You are on page 1of 4

Averaged-Circuit Modeling of Line-Commutated

Rectifiers for Transient Simulation Programs


Sina Chiniforoosh1, Ali Davoudi2, and Juri Jatskevich1
1

University of British Columbia, Vancouver, Canada; 2University of Illinois at Urbana-Champaign, USA


1
{sinach, jurij}@ece.ubc.ca; 2davoudi2@uiuc.edu

Abstract Dynamic average-value models (AVM) for linecommutated rectifier circuits are generally formulated in a
state-space form and hence are straightforward to implement in
state-variable-based simulation languages. In nodal-analysisbased languages, however, developing AVMs requires additional
effort to reformulate the models and interface them with the
external circuit networks. This paper proposes an averagedcircuit model for representing three-phase line-commutated
rectifier in nodal-analysis-based simulation languages. The
model derivation and its interface with the network are
presented. The proposed model is verified against a detailed
switch-level implementation of the system.

I.

INTRODUCTION

With the development of modern simulation tool, detailed


models of line-commutated converters, where switching of all
devices is implemented in detail, may be readily built in both
the state-variable- [1]-[2] and nodal-analysis-based [3]-[5]
simulation languages. It is therefore possible to develop
models of larger systems from a number of smaller
subsystems/modules that can be used for the simulation of
systems transients. However, the use of detailed switching
models leads to significant increase of the required computing
time, which in turn limits the size of the system that can be
practically simulated. Moreover, the switching models are
discontinuous and hence difficult to use for extracting the
small-signal characteristics of various modules for the systemlevel analysis.
The above challenges have led to development of the socalled dynamic average-value models (AVMs) which
approximate the original system by neglecting or
averaging the effects of fast switching within a prototypical
switching interval. These models are computationally efficient
and could run orders of magnitudes faster than the original
switching models enabling efficient simulation of systems
transients. Additionally, since AVMs are time-invariant, they
can be linearized about any desired operating point for smallsignal analysis, i.e., obtaining local transfer functions.
Dynamic average-value models for line-commutated
rectifiers are generally developed using two main approaches,
i.e., analytical [6]-[8] and parametric [9]-[11]. Regardless of
the approach, the final model is typically formulated in the

978-1-4244-5309-2/10/$26.00 2010 IEEE

state-space form. Therefore, implementation of the AVMs in


state-variable-based programs, where the discretization of the
differential equations is done at the system level, is
straightforward. However, in nodal-analysis-based programs
(the EMTP-type programs [3] and its many variants) the
discretization is done at the component level making modeling
approaches more challenging. Consequently, developing
AVMs for representing line-commutated rectifiers in the
nodal-analysis-based programs requires additional effort to
reformulate the models and interface them with the external
circuit networks. This has not yet received the needed
attention in the literature.
In this paper, an averaged-circuit model is developed for
representing three-phase line-commutated rectifiers in a nodalanalysis-based program. The final model is developed in the
form of an averaged equivalent circuit for the switching
rectifier and is interfaced with the external ac circuit-network
using the so-called indirect method of interfacing [12], [13].
II.

MOTIVATION FOR AVERGAE-VALUE MODELING

In the nodal-analysis-based programs, the underlying


solution approach is based on discretizing the differential
equations for each circuit component using a particular
integration rule. The EMTP [3] uses an implicit trapezoidal
rule for discretization and formulating the network nodal
equation that has the following general form

GVn = I h .

(1)

Here, G is the network nodal conductance matrix, and the


vector I h includes the so-called history current sources and
independent current sources injected into the nodes. The nodal
voltages Vn are unknown and calculated by solving (1) at
every time step.
Line-commutated rectifiers repeatedly introduce changes
into the topology of the circuit. As shown in Fig. 1 (a),
representing converters by detailed switching models makes
the overall circuit-network time-variant and repeatedly
changes the G matrix. In a typical algorithm accounting for
topological changes [14], [15], at each time step, possible
changes in the topology of the network are detected.
Whenever a change is detected, the network conductance

2318

matrix G is updated and re-factorized, and the solution of the


network for the instant of switching is recalculated using
interpolations/extrapolations. This imposes a burden on the
overall solution.

!"


Ldc

Lc

ebs

rdc

S3

vab_s

  

=>?

