2012.

[56] T. T. GmbH, "Grid Code - extra high voltage," October 2010.

[57] ENTSO-E, "Requirements for Grid Connection Applicable to all Generators," June 2012.

[58] T. M. Haileselassie, PhD Thesis: Control, Dynamics and Operation of Multi-terminal VSC-

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
HVDC Transmission systems, : Norwegian University of Science and Technology (NTNU),
2012.

[59] Vladimir Blasko and Vikram Kaura, "Anew Mathematical Model and Control of a Three-Phase
AC-DC Voltage Source Converter," IEEE Transactions on power electronics, VOL. 12, NO. 1,
Jan, 1997.

[60] J. G. Slootweg, PhD thesis: Wind Power Modeling and Impact on Power System Dynamics,
Delft: TU Delft, 2003.

[61] Olympo Anaya-Lara, F. Michael Hughes, Nicolas Jenkins and Goran Strbac, "Influence of Wind
Farms on Power System Dynamics and Tranient Stability," Wind Engineering , pp. Vol. 30,
Mo.2, pp 107-127, 2006.

[62] A. A van der Meer, R. L. Hendriks,W. L. Kling, "Combined stability and electro-magnetic
transients simulation of offshore wind power connected through multi-terminal VSC-HVDC,"
Power and Energy Society General Meeting, vol., no., pp.1,7, 25-29 July 2010 2010.

[63] T. W. Shire, "VSC-HVDC based Network Reinforcement," May 2009.

[64] M. Ndreko, A. A. van der Meer, J. Bos, M. Gibescu, K. Jansen and M.A.A.M van der Meijden,
"Transient Stability Analysis of an Onshore Power System With Multi-Terminal Offshore VSC-
HVDC: A Case study for The Netherlands," PES General Meeting, p. Accepted for publication,
July 2013.

[65] Jef Berteen, Stijn Cole and Ronnie Belmans, "Generalized Steady-State VSC MTDC Model for
Sequential AC/DC Power Flow Algorithms," in IEEE Transactions on Power Systems, vol.27,
no.2, pp.821,829, May 2012.

[66] Paulo Chainho, Arjen A. Van der Meer, Madeleine Gibescu and Mart A.M.M. Van der Meijden,
"General Modeling of Multi-Terminal VSC-HVDC Systems for Transient Stability Studies".

[67] Jun Linag, Oriol Gomis-Bellmunk, Janaka Ekanayake and Nicholas Jenkins, "Control of mulit-
terminal VSC-HVDC transmission for offshore wind power".

[68] Hong-seok Song and Kwanghee Nam., "Dual current control scheme for PWM converter under
unbalanced input voltage conditions," Industrial Electronics, IEEE Transactions on 46.5 (1999):
953-959..

[69] Zheng Chao, Zhou Xiaoxin and Li Ruomei, "Dynamic Modeling and Transient Simulation for
VSCbased HVDC in Multi-Machine System," in Power System Technology, 2006. PowerCon
2006. International Conference on.

[70] Saeedifard, Maryam, and Reza Iravani., "Dynamic performance of a modular multilevel back-to-
back HVDC system.," Power Delivery, IEEE Transactions on 25, no. 4 (2010): 2903-2912..

[71] Nilanjan Ray Chaudhuri, Rajat Majumber, Balarko Chaundhuri, Jiuling Pan and Reynaldo
Nuqui, "Modeling and Stability Analysis of MTDC Grids for offshore Wnd Farms: A case study
on the North Sea Benchmark System," in Power and Energy Society General Meeting, 2011
IEEE.

[72] Xu, Lie, Bjarne R. Andersen, and Phillip Cartwright., "VSC transmission operating under

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
unbalanced AC conditions-analysis and control design," Power Delivery, IEEE Transactions on
20.1 (2005): 427-434..

[73] Erlich, István, Tobias Neumann, Fekadu Shewarega, Peter Schegner, and Jorg Meyer., "Wind
turbine negative sequence current control and its effect on power system protection.," In Power
and Energy Society General Meeting (PES), 2013 IEEE, pp. 1-5. IEEE, 2013..

