24

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 92

Massachusetts Institute of Technology Department of Electrical Engineering and Computer Science

6.685 Electric Machinery Class Notes 9: Synchronous Machine Simulation Models c 2005 James L.
Kirtley Jr. 1 Introduction In this document we develop models useful for calculating the dynamic
behavior of synchronous machines. We start with a commonly accepted picture of the synchronous
machine, assuming that the rotor can be fairly represented by three equivalent windings: one being
the field and the other two, the d- and q- axis “damper” windings, representing the effects of rotor
body, wedge chain, amortisseur and other current carrying paths. While a synchronous machine is
assumed here, the results are fairly directly applicable to induction machines. Also, extension to
situations in which the rotor representation must have more than one extra equivalent winding per
axis should be straightforward. 2 Phase Variable Model To begin, assume that the synchronous
machine can be properly represented by six equivalent windings. Four of these, the three armature
phase windings and the field winding, really are windings. The other two, representing the effects of
distributed currents on the rotor, are referred to as the “damper” windings. Fluxes are, in terms of
currents: " λph # " L M I = ph λR MT LR # " ph IR # (1) where phase and rotor fluxes (and, similarly,
currents) are: λph = λa λb λc (2) λR = λf λkd (3) λkq There are three inductance sub-
matrices. The first o f these describes armature winding induc tances: c L = La Lab La Lab Lb
Lbc ) ph Lac bc Lc (4 L 1 where, for a machine that may have some saliency: La = La0 + L2 cos 2θ
(5) 2π Lb = La0 + L2 cos 2(θ − ) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3
Lbc = Lab0 + L2 cos 2θ (9) π Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions,
we have assumed a particular form for the mutual in ductances. This is seemingly restrictive,
because it constrains the form of phase- to- phase mutual inductance variations with rotor position.
The coefficient L2 is actually the same in all six of these last expressions. As it turns out, this
assumption does not really restrict the accuracy of the model very much. We will have more to say
about this a bit later. The rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0
0 0 Lkq (11) And the stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ
M = M cos(θ − 2π 3 ) Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 )
−Lakq sin(θ + 2π 3 ) (12) 3 Park’s Equations The first step in the development of a suitable model
is to transform the armature winding variables to a coordinate system in which the rotor is
stationary. We identify equivalent armature windings in the direct and quadrature axes. The direct
axis armature winding is the equivalent of one of the phase windings, but aligned directly with the
field. The quadrature winding is situated so that its axis leads the field winding by 90 electrical
degrees. The transformation used to map the armature currents, fluxes and so forth onto the direct
and quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms.

Massachusetts Institute of Technology Department of Electrical Engineering and Computer Science


6.685 Electric Machinery Class Notes 9: Synchronous Machine Simulation Models c 2005 James L.
Kirtley Jr. 1 Introduction In this document we develop models useful for calculating the dynamic
behavior of synchronous machines. We start with a commonly accepted picture of the synchronous
machine, assuming that the rotor can be fairly represented by three equivalent windings: one being
the field and the other two, the d- and q- axis “damper” windings, representing the effects of rotor
body, wedge chain, amortisseur and other current carrying paths. While a synchronous machine is
assumed here, the results are fairly directly applicable to induction machines. Also, extension to
situations in which the rotor representation must have more than one extra equivalent winding per
axis should be straightforward. 2 Phase Variable Model To begin, assume that the synchronous
machine can be properly represented by six equivalent windings. Four of these, the three armature
phase windings and the field winding, really are windings. The other two, representing the effects of
distributed currents on the rotor, are referred to as the “damper” windings. Fluxes are, in terms of
currents: " λph # " L M I = ph λR MT LR # " ph IR # (1) where phase and rotor fluxes (and, similarly,
currents) are: λph = λa λb λc (2) λR = λf λkd (3) λkq There are three inductance sub-
matrices. The first o f these describes armature winding induc tances: c L = La Lab La Lab Lb
Lbc ) ph Lac bc Lc (4 L 1 where, for a machine that may have some saliency: La = La0 + L2 cos 2θ
(5) 2π Lb = La0 + L2 cos 2(θ − ) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3
Lbc = Lab0 + L2 cos 2θ (9) π Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions,
we have assumed a particular form for the mutual in ductances. This is seemingly restrictive,
because it constrains the form of phase- to- phase mutual inductance variations with rotor position.