S4

S5

S6

 


 

 



Rload

vdc

Lc

ecs

Fig. 2. Three-phase line-commutated rectifier system.

i qs = i qs ,com + i qs,cond ,

#



Fig. 1. Implementation of line-commutated rectifiers in nodal-analysis-based


progrmas: (a) embedded switching models; (b) interfaced AVMs.

The use of dynamic average-value models eliminates the


switching form the system. However, the AVMs are generally
nonlinear and typically include inter-dependant current and/or
voltage sources that relate the ac and dc ports of the converter
circuit. Moreover, for the purpose of derivation, the ac
variables are normally transformed to an appropriate qd
reference frame. Hence, the ac ports of the AVMs are
represented in qd coordinates, whereas the external ac
network is normally represented in abc phase coordinates,
which additionally complicates replacing of the switching
models with their AVMs. The approach developed here is
based on interfacing the AVMs with the external circuit
network as shown in Fig. 1 (b).
III.

=>?

S2

ias

  

S1

Lc

eas

 



!"


vas

 






 

!"


idc

2 3 idc

iqs,com =

iqs,cond =

ids,com =

3 vqs

[ (

[cos (2 ) 4 cos + 3],

4 Lce

2 3 idc

( )]

5
5
sin 6 + sin 6

2 3 idc

3 vqs

i ds = i ds,com + i ds,cond , (3)

[ ( ) (

)]

[ ( )

)]

7
5 ,
sin 6 sin + 6

5
5
cos 6 cos 6

[sin (2 ) 4 sin + 2 ],

4 Lce

[ (

2 3 idc

( )]

5
7 ,
cos + 6 cos 6
where is the so-called commutation angle defined as
ids,cond =

(4)

(5)

(6)

(7)

DEVELOPING THE AVERAGE-VALUE MODEL

A. State-Space Model
Let us consider the three-phase line-commutated rectifier
system depicted in Fig. 2. The dynamic average model for this
system is typically considered for Mode 1 where each
switching interval is divided into two sub-intervals, namely
commutation and conduction [16]. For the purpose of
derivation, the ac variables are represented in the so-called
qd converter reference frame in which the d-axis component
of the input voltage is identically zero [6]. Also, it is assumed
that the dc bus current does not change within a switching
interval [6]. Following the approach set forth in [6], the
dynamics of the dc bus may be represented by the following
state equation

3 3

3
vqs rdc + Lc e idc vdc , (2)

2 Lc e idc
3 vqs

(8)

B. Averaged-Circuit Model
Equations (2)-(8) represent a state-space average-value
model for the switching circuit of Fig. 2. This has been
illustrated in the block diagram of Fig 3. In this figure, the ac
and dc networks are connected through a nonlinear timeinvariant block which replaces the switching network of Fig.
2. The inputs to this block are the qd components of the ac
input voltage and the voltage of the dc bus. The outputs
include the dc-bus current and the qd components of the ac
phase currents.

Ldc + 2 Lc

didc
=
dt

= cos 1 1



x=





x  (x , v ,  )

is the angular electrical frequency of the source. The bar


symbol has been used to denote the so-called fast average of
variables over the switching interval.

i







The algebraic equations describing the ac side are


obtained by averaging the q- and d- component currents in the
commutation and conduction sub-intervals, and combining the
result:

where vqs is the q component of the ac input voltage and e

= g (x , v

,

)






Fig. 3. Block diagram of the state-space average-value model.

The developed state model has to be discretized according


to the trapezoidal rule (as used in the EMTP) and interfaced

2319

with the external circuit-network. Omitting the intermediate


steps due to limited space, the final equation is:

1
Req

idc (t ) =

3 3

v qs (t ) v dc (t ) + ih,dc (t ) ,

(9)

2 Lrec
,
t

(10)

where
Req = Rrec +

Rrec = rdc +

Lce ,

Lrec = Ldc + 2 Lc .

(11)

Here, the discretization time step is denoted by t .The socalled history term, ih, dc (t ) , is a function of the network
variables at the previous time step. In particular,
2R
ih, dc (t ) = 1 rec idc (t t )

Req

(12)

1 3 3
+
vqs (t t ) vdc (t t ) .