[74] Georgios Tsourakis, Sotirios Nanou and Costas Vournas, "A power system stabilizer for variable
speed Wind generators," in 18th IFAC World Congress, Milano, 2011.

[75] Tatsuhito Nakajama and Shoichi Irokawa, "A control system for HVDC Trnasmission by
Voltage Sourced Converters".Power Engineering Society Summer Meeting, 1999. IEEE , vol.2,
no., pp.1113-1119 vol.2, 1999.

[76] Yan Liu, Zhe Chen, "A Flexible Power Control Method of VSC-HVDC Link for the
Enhancement of Effective Short-Circuit Ratio in a Hybrid Multi-Infeed HVDC System".Power
Systems, IEEE Transactions on , vol.28, no.2, pp.1568,1581, May 2013.

[77] Moawwad, A.; El Moursi, M.S.; Weidong Xiao, "A Novel Transient Control Strategy for VSC-
HVDC Connecting Offshore Wind Power Plant," Sustainable Energy, IEEE Transactions on ,
vol.5, no.4, pp.1056,1069,, Oct.2014.

[78] Moawwad, Ahmed, El Moursi, Mohamed Shawky, and Weidong Xiao., "A novel transient
control strategy for VSC-HVDC connecting offshore wind power plant.," Sustainable Energy,
IEEE Transactions on 5, no. 4 (2014): 1056-1069..

[79] CIGRE, "Analysis and Damping Oscillations in the UCTE/CENTREL Power System," 2000.

[80] Joshua K. Wang, Robert M. Gardner and Yilu Liu, "Analysis of System Oscillations using wide
Area measurements".

[81] J. Dai, Y. Phulpin, A. Sarlette, D. Ernst, "Coordinated primary frequency control among non-
synchronous systems connected by a multi-terminal high-voltagae direct current grid".IET
Generation, Transmission & Distribution, 2012, vol.6, Iss.2, pp.99-108.

[82] Rusejla Sdaikovic, "Damping controller design for power system oscillations," ABB.

[83] R. Preece, A. M. Almutairi, O. Marjanovic and J. V. Milanovic, "Damping of Electromechanical


Oscillations by VSC-HVDC Active power Modulation with Supplementary WAMS Based
Modal LQG Controller," in Power and Energy Society General Meeting, 2011 IEEE , vol., no.,
pp.1-7, 24-29 July 2011.

[84] Margaris, I.D.; Papathanassiou, S.A.; Hatziargyriou, N.D.; Hansen, A.-D.; Sorensen, P.,
"Frequency Control in Autonomous Power Systems With High Wind Power
Penetration".Sustainable Energy, IEEE Transactions on , vol.3, no.2, pp.189,199, April 2012.

[85] Ruttledge, L.; Miller, N.W.; O'Sullivan, J.; Flynn, D, "Frequency Response of Power Systems
With Variable Speed Wind Turbines".Sustainable Energy, IEEE Transactions on , vol.3, no.4,
pp.683,691, Oct. 2012.

[86] Mario Ndreko, Marjan Popov, Jose L. Rueda-Torres, Mart A. M. M. van der Meijden, "Impact of
Offshore Wind and Conventional Generation Outages on the Dynamic Performance of AC-DC

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Transmission Systems," in IEEE PowerTech 2015, Eindhoven, Netherlands, 2015.

[87] C. M. Diaz Cano, "Impact Of Offshore Wind Power on the Stability of the Dutch Electrical
Grid," MSc Thesis Report - TU Delft, 2010.

[88] G. J. Lim T. T. Lie, G. B. Shrestha and K. L. Lo, "Impementation of coordinated multiple facts
controller for damping pscillations," Elevier- Electrical Power Systems, 2000.

[89] H. Latorre and M. Ghandhari, "Improvement of Power System Stability by Using VSC- HVDC,"
Electrical Power & Energy Systems, vol. 33, NO.2, pp. 332-339, Feb, 2011.

