You are on page 1of 11

584 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO.

3, SEPTEMBER 2005

A Synchronous Machine Model With Saturation and


Arbitrary Rotor Network Representation
Dionysios C. Aliprantis, Member, IEEE, Scott D. Sudhoff, Senior Member, IEEE, and Brian T. Kuhn, Member, IEEE

Abstract—This paper addresses equivalent circuit and magnetic , Order of -axes rotor network minimal
saturation issues associated with synchronous machine modeling. polynomials.
In the proposed synchronous machine model, the rotor equivalent Number of poles.
circuits are replaced by arbitrary linear networks. This allows for
elimination of the equivalent circuit parameter identification pro- Electric power supplied to coupling field
cedure since the measured frequency response may be directly em- (in watts).
bedded into the model. Magnetic saturation is also represented in Characteristic polynomial of transfer func-
both the - and -axis. The model is computationally efficient and tion matrix .
suitable for dynamic time-domain power system studies. Effective resistance of an average-value
Index Terms—Electric machines, modeling, power system exciter-rectifier system representation (in
stability, realization theory, rotating machine transient analysis, ohms).
synchronous generator transient analysis, synchronous machines, Field winding resistance (in ohms).
transfer functions.
Stator winding resistance (in ohms).
Complex frequency (in radians per
, , Rotor network state equation matrices second).
. Electromagnetic torque ( m).
, Column vectors of the matrix . Prime-mover torque ( m).
, Row vectors of the matrix . Voltage across the field side of the -axis
Voltage-behind-reactance of an average- rotor network (in volts).
value exciter-rectifier system representa- Field winding voltage (in volts).
tion (in volts). , -axes magnetizing branch voltages (in
Field winding current (in amperes). volts).
, -axes magnetizing branch current (in Armature voltage (in
amperes). volts).
, Current flowing into the armature side of Coupling field energy (in Joules).
the rotor networks (in amperes). , -axes rotor network states.
Stator windings current -axis rotor two-port network transfer
(in amperes). function matrix ( ).
Moment of inertia ( ). -axis rotor network transfer function
Park’s transformation matrix. ( ).
Stator winding leakage inductance (in Hen- Constant related to the -axis rotor network
ries). admittance at dc ( ).
Transient inductance of an average-value , , Elements (transfer functions) of .
exciter-rectifier system representation (in Incremental inverse magnetizing induc-
Henries). tance matrix.
Minimal polynomial of transfer function , Inverse magnetizing inductances ( ).
matrix . , Saliency-dependent magnetizing path
Field-to-armature turns ratio. characteristic constants.
, Number of -axes rotor network states. , , , Constants of the -axis rotor network
transfer functions .
, Constants of the -axis rotor network
Manuscript received September 17, 2003; revised March 31, 2004. This
work was supported by the “Naval Combat Survivability” Effort under Grant transfer functions ,
N00024-02-NR-60427. Paper no. TEC-00249-2003. .
D. C. Aliprantis is with the Greek Armed Forces (e-mail: aliprantis@ Electrical rotor position (in radians).
alumni.purdue.edu).
S. D. Sudhoff is with the Department of Electrical and Computer Engi- Mechanical rotor position (in radians).
neering, Purdue University, West Lafayette, IN 47907-1285 USA (e-mail: Stator windings leakage flux linkage
sudhoff@ecn.purdue.edu). ).
B. T. Kuhn is with SmartSpark Energy Systems, Inc., Champaign, IL 61820
USA (e-mail: b.kuhn@smartsparkenergy.com). Effective magnetizing flux linkage .
Digital Object Identifier 10.1109/TEC.2005.845455 , -axes magnetizing flux linkage .
0885-8969/$20.00 © 2005 IEEE

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
ALIPRANTIS et al.: SYNCHRONOUS MACHINE MODEL WITH SATURATION AND ARBITRARY ROTOR NETWORK REPRESENTATION 585

Stator windings flux linkage


.
Electrical rotor speed (in radians per
second).
Mechanical rotor speed (in radians per
second).