The coefficient L2 is actually the same in all six of these last expressions. As it turns out, this
assumption does not really restrict the accuracy of the model very much. We will have more to say
about this a bit later. The rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0
0 0 Lkq (11) And the stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ
M = M cos(θ − 2π 3 ) Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 )
−Lakq sin(θ + 2π 3 ) (12) 3 Park’s Equations The first step in the development of a suitable model
is to transform the armature winding variables to a coordinate system in which the rotor is
stationary. We identify equivalent armature windings in the direct and quadrature axes. The direct
axis armature winding is the equivalent of one of the phase windings, but aligned directly with the
field. The quadrature winding is situated so that its axis leads the field winding by 90 electrical
degrees. The transformation used to map the armature currents, fluxes and so forth onto the direct
and quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machinery Class Notes 9: Synchronous Machine
Simulation Models c 2005 James L. Kirtley Jr. 1 Introduction In this document we develop models
useful for calculating the dynamic behavior of synchronous machines. We start with a commonly
accepted picture of the synchronous machine, assuming that the rotor can be fairly represented by
three equivalent windings: one being the field and the other two, the d- and q- axis “damper”
windings, representing the effects of rotor body, wedge chain, amortisseur and other current
carrying paths. While a synchronous machine is assumed here, the results are fairly directly
applicable to induction machines. Also, extension to situations in which the rotor representation
must have more than one extra equivalent winding per axis should be straightforward. 2 Phase
Variable Model To begin, assume that the synchronous machine can be properly represented by six
equivalent windings. Four of these, the three armature phase windings and the field winding, really
are windings. The other two, representing the effects of distributed currents on the rotor, are
referred to as the “damper” windings. Fluxes are, in terms of currents: " λph # " L M I = ph λR MT LR
# " ph IR # (1) where phase and rotor fluxes (and, similarly, currents) are: λph = λa λb λc
(2) λR = λf λkd (3) λkq There are three inductance sub- matrices. The first o f these
describes armature winding induc tances: c L = La Lab La Lab Lb Lbc ) ph Lac bc Lc (4 L 1
where, for a machine that may have some saliency: La = La0 + L2 cos 2θ (5) 2π Lb = La0 + L2 cos 2(θ −
) (6) 3 2π Lc = La0 + L2 cos 2(θ + ) (7) 3 π Lab = Lab0 + L2 cos 2(θ − ) (8) 3 Lbc = Lab0 + L2 cos 2θ (9) π
Lac = Lab0 + L2 cos 2(θ + ) (10) 3 Note that, in this last set of expressions, we have assumed a
particular form for the mutual in ductances. This is seemingly restrictive, because it constrains the
form of phase- to- phase mutual inductance variations with rotor position. The coefficient L2 is
actually the same in all six of these last expressions. As it turns out, this assumption does not really
restrict the accuracy of the model very much. We will have more to say about this a bit later. The
rotor inductances are relatively simply stated: Lf Lfkd 0 L = R Lfkd Lkd 0 0 0 Lkq (11) And the
stator- to- rotor mutual indu ctances are: M cos θ Lakd cos θ −Lakq sin θ M = M cos(θ − 2π 3 )
Lakd cos(θ − 2π 3 ) −Lakq sin(θ − 2π 3 ) M cos(θ + 2π 3 ) Lakd cos(θ + 2π 3 ) −Lakq sin(θ + 2π 3 )
(12) 3 Park’s Equations The first step in the development of a suitable model is to transform the
armature winding variables to a coordinate system in which the rotor is stationary. We identify
equivalent armature windings in the direct and quadrature axes. The direct axis armature winding is
the equivalent of one of the phase windings, but aligned directly with the field. The quadrature
winding is situated so that its axis leads the field winding by 90 electrical degrees. The
transformation used to map the armature currents, fluxes and so forth onto the direct and
quadrature axes is the celebrated Park’s Transformation, named after Robert H. Park, an early
investigator into transient behavior in synchronous machines. The mapping takes the form: ud a q
u u = udq = Tuph = T 1 u ub ( 3) 0 uc Where the transformation and it s inve rse are:
cos θ cos(θ − 2π 3 ) cos(θ + 2π 2 3 ) T = − sin θ − sin(θ − 2π ) 3 3 − sin(θ + 2π 3 ) 1 1 1 2 2 2
(14) 2 1 T cos θ − sin θ −1 = cos(θ − 2π 3 ) − sin(θ − 2π 3 ) 1 (15) cos(θ + 2π ) − sin(θ + 2π 3 3 ) 1
This transformation maps balanc ed sets of phase currents into cons tant currents in the d-q
frame. That is, if rotor angle is θ = ωt + θ0, and phase currents are: Ia = I cos ωt 2π Ib = I cos(ωt − ) 3
2π Ic = I cos(ωt + ) 3 Then the transformed set of currents is: Id = I cos θ0 Iq = −I sin θ0 Now, we apply
this transformation to (1) to express fluxes and currents in the armature in the d-q reference frame.
To do this, extract the top line in (1): λph = L Iph + MIR (16) ph The transformed flux is obtained by
premultiplying this whole expression by the transformation matrix. Phase current may be obtained
from d-q current by multiplying by the inverse of the transformation matrix. Thus: λ − d = TL T 1 q Idq
+ TMIR (17) ph The same process carried out for the lower line of (1) yields: λ T R = M T −1 Idq + L IR
(18) R Thus the fully transformed version of (1) is: " λdq λR # L = " dq LC 3 2 L T L C R # " Idq IR # (19)
If the conditions of (5) through (10) are satisfied, the inductance submatrices of (19) wind up being
of particularly simple form. (Please note that a substantial amount of algebra has been left out here!)
L L = dq d 0 0 0 Lq 0 (20) 0 0 L0 L =C M Lakd 0 0 0 Lakq 0 0 0 (21) 3 Note that
(19) through (21) express three separate sets of apparently independent flux/current relationships.