[90] J.T.G Pierik, "NSTG Wind farm locations and development," ECN, 2009.

[91] Ooos, Etienne Veilleux and Boon-Teck, "Power Flow Analysis in Multi-terminal HVDC Grid,"
in Power Systems Conference and Exposition (PSCE), 2011 IEEE/PES .

[92] Pinaki Mitra and Lindong Zhang, "Real-Time Simulation of Wind Connected HVDC grid," 4-5
December 2012.

[93] Oriol Gomis Bellmunt, Jun Liang, Janaka Ekanaka, Rosemary King and Nicholas Jenkins,
"Topologies of multi-teminal VSC-HVDC transmission for large offshore wind farms," Electric
Power Systems Research - Science Direct, 2011.

[94] Lie Xu, Liandzhong Yao and Yi Wang, "The Role of Multiterminal HVDC for Wind Power
Transmission and AC Network Support," in Power and Energy Engineering Conference
(APPEEC), 2010 Asia-Pacific.

[95] H. W. a. F. K. Christian Feltes, "Fault Ride-Through of DFIG-based Wind Farms connected to


the Grid through VSC-based HVDC Link".

[96] S. M. F. Rodrigues, "Dynamic Modeling and Control of VSC-based Multiterminal DC


networks," Master Thesis.

[97] W. Wang, M. Barnes, O. Marjanovic, "Droop control modelling and analysis of multi-terminal
VSC-HVDC for offshore wind farms".

[98] Chaudhuri, Nilanjan Ray; Domahidi, Alexander; Chaudhuri, Balarko; Majumder, Rajat; Korba,
Petr; Ray, Swakshar; Uhlen, jetil; , “Power oscillation damping control using wide-area
signals: A case study on Nordic equivalent system,” Transmission and Distribution Conference
and Exposition, IEEE PES, pp. vol., no., pp.1-8, 19-22 , April 2010.

[99] M.Ndreko, A.van der Meer, M. Gibescu, B. Rawn and M.A.M.M van der Meijden, “Damping
Power System Oscillations by VSC-Based HVDC Networks: A North Sea Grid Case Study,” in
Wind Energy Integration workshop, London, Oct. 2013, 2013.

[100] R. Teixeira Pinto, S. F Rodrigues, P. Bauer, J. Pierik, "Grid code compliance of VSC-HVDC in
offshore multi-terminal DC networks," in Industrial Electronics Society, IECON 2013 - 39th
Annual Conference of the IEEE , vol., no., pp.2057,2062, 10-13 Nov. 2013.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
This report is restricted to be used only within TU Delft for the
course EE4545 Power Systems of the Future
Appendix A – Benchmark Power System Model Parameters

Table B1. Impedances of the benchmark model calculated in 1000MVA base [2]
From Bus To Bus R (pu) X(pu) B(pu) B/2

1 8 0 0,02770083 0 0
2 8 0 0,22160665 0 0
3 4 0,0027701 0,03400277 0,1455162 0,072758
3 8 0,0826177 0,79390582 0 0
4 7 0,0006233 0,00886427 0,1302733 0,065137
4 8 0,0448061 0,44577562 0,171823 0,085912
5 6 0,0117729 0,18123269 0,2131272 0,106564
5 7 0,0162742 0,20914127 0,21834 0,10917
6 7 0,0026316 0,03379501 0,2169985 0,108499

1 2 0,0495868 0,12002066 0,1367784 0,068389

Table B2.HVAC and HVDC offshore cable parameters for given rated voltages at figure 25 & 31
CABLES CABLES
HVAC HVDC
R (Ohms/km) 0,048 R (Ohms/km) 0.03
L(mH/km) 0,37 L(mH/Km) 0.2
C (nF/km) 180 C (nF/km) 220
Cdc_VSC
I rated (A) 1055 (uF) 50.0
T max [oC] 90

km 50 km 50

Table B3. 6th order synchronous generators dynamic [2]