I. INTRODUCTION

M ATHEMATICAL models of physical devices are ab-


stractions of reality. Their purpose is to portray, with
sufficient accuracy and simplicity, the important characteristics
of the systems under consideration. For synchronous machines,
relatively low-order models may be employed to predict their
steady-state or dynamic behavior, with the orthogonal -axes Fig. 1. IEEE Standard model of type 3.3.
theory [1] used in the majority of cases. Over the years, a
plethora of models has been proposed to represent phenomena
related to the distributed circuit behavior of rotor windings,
magnetic saturation, and saliency-related issues. However,
the common thread among them is that they were focused
on—and, thus, limited by—the specific problem at hand.
This work provides an integrated perspective on synchronous
machine modeling, realizing the need for a generalized formu-
lation that is compatible with modern computational tools.
Synchronous machines are usually employed as generators
connected to a power system. In the current IEEE Standard [2], it
is acknowledged that synchronous machines may be accurately
modeled by two lumped-parameter equivalent circuits repre-
senting the - and the -axis, as shown in Fig. 1. The number of
rotor “damper” branches is selected in accordance with the rotor Fig. 2. Proposed synchronous machine model.
design. In low-order models, these branches correspond to the
actual amortisseur windings; higher order models utilize these If magnetizing path saturation is to be considered—either by ad-
branches to represent the distributed effects in the rotor iron. justing the magnetizing branch inductance or by replacing it by
The rotor-related part of the equivalent circuits proposed in a nonlinear element—the physical significance of these elements
the standard differs from other “conventional” equivalent cir- should be preserved [13]. Although some physical significance
cuits [1], [3] since it contains the “differential leakage” induc- may be attributed to the parameters of models with up to three
tance . Its use was proposed by Canay in his seminal paper “damper” circuits [14], it is gradually lost as the order is further
of 1969 [4], to signify unequal coupling between the rotor wind- increased, as is recommended for accurate transient stability or
ings and the stator, and between the rotor circuits themselves. A subsynchronous resonance studies [15].
variety of other circuit structures has also been proposed in the The purpose of this work is to address these equivalent circuit
literature [5]–[8]. issues. In the model that is set forth herein, the equivalent-circuit
New circuits having equal to zero may be obtained with structure of the rotor is replaced by a completely arbitrary linear
an appropriate transformation, as was shown by Canay in his circuit, a two-port network for the -axis, and a single branch for
original paper, and again by Kirtley in [9]. This corresponds to the -axis, as shown in Fig. 2. This approach has been adopted
forcing equal magnetic coupling between all rotor circuits and from previous modeling work on induction machines [16], em-
the stator winding—often called the “equal mutuals” base. How- phasizing the importance of the rotor’s actual input–output be-
ever, the equal mutuals base is restricted to the case of three cou- havior rather than the debatable physical meaning of the equiv-
pled circuits; that is, when the rotor has, at most, a field winding alent circuit parameters. It offers the advantage that once the
and one damper [10]. The most unconstrained form of equivalent transfer functions of these linear circuits are determined—using
circuit is one with differential leakage inductances between all standstill frequency response tests, for example, they may be
damper circuits. Important theoretical results about equivalent immediately incorporated in the model [17]. It does not require
circuits were published in [11] and [12] where it was shown that the tedious and time-consuming process of equivalent circuit pa-
there is no unique ladder network representation of a given rameter identification to which numerous studies have been de-
rotor two-port network with a prescribed impedance matrix. voted [18], [19]. Indeed, this problem is highly nonlinear, and
A further disadvantage of using a mathematical transformation the solution algorithms often face convergence issues, accentu-
to ensure a specific network structure (one with no differential ated by the fact that there is no unique solution. Furthermore, for
leakage) is that it causes the new values of the stator leakage and modern computer simulation software, such as Matlab/Simulink
magnetizing inductance to lose their original physical meaning. [20] or ACSL [21], the entry of a specific equivalent circuit

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
586 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO. 3, SEPTEMBER 2005

structure is not required. Rather, it is more efficient to directly B. Voltage Equations


provide a state-space representation of the system to be simu- The stator voltage equations may be expressed in vari-
lated. The proposed model possesses such a form. ables as
Apart from equivalent circuit-related problems, synchronous
machine research has also focused on the accurate incorporation
of magnetic saturation, which has been shown to considerably (3)
affect their operating characteristics [22]. Different methods to
model this complicated phenomenon were analyzed in a variety where , , and denote stator winding (phase-to-
of publications, such as [23]–[31]. An alternative approach is to neutral) voltages, currents flowing into the machine terminals,
employ artificial neural networks to represent the magnetizing and flux linkages, respectively, and is the stator winding re-
path saturation [32]–[34], or the variation of rotor parameters sistance. Transforming (3) to the rotor reference frame yields
[35], [36].
The form of the proposed model readily lends itself to the mod-
(4)
ification of the magnetizing inductances for saturation (and cross-
saturation) modeling in both axes. Recent evidence suggests that
lumping saturation effects in the magnetizing branch is a reason- where . It will be assumed hereafter that
able assumption [37], [38] and this hypothesis is further corrob- the zero-sequence variables can be neglected.
orated by our results. Although the leakage flux paths may satu- The state equations of the -axis two-port network (Fig. 2)
rate due to excessive current or by high magnetizing flux levels, may be expressed by a linear system of order as
synchronous generators are usually operated close to their nom-
inal values of flux, so the leakage inductance may be assumed
to remain constant for a wide range of studies. A derivation of
the model’s torque equation, and restrictions on the magnetizing
inductances arising from the assumption of a lossless (conser- (5)
vative) coupling field will be presented in later sections.
(6)
II. PROPOSED SYNCHRONOUS MACHINE MODEL
From a computer simulation point of view, the proposed where , , , and
model is of the voltage-in, current-out type; the stator and . The contents of and were written as the column
field winding voltages are the inputs, while the stator and field vectors , , and the row vectors , . The magne-
winding currents are the outputs. The model is computationally tizing branch voltages are equal to the derivatives of the corre-
efficient in the sense that it is noniterative at each time step, sponding magnetizing flux linkages, which will be computed in
and uses only a minimum number of states obtained from a the ensuing analysis.
minimal realization of the measured input–output behavior. It The voltage may be eliminated from the equations, since
is especially suitable for time-domain dynamic simulations of it is related to the field voltage and current by
power systems, as well as for the design and optimization of . Using this, the following state equation is obtained:
control schemes.

