Sie sind auf Seite 1von 9

2046 IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO.

5, OCTOBER 2014

Dynamic Average-Value Modeling of


CIGRE HVDC Benchmark System
IEEE Task Force on Dynamic Average Modeling

H. Atighechi, S. Chiniforoosh, J. Jatskevich, TF Chair, A. Davoudi, J. A. Martinez, M. O. Faruque, V. Sood,


M. Saeedifard, J. M. Cano, J. Mahseredjian, D. C. Aliprantis, and K. Strunz

Abstract—High-voltage direct-current (HVDC) systems play an developed dynamic average models are shown to have computa-
important role in modern energy grids, whereas efficient and ac- tional advantages.
curate models are often needed for system-level studies. Due to the
Index Terms—Average-value modeling, CIGRE HVDC bench-
inherent switching in HVDC converters, the detailed switch-level
mark, electromagnetic transients, inverter, line-commutated con-
models are computationally expensive for the simulation of
verter, rectifier.
large-signal transients and hard to linearize for small-signal
frequency-domain characterization. In this paper, a dynamic
average-value model (AVM) of the first CIGRE HVDC bench-
mark system is developed in a state-variable-based simulator, I. INTRODUCTION
such as Matlab/Simulink, and nodal-analysis-based electromag-

H
netic transient program (EMTP), such as PSCAD/EMTDC. The VDC SYSTEMS are frequently employed for long-dis-
12-pulse converters in the HVDC system are modeled with a set
tance transmission due to their lower electrical losses.
of nonlinear algebraic functions that are extracted numerically.
The results from the average-value models are compared with For shorter distances, the higher costs of the power-electronic
the results of the detailed simulation to verify the accuracy of the equipment may still be justified due to other benefits achieved
AVMs in predicting the large-signal time-domain transients. The by HVDC, such as improved system stability and interconnec-
tion between unsynchronized ac systems.
Design and analysis of power-electronic-based systems, such
Manuscript received August 20, 2012; revised July 31, 2013 and January 13, as HVDC transmission, rely extensively on modeling and com-
2014; accepted July 16, 2014. Date of publication August 11, 2014; date of
puter simulation. The procedures for design and analysis of such
current version September 19, 2014. Paper no. TPWRD-00872-2012.
Task Force on Dynamic Average Modeling is with the Working Group on complex systems typically involve a large number of time-do-
Modeling and Analysis of System Transients Using Digital Programs, General main transient studies as well as analysis in frequency domain
Systems Subcommittee, T&D Committee, IEEE Power & Energy Society.
for control purposes.
Task Force Members: S. A. Abdulsalam, D. C. Aliprantis, U. Annakkage,
H. Atighechi, J. Belanger, J. M. Cano, S. Chiniforoosh, A. Davoudi, The detailed modeling of HVDC transmission system,
V. Dinavahi, O. Faruque, S. Filizadeh, D. Goldsworthy, A. Gole, R. Iravani, wherein the switching of all devices is represented in full
J. Jatskevich (Chair), R. Jayasinghe, H. Karimi, M. Kuschke, A. St. Leger,
detail, may be readily accomplished in the state-variable-based
J. Mahseredjian, J. A. Martinez, N. Nair, L. Naredo, T. Noda, J. N. Paquih,
J. Peralta, A. Ramirez, A. Rezaei-Zare, M. Rioual, M. Saeedifard, K. Schoder, (SV-based) [1]–[7] and Electromagnetic Transient (EMT)
V. Sood, K. Strunz, A. VanDerMeer, X. Wang, A. Yazdani. programs (EMTP type) [8]–[12]. Although the efficiency of
H. Atighechi, S. Chiniforoosh, and J. Jatskevich are with the Electrical and
detailed models can be increased (e.g., [13]), the detailed
Computer Engineering Department, University of British Columbia, Vancouver,
BC V6T 1Z4 Canada. models are generally computationally expensive, and require
A. Davoudi is with the Electrical and Computer Engineering Department, significant simulation time especially for system-level studies.
University of Texas, Arlington, TX 76011 USA.
Moreover, due to switching, these models are discontinuous
J. A. Martinez is with the Department Eng. Electrica-ETSEIB, Universitat
Politecnica de Catalunya, Barcelona 08028, Spain. and not suitable for small-signal frequency-domain analysis.
M. O. Faruque is with the Center for Advanced Power Systems, Florida State It is therefore desirable to develop equivalent models that do
University, Tallahassee, FL 32310 USA.
not include switching and may be used to predict the system’s
V. Sood is with the Faculty of Engineering and Applied Science, University
of Ontario Institute of Technology, Oshawa, ON L1H 7K4 Canada. slower transient behavior and steady-state characteristics.
M. Saeedifard is with the Electrical and Computer Engineering Department, Developing steady-state and dynamic equivalent models
Purdue University, West Lafayette, IN 47907 USA.
for HVDC transmission systems has been of great interest in
J. M. Cano is with the Universidad de Oviedo, Ingeniería Eléctrica, Gijón
33204, Spain. the power system research community. Steady-state models
J. Mahseredjian is with the Ecole Polytechnique, Montreal, QC H3C 3A7 for HVDC systems were first proposed in [14] and further
Canada.
improved in [15] and [16]. The effects of harmonics and inter-
D. C. Aliprantis is with the Electrical and Computer Engineering Department,
Iowa State University, Ames, IA 50011 USA. harmonics in steady-state HVDC models have been discussed
K. Strunz is with the Technische Universität Berlin, Elektrotechnik und In- in the literature [17]. The impedance mapping and equivalent
formatik, Berlin 10587, Germany.
circuit of HVDC systems have been considered in [18]. More
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. recent improvements have been made in [19] by considering
Digital Object Identifier 10.1109/TPWRD.2014.2340870 the commutation subinterval. The conventional steady-state