These may be re-cast into the following form: λd Ld Lakd M Id λ d k = 3 2 Lakd Lkd
Lfkd 3 Ikd (22) λf 2M Lfkd Lf If " λq # = " Lq Lakq Iq 3 (23) λkq 2 Lakq Lkq # " Ikq # λ0 =
L0I0 (24) Where the component inductances are: 3 Ld = La0 − Lab0 + L2 (25) 2 3 Lq = La0 − Lab0 − L2
(26) 2 L0 = La0 + 2Lab0 (27) Note that the apparently restrictive assumptions embedded in (5)
through (10) have resulted in the very simple form of (21) through (24). In particular, we have three
mutually independent sets of fluxes and currents. While we may be concerned about the
restrictiveness of these expressions, note that the orthogonality between the d- and q- axes is not
unreasonable. In fact, because these axes are orthogonal in space, it seems reasonable that they
should not have mutual flux linkages. The principal consequence of these assumptions is the de-
coupling of the zero-sequence component of flux from the d- and q- axis components. We are not in
a position at this time to determine the reasonableness of this. However, it should be noted that
departures from this form (that is, coupling between the “direct” and “zero” axes) must be through
higher harmonic fields that will not couple well to the armature, so that any such coupling will be
weak. Next, armature voltage is, ignoring resistance, given by: d d V −1 ph = λph = T λ dt dt dq (28)
and that the transformed armature voltage must be: V dq = TV ph d = T (T −1λ dt dq) d d = λdq + (T T
−1 )λdq (29) dt dt A good deal of manupulation goes into reducing the second term of this, resulting
in: d T T −1 = dt 0 − dθ 0 dt ( dθ dt 0 0 0 0 0 30) 4 This expresses the speed voltage that
arises from a coordinate transformation. The two voltage/flux relationships that are affected are: dλd
Vd = − ωλq (31) dt dλq Vq = + ωλd (32) dt where we have used dθ ω = (33) dt 4 Power and Torque
Instantaneous power is given by: P = VaIa + VbIb + VcIc (34) Using the transformations given above,
this can be shown to be: 3 3 P = VdId + VqIq + 3V0I0 (35) 2 2 which, in turn, is: 3 3 dλd dλq dλ0 P = ω
(λdIq − λqId) + ( Id + Iq) + 3 I0 (36) 2 2 dt dt dt Then, noting that electrical speed ω and shaft speed Ω
are related by ω = pΩ and that (36) describes electrical terminal power as the sum of shaft power
and rate of change of stored energy, we may deduce that torque is given by: 3 T = p(λdIq − λqId) (37)
2 5 Per-Unit Normalization The next thing for us to do is to investigate the way in which electric
machine system are nor malized, or put into what is called a per-unit system. The reason for this
step is that, when the voltage, current, power and impedance are referred to normal operating
parameters, the behavior characteristics of all types of machines become quite similar, giving us a
better way of relating how a particular machine works to some reasonable standard. There are also
numerical reasons for normalizing performance parameters to some standard. The first step in
normalization is to establish a set of base quantities. We will be normalizing voltage, current, flux,
power, impedance and torque, so we will need base quantities for each of these. Note, however, that
the base quantities are not independent. In fact, for the armature, we need only specify three
quantities: voltage (VB), current (IB) and frequency (ω0). Note that we do not normalize time nor
frequency. Having done this for the armature circuits, we can derive each of the other base
quantities: 5 • Base Power 3 PB = VBIB 2 • Base Impedance VB ZB = IB • Base Flux VB λB = ω0 • Base
Torque p TB = PB ω0 Note that, for our purposes, base voltage and current are expressed as peak
quantities. Base voltage is taken on a phase basis (line to neutral for a “wye” connected machine),
and base current is similarly taken on a phase basis, (line current for a “wye” connected machine).
Normalized, or per-unit quantities are derived by dividing the ordinary variable (with units) by the
corresponding base. For example, per-unit flux is: λ ω0λ ψ = = (38) λB VB In this derivation, per- unit
quantities will usually be designated by lower case letters. Two notable exceptions are flux, where we
use the letter ψ, and torque, where we will still use the upper case T and risk confusion. Now, we
note that there will be base quantities for voltage, current and frequency for each of the different
coils represented in our model. While it is reasonable to expect that the frequency base will be the
same for all coils in a problem, the voltage and current bases may be different. We might write (22)
as: ω0I L ω0I ω I dB kB V L 0 fB ψ V d akd d db db Vdb M id ω I ψ = ω0IdB 3L ω0IkB kd 2
akd L 0 fB V kd Lfkd kb Vkb V ikd ) kdb (39 ψ f ω0IdB 3 ω0IkB ω I0 fB if V 2M fkd Lf fb V L
fb Vfb where i = I/IB denotes per-unit, or normalized current. Note that (39) may be
written in simple form: ψd xd xakd xad id ψkd = d xakd xkd xfk ikd (40) ψf xad
xfkd xf if It is important to note that (40) assumes reciprocity in the normalized system. To
wit, the following expressions are implied: IdB xd = ω0 Ld (41) VdB 6 IkB xkd = ω0 Lkd (42) VkB IfB xf =
ω0 Lf (43) VfB IkB xakd = ω0 Lakd VdB 3 IdB = ω0 Lakd (44) 2 VkB IfB xad = ω0 M VdB 3 IdB = ω0 M
(45) 2 VfB IkB xfkd = ω0 Lfkd Vfb IfB = ω0 Lfkd (46) Vkb These in turn imply: 3 VdBIdB = VfBIfB (47) 2 3
VdBIdB = VkBIkB (48) 2 VfBIfB = VkBIkB (49) These expressions imply the same power base on all of
the windings of the machine. This is so because the armature base quantities Vdb and Idb are stated
as peak values, while the rotor base quantities are stated as DC values. Thus power base for the
three- phase armature is 3 2 times the product of peak quantities, while the power base for the rotor
is simply the product of those quantities. The quadrature axis, which may have fewer equivalent
elements than the direct axis and which may have different numerical values, still yields a similar
structure. Without going through the details, we can see that the per-unit flux/current relationship
for the q- axis is: " ψq ψkq # x = " q xakq iq (50) xakq xkq # " ikq # The voltage equations, including
speed voltage terms, (31) and (32), may be augmented to reflect armature resistance: dλd Vd = −
ωλq + RaId (51) dt dλq Vq = ωλd + + RaIq (52) dt The per-unit equivalents of these are: 1 dψd ω vd =
− ψq + raid (53) ω0 dt ω0 7 ω 1 dψq vq = ψd + + raiq (54) ω0 ω0 dt Where the per-unit armature
resistance is just ra = Ra ZB Note that none of the other circuits in this model have speed voltage
terms, so their voltage expressions are exactly what we might expect: 1 dψf vf = + rf if (55) ω0 dt 1
dψkd vkd = + rkdikd (56) ω0 dt 1 dψkq vkq = + rkqikq (57) ω0 dt 1 dψ0 v0 = + rai0 (58) ω0 dt It should
be noted that the damper winding circuits represent closed conducting paths on the rotor, so the
two voltages vkd and vkq are always zero. Per-unit torque is simply: Te = ψdiq − ψqid (59) Often, we
need to represent the dynamic behavior of the machine, including electromechanical dynamics
involving rotor inertia. If we note J as the rotational inertia constant of the machine system, the rotor
dynamics are described by the two ordinary differential equations: 1 dω J = T e + T m (60) p dt dδ = ω
− ω0 (61) dt where T e and T m represent electrical and mechanical torques in “ordinary” variables.