G1 G2 G3 G4 G5 G6 G7
T'do (> 0) 0,692 1,008 0,933 0,721 1,002 1,004 0,69
T''do (> 0) 0,023 0,023 0,036 0,019 0,032 0,065 0,03
T'qo (> 0) 0,692 1,008 0,933 0,721 1,002 1,004 0,69
T''qo (> 0) 0,064 0,064 0,036 0,02 0,047 0,071 0,03
Inertia H 4,237 4,464 4,358 5,562 5,474 4,879 4,07
Speed Damping D 0 0 0 0 0 0 0
Xd 2,36 2,116 2,001 2,47 2,235 2,158 2,51
Xq 2,261 2,043 1,937 2,301 2,113 2,026 2,45
X'd 0,297 0,312 0,293 0,278 0,269 0,3 0,3
X'q 0,297 0,312 0,293 0,278 0,609 0,3 0,3
X''d = X''q 0,209 0,25 0,219 0,203 0,209 0,227 0,23
Xl 0,167 0,187 0,167 0,169 0,167 0,237 0,17
S(1.0) 0,03 0,03 0,03 0,03 0,03 0,03 0,03
S(1.2) 0,4 0,4 0,4 0,4 0,4 0,4 0,4

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Table B4.Parameters for standard IEEE excitation system and governor models used in the benchmark model

SEXS TGOV1
TA/TB 0,1 R 0,05
TB (> 0) 10 T1 (>0)(sec) 0,5
K 20 V MAX 1
TE 0,1 V MIN 0
EMIN 0 T2 (sec) 1,5
EMAX 3 T3 (>0)(sec) 5
Dt 0

Figure B1. Block diagram of the TGOV1 governor model as used in the Benchmark power system model in
PSS®E

Figure B2.Block diagram of the TGOV1 governor models as used in the Benchmark power system model in
PSS®E

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Figure B3.Active power and reactive power of the synchronous generators in the benchmark model

6000
Active Power

MW /MVar/MVA
5000
Sb
4000
3000 Reactive Power

2000
1000
0
1 2 3 4 5 6 7
-1000
Gi, at bus 100i, i=1,...,7

Figure B4.Active power and reactive power of the loads

2500 Pload
Qload
MW /MVar

2000

1500

1000

500

0
1 2 3 4 5 6 7 8
Bus 100i, i, i=1,...,8

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
This report is restricted to be used only within TU Delft for the
course EE4545 Power Systems of the Future
Appendix B - PSS/E type 3 (DFIG) variable speed wind turbine sub-
models [Source: PSS/E User Manual]

Figure E1. Generator Electrical Model.

cos   sin  
T 
 sin  cos  

Figure E2. Reactive power control model

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Figure E3. Active Power (Torque ) Control model

Figure E4. Linear aerodynamic model

Figure E5 Wind turbine two-mass model

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Figure E6 Pitch control model

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
Appendix C – Detailed differential equations of the VSC model
In this section a detailed description of the mathematical model of the VSC will be given. As
it has already been mentioned, the most important element of the VSC is the phase reactor. In
order to start with the mathematical model of the phase reactor, let us first consider the very
simple equivalent model. As it has already been discussed, in the time-averaged modelling
approach, the VSC is modelled from the AC side as control voltage source. Thus, u c,abc
represents the instantaneous three phase voltages of the VSC behind the phase reactor.
Similarly, us,abc refers to the three phase instantaneous value of the grid voltage at the PCC
[9].

r L iabc

AC
uc ,abc AC us ,abc

Figure A1 . Phase reactor equivalent circuit

In this very simple model, we can write the Kirchhoff law in the stationary abc reference
frame with the instantaneous values of voltages and currents we have:

diabc
uc ,abc  us ,abc  riabc  L
dt
By making use of the Clark transformation we can write the same differential equation in the
fictitious αβ stationary reference frame. The αβ reference phrase is an equivalent two phase
system, where the three phase AC quantities are represented by equivalent set of two-phase
quantities which produce the same an identical space vector (called virtual flux).