A. Notation (7)
In order to assist the reader, the paper’s nomenclature is de-
fined. Throughout this work, matrix and vector quantities ap- Similarly, the -axis state equations are
pear in bold font. The primed rotor quantities denote referral to
the stator through the turns ratio, which is defined as the ratio
of field-to-armature turns [39]. The analysis (8)
takes place in the rotor reference frame; the often used “ ” su- (9)
perscript [1] is omitted for convenience. The electrical rotor po-
sition and electrical rotor speed are times the me- where , , ,
chanical rotor position and mechanical speed, , where , and is the order of the -axis system.
is the number of poles. The transformation of stationary It will be useful to derive expressions for the current deriva-
to variables in the rotor reference frame is defined by [1] tives. From (6)–(9)
(1)

where

(10)

(2) (11)

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
ALIPRANTIS et al.: SYNCHRONOUS MACHINE MODEL WITH SATURATION AND ARBITRARY ROTOR NETWORK REPRESENTATION 587

C. Rotor Transfer Functions The formulation of the -axis transfer function is more
In the frequency domain, the rotor currents and voltages are straightforward due to the absence of a field winding. In partic-
related by the transfer functions ular, the -axis transfer function may be expressed as

(22)

In this case, poles and zeros at the origin are not allowed.
(12)
D. Realization Theory
(13) The rotor transfer functions are the starting point for the com-
where and the tilde are used to denote phasor quanti- putation of the time-domain state matrices, as used in (7) and
ties. The elements of ( , , ), and (8). The problem is to determine an internal, state-space descrip-
are strictly proper rational polynomial functions of the complex tion of a linear system, given its external, input–output descrip-
frequency of the form tion. This is the subject of system realization theory [40].
Numerous algorithms exist for deriving a realization. How-
(14) ever, for increased computational efficiency, it is desirable to
obtain a system of the least possible order, a so-called minimal
The coefficients and may not be simultaneously equal to realization. In contrast to the single-input/single-output case,
zero since that would result in a pole-zero cancellation at the where it is rather straightforward to obtain a realization, the mul-
origin. This is a representation of a most general form, but a tiple-input/multiple-output case (like the two-port -axis rotor
simplified version may be obtained for the -axis if the physics system) is more complicated. The difficulty arises when deter-
of the rotor are taken into account. mining the minimal realization order. Consider a transfer func-
To this end, by applying Faraday’s law to the field winding tion written in the form , where is a
(i.e., ), it can be seen that there is no matrix of polynomials and is the minimal polynomial of
dc voltage drop besides the ohmic drop of the winding’s resis- .1 The roots of constitute a subset of the eigenvalues
tance (which is external to the two-port network). This implies of the minimal realization; hence, a minimal polynomial with
that there exists a direct path for dc current between the primary two roots may correspond to a system with a second-, third-, or
and secondary sides of the two-port network. fourth-order minimal realization.
Under this assumption, the constants of the polynomials In the case of the proposed model, the transfer function matrix
will all have the same absolute value, and the system will pos- entries are the functions (19)–(21), the degree of the minimal
sess a pole at the origin. To see this, assume that the field side polynomial is , and the order of the minimal realization is
of the admittance block is short circuited so that . This is proved using the following theorem:
(15) The order of any minimal realization is equal to the de-
(16) gree of the characteristic polynomial of the transfer
function matrix [40, p. 397].2
As the frequency approaches zero , the hypothesis that It remains to compute , given the functional forms
at low frequency the two-port network behaves as an ideal series (19)–(21).
inductor implies that , so Let us write the -axis transfer function as
(17)
(23)
Since the denominator of (14) corresponds to the least common
where
denominator of all elements of , the element of
is equal to of . Similarly, setting and letting (24)
the frequency approach zero yields
It is assumed that no common factors exist between ,
(18)
, , and . To obtain , it is necessary to
Thus, the element of is equal to of . This compute all nonzero minors of . The first-order minors
observation, coupled with the fact that these admittances be- are the entries of , and their monic least common denom-
come infinite as (again because of the assumption that inator is (by definition) the minimal polynomial . The
the network acts as an ideal series inductance at sufficiently low second-order minor is equal to the determinant
frequency), implies that the -axis transfer functions must have
(25)
the following forms:
Taking into account the specified forms of the transfer functions
(19) (19)–(21), it is readily shown that the numerator of has a
1The minimal polynomial is defined herein as the monic least common de-
(20) nominator of the transfer function entries. A polynomial is monic when the
highest order term has a coefficient of one.
2The characteristic polynomial p (s) is defined as the monic least common
(21)
denominator of all nonzero minors of (s). Y
Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
588 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO. 3, SEPTEMBER 2005

root at the origin, which cancels out with one of the two zero the electric power supplied to the coupling field may be ex-
roots of . Hence pressed after manipulation as

(26)