The angle δ represents rotor phase angle with respect to some synchronous reference. It is
customary to define an “inertia constant” which is not dimensionless but which nevertheless fits into
the per-unit system of analysis. This is: Rotational kinetic energy at rated speed H ≡ (62) Base Power
Or: 1 2 J ω0 p 2 Jω0 H = = (63) PB 2pTB Then the per-unit equivalent to (60) is: 2H dω = Te + Tm
(64) ω0 dt where now we use Te and Tm to represent per-unit torques. 8 6 Equal Mutual’s Base In
normalizing the differential equations that make up our model, we have used a number of base
quantities. For example, in deriving (40), the per-unit flux- current relationship for the direct axis, we
used six base quantities: VB, IB, VfB, IfB, VkB and IkB. Imposing reciprocity on (40) results in two
constraints on these six variables, expressed in (47) through (49). Presumably the two armature base
quantities will be fixed by machine rating. That leaves two more “degrees of freedom” in selection of
base quantities. Note that the selection of base quantities will affect the reactance matrix in (40).
While there are different schools of thought on just how to handle these degrees of freedom, a
commonly used convention is to employ what is called the equal mutuals base system. The two
degrees of freedom are used to set the field and damper base impedances so that all three mutual
inductances of (40) are equal: xakd = xfkd = xad (65) The direct- axis flux- current relationship
becomes: ψd xd xad xad id ψkd = xad xkd xad ik ψf xad xad xf d if
(66) 7 Equivalent Circuit id i r ✲ f f l a xal x rf ✛ ∧ ∧ ∧ ∧ ∧ ∧ ∨ ∨ ∩∩∩∩ ∩∩∩∩ ∨ ∨ ⊃ + + ⊃x + ⊃ kdl
⊃ ⊃ (ω0v ⊃ d + ωψq) ψ xad v d ⊃ f ⊃ <> - < r - > kd - < Figure 1: D- Axis Equivalent Circuit The flux-
current relationship of (66) is represented by the equivalent circuit of Figure 1, if the “leakage”
inductances are defined to be: xal = xd − xad (67) xkdl = xkd − xad (68) xf l = xf − xad (69) 9 Many of
the interesting features of the electrical dynamics of the synchronous machine may be discerned
from this circuit. While a complete explication of this thing is beyond the scope of this note, it is
possible to make a few observations. The apparent inductance measured from the terminals of this
equivalent circuit (ignoring resis tance ra) will, in the frequency domain, be of the form: ψd(s) Pn(s)
x(s) = = xd (70) id(s) Pd(s) Both the numerator and denominator polynomials in s will be second
order. (You may convince yourself of this by writing an expression for terminal impedance). Since this
is a “diffusion” type circuit, having only resistances and inductances, all poles and zeros must be on
the negative real axis of the “s-plane”. The per-unit inductance is, then: (1 + T ′ d s)(1 + T ′′ x(s) = x d s)
d ) (1 + ′ ′′ (71 Tdos)(1 + Tdos) The two time constants T ′ d and T ′′ d are the reciprocals of the zeros
of the impedance, which are the poles of the admittance. These are called the short circuit time
constants. The other two time constants T ′ do and T ′′ do are the reciprocals of the poles of the
impedance, and so are called the open circuit time constants. We have cast this thing as if there are
two sets of well- defined time constants. These are the transient time constants T ′ d and T ′ do, and
the subtransient time constants T ′′ d and T ′′ do. In many cases, these are indeed well separated,
meaning that: T ′ d ≫ T ′′ d (72) T ′ do ≫ T ′′ do (73) If this is true, then the reactance is described by
the pole-zero diagram shown in Figure 2. Under this circumstance, the apparent terminal inductance
has three distinct values, depending on frequency. These are the synchronous inductance, the
transient inductance, and the subtransient inductance, given by: ′ T ′ xd = x d d (74) T ′ do ′ x ′′ = x ′ T ′
d d d T ′′ do T ′ T ′′ = x d d d T ′ do T ′′ (75) do A Bode Plot of the terminal reactance is shown in Figure
3. If the time constants are spread widely apart, they are given, approximately, by: T ′ xf do = (76)
ω0rf ′′ xkdl + xf l||xad Tdo = (77) ω0rkd 10 ❝ ❝ × × 1 Td” 1 T ′ d 1 Tdo” 1 T ′ do Figure 2: Pole-Zero
Diagram For Terminal Inductance ❅ ❅❅ ❅ ❅❅ 1 T ′ do 1 T ′ d 1 Tdo” 1 log |x(jω)| log ω Td” Figure
3: Frequency Response of Terminal Inductance Finally, note that the three reactances are found
simply from the model: xd = xal + xad (78) x ′ d = xal + xad||xf l (79) x ′′ d = xal + xad||xf l||xkdl (80) 8
Statement of Simulation Model Now we can write down the simulation model. Actually, we will
derive more than one of these, since the machine can be driven by either voltages or currents.
Further, the expressions for permanent magnet machines are a bit different. So the first model is one
in which the terminals are all constrained by voltage. The state variables are the two stator fluxes ψd,
ψq, two “damper” fluxes ψkd, ψkq, field flux ψf , and rotor speed ω and torque angle δ. The most
straightforward way of stating the model employs currents as auxiliary variables, and these are: −1
id xd xad xad ψd i = x x x ψ kd if ad kd ad xad ad xf kd x ψf (81) " iq ikq
# = " −1 xq xaq ψq (82) xaq xkq # " ψkq # 11 Then the state equations are: dψd = ω0vd + ωψq −
ω0raid (83) dt dψq = ω0vq − ωψd − ω0raiq (84) dt dψkd = −ω0rkdikd (85) dt dψkq = −ω0rkqikq (86)
dt dψf = ω0vf − ω0rf if (87) dt dω ω0 = (Te + Tm) (88) dt 2H dδ = ω − ω0 (89) dt and, of course, Te =
ψdiq − ψqid 8.1 Statement of Parameters: Note that often data for a machine may be given in terms
of the reactances xd, x ′ d , x ′′ d , T ′ do and T ′′ do, rather than the elements of the equivalent circuit
model. Note that there are four inductances in the equivalent circuit so we have to assume one.