0885-8977 © 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
ATIGHECHI et al.: DYNAMIC AVERAGE-VALUE MODELING OF CIGRE HVDC BENCHMARK SYSTEM 2047

Fig. 1. CIGRE HVDC benchmark system circuit diagram.

models discussed in [14]–[19] are straightforward to imple- developed AVMs are verified against a detailed implementa-
ment. A simplified model [20] considers the dc- and ac-side tion of the system by means of simulation studies. Due to lim-
dynamics with some approximations. A dynamic model for ited space, this paper focuses on the time-domain transients,
the rectifier side of the hybrid HVDC system can be found in whereas it is envisioned that a follow-up companion paper will
[21], where the inverter side is modeled as a voltage or current present detailed guidelines and examples of using the developed
source. In order to predict the HVDC system’s stability, the AVM for the small-signal and frequency-domain analysis of the
small-signal model has been developed in [22]–[26]. Such lin- CIGRE benchmark HVDC system.
earized models are advantageous for controller design, analysis
of interaction among the subsystems, subsynchronous torsional
II. SYSTEM DETAILED MODEL
interaction phenomenon, etc. [25]. These small-signal models
are typically valid only within a small range about an operating The CIGRE HVDC benchmark system, first proposed in
point for which they were established. [29], is a monopolar 500 kV, 1000 MW dc link which em-
The dynamic average-value models (AVMs) have evolved ploys 12-pulse converters on the rectifier and inverter sides.
in order to overcome the challenges due to switching phe- The detailed modeling of this system in PSCAD/EMTDC
nomena in power-electronic-based systems. In such models, and Matlab/Simulink has been presented in [34]. The circuit
the effects of switching are neglected or “averaged” over diagram of this system is illustrated in Fig. 1, and the cor-
a prototypical switching interval. The resulting AVMs are responding parameters are summarized in Appendix B. The
continuous and can potentially run orders of magnitude faster system is composed of two ac sides, each represented by the
than the original detailed models. Such models are therefore equivalent supplying/receiving network. The rectifier and in-
very useful for studying the system-level transients, wherein verter ac networks are represented by R-L circuits to represent
the details of switching can be omitted. However, since the weak grids on each side. In order to absorb the harmonics
switching details are removed (averaged) in these models, such generated by the converter as well as providing the converter
models cannot be used to study the commutation failures or with reactive power, multiharmonic ac filters are placed on both
any similar phenomena that are related to the converter valves’ ac sides. The dc subsystem is represented by the T-equivalent
switching instances. When implemented in SV-based pro- circuit of the transmission line combined with the smoothing
grams, the AVMs may be numerically linearized for subsequent reactors that are connected at both sides in order to reduce the
small-signal analysis. ripple in dc current.
A summary of promising approaches and their application to The 12-pulse converters are constructed by a series connec-
modeling and analysis of power systems transients can be found tion of two 6-pulse converters. At each ac side, (rectifier and
in recent reports by the IEEE Task Force on Dynamic Average inverter), the 3-phase transformers with Y- and -connected
Modeling [27] and [28]. In order to provide a common refer- secondary windings are employed to produce the desired 30
ence system for HVDC transmission studies, the CIGRE bench- shifted voltages (equivalent to 6-phase) that feed the two 6-pulse
mark HVDC system was proposed in 1985 [29] and has been converters, respectively, thus resulting in a 12-pulse operation.
extensively used in many studies [30], [31]. This benchmark As depicted in Fig. 1, each converter is controlled through the
system is considered here to demonstrate the dynamic average respective control module.
modeling methodology and its advantages. The average-value In the considered CIGRE HVDC benchmark system, the ac
models for the rectifier and inverter sides are set forth and de- network is weak (short-circuit ratio (SCR) of 2.5 at a rated fre-
veloped based on the parametric approach [32], [33]. To demon- quency of 50 Hz) [34], which also presents some challenges for
strate the effectiveness of the methodology in SV-based and the controls. It should be noted that the (conventional) HVDC
EMTP-type software packages, the AVMs are implemented and classic system is normally limited to networks with SCRs of at
demonstrated in Matlab/Simulink and PSCAD/EMTDC. The least around 2 [35]. This limitation comes from the voltage and
2048 IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 5, OCTOBER 2014

lower limit for this angle is determined based on the protection


constraints to allow enough time for the valves to turn off be-
fore the end of a half-cycle. The -control (PI-3) adjusts the
value of extinction angle. However, the inverter control may
be switched from the -control strategy to the current-control
mode that limits the current to some specified reference value.
The firing angle for the inverter is then established based on the
output of these control modules, and the respective gate signals
are constructed by the pulse generators depicted in Fig. 1.