The degree of the characteristic polynomial—the order of the (33)


minimal realization—is thus . An algorithm to
obtain a minimal state-space realization from the transfer func- The electrical energy provided to the coupling field is partly
tion matrix is described in the Appendix. stored, and the rest is transferred to the mechanical system. If
The order of the -axis realization, which is a single- denotes the coupling field energy and the EM torque,
input/single-output system, is . then the power balance may be written as

E. Leakage and Magnetizing Path Magnetics (34)


The stator flux linkage is separated into leakage and magne-
tizing flux terms as where the product is positive when energy is supplied
to the rotor. Equating (33) and (34), and solving for the rate of
(27) change of the field energy yields

where is the stator leakage flux, and is the magne-


tizing flux. The same value of leakage inductance is used
for both the - and the -axis.
The functional forms of main path saturation in the proposed (35)
model are similar to [30]. Since in the general case of a salient
rotor machine, the magnetizing magnetomotive force (mmf) and Hence, the differential change of the coupling field energy may
flux vectors are not aligned, the following relations may be as- be written as
sumed for the magnetizing path:
(36)
(28)
(29) where
The magnitude of the effective magnetizing flux vector is de- (37)
fined by
(38)
(30)
(39)
where is a saliency-dependent parameter. Differentiating (28),
(29) with respect to time yields
The change of coupling field energy from an initial state
to an arbitrary final state is
(31)
obtained by integrating (36)
where

(32)

(40)
The “ ” subscript denotes incremental value.

F. Torque Equation Since the field is assumed to be conservative, the integration may
The proposed model’s electromagnetic (EM) torque equation be performed over an arbitrary trajectory. Assume that the initial
may be derived by examining the energy balance of the ma- energy is , integrate the first term from to , while
chine’s coupling field [1]. The coupling field is created by the the fluxes are maintained at zero—which forces to be zero
magnetizing flux, which links both stator and rotor circuits. as well. This transition does not change the field energy. Then,
The electric power supplied to the coupling field is equal to consecutively integrate each flux from zero to an arbitrary final
the input power, minus the power lost in the stator and field re- value, while keeping and the other flux constant. Recall that
sistances, minus the power that supplies the stator leakage field, the transformation to the rotor reference frame eliminates the de-
minus the power that is dissipated or stored inside the rotor ad- pendence of the magnetizing inductances from the rotor position.
mittance block. Using the equivalent circuit of Fig. 2 and (27), The magnetizing currents are independent of —as in (28) and

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
ALIPRANTIS et al.: SYNCHRONOUS MACHINE MODEL WITH SATURATION AND ARBITRARY ROTOR NETWORK REPRESENTATION 589

(29). Therefore, the final value of field energy Next, the time derivatives of the stator flux linkages are cal-
is independent of the angular position , that is culated from (4)

(41) (49)

This observation coupled with (37) yields the following well- (50)
known expression for the EM torque:
as functions of state variables and model inputs (the stator volt-
(42) ages). However, they are only evaluated as an intermediate cal-
culation; they are not integrated since are not states.
For generator action (and ), the torque will be negative. The differentiation with respect to time of (27) yields an al-
ternate expression for the derivatives of the stator flux linkages
G. Restrictions on the Inverse Magnetizing Inductances
[cf. (49) and (50)]
In (28) and (29), the inverse magnetizing inductances were
defined as any arbitrary function of flux. However, to be con- (51)
sistent with the assumption of a lossless coupling field, certain
modeling restrictions must be imposed [41], [42]. (52)
Specifically, the coupling field’s energy expression, which in
view of (41), has become
The derivatives of the rotor and magnetizing currents may be
evaluated using (10) and (11) and (31), so

(43)

must satisfy the requirements of a conservative field. A neces-


sary and sufficient condition for this is [43]
(53)
(44)

Substitution of (28) and (29) into (44) yields

(45) (54)

Canceling common terms and integrating both sides yields The following linear system of equations may therefore be
formulated:
(46)

This restriction has to be enforced during the magnetizing char-


acteristics’ curve-fitting procedure. Note that it also renders
the incremental inverse inductance matrix symmetric. The
method for determining the constants and is described in
detail in [17].
(55)
H. Model Integration
The state variables are selected as . The
goal of the ensuing analysis is the formulation of equations for
the time derivatives of the state variables.
First, note that both the rotor and magnetizing currents have
(56)
been previously expressed as functions of the states: the rotor
currents are given in terms of the rotor admittance states from
(6) and (9); the magnetizing currents depend on the magnetizing where the quantities and are obtained
flux states, as seen from (28)–(30). Hence, the stator currents from (49) and (50). The solution of (55) and (56) yields the
time derivatives of two of the state variables (the magnetizing
(47) flux linkages and ). Apart from being integrated them-
(48) selves, these derivatives are also inserted in (7) and (8)—as
the magnetizing voltages and —for calculating the
are also functions of state variables. derivatives of the remaining state variables and .