There is no loss in generality in doing so. Usually one assumes a value for the stator leakage
inductance, and if this is done the translation is straightforward: xad = xd − xal x ′ ad(x − xal) x = d f l
xad − x ′ d + xal 1 xkdl = 1 x ′′− − 1 − 1 d xal xad xfl xf l + xad rf = ω0T ′ do xkdl + xad||xf l rkd = ω ′′
0Tdo 8.2 Linearized Model Often it becomes desirable to carry out a linearized analysis of machine
operation to, for example, examine the damping of the swing mode at a particular operating point.
What is done, then, is to assume a steady state operating point and examine the dynamics for
deviations from that operating point that are “small”. The definition of “small” is really “small
enough” that everything important appears in the first-order term of a Taylor series about the steady
operating point. Note that the expressions in the machine model are, for the most part, linear. There
are, however, a few cases in which products of state variables cause us to do the expansion of the 12
Taylor series. Assuming a steady state operating point [ψd0 ψkd0 ψf0 ψq0 ψkq0 ω0 δ0], the
first order (small-signal) variations are described by the following set of equations. First, since the
flux-current relationship is linear: id1 −1 xd xad xad ψd1 ik d1 = x ad xkd xad ψ
(90) if1 xad xad xf −1 kd1 ψf1 " iq1 # x = " xq aq ψ ikq1 xaq xkq # " q1 ) ψkq # (91 1
Terminal voltage will be, for operation against a voltage source: Vd = V sin δ Vq = V cos δ Then the
differential equations governing the first-order variations are: dψd1 = ω0V cos δ0δ1 + ω0ψq1 +
ω1ψq0 − ω0raid1 (92) dt dψq1 = −ω0V sin δ0δ1 − ω0ψd1 − ω1ψd0 − ω0raiq1 (93) dt dψkd1 =
−ω0rkdikd1 (94) dt dψkq1 = −ω0rkqikq1 (95) dt dψf1 = −ω0rf if1 (96) dt dω1 ω0 = (Te1 + Tm1) (97) dt
2H dδ1 = ω1 (98) dt Te = ψd0iq1 + ψd1iq0 − ψq0id1 − ψq1id0 8.3 Reduced Order Model for
Electromechanical Transients In many situations the two armature variables contribute little to the
dynamic response of the machine. Typically the armature resistance is small enough that there is
very little voltage drop across it and transients in the difference between armature flux and the flux
that would exist in the “steady state” decay rapidly (or are not even excited). Further, the relatively
short armature time constant makes for very short time steps. For this reason it is often convenient,
particularly when studying the relatively slow electromechanical transients, to omit the first two
differential equations and set: ψd = vq = V cos δ (99) ψq = −vd = −V sin δ (100) The set of differential
equations changes only a little when this approximation is made. Note, however, that it can be
simulated with far fewer “cycles” if the armature time constant is short. 13 9 Current Driven Model:
Connection to a System The simulation expressions developed so far are useful in a variety of
circumstances. They are, however, difficult to tie to network simulation programs because they use
terminal voltage as an input. Generally, it is more convenient to use current as the input to the
machine simulation and accept voltage as the output. Further, it is difficult to handle unbalanced
situations with this set of equations. An alternative to this set would be to employ the phase currents
as state variables. Effectively, this replaces ψd, ψq and ψ0 with ia, ib, and ic. The resulting model will,
as we will show, interface nicely with network simulations. To start, note that we could write an
expression for terminal flux, on the d- axis: ψd = x ′′ xad||xkdl xad||xf l d id + ψf + ψkd (101)
xad||xkdl + xf l xad||xf l + xkdl and here, of course, x ′′ d = xal + xad||xkdl||xf l This leads us to
define a “flux behind subtransient reactance”: ′′ xadxkdlψf + xadxf lψkd ψd = (102) xadxkdl + xadxf l +
xkdlxf l So that ψd = ψ ′′ d + x ′′ d id On the quadrature axis the situation is essentially the same, but
one step easier if there is only one quadrature axis rotor winding: xaq ψq = x ′′ q iq + ψkq (103) xaq +
xkql where x ′′ q = xal + xaq||xkql Very often these fluxes are referred to as “voltage behind
subtransient reactance, with ψ ′′ ′′ d = eq and ψ ′′ q = −e ′′ d . Then: ψd = x ′′ d id + e ′′ q (104) ψq = x ′′
q iq − e ′′ d (105) Now, if id and iq are determined, it is a bit easier to find the other currents required