III. DYNAMIC AVM


In this paper, the AVM is developed in both Matlab/Simulink
and PSCAD/EMTDC. In order to model a complex system,
such as the CIGRE HVDC benchmark, it is necessary to divide
it up into smaller parts. Hence, the main circuit of HVDC
system is divided into seven subsystems: two line-commutated
converters (rectifier and inverter sides), which are modeled
using the parametric AVM approach; two ac subsystems; two
controller subsystems; and the dc transmission-line subsystem.
For convenience, each subsystem is implemented as a separate
module. These modules are then interconnected to form the
overall HVDC system.
The parametric AVM approach is generally independent of
the simulator platform. Dynamic AVM assumes that all ac and
Fig. 2. CIGRE HVDC benchmark control subsystem: (a) Rectifier control and
dc variables can be replaced by their fast averages over a proto-
(b) inverter control. typical switching interval as follows:

(1)
power stability issues between the converters and the ac net-
works [35]–[37]. To reduce this effect, sometimes a static or
synchronous compensator might be added to the converter ter- where may represent voltage or current. Due to the
minals, or the capacitor-commutated converter (developed by 12-pulse operation, , where represents the
ABB [35]) may also be considered. However, further details on period of the ac waveform as depicted in Fig. 3. This definition
this issue are out of the scope for this paper. of fast average directly applies to the dc-side variables, but in
For the purpose of this paper, in Fig. 1, the power is assumed order to obtain the fast average of the ac variables, they must be
to flow from left to right, which determines the rectifier and in- transferred into a synchronous rotating reference frame. The
verter sides, respectively. But otherwise, the controllers can re- switching interval and primary voltages and currents of
verse the power flow by appropriately changing the firing an- the transformer in the rectifier side are illustrated in Fig. 3. As
gles on each side and changing the ac voltage magnitudes of seen in Fig. 3, the transformed ac voltages and currents
the two sources. The details of the rectifier and inverter con- appear as dc in the synchronous reference frame and
trollers are depicted in Fig. 2, and the corresponding param- have ripple with a period of .
eters are summarized in Appendix C. For the rectifier side, a
proportional-plus-integral (PI) current control (PI-1) is realized A. Rectifier Dynamic AVM
with the reference denoted by . This reference cur- The development of an AVM requires relating the ac and
rent is provided by the inverter controller. At the inverter side, dc quantities of each rectifier (and inverter). The classical
the controller is composed of a voltage-dependent current con- approach is based on a number of simplifying assumptions
troller (PI-2) and the extinction angle -control, respectively. (ideal voltage sources, no commutating resistance, only com-
The reference value for the rectifier controller is mutating inductance, constant dc current, etc.) as presented in
determined by the inverter controller in order to ensure lim- [38, Ch. 11] and [39, Ch. 10]. In this section, the parametric
ited current during undervoltage conditions. Under normal op- AVM methodology introduced in [32] for a 6-pulse diode
eration, the externally provided reference value is sent rectifier (see Appendix A) is considered instead as a more
to the output. This value determines the power flow through general approach that does not relay on similar simplifications.
the HVDC system. Under fault conditions, the reference value This approach is further extended to the 12-pulse rectifier
is limited based on the dc-link voltage and cur- side of the CIGRE HVDC benchmark system. Herein, the
rent at the inverter side, as shown in Fig. 2. 12-pulse rectifier is composed of two 6-pulse line-commutated
During normal operation, the inverter is controlled using the converters. In the AVM developed here, each converter may
extinction angle , which determines the difference between be modeled separately to represent the corresponding dc- and
the turning-off angle of the valves and the angle of 180 . The ac-side variables. The ac-side converter voltages and currents
ATIGHECHI et al.: DYNAMIC AVERAGE-VALUE MODELING OF CIGRE HVDC BENCHMARK SYSTEM 2049

Fig. 4. Block diagram depicting implementation of the AVM rectifier side:


(a) original AVM with two rectifiers separately and (b) collapsed AVM.