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
590 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO. 3, SEPTEMBER 2005

I. Model Summary
The proposed model is of the voltage-in, current-out type; in-
puts are the armature and field winding voltages, as well as the
rotor angular position and speed, whereas outputs are the arma-
ture and field winding currents, and the machine’s developed
EM torque. The model’s computational structure may be sum-
marized as follows.
1) Transform the stator voltages to the rotor reference
frame using (1).
Fig. 3. Schematic of experimental setup; the brushless synchronous generator
2) Compute the magnetizing currents, using (28) and (29). is feeding a nonlinear rectifier load.
3) Compute the rotor currents using (6) and (9).
4) Compute the -axis armature currents using (47) and where is the voltage behind the effective resistance and
(48). the transient inductance ; the primes denote that all quanti-
5) Calculate the armature currents by applying the in- ties have been referred to the stator. These voltage drops may be
verse of transformation (1). incorporated in the proposed model in a straightforward manner.
6) Calculate the EM torque from (42). In particular, the voltage at the field side of the -axis admit-
7) Compute the derivatives of the stator flux linkages, using tance block may be written as
(27), (49), and (50).
8) Compute the incremental inverse inductance matrix (58)
(32).
9) Substitute known quantities into (55) and (56), and solve Using (5) and (6), this equation becomes
the 2 2 linear system for the derivatives of the magne-
tizing flux linkages.
10) Compute the derivatives of the rotor network states using
(59)
(7) and (8).
11) Interface the synchronous machine model with the rest and solving for yields
of the system that is simulated, numerically integrate the
state equations, and repeat from step 1).
The model is not computationally intensive, since it only in- (60)
volves simple numerical computations at each time step. Thus, where . After the substitution of (60)
it can be readily implemented in a dynamic simulation environ- into (5), the following modified state equation is obtained:
ment, such as Matlab/Simulink [20].
In order to initialize the model from a load-flow study—
where the machine is normally connected to a generator
bus or a swing bus, the model states may be (61)
calculated by solving a set of nonlinear equations. Specifically,
where denotes the identity matrix of dimension .
it can be seen from (7), (8), and the proposed realization (75)
The derivatives of the magnetizing flux linkages are given by the
that in the steady-state, all rotor network states must be equal
solution of the linear system of equations formed by (56) and
to zero; the exception is , which is directly related to the
field winding voltage through . The
mathematical expressions for the prespecified quantities ( ,
, or ) may be readily manipulated and written in terms of
and . The nonlinear system may then be solved for the
magnetizing flux states.

J. External Impedance Incorporation


The model is flexible enough to allow the incorporation of
an external resistance and inductance, connected in series with (62)
the field winding. For example, such would be the case of a
brushless excitation system, when the detailed model of the ex- which was obtained by a procedure analogous to the one de-
citer-rectifier system is replaced by a nonlinear average-value scribed in the previous section.
model. Essentially, this simplification produces a voltage be-
hind reactance representation of the exciter. As was shown in III. EXPERIMENTAL VALIDATION
[44], the effective voltage at the generator field has the form
The experimental setup is depicted in Fig. 3 and contains
a Leroy–Somer brushless synchronous generator, model LSA
(57) 432L7. This is a salient four-pole machine rated for 59 kW,

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
ALIPRANTIS et al.: SYNCHRONOUS MACHINE MODEL WITH SATURATION AND ARBITRARY ROTOR NETWORK REPRESENTATION 591

600 V, at 1800 r/min. The generator’s prime mover is a Dyne


Systems 110-kW, 590 m, vector-controlled induction motor-
based dynamometer that is programmed to maintain constant
rated speed. The voltage regulator uses a proportional-integral
control strategy to maintain the commanded voltage (560 V
line-to-line, fundamental rms) at the generator terminals; the
brushless exciter’s field current is controlled with a hysteresis
modulator. The rotor’s angular position and speed are measured
with an optical position encoder that is fitted on the machine
shaft. The generator is loaded with an uncontrolled rectifier that
feeds a resistive load through an filter.
The complete methodology for characterizing the main syn-
chronous machine is described in detail in [17]. The obtained
parameters and functional forms are as follows: ,
, ,

(63)

(64)

(65)

(66)

(67)

(68)

where and , . Note that


the expression for is valid for
(which is above the knee of the saturation curve). For greater
values, the rational function diverges, and an appropriate con-
tinuation is defined so that the flux linkage increases linearly Fig. 4. Plots of the commanded and actual line-to-line voltage “envelope,”
with current (with the same slope as at ). The load parame- which is computed from the synchronous reference frame voltages v =
[3(v + v )] . Each of the seven trapezoid shaped blocks is characterized
ters are mH, F, . With by a different slope (the same for rise and fall) and peak voltage: (1) 20 000 V/s,
this load, the machine operates at the region of the knee of the 560 V; (2)–(4) 2000 V/s, 560 V, 420 V, 280 V, respectively; (5)–(7) 400 V/s,
magnetizing curve ( [17]). The brushless ex- 560 V, 420 V, 280 V, respectively. (Note: the above voltage values correspond
to rms quantities.)
citer is represented by an average-value model that incorporates
magnetic hysteresis using Preisach’s theory; this model is dis-
cussed in detail in [45] and a forthcoming publication [46]. The
prime mover model’s output is the mechanical torque, which is
inserted in the well-known equation of motion. The mechan-
ical rotor speed may thus be computed by numerically inte-
grating . The electrical rotor po-
sition is then calculated by integrating the rotor’s speed (i.e.,
). The prime mover’s model is documented
in [45].
For the first experiment, the switch is closed so that the
total resistive load is . The generator’s voltage Fig. 5. Shaft speed variation around the commanded value of 1800 r/min.
reference is modified according to the profile shown in Fig. 4.
This series of commanded voltage steps creates an extended
period of significant disturbances, and tests the validity of for mechanical speed (Fig. 5) are in excellent agreement with
the model for large-transients simulations. The actual voltage the experimental results. Finally, detailed voltage and current
exhibits an overshoot, which is more pronounced for the faster waveforms are depicted in Fig. 6. Since the proposed model
slew rate steps. Moreover, due to the exciter’s magnetically is based on the -axes theory, higher harmonics attributable
hysteretic behavior, it does not fall to zero. The predictions to the machine’s design are not represented. This reflects on