Historically, this two phase reference frame is created to represent an ideal two phase
machine, where the two axis are defined such as the currents in the fictitious coils, located at
each axis, would produce the same mmf. This transformation can be achieved by multiplying
the vector of the instantaneous three-phase voltages with the given matrix:

 1 1 
1  
2 2 
u    ua 
u   3 3 
    0 2
 ub
2  
(A.1)
uo    u 
1 1 1  c
2 2 2 

For a symmetrical three phase voltages, as in our case study, uo is equal to zero. Now, in the
αβ reference frame, it is possible to write the differential equation which describes the
dynamic response of the phase reactor such as:

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
diabc di
uc ,abc  us ,abc  riabc  L Clark
Transformation
 uc ,  u s ,  ri  L (A.2)
dt dt

 uabc

b u
u

t
a

Figure A2. Stationary abc reference frame with respect to the stationary αβ reference frame

q Uc
uc s
d

ucd
Us
ucq id

s t

uc
I pr
iq

Figure A3. The dq reference frame with respect to the αβ reference frame

The common approach which is followed for the control of the VSC is the d-q vector control
scheme. This d-q control method uses a synchronously rotating reference frame which its d-
axis is aligned with vector of the voltage at the point of common coupling, as shown in figure

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
A3. Any vector which is written in the αβ stationary reference frame can further be expressed
in any other two-phase reference frame (such as the d-q frame), as shown in figure A3.
The transformation between the d-q and αβ reference frame is given such as:

us ,  e jt us ,dq


uc ,  e jt uc ,dq (A.3)
jt
i  e idq

Given the equations A3, it is not difficult from A2 to write the differential equations in the dq
reference frame as:
di jt jt jt
d  e jt idq 
uc ,  us ,  ri  L  e uc ,dq  e us ,dq  re idq  L
dt dt
d  idq 
 e jt uc ,dq  e jt us ,dq  re jt idq  j Le jt idq  e jt L
dt
d  idq 
 uc ,dq  us ,dq  ridq  j Lidq  L (A.4)
dt
d  idq 
 (uc ,d  juc ,q )  (us , d  jus , q )  r (id  jiq )  j L(id  jiq )  L
dt
d  id  jiq 
 (uc ,d  juc ,q )  (us , d  jus , q )  r (id  jiq )   L( jid  iq )  L
dt

Thus, a set of two differential equations is finally achieved each for one axis of the d-q
reference frame.

 d  id 
uc ,d  us ,d  rid   Liq  L uc ,d  us ,d  rid   Liq  sLid
dt Ls
   (A.5)
 u  u  ri   Li  L d  id  uc ,q  us ,q  riq   Lid  sLid
id (0) id (0)  0

 c ,q s ,q q d
dt

If we now consider the Laplace transformation of these two ordinary differential equations we
can obtain the block diagram of the phase reactor in the d-q reference frame.

uc ,d  us ,d  rid   Liq  sLid uc ,d  us ,d   Liq   r  sL  id



uc ,q  us ,q  riq   Liq  sLid uc ,q  us ,q   Lid   r  sL  iq
uc ,d  us ,d   Liq
id  (A.6)
 r  sL 

uc ,q  us ,q   Lid
iq 
 r  sL 

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future
usd Phase
Reactor
ucd +
-
1 id
Σ
𝑟 + 𝐿𝑠
+

 Liq
 Lid

ucq -
iq
+ 1
Σ 𝑟 + 𝐿𝑠
-

usq

Figure A4 . Block diagram of the phase reactor, open loop system of the VSC model

Up to this point, it has been shown that the model of the VSC in the AC side is simply the
model of the phase reactor. It becomes clear that given the system voltage, it is possible to
control the current flowing through the phase reactor. Indeed this is true but such an approach
is never followed in practice. The open loop system illustrates high oscillatory behavior.
Thus, in order to achieve a more robust operation, with fast rise-up/down time a current
feedback loop is introduced.

This report is restricted to be used only within TU Delft for the


course EE4545 Power Systems of the Future

You might also like