Fig. 3. Typical voltages and currents of a 12-pulse rectifier transformer primary


winding.
voltages and currents are combined into dynamic impedances
defined as

for both 6-pulse rectifiers can be expressed in the converter


reference frame [38]. (5)
The parametric AVM assumes that the rectifier may be
viewed as a multiport switching cell, wherein the transformed
Assuming that AVM parameters for each rectifier are ex-
ac-side and dc-side variables can be related through some alge-
tracted and saved as lookup tables, the AVM of the rectifier side
braic functions. Herein, since there are two 6-pulse bridges, the
of the CIGRE HVDC benchmark system can be implemented
two sets of parametric functions relating the ac and dc voltages
according to the diagram depicted in Fig. 4. For example, if
are defined as
the two AVMs are constructed (one for each bridge), then
the system can be implemented according to Fig. 4(a), where
(2) each bridge is modeled as an algebraic module that relates
the corresponding dc- and ac-side variables transferred to a
Considering the series connection of the rectifier bridges, the ac- rotating reference frame. Alternatively, taking into account
and dc-side currents are also related as the symmetry between the bridges and their series connection,
the two submodels can be collapsed into one as depicted in
(3) Fig. 4(b). In particular, assuming that the dc voltages for both
rectifiers have the same average value, we have
where , , and are algebraic functions of loading con- (6)
ditions and the thyristor firing angle, and “conv1” and “conv2”
refer to the converter reference frame for each converter, respec- Therefore, (2) and (3) can be simplified as
tively. The power factor angles between vectors and
for each converter are defined as (7)
(8)
(4)
where and
. Following the same approach, the equivalent power factor
angle may be defined as
Here, the superscript “ ” refers to the sending-end equivalent
voltage.
Conditions (2) and (3) hold for any operating point, since (9)
functions , , and exist for the entire range of
operation. The considered approach relies on calculating func- The described parametric approach has an advantage that
tions , , and numerically by running the detailed (2)–(9) are all algebraic and explicit. Specifically, (2)–(5) are
simulation similar to conventional 6-pulse converters [32]. The written, considering both converter bridges separately, and
loading condition of each rectifier may be specified in terms of (6)–(9) combine the voltages and currents assuming series
terminal currents and voltages. For compactness, the terminal connection. Using (7)–(9), the collapsed AVM of the rectifier
2050 IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 5, OCTOBER 2014

side can be implemented as shown in Fig. 4(b). The equivalent


dynamic impedance is defined similarly, using appropriate
equivalent dc voltage and ac currents in (5). The firing angle
is also provided by the respective controller (see Fig. 2). It is
worth mentioning that the collapsed AVM of Fig. 4(b) provides
more simplicity in comparison to the general formulation of
(2)–(4) when the rectifiers are symmetrical with respect to each Fig. 5. Block diagram depicting implementation of the AVM inverter side.
other.
The pulse generator module in Fig. 1 produces the neces-
sary thyristor firing pulses which are generally determined using
the filtered converter ac voltages or their fundamental compo-
nents. For example, the generation of firing pulses may be im-
plemented based on the zero-crossing of the respective ac volt-
ages (fundamental components). Any such filters will generally
have some response delay which should be taken into account
when constructing the AVM where all voltages and currents are
continuous and appear as dc in an appropriate synchronous ref-
erence frame.