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
592 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO. 3, SEPTEMBER 2005

IV. CONCLUSION
This work presents an integrated perspective on synchronous
machine modeling, using arbitrary transfer function representa-
tions that replace the rotor’s equivalent circuit structures. This
approach offers several advantages, such as the direct incorpo-
ration of frequency response results into the model—without
further consideration of equivalent circuit parameter identi-
fication—and accurate representation of magnetic saturation
effects. The model retains the computational efficiency of the
-axes theoretical framework and is suitable for small- and
large-signal time-domain simulations of power systems.

APPENDIX
REALIZATION ALGORITHM
The following algorithm produces a realization with matrix
diagonal [40]. It is valid only for the case where the roots
Fig. 6. Steady-state voltage and current waveforms (560 V, line-to-line, rms).
of the minimal polynomial are distinct. It is assumed that the
transfer function matrix elements are (19)–(21).
1) Compute the roots of the minimal polynomial

(69)

2) Expand into partial fractions

(70)

The 2 2 residue matrices may be computed by

(71)

Fig. 7. Line-to-line voltage “envelope.” (Lowpass filtered;  = 2:65 ms.) and are of full rank; however

(72)

is a matrix of rank 1.
3) Write

(73)

for example, by computing the LU decomposition. The


matrices , are 2 2. may be factored as
Fig. 8. Shaft speed variation around the commanded value of 1800 r/min.

(74)
the voltage waveforms of Fig. 4, wherein the experimental
waveform contains more ripple than the simulated waveform. 4) The realization is given by
However, the harmonics that are caused by the nonlinearity of
the load are predicted accurately.
The second experiment involves sudden load changes. Ini-
tially, the switch is open; at , it is closed, and at ..
.
, it is opened again. In Fig. 7, a lowpass-filtered version
of the line-to-line voltage “envelope” is depicted. On average,
the simulated and experimental waveforms are similar. As dis-
cussed above, the experimental voltage includes higher-order (75)
..
harmonics caused by slot effects. The mechanical speed wave- .
forms are illustrated in Fig. 8.

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
ALIPRANTIS et al.: SYNCHRONOUS MACHINE MODEL WITH SATURATION AND ARBITRARY ROTOR NETWORK REPRESENTATION 593

REFERENCES [27] F. P. de Mello and L. N. Hannett, “Representation of saturation in syn-