in the simulation. Note we can write: " ψkd # = " xkd xad # " ikd # x + a f " d ) f xad xf i xad # id (106 ψ
and this inverts easily: " −1 ikd # = " xkd xad # " ψkd # x − a x " d ) if xad f ψf x d # id (107 a ! 14 The
quadrature axis rotor current is simply: 1 xaq ikq = ψkq − iq (108) xkq xkq The torque equation is the
same, but since it is usually convenient to assemble the fluxes behind subtransient reactance, it is
possible to use: Te = e ′′ q iq + e ′′ d id + (x ′′ d − x ′′ q )idiq (109) Now it is necessary to consider
terminal voltage. This is most conveniently cast in matrix notation. The vector of phase voltages is:
vph = va vb vc (110) Then, with similar notation for phase flux, te rmina l voltage is,
ignoring armature resistance: 1 dψph vph = ω0 dt 1 d = ω0 dt n T −1ψdqo (111) Note that we may
define the transformed vector of fluxes to be: ψ = x ′′idq + e ′′ (112) dq where the matrix of
reactances shows orthogonality: x ′′ d 0 0 x ′′ = 0 x ′′ q 0 3 0 0 x (11 ) 0 and the vector of
internal fluxes is: e ′′ = e ′′ q −e ′′ d 0 (114) Now, of course, i = dq Tiph, so that we may
re-cast (111) as: 1 d vph = n T −1x ′′Ti ′′ h + T −1 p e o (115) ω0 dt Now it is necessary to make one
assumption and one definition. The assumption, which is only moderately restrictive, is that
subtransient saliency may be ignored. That is, we assume that x ′′ d = x ′′ q . The definition separates
the “zero sequence” impedance into phase and neutral components: 15 x0 = x ′′ d + 3xg (116) Note
that according to this definition the reactance xg accounts for any impedance in the neutral of the
synchronous machine as well as mutual coupling between phases. Then, the impedance matrix
becomes: x ′′ = x ′′ d 0 0 0 0 x ′′ d 0 + 0 0 0 0 0 (11 0 0 x ′′ d 0 0 3xg 7) In compact
notation, this is: x ′′ = x ′′ d I + x (118) g where I is the identity matrix. Now the vector of phase
voltages is: 1 d vph = ′ ω dt n x ′ − d i 1 −1 ′′ ph + T x Tiph + T e g 0 o (119) Note that in (119), we have
already factored out the multiplication by the identity matrix. The next step is to carry out the matrix
multiplication in the third term of (119). This operation turns out to produce a remarkably simple
result: T −1x T = xg g 1 1 1 1 1 1 1 1 1 (120) The impact of this is that each of the three phase
volta ges has the same term, and that is related to the time derivative of the sum of the three
currents, multiplied by xg. The third and final term in (119) describes voltages induced by rotor
fluxes. It can be written as: 1 d 1 d 1 e ′′ T −1 d e ′′ = T −1 e ′′ + T −1 (121) ω0 dt ω0 dt ω0 dt Now, the
time derivative of n the inve o rse transfo n rm is o : − sin(θ) − cos(θ) 0 1 d T −1 ω = − sin(θ − 2π ) −
cos(θ − 2π 3 3 ) 0 (122) ω0 dt ω0 − sin(θ + 2π ) − cos(θ + 2π 3 3 ) 0 Now the three phase
voltages can be extracted from all of this matrix algebra: x ′′ d dia xg d va = + (i ′ a + i ′ b + ic) + e ω0
dt ω0 dt a (123) x ′′ d dib xg d vb = + (i ′′ a + ib + ic) + e ω0 dt ω0 dt b (124) x ′′ vc = d dic xg d + (ia + ib
+ ic) + e ′′ c (125) ω0 dt ω0 dt 16 Where the internal voltages are: e ′′ ω = − (e ′′ sin(θ) − e ′′ a ω q d
cos(θ)) 0 1 de′′ q 1 de′′ + cos(θ) + sin(θ) d (126) ω0 dt ω0 dt e ′′ ω 2 b = − (e ′′ 2π ω q sin(θ − ) − e ′′ π 3
d cos(θ − )) 0 3 1 2π de′′ 1 2π de′′ q + cos(θ − ) + sin(θ − ) d (127) ω0 3 dt ω0 3 dt e ′′ ω 2π π c = − ( ′′ 2
eq sin(θ + ) − e ′′ d cos(θ + )) ω0 3 3 1 2π de′′ 2π ′ q 1 de ′ + cos(θ + ) + sin(θ + ) d (128) ω0 3 dt ω0 3 dt
This set of expressions describes the equivalent circuit shown in Figure 4. e ′′ i✲a x ′′ a d va ∩∩∩∩ ✗
✖+ − ✔ ′ i ′ b ✲b x ′ e ′ ✕ d v ∩∩ ∩ ✔x ∩ ✗ g b − x ′ ✖+ ∩∩∩∩ e ′′ ′ c c ✕ ✲i d vc ∩∩∩∩ ✗+ − ✔ ✖✕
Figure 4: Equivalent Network Model 10 Restatement Of The Model The synchronous machine model
which uses the three phase currents as state variables may now be stated in the form of a set of
differential and algebraic equations: dψkd = −ω0rkdikd (129) dt dψkq = −ω0rkqikq (130) dt dψf =
−ω0rf if (131) dt dδ = ω − ω0 (132) dt 17 dω ω0 = dt 2H T + e ′′ ′ q iq + e ′ m d id (133) where: " − i
d # " 1 k x x = kd ad ψkd x − ad i i d f xad xf # " ψf # " xad # ! and 1 xaq ikq = ψkq − iq xkq xkq (It is
assumed here that the difference between subtransient reactances is small enough to be neglected.)