B. Inverter Dynamic AVM


The inverter side of the CIGRE HVDC benchmark can be Fig. 6. Block diagram depicting implementation and interfacing of the
inverter-side ac subsystem: (a) using the original circuit and (b) using the
modeled using an approach similar to the rectifier side. As de- transformed circuit.
picted in Fig. 1, the 12-pulse inverter is composed of two 6-pulse
bridges connected in series. To denote the variables related to
the inverter side (receiving end), the superscript “ ” will be other modules of the overall system. The interfacing problem
used. If the inverters’ AVMs are represented separately, one may is, in fact, similar to that observed with the models of rotating
consider two converter reference frames with the angular differ- electrical machines [39], where it is possible to have direct
ence of 30 . Assuming that there are no energy-storing elements and indirect interfacing. Based on the methodology described
in the inverter circuit, the algebraic relationships relating the dc in [39], the desired interfacing for the AVM of Fig. 5 can be
and ac variables similar to (2)–(4) can be readily established. achieved using snubber circuits (typically large resistors or
The corresponding implementation will be similar to Fig. 4(a), small capacitors). Equivalent interfacing may also be achieved
except reverse direction of the power and variables will be re- if the ac-side inductor branches are represented using the proper
quired. For the inverter side, the -control is also required to transfer function with a very fast pole [32].
calculate the extinction angle. In the detailed simulation, the ex-
tinction angle is calculated based on voltages and currents on
C. AC Subsystems
the valves of the bridges. However, in the AVM, the switching
variables are replaced by their respective fast averages. There- The ac subsystems at the rectifier and inverter sides are com-
fore, is extracted by running a detailed simulation similar to posed of equivalent ac networks, the ac filters, and the phase-
, , and , and is saved for future use. shifting transformers. For the purpose of dynamic average mod-
Similar to the rectifier side, due to the symmetry in the eling, the ac subsystem can be modeled using approaches as de-
system, the two AVMs of the inverters can be collapsed and picted in Fig. 6. In the first approach shown in Fig. 6(a), the
represented using simplified equivalent relationships similar to ac side is modeled as a network circuit in physical vari-
(6)–(9). Implementation of the equivalent (collapsed) inverter ables. In this case, interfacing the ac subsystem model with the
AVM is shown in Fig. 5. The firing angle is also provided AVM of the converter may require interfacing circuitry [28].
by the respective controller (see Fig. 2). For the inverter AVM This approach is suitable for either PSCAD/EMTDC or Sim-
shown in Fig. 5, as opposed to the rectifier AVM shown in PowerSystems in Matlab/Simulink. However, the ac subsystem
Fig. 4, the inverter ac currents and dc voltage are can also be implemented in transformed synchronous co-
assumed to be the outputs of the inverter average switching ordinates/variables as a -circuit shown in Fig. 7. If this is
cell, whereas the ac voltages and dc current are possible, then the interface of the converter AVM with the ac
the inputs, respectively. Here, the inverter operating condition network becomes simpler as depicted in Fig. 6(b), since no
is determined in terms of an equivalent dynamic admittance to transformations of the interfacing variables would be re-
defined as quired. Moreover, due to the symmetry of the transformer and
the collapsed AVM representation of the converter, the trans-
(10) former is represented here only in terms of its equivalent series
impedance as shown in Fig. 7. The ac subsystem of the rectifier
side is very similar and can be modeled using any of the ap-
Interfacing the AVM depicted in Fig. 5 may require special proaches depicted in Figs. 6 and 7, which is not shown due to
consideration to achieve the input–output compatibility with limited space.
ATIGHECHI et al.: DYNAMIC AVERAGE-VALUE MODELING OF CIGRE HVDC BENCHMARK SYSTEM 2051

E. Control Subsystem
Since dynamic average modeling is only changing the im-
plementation of the switching part of the system, the developed
AVM of the entire CIGRE HVDC benchmark should be valid
for a wide range of operating conditions as well as different con-
trollers. Thus, the control system depicted in Fig. 2 is considered
here for the AVM as well as for all subsequent studies for con-
sistency with the original detailed model of the system.

IV. COMPUTER STUDIES


In general, the detailed switching models may be developed
in many commonly available simulation packages, such as
MATLAB/SimPowerSystems and PSCAD/EMTDC. With
proper care to the implementation details, time-step and ac-
curacy settings, the transient responses obtained by different
packages should be sufficiently similar and even identical.
However, a good AVM should follow the transient response
of the reference detailed model (regardless of the package
in which it was developed) with acceptable accuracy. For
the purpose of this paper, the detailed model of the CIGRE
HVDC benchmark system has been implemented in SimPow-
erSystem toolbox (using discrete blocks for faster simulation)
in Matlab/Simulink. The AVM of the same system is also
implemented in PSCAD/EMTDC and Matlab/Simulink using
standard library blocks. In order to verify the accuracy of the
proposed AVMs, the results of the detailed model are compared
with the results of the developed AVMs in two transient studies.

A. Change in DC Reference Current


In the first study, a fast change in the dc current reference is
applied. First, it is assumed that the HVDC system is operating
in a steady state and under nominal conditions with the dc cur-
rent of 2000A. At , the command value of the dc current
is reduced to 1600A. After , the current ref-
erence is set back to 2000A. The transient responses produced
by the detailed switching model and the AVMs implemented in
Fig. 7. Inverter-side ac subsystem implementation using transformed
circuits. Matlab/Simulink and PSCAD/EMTDC are shown in Fig. 8. As
it can be seen in Fig. 8, the reduction in the reference current on
the rectifier side causes the current drop on both the rectifier and
D. DC Subsystem inverter sides, which also results in respective increase in the dc
voltage. The ac voltages on both sides are also depicted in Fig. 8
For the purpose of modeling in this paper, the dc line is rep- (last two subplots), wherein the corresponding changes are also
resented as an equivalent T circuit with two smoothing induc- visible but less pronounced due to the scale. As expected, the re-
tors (one on each side), and one equivalent capacitor, as de- sults of AVMs in Matlab/Simulink and PSCAD/EMTDC follow
picted in Fig. 1. If the dc subsystem is implemented as a circuit, the transient response of the detailed simulation with good ac-
it can be readily realized in most transient simulators, such as curacy but without the switching.
PSCAD/EMTDC and SimPowerSystems in Matlab/Simulink, As the switching is averaged over a prototypical switching in-
wherein appropriate interfacing with the rectifier and inverter terval, the AVMs can be run with larger time-step without losing
AVMs will be required [39]. In addition, the dc network could the accuracy of results and therefore execute faster than the cor-
be modeled in state-space form using proper transfer functions responding detailed model. In order to demonstrate this point
to represent the inductors and capacitors as explained in [32]. for the same study, both AVMs and detailed simulation are run
For example, for interfacing with the rectifier AVM, which re- with the maximum possible time-step to achieve acceptable ac-
quires the input dc voltage as an input, the smoothing in- curacy. The detailed simulation is run using a fixed-step discrete
ductor in the rectifier dc side can be represented using a proper solver as the model is developed using discrete blocks from the
transfer function as in [32]. Alternatively, a snubber circuit (typ- SimPowerSystem library. For this model, the time step of at
ically a large resistor or small capacitor) can be added to the dc least 10 was required to maintain the accuracy. The AVM
terminal of the rectifier [39]. implemented in PSCAD/EMTDC required a 110 time step.
2052 IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 5, OCTOBER 2014