chronous machines,” IEEE Trans. Power Syst., vol. 1, no. 4, pp. 8–18,
[1] P. C. Krause, O. Wasynczuk, and S. D. Sudhoff, Analysis of Electric Nov. 1986.
Machinery. New York: IEEE Press, 1995. [28] J. O. Ojo and T. A. Lipo, “An improved model for saturated salient pole
[2] Guide for Synchronous Generator Modeling Practices in Stability Ana- synchronous motors,” IEEE Trans. Energy Convers., vol. 4, no. 1, pp.
lyzes, IEEE Std. 1110, Nov. 1991. 135–142, Mar. 1989.
[3] C.-M. Ong, Dynamic Simulation of Electric Machinery Using Matlab/ [29] J. Tamura and I. Takeda, “A new model of saturated synchronous ma-
Simulink. Englewood Cliffs, NJ: Prentice-Hall, 1998. chines for power system transient stability simulations,” IEEE Trans.
[4] I. M. Canay, “Causes of discrepancies on calculation of rotor quanti- Energy Convers., vol. 10, no. 2, pp. 218–224, Jun. 1995.
ties and exact equivalent diagrams of the synchronous machine,” IEEE [30] S.-A. Tahan and I. Kamwa, “A two-factor saturation model for syn-
Trans. Power App. Syst., vol. PAS-88, no. 7, pp. 1114–1120, Jul. 1969. chronous machines with multiple rotor circuits,” IEEE Trans. Energy
[5] R. P. Schulz, W. D. Jones, and D. N. Ewart, “Dynamic models of turbine Convers., vol. 10, no. 4, pp. 609–616, Dec. 1995.
generators derived from solid rotor equivalent circuits,” IEEE Trans. [31] K. A. Corzine, B. T. Kuhn, S. D. Sudhoff, and H. J. Hegner, “An im-
Power App. Syst., vol. PAS-92, no. 3, pp. 926–933, May/Jun. 1973. proved method for incorporating magnetic saturation in the q-d syn-
[6] H. Bissig, K. Reichert, and T. S. Kulig, “Modeling and identification chronous machine model,” IEEE Trans. Energy Convers., vol. 13, no.
of synchronous machines, a new approach with an extended frequency 3, pp. 270–275, Sep. 1998.
range,” IEEE Trans. Energy Convers., vol. 8, no. 2, pp. 263–271, Jun. [32] H. Tsai, A. Keyhani, J. A. Demcko, and D. A. Selin, “Development of a
1993. neural network based saturation model for synchronous generator anal-
[7] I. M. Canay, “Determination of the model parameters of machines from ysis,” IEEE Trans. Energy Convers., vol. 10, no. 4, pp. 617–624, Dec.
the reactance operators x (p); x (p) (evaluation of standstill frequency 1995.
response test),” IEEE Trans. Energy Convers., vol. 8, no. 2, pp. 272–279, [33] S. Pillutla and A. Keyhani, “Neural network based saturation model for
Jun. 1993. round rotor synchronous generator,” IEEE Trans. Energy Convers., vol.
[8] A. Keyhani and H. Tsai, “Identification of high-order synchronous gen- 14, no. 4, pp. 1019–1025, Dec. 1999.
erator models from SSFR test data,” IEEE Trans. Energy Convers., vol. [34] H. B. Karayaka, A. Keyhani, B. L. Agrawal, D. A. Selin, and G. T. Heydt,
9, no. 3, pp. 593–603, Sep. 1994. “Identification of armature, field, and saturated parameters of a large
[9] J. L. Kirtley, “On turbine-generator rotor equivalent circuits,” IEEE steam turbine-generator from operating data,” IEEE Trans. Energy Con-
Trans. Power Syst., vol. 9, no. 1, pp. 262–271, Feb. 1994. vers., vol. 15, no. 2, pp. 181–187, Jun. 2000.
[10] M. R. Harris, P. J. Lawrenson, and J. M. Stephenson, Per-Unit Sys- [35] S. Pillutla and A. Keyhani, “Neural network based modeling of round
tems With Special Reference to Electrical Machines. Cambridge, U.K.: rotor synchronous generator rotor body parameters from operating data,”
Cambridge Univ. Press, 1970. IEEE Trans. Energy Convers., vol. 14, no. 3, pp. 321–327, Sep. 1999.
[11] J. Verbeeck, R. Pintelon, and P. Guillaume, “Determination of syn- [36] H. B. Karayaka, A. Keyhani, G. T. Heydt, B. L. Agrawal, and D. A. Selin,
chronous machine parameters using network synthesis techniques,” “Neural network based modeling of a large steam turbine-generator rotor
IEEE Trans. Energy Convers., vol. 14, no. 3, pp. 310–314, Sep. 1999. body parameters from on-line disturbance data,” IEEE Trans. Energy
[12] J. Verbeeck, R. Pintelon, and P. Lataire, “Relationships between param- Convers., vol. 16, no. 4, pp. 305–311, Dec. 2001.
eter sets of equivalent synchronous machine models,” IEEE Trans. En- [37] J. Verbeeck, R. Pintelon, and P. Lataire, “Influence of saturation on esti-
ergy Convers., vol. 14, no. 4, pp. 1075–1080, Dec. 1999. mated synchronous machine parameters in standstill frequency response
[13] I. Kamwa and P. Viarouge, “On equivalent circuit structures for empir- tests,” IEEE Trans. Energy Convers., vol. 15, no. 3, pp. 277–283, Sep.
ical modeling of turbine-generators,” IEEE Trans. Energy Convers., vol. 2000.
9, no. 3, pp. 579–592, Sep. 1994. [38] N. Dedene, R. Pintelon, and P. Lataire, “Estimation of a global syn-
[14] I. M. Canay, “Physical significance of sub-subtransient quantities in dy- chronous machine model using a multiple-input multiple-output esti-
namic behavior of synchronous machines,” Proc. Inst. Elect. Eng. B, vol. mator,” IEEE Trans. Energy Convers., vol. 18, no. 1, pp. 11–16, Mar.
135, no. 6, pp. 334–340, Nov. 1988. 2003.
[15] P. L. Dandeno and M. R. Iravani, “Third order turboalternator elec- [39] Test Procedures for Synchronous Machines, IEEE Std. 115, Dec. 1995.
trical stability models with applications to subsynchronous resonance [40] P. J. Antsaklis and A. N. Michel, Linear Systems. New York: McGraw-
studies,” IEEE Trans. Energy Convers., vol. 10, no. 1, pp. 78–86, Mar. Hill, 1997.
1995. [41] J. A. Melkebeek and J. L. Willems, “Reciprocity relations for the mutual
[16] S. D. Sudhoff, D. C. Aliprantis, B. T. Kuhn, and P. L. Chapman, “An inductances between orthogonal axis windings in saturated salient-pole
induction machine model for predicting inverter-machine interaction,” machines,” IEEE Trans. Ind. Appl., vol. 26, no. 1, pp. 107–114, Jan./Feb.
IEEE Trans. Energy Convers., vol. 17, no. 2, pp. 203–210, Jun. 2002. 1990.
[17] D. C. Aliprantis, S. D. Sudhoff, and B. T. Kuhn, “Experimental charac- [42] P. W. Sauer, “Constraints on saturation modeling in AC machines,” IEEE
terization procedure for a synchronous machine model with saturation Trans. Energy Convers., vol. 7, no. 1, pp. 161–167, Mar. 1992.
and arbitrary rotor network representation,” IEEE Trans. Energy Con- [43] G. B. Thomas, Calculus and Analytic Geometry, 4th ed. Reading, MA:
vers., to be published. Addison-Wesley, 1968.
[18] L. Salvatore and M. Savino, “Experimental determination of syn- [44] S. D. Sudhoff, K. A. Corzine, H. J. Hegner, and D. E. Delisle, “Transient
chronous machine parameters,” Proc. Inst. Elect. Eng., B, vol. 128, no. and dynamic average-value modeling of synchronous machine fed load-
4, pp. 212–218, Jul. 1981. commutated converters,” IEEE Trans. Energy Convers., vol. 11, no. 3,
[19] I. M. Canay, “Modeling of alternating-current machines having multiple pp. 508–514, Sep. 1996.
rotor circuits,” IEEE Trans. Energy Convers., vol. 8, no. 2, pp. 280–296, [45] D. C. Aliprantis, “Advances in electric machine modeling and evolu-
Jun. 1993. tionary parameter identification,” Ph.D. dissertation, Purdue University,
[20] The MathWorks, Inc., Simulink Reference, Natick, MA, 2002. West Lafayette, IN, Dec. 2003.
[21] AEgis Technologies Group, Inc., Advanced Continuous Simulation Lan- [46] D. C. Aliprantis, S. D. Sudhoff, and B. T. Kuhn, “A brushless exciter
guage (ACSL) Reference Manual, Huntsville, AL, 1999. model incorporating multiple rectifier modes and Preisach’s hysteresis
[22] R. G. Harley, D. J. N. Limebeer, and E. Chirricozzi, “Comparative theory,” IEEE Trans. Energy Convers., to be published.
study of saturation methods in synchronous machine models,” Proc.
Inst. Elect. Eng. B, vol. 127, no. 1, pp. 1–7, Jan. 1980.
[23] G. R. Slemon, “Analytical models for saturated synchronous ma-
chines,” IEEE Trans. Power App. Syst., vol. PAS-90, no. 2, pp. 409–417,
Mar./Apr. 1971.
Dionysios C. Aliprantis (M’04) received the elec-
[24] J. E. Brown, K. P. Kovács, and P. Vas, “A method of including the effects
trical and computer engineering diploma from the
of main flux path saturation in the generalized equations of A.C. ma-
National Technical University of Athens, Athens,
chines,” IEEE Trans. Power App. Syst., vol. PAS-102, no. 1, pp. 96–103,
Greece, in 1999. He received the Ph.D. degree in
Jan. 1983.
electrical and computer engineering from Purdue
[25] R. S. Ramshaw and G. Xie, “Nonlinear model of nonsalient syn-
University, West Lafayette, IN, in 2003.
chronous machines,” IEEE Trans. Power App. Syst., vol. PAS-103, no.
7, pp. 1809–1815, Jul. 1984. Currently, he is serving in the armed forces of
[26] G. Xie and R. S. Ramshaw, “Nonlinear model of synchronous machines Greece. His interests include the modeling and
with saliency,” IEEE Trans. Energy Convers., vol. 1, no. 3, pp. 198–204, simulation of electric machines and power systems,
Sep. 1986. and evolutionary optimization methods.