Req

The AVM described by (3)-(12) is of the form depicted in


Fig. 4, where the dc and ac sides have been represented by
inter-dependent current sources. The dc port includes a
Norton-equivalent conductance-current-source pair. The acside current sources are represented as nonlinear functions of
the dc bus current idc and ac input voltage vqs .

between the ac network and the input port of the AVM in each
phase. The resistor must be chosen sufficiently large to reduce
the interfacing error. The value of the compensating current
source at each time step is then calculated as
icomp _ x (t ) =

V x (t t )
,
rz

(13)

where x is either of the abc phases. The final averaged


model with interfacing circuitry, which includes compensating
resistor rz and current source icomp is illustrated in Fig. 5.

L )



L )





 (

,


L )





 



 

 



 

The time-step delay in this method of interfacing may


cause unfavorable numerical oscillations and convergence
problems. To mitigate this problem, the use of interfacing
circuitry has been proposed for the machine models in [13]. A
modified circuit includes a resistor rz and the so-called
compensating current source icomp . This circuit is inserted

 

illustrated in Fig. 5.

 

dependent current sources at the ac side are obtained using the


solution of the network at the previous time step. These
current sources are then transformed back to abc , and
injected into the ac network as iinj _ a , iinj _ b , and iinj _ c as is

+


becomes a straightforward operation and the final model is


conveniently interfaced with the ac network in abc phase
variables (coordinates). In particular, the values of the input
three-phase voltages are transformed into the converter
reference frame using the appropriate transformation matrix
[6]. The value of the dependent current source at the dc side is
then obtained as a function of vqs . Similarly, the values of the



Fig. 4. Descretized averaged-circuit model.

  


The presence of nonlinear dependent sources, as well as


the ac side being represented in qd variables, still makes it
challenging to implement this model directly in a nodalanalysis-based program wherein the external ac network is
represented in physical phase variables, i.e., abc . Similar to
interfacing the qd machine models with the abc networks
[12], [13], a simple solution is to introduce a time-step delay
between the dc and ac subsystems. This indirect interfacing
method has been used to interface all the qd machine models
in the PSCAD/EMTDC software [5]. Adopting this method,
the ac and dc subsystems become decoupled and the
dependent current sources turn into independent sources. The
values of these sources are calculated, similar to the history
sources, according to the solution of the network at the
previous time step t t . Also, since the solution of the
network at the previous time steps are readily available, the
transformation of the network variables between abc and qd

 




 



 

Fig. 5. Compensated equivalent-circuit model of interfacing AVM.

IV. MODEL VRIFICATION


To demonstrate the effectiveness of the averaged-circuit
model formulated in the previous section, simulation studies
are carried out as follows. The detailed (switch-level) model
of the system (Fig. 2) has been implemented in
PSCAD/EMTDC [5]. The system parameters are given in the
Appendix. The averaged-circuit model is implemented using
an EMTP-type algorithm. Initially, at t = 0 , the system is at
zero initial conditions when the three-phase input source is
switched on. Then, at t = 0.02 s , the load resistance is
switched from its initial value, Rload = 2 , to Rload = 1 .
Finally, at t = 0.04 s , the load resistance is switched back to
2 . The resulting voltage and current at the dc bus, idc ,
vdc , as predicted by the detailed and average-value models

2320

are shown in Fig. 6. The waveforms of the phase a current ias


predicted by the detailed and average-value models are
superimposed in Fig. 7. Studies of Figs. 6 and 7 are obtained
using the typical EMTP time step of 50 s . These figures
demonstrate a close match between the detailed and averagevalue models of the system.
Next, the effect of increasing the time-step size is
evaluated. For this purpose, the results of the previous study
(Figs. 6 and 7) as predicted by the AVM are chosen as the
reference solution (labeled as Ref in the following figures).
The same study is carried out with time-step of 500 s , and
the results are superimposed in Figs. 8 and 9. As seen in this
figure, the accuracy of the AVM remains very good even at
such a large time-step. This also demonstrates an order of
magnitude increase of the simulation efficiency.




@?






 




 

@?


 




[2]

[5]
[6]
[7]







 

Fig. 7. Input phase current predicted by the models with



@?



@?

 

[10]

 

 






[12]

 
 

 



Fig. 8. DC bus current and voltage predicted with





 

[14]



 

[13]

500 s time step.