TABLE I
COMPARISON OF SIMULATION SPEED FOR DIFFERENT MODELS FOR TIME
INTERVAL FROM TO

Fig. 8. Currents and voltages resulting from a change in dc reference current


as predicted by various models.

The developed AVM in MATLAB/Simulink can run using fixed


and variable step solvers. However, since the fixed step solvers
are explicit, they are typically not suitable for solving stiff sys-
tems (which is the case here). Therefore, when using the fixed
step solver, the time-step should be chosen very small, which
makes the simulation computationally less efficient. Therefore,
the AVM in Matlab/Simulink is run using a variable-step solver Fig. 9. Currents and voltages during a short circuit in dc side as predicted by
ODE23tb which is implicit and dynamically adjusts the time- various models.
step size. For this model, the maximum time step was set to 2
ms, and the average step size used by the solver was 1.6 ms. The
corresponding CPU time for different models is summarized in B. Short Circuit in DC Side
Table I. To remove the overhead due to initialization in each In the following study, a short circuit happens in the dc line.
program, the CPU time is measured from to . The HVDC system is again assumed to operate in a steady state
As it can be seen in Table I, the AVMs run much faster than the and under the nominal conditions with the dc current of 2000A.
detailed simulation. Then, at , a resistive short circuit (0.4 p.u.) is applied
ATIGHECHI et al.: DYNAMIC AVERAGE-VALUE MODELING OF CIGRE HVDC BENCHMARK SYSTEM 2053

TABLE II simulation in large-signal time-domain transients including


CIGRE HVDC SYSTEM PARAMETERS USED FOR SIMULATION STUDIES the dc reference current command change and a short circuit
implemented on the dc transmission line.
This paper will be useful to researchers and practicing engi-
neers in the area of electric power and energy systems who are
dealing with applications of power-electronic components and
modules as well as using modern digital simulation tools.

APPENDIX

A. Parametric Average-Value Modeling [32]


Assuming that there are no energy storing elements in con-
in the middle of the dc line. The short circuit is then cleared at verter circuit, the average values of the dc and ac side variables
and the system recovers following a transient. The for any operating point can be related through some algebraic
dc currents and voltages, and the phase voltage of the trans- functions as
formers primary sides as predicted by various models are shown
in Fig. 9. As can be seen from Fig. 9, during the short circuit, (A1)
the dc currents of the rectifier and inverter drop significantly due
and
to the controller action that limits the value of the fault currents
by reducing the dc reference current and also increasing and de-
creasing the applied firing angles of the rectifier and inverter
sides, respectively. During the short circuit, as the dc voltage (A2)
drops, the output current of the rectifier is limited to the min-
imum value due to the voltage-dependent current control on the The power factor angle between vectors and is
rectifier side. However, on the inverter side, the firing angle is defined as
reduced to the minimum value and therefore inverter current is
close to zero. The dc voltages at the inverter and rectifier ends of (A3)
the transmission line are also undergoing significant drop. After
the fault is cleared, the system recovers and following a tran-
where the superscript “ ” refers to the synchronous reference
sient response all voltages and currents return to their pre-fault
frame aligned with an equivalent source.
values as can be seen in Fig. 9. Throughout the whole transient
The rectifier operating point may be specified in terms of ter-
study, which is determined by the HVDC system dynamics and
minal voltages and currents that define the instantaneous dy-
the action of its both controllers, the results predicted by the
namic impedance of the converter switching cell as
AVM remain in very good agreement with respect to the refer-
ence solution provided by the detailed switching model.
(A4)
The dynamic AVMs, based on either the analytical or para-
metric approaches, are only capable of predicting the transients
in the fundamental component of the ac voltages and currents
due to the fundamental assumptions and approach in which they B. CIGRE HVDC Benchmark System Parameters
were derived/established. Therefore, the AVMs will predict the
See Table II.
transients that would appear in the fundamental component of
the ac voltages and currents and the transients in the dc side (e.g. C. Controller Parameters
see Figs. 8 and 9). The AVMs however will not predict any phe-
nomena related to switching (including harmonics that are due , , , ,
to switching). , , , ,
, , , ,