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.
594 IEEE TRANSACTIONS ON ENERGY CONVERSION, VOL. 20, NO. 3, SEPTEMBER 2005

Scott D. Sudhoff (SM’01) received the B.S. (Hons.), Brian T. Kuhn (M’93) received the B.S. and M.S.
M.S., and Ph.D. degrees in electrical engineering degrees in electrical engineering from the University
from Purdue University, West Lafayette, IN, in 1988, of Missouri-Rolla in 1996 and 1997, respectively.
1989, and 1991, respectively. He was a Research Engineer at Purdue University,
Currently, he is a Full Professor at Purdue Univer- West Lafayette, IN, from 1998 to 2003. Currently, he
sity. From 1991 to 1993, he was Part-Time Visiting is a Senior Engineer with SmartSpark Energy Sys-
Faculty with Purdue University and as a Part-Time tems, Inc., Champaign, IL. His research interests in-
Consultant with P. C. Krause and Associates, West clude power electronics and electrical machinery.
Lafayette, IN. From 1993 to 1997, he was a Faculty
Member at the University of Missouri-Rolla. He has
authored many papers. His interests include electric
machines, power electronics, and finite-inertia power systems.

Authorized licensed use limited to: Purdue University. Downloaded on August 7, 2009 at 08:06 from IEEE Xplore. Restrictions apply.

You might also like