The network interface equations are, from the network to the machine: 2π 2π id = ia cos(θ) + ib cos(θ
− ) + ic cos(θ + ) (134) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (135) 3 3 and, in the reverse
direction, from the machine to the network: ′′ ω ea = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′ ′ q 1 de′
+ cos(θ) + sin(θ) d (136) ω0 dt ω0 dt e ′′ ω 2π b = − (e ′′ 2π q sin(θ − ) − e ′′ ω0 3 d cos(θ − )) 3 1 2π de′′
1 2 d ′′ q π e + cos(θ − ) + sin(θ − ) d (137) ω0 3 dt ω0 3 dt ω e ′′ π = − (e ′′ 2 sin(θ + ) − e ′′ 2π c ω q 0 3
d cos(θ + )) 3 1 2π de′′ 1 2π de′′ q + cos(θ + ) + sin(θ + ) d (138) ω0 3 dt ω0 3 dt And, of course, θ =
ω0t + δ (139) e ′′ q = ψ ′′ d (140) e ′′ d = −ψ ′′ q (141) ′′ xadxkdlψf + xadxf lψkd ψd = (142) xadxkdl +
xadxf l + xkdlxf l ψ ′′ xaq q = ψkq (143) xaq + xkql 18 11 Network Constraints This model may be
embedded in a number of networks. Different configurations will result in different constraints on
currents. Consider, for example, the situation in which all of the terminal voltages are constrained,
but perhaps by unbalanced (not entirely positive sequence) sources. In that case, the differential
equations for the three phase currents would be: x ′′ d di ′ a + 2 g = (va − e ′′ x ′ xg x a ) d − (v − e ′′ b )
+ (vc − e ′′) ω dt x ′′ + 3x b c x ′′ (144) 0 d g d + 3xg x ′′ d di ′ x ′′ b g = (v − e ′ ) d + 2x − (v − e ′′ xg b b ′′
a a ) + (vc − e ′′ ω0 dt xd + 3x c ) g (145) x ′′ d + 3xg x ′′ d di ′′ c = (v − e ′′ x ) d + 2xg − (v − e ′′) + (v − e
′′ c b xg c ′′ b a a ) ′′ (146) ω0 dt xd + 3xg xd + 3xg 12 Example: Line-Line Fault We are not, however,
constrained to situations defined in this way. This model is suitable for embedding into network
analysis routines. It is also possible to handle many different situations directly. Consider, for
example, the unbalanced fault represented by the network shown in Figure 5. This shows a line-line
fault situation, with one phase still connected to the network. ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d ∩∩∩∩ x ′′ d
✖✕ ✗✔+ − e ′′ c ✖✕ ✗✔+ − e ′′ b ✖✕ ✗✔+ − e ′′ a ∩∩∩∩ xg i✲a ✲ib va ∧ ∧ ∧ ∨ ∨ ra ∧ ∧ ∧ ∨ ∨ ra ∧
∧ ∧ ∨ ∨ ra Figure 5: Line-Line Fault Network Model In this situation, we have only two currents to
worry about, and their differential equations would be: dib ω0 = ′′(e ′′ − e ′′ − 2raib) (147) dt 2x c b d
dia ω0 = ′′ (v − ′′ a e − raia) (148) dt xd + x a g and, of course, ic = −ib. Note that here we have
included the effects of armature resistance, ignored in the previous section but obviously important
if the results are to be believed. 19 13 Permanent Magnet Machines Permanent Magnet machines
are one state variable simpler than their wound-field counterparts. They may be accurately viewed
as having constant field current. Assuming that we can define the internal (field) flux as: ψ0 = xadif0
(149) 13.1 Model: Voltage Driven Machine We have a reasonably simple expression for the rotor
currents, in the case of a voltage driven machine: " id x kd # = i " d xad xad xkd #−1 " ψd − ψ0 ψkd −
ψ0 # (150) " −1 iq # x = " q xaq ikq xaq xkq # " ψq ψkq # (151) The simulation model then has six
states: dψd = ω0vd + ωψq − ω0raid (152) dt dψq = ω0vq − ωψd − ω0raiq (153) dt dψkd = −ω0rkdikd
(154) dt dψkq = −ω0rkqikq (155) dt dω ω0 = (ψdiq − ψqid + Tm) (156) dt 2H dδ = ω − ω0 (157) dt
13.2 Curent-Driven Machine Model In the case of a current-driven machine, rotor currents required
in the simulation are: 1 ikd = (ψkd − xadid − ψ0) (158) xkd 1 ikq = (ψkq − xaqiq) (159) xkq Here, the
“flux behind subtransient reactance” is, on the direct axis: ′′ xkdlψ0 + xadψkd ψd = (160) xad + xkdl
and the subtransient reactance is: x ′′ d = xal + xad||xkdl (161) 20 On the quadrature axis, ′′ xadψkq
ψq = (162) xad + xkql and x ′′ q = xal + xaq||xkql (163) In this case there are only four state equations:
dψkd = −ω0rkdikd (164) dt dψkq = −ω0rkqikq (165) dt dω ω0 = e ′′ e ′′ q iq + d id + Tm (166) dt 2H
dδ = ω − ω0 (167) dt The interconnections to and from the network are the same as in the case of a
wound-field machine: in the “forward” direction, from network to machine: 2π 2π id = ia cos(θ) + ib
cos(θ − ) + ic cos(θ + ) (168) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − ) − ic sin(θ + ) (169) 3 3 and, in the
reverse direction, from the machine to the network: e ′′ ω a = − (e ′′ q sin(θ) − e ′′ ω d cos(θ)) 0 1 de′′
1 de′′ q + cos(θ) + sin(θ) d (170) ω0 dt ω0 dt e ′′ ω 2π 2π b = − (e ′′ q sin(θ − ) − e ′′ c ω d os(θ − )) 0 3 3
1 2π de′′ π ′′ q 1 2 de + cos(θ − ) + sin(θ − ) d (171) ω0 3 dt ω0 3 dt ′′ ω ′′ 2π ′ π ec = − (eq sin(θ + ) − e ′
2 ω 3 d cos(θ + )) 0 3 1 2π de′′ π de ′ q 1 2 ′ + cos(θ + ) + sin(θ + ) d (172) ω0 3 dt ω0 3 dt 13.3 PM
Machines with no damper PM machines without much rotor conductivity may often behave as if
they have no damper winding at all. In this case the model simplifies even further. Armature currents
are: 1 id = (ψd − ψ0) (173) xd 1 iq = ψq (174) xq 21 The state equations are: dψd = ω0vd + ωψq −
ω0raid (175) dt dψq = ω0vq − ωψd − ω0raiq (176) dt dω ω0 = (ψdiq − ψqid + Tm) (177) dt 2H dδ = ω
− ω0 (178) dt 13.4 Current Driven PM Machines with no damper In the case of no damper the
machine becomes quite simple. There is no “internal flux” on the quadrature axis. Further, there are
no time derivatives of the internal flux on the d- axis. The only machine state equations are
mechanical: dω ω0 = (ψ0iq + Tm) (179) dt 2H dδ = ω − ω0 (180) dt The “forward” network interface
is as before: 2π 2π id = ia cos(θ) + ib cos(θ − ) + ic cos(θ + ) (181) 3 3 2π 2π iq = −ia sin(θ) − ib sin(θ − )
− ic sin(θ + ) (182) 3 3 and, in the reverse direction, from the machine to the network, things are a bit
simpler than before: e ′′ ω a = − ψ0 sin(θ) (183) ω0 e ′′ ω 2π b = − ψ0 sin(θ − ) (184) ω0 3 e ′′ ω 2π c =
− ψ0 sin(θ + ) (185) ω0 3 (186) 22 MIT OpenCourseWare http://ocw.mit.edu 6.685 Electric Machines
Fall 2013 For information about citing these materials or our Terms of Use, visit:
http://ocw.mit.edu/terms. Massachusetts Institute of Technology Department of Electrical
Engineering and Computer Science 6.685 Electric Machines Class Notes 1: Electromagnetic Forces c
2003 James L. Kirtley Jr. 1 Introduction Stator Stator Conductors Rotor Rotor Conductors Bearings
Shaft End Windings Air Gap Figure 1: Form of Electric Machine This section of notes discusses some
of the fundamental processes involved in electric machinery. In the section on energy conversion
processes we examine the two major ways of estimating elec tromagnetic forces: those involving
thermodynamic arguments (conservation of energy) and field methods (Maxwell’s Stress Tensor).