V. CONCLUSIONS
REFERENCES
In this paper, the dynamic average-value model of the
[1] “ACSLX, Advanced Continuous Simulation Language, User’s Guide”
first CIGRE HVDC benchmark system has been developed Ver. 2.4 The AEgis Technologies Group, Inc., 2008. [Online]. Avail-
and demonstrated in both state-variable-based simulation able: http://www.acslsim.com
languages, such as Matlab/Simulink, as well as EMTP-type [2] “EASY5 Engineering Software for the Design, Analysis and Simula-
tion” MSC SimEnterprise, Inc., 2008. [Online]. Available: http://www.
packages, such as PSCAD/EMTDC. Due to removing the mscsoftware.com
switching, the AVM is continuous and computationally ef- [3] “EUROSTAG: Software for Simulation of Large Electric Power Sys-
ficient, which makes it suitable for system-level studies. In tems” Tractebel Energy Engineering, 2008. [Online]. Available: http://
www.eurostag.be
general, the AVM is independent of the simulation package. [4] “Simulink Dynamic System Simulation Software Users Manual” Math-
The accuracy of the AVMs is verified against the detailed Works Inc., 2008. [Online]. Available: http://www.mathworks.com
2054 IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 5, OCTOBER 2014

[5] “SimPowerSystems: Model and Simulate Electrical Power Systems [25] P. F. Toledo, L. Angquist, and H. P. Nee, “Frequency domain model
User’s Guide” The MathWorks Inc., 2006. [Online]. Available: of an HVDC link with a line-commutated current-source converter.
http://www.mathworks.com Part II: Varying overlap,” IET Trans. Gen., Transm. Distrib., vol. 3,
[6] “Piecewise Linear Electrical Circuit Simulation (PLECS) User pp. 757–770, Aug. 2009.
Manual” ver. 1.4, Plexim GmbH, 2008. [Online]. Available: [26] P. F. Toledo, L. Angquist, and H. P. Nee, “Frequency domain model of
www.plexim.com an HVDC link with a line-commutated current-source converter. Part I:
[7] “Automated State Model Generator (ASMG) Reference Manual” ver. Fixed overlap,” IET Trans. Gen., Transm. Distrib., vol. 3, pp. 771–782,
2, P C Krause&Associates, 2003. [Online]. Available: www.pcka.com Aug. 2009.
[8] “MicroTran Reference Manual” MicroTran Power System Analysis [27] S. Chiniforoosh, J. Jatskevich, A. Yazdani, V. Sood, V. Dinavahi, J. A.
Corp., 1997. [Online]. Available: http://www.microtran.com Martinez, and A. Ramirez, “Definitions and applications of dynamic
[9] “Alternative Transients Programs” ATP-EMTP, ATP User Group, average models for analysis of power systems,” IEEE Trans. Power
2007. [Online]. Available: http://www.emtp.org Del., vol. 25, no. 4, pp. 2655–2669, Oct. 2010.
[10] “PSCAD/EMTDC,” Ver. 4.0 On-Line Help Manitoba HVDC Research [28] S. Chiniforoosh, H. Atighechi, A. Davoodi, J. Jatskevich, A. Yazdani,
Centre and RTDS Technologies Inc., Winnipeg, MB, Canada, 2005. S. Filizadeh, M. Saeedifard, J. A. Martinez, V. Sood, K. Strunz, J. Mah-
[11] “Electromagnetic Transient Program” EMTP RV, Powersys Solu- seredjian, and V. Dinavahi, “Dynamic average modeling of front-end
tions, Inc., 2011. [Online]. Available: http://www.emtp.com diode rectifier loads considering discontinuous conduction mode and
[12] “Resistive Companion Modeling and Simulation for the Virtual Test unbalanced operation,” IEEE Trans. Power Del., vol. 27, no. 1, pp.
Bed (VTB) Modeling Guide” Univ. South Carolina, 2003. [Online]. 421–429, Jan. 2012.
Available: http://vtb.ee.sc.edu [29] J. D. Ainsworth, “Proposed benchmark model for study of HVDC con-
[13] K. Strunz and E. Carlson, “Nested fast and simultaneous solution for trols by simulator or digital computer,” presented at the CIGRE SC-14
time-domain simulation of integrative power-electric and electronic Colloq. HVDC With Weak AC Systems, Maidstone, U.K., Sep. 1985.
systems,” IEEE Trans. Power Del., vol. 22, no. 1, pp. 277–287, Jan. [30] M. Szechtman, T. Wess, and C. V. Thio, “First benchmark model for
2007. HVDC control studies,” Electra, pp. 54–67, Apr. 1991.