=I

[9]

[11]
 

50s time step.






[8]








Fig. 9. Input phase current predicted by the models with

parameters:

Vline rms = 208V ,

e = 2 60 rad s

rs = 0.1 , Lc = 0.095 mH , rdc = 0.5 , Ldc = 1.33 mH .

=I




 

50s time step.

Fig.6. DC bus current and voltage predicted with

System

[4]

   
 

APPENDIX

[3]

CONCLUSION

In this paper, a new averaged-circuit model was developed


for representing three-phase line-commutated rectifiers in
nodal-analysis-based simulation programs such as EMTP. The
proposed AVM is conveniently interfaced with the external
circuit-network through its Norton-equivalent conductancecurrent-source. The AVM is continuous (does not switch) and
can execute with much larger time step, which improves the
simulation efficiency for the system-level transient studies.
The simulation results demonstrate an excellent match
between the detailed switch-level implementation of the
system and the proposed average model.

[1]

   
 

V.

[15]

500 s time step.


[16]

2321

REFERENCES
Simulink Dynamic System Simulation Software, Users Manual,
MathWorks Inc., 2008. Available: http://www.mathworks.com.
Piecewise Linear Electrical Circuit Simulation (PLECS), User Manual,
Version 1.4, Plexim GmbH, 2008. Available: www.plexim.com.
H. W. Dommel, EMTP Theory Book, MicroTran Power System
Analysis Corp., 1992.
MicroTran Reference Manual, MicroTran Power System Analysis
Corp., 1997. Available: http://www.microtran.com.
PSCAD/EMTDC V4.0 On-Line Help, Manitoba HVDC Research
Centre and RTDS Technologies Inc., 2005.
P. C. Krause, O. Wasynczuk, and S. D. Sudhoff, Analysis of Electric
Machinery and Drive Systems, IEEE Press, 2002.
S. D. Sudhoff and O. Wasynczuk, Analysis and average-value
modeling of line-commutated converter-synchronous machine
systems, IEEE Trans. Energy Conversion, vol. 8, no. 1, pp. 92-99,
Mar. 1993.
H. Zhu, New multi-pulse diode rectifier average models for ac and dc
power systems studies, PhD Dissertation, Virginia Polytechnic
Institute and State University, 2005.
I. Jadric, D. Borojevic, and M. Jadric, Modeling and control of a
synchronous generator with an active DC load, IEEE Trans. Power
Electronics, vol. 15, no. 2, pp. 303311, Mar. 2000.
J. Jatskevich, S. D. Pekarek, and A. Davoudi, Parametric averagevalue model of synchronous machine-rectifier systems, IEEE Trans.
Energy Conversion, vol. 21, no. 1, pp. 9-18, Mar. 2006.
J. Jatskevich, S. D. Pekarek, and A. Davoudi, Fast procedure for
constructing an accurate dynamic average-value model of synchronous
machine-rectifier systems, IEEE Trans. Energy Conversion, vol. 21,
no. 2, pp. 435-441, Jun. 2006.
D.A. Woodford, A.M. Gole, and R.W. Menzies, Digital Simulation of
DC Links and AC Machines, IEEE Trans. Power Apparatus and
Systems, vol. PAS-102, no. 6, pp. 1616-1623, Jun. 1983.
A.M. Gole, R.W. Menzies, H.M. Turanli, and D.A. Woodford,
Improved interfacing of electrical machine models to electromagnetic
transients programs,
IEEE Trans. Power Apparatus and
Systems, vol. PAS-103, no. 9, pp. 24462451, Sep. 1984.
K. Strunz, L. Linares, J. R. Marti, O. Huet, and X. Lombard, Efficient
and accurate representation of asynchronous network structure
changing phenomena in digital real time simulators, IEEE Trans.
Power Systems, vol. 15, no. 2, pp. 586-592, May 2000.
L. R. Linares and J. R. Marti, A resynchronization algorithm for
topological changes in real time fast transients simulation, in Proc.
14th Power Systems Computation Conference (PSCC02), Sevilla,
Spain, June 2002.
R. M. Davis, Power Diode and Thyristor Circuits, Cambridge at the
University Press, 1971.

You might also like