[14] C. E. Grund, T. H. Lee, S. R. Lightfoot, R. J. Piwko, R. V. Pohl, [31] M. Szechtman, T. Wess, and C. V. Thio, “A benchmark model for
K. Mortensen, R. J. Newell, J. Reeve, and D. A. Woodford, “Func- HVDC system studies,” in Proc. Int. Conf., AC and DC Power Transm.,
tional model of two-terminal HVDC systems for transient and steady- 1991, pp. 374–378.
state stability,” IEEE Trans. Power App. Syst., vol. PAS-103, no. 6, pp. [32] J. Jatskevich, S. D. Pekarek, and A. Davoudi, “Parametric average-
1249–1255, Jun. 1984. value model of a synchronous machine-rectifier system,” IEEE Trans.
[15] Y. X. Ni, “A simplified two-terminal HVDC model and its use in direct Energy Convers., vol. 21, no. 1, pp. 9–18, Mar. 2006.
transient stability assessment,” IEEE Trans. Power Syst., vol. PWRS-2, [33] J. Jatskevich, S. D. Pekarek, and A. Davoudi, “Fast procedure for con-
no. 4, pp. 1006–1012, Nov. 1987. structing an accurate dynamic average-value model of synchronous
[16] B. K. Johnson, “HVDC models used in stability studies,” IEEE Trans. machine-rectifier system,” IEEE Trans. Energy Convers., vol. 21, no.
Power Del., vol. 4, no. 2, pp. 1153–1163, Apr. 1989. 2, pp. 435–441, Jun. 2006.
[17] L. Hu and R. Yacamini, “Calculation of harmonics and interharmonics [34] M. O. Faruque, Y. Zhang, and V. Dinavahi, “Detailed modeling of
in HVDC schemes with low DC side impedance,” Proc. Inst. Elect. CIGRE HVDC benchmark system using PSCAD/EMTDC and PSB/
Eng., Gen., Transm. Distrib., pp. 469–476, 1993. SIMULINK,” IEEE Trans. Power Del., vol. 21, no. 1, pp. 378–387,
[18] L. Hu, “Sequence impedance and equivalent circuit of HVDC system,” Jan. 2006.
IEEE Trans. Power Syst., vol. 13, no. 2, pp. 354–360, May 1998. [35] P. Fisher, L. Angquist, and H. P. Nee, “A new control scheme for
[19] R. Kumar and T. Leibfried, “Analytical modeling of HVDC trans- an HVDC transmission link with capacitor-commutated converters
mission system converter using Matlab/Simulink,” in Proc. IEEE Ind. having the inverter operating with constant alternating,” in Proc.
Comm. Power Syst. Tech. Conf., 2005, pp. 140–146. CIGRE Int. Conf. Large High Voltage Elect. Syst., 2012, pp. 1–11.
[20] R. M. Brandt, U. D. Annakkage, D. P. Brandt, and N. Kshatriya, “Vali- [36] IEEE Guide for Planning DC Links Terminating at AC Locations
dation of a two-time step HVDC transient stability simulation model in- Having Low Short-Circuit Capacities, IEEE Standard 1204-1997.
cluding detailed HVDC controls and DC line L/R dynamics,” in Proc. [37] IEEE Guide for Planning DC Links Terminating at AC Locations
IEEE Power Eng. Soc. Gen. Meeting, 2006, pp. 1–6. Having Low Short-Circuit Capacities, IEEE Standard 1204-1997, Jun.
[21] H. Zhou, G. Yang, and J. Wang, “Modeling, analysis, and control for 1997.
the rectifier of hybrid HVDC systems for DFIG-based wind farms,” [38] P. C. Krause, O. Wasynczuk, and S. D. Sudhoff, Analysis of Electric
IEEE Trans. Energy Convers., vol. 26, no. 1, pp. 340–353, Mar. 2011. Machinery and Drive Systems, 2nd ed. Piscataway, NJ, USA: IEEE/
[22] D. Jovcic, N. Pahalawaththa, and M. Zavahir, “Analytical modeling of Wiley, 2002.
HVDC-HVAC systems,” IEEE Trans. Power Del., vol. 14, no. 2, pp. [39] P. Kundur, Power System Stability and Control. New York, USA:
506–511, Apr. 1999. McGraw-Hill, 1994.
[23] C. Osauskas and A. Wood, “Small-signal dynamic modeling of HVDC [40] L. Wang, J. Jatskevich, V. Dinavahi, H. W. Dommel, J. A. Martinez,
systems,” IEEE Trans. Power Del., vol. 18, no. 1, pp. 220–225, Jan. K. Strunz, M. Rioual, G. W. Chang, and R. Iravani, “Methods of in-
2003. terfacing rotating machine models in transient simulation programs,”
[24] R. K Pandey, “Stability analysis of AC/dc system with multirate dis- IEEE Trans. Power Del., vol. 25, no. 2, pp. 891–903, Apr. 2010.
crete-time HVDC converter model,” IEEE Trans. Power Del., vol. 23,
no. 1, pp. 311–318, Jan. 2008.

Das könnte Ihnen auch gefallen