But first it is appropriate to introduce the topic by describing a notional rotating electric machine.
Electric machinery comes in many different types and a strikingly broad range of sizes, from those
little machines that cause cell ’phones and pagers to vibrate (yes, those are rotating electric
machines) to turbine generators with ratings upwards of a Gigawatt. Most of the machines with
which we are familiar are rotating, but linear electric motors are widely used, from shuttle drives in
weaving machines to equipment handling and amusement park rides. Currently under development
are large linear induction machines to be used to launch aircraft. It is our purpose in this subject to
develop an analytical basis for understanding how all of these different machines work. We start,
however, with a picture of perhaps the most common of electric machines. 2 Electric Machine
Description: Figure 1 is a cartoon drawing of a conventional induction motor. This is a very common
type of electric machine and will serve as a reference point. Most other electric machines operate in
1 a fashion which is the same as the induction machine or which differ in ways which are easy to
reference to the induction machine. Most (but not all!) machines we will be studying have essentially
this morphology. The rotor of the machine is mounted on a shaft which is supported on some sort of
bearing(s). Usually, but not always, the rotor is inside. I have drawn a rotor which is round, but this
does not need to be the case. I have also indicated rotor conductors, but sometimes the rotor has
permanent magnets either fastened to it or inside, and sometimes (as in Variable Reluctance
Machines) it is just an oddly shaped piece of steel. The stator is, in this drawing, on the outside and
has windings. With most of the machines we will be dealing with, the stator winding is the armature,
or electrical power input element. (In DC and Universal motors this is reversed, with the armature
contained on the rotor: we will deal with these later). In most electrical machines the rotor and the
stator are made of highly magnetically permeable materials: steel or magnetic iron. In many common
machines such as induction motors the rotor and stator are both made up of thin sheets of silicon
steel. Punched into those sheets are slots which contain the rotor and stator conductors. Figure 2 is a
picture of part of an induction machine distorted so that the air-gap is straightened out (as if the
machine had infinite radius). This is actually a convenient way of drawing the machine and, we will
find, leads to useful methods of analysis. Stator Core Stator Conductors In Slots Rotor Conductors In
Slots Air Gap Figure 2: Windings in Slots What is important to note for now is that the machine has
an air gap g which is relatively small (that is, the gap dimension is much less than the machine radius
r). The air-gap also has a physical length l. The electric machine works by producing a shear stress in
the air-gap (with of course side effects such as production of “back voltage”). It is possible to define
the average air gap shear stress, which we will refer to as τ . Total developed torque is force over the
surface area times moment (which is rotor radius): T = 2πr2 ℓ < τ > Power transferred by this device
is just torque times speed, which is the same as force times 2 surface velocity, since surface velocity
is u = rΩ: Pm = ΩT = 2πrℓ < τ > u If we note that active rotor volume is , the ratio of torque to volume
is just: T = 2 < τ > Vr Now, determining what can be done in a volume of machine involves two things.
First, it is clear that the volume we have calculated here is not the whole machine volume, since it
does not include the stator. The actual estimate of total machine volume from the rotor volume is
actually quite complex and detailed and we will leave that one for later. Second, we need to estimate
the value of the useful average shear stress. Suppose both the radial flux density Br and the stator
surface current density Kz are sinusoidal flux waves of the form: Br = √ 2B0 cos (pθ − ωt) Kz = √ 2K0
cos (pθ − ωt) Note that this assumes these two quantities are exactly in phase, or oriented to ideally
produce torque, so we are going to get an “optimistic” bound here. Then the average value of
surface traction is: 1 2π < τ >= Z BrKzdθ = B0K0 2π 0 The magnetic flux density that can be developed
is limited by the characteristics of the magnetic materials (iron) used. Current densities are a function
of technology and are typically limited by how much effort can be put into cooling and the
temperature limits of insulating materials. In practice, the range of shear stress encountered in
electric machinery technology is not terribly broad: ranging from a few kPa in smaller machines to
about 100 kPa in very large, well cooled machines. It is usually said that electric machines are torque
producing devices, meaning tht they are defined by this shear stress mechanism and by physical
dimensions. Since power is torque times rotational speed, high power density machines necessarily
will have high shaft speeds. Of course there are limits on rotational speed as well, arising from
centrifugal forces which limit tip velocity. Our first step in understanding how electric machinery
works is to understand the mechanisms which produce forces of electromagnetic origin. 3 Energy
Conversion Process: In a motor the energy conversion process can be thought of in simple terms. In
“steady state”, electric power input to the machine is just the sum of electric power inputs to the
different phase terminals: Pe = Xviii i Mechanical power is torque times speed: Pm = TΩ 3 Mechanical
Electro Converter Electric Power In Mechanical Power Out Losses: Heat, Noise, Windage,... Figure 3:
Energy Conversion Process And the sum of the losses is the difference: Pd = Pe − Pm It will
sometimes be convenient to employ the fact that, in most machines, dissipation is small enough to
approximate mechanical power with electrical power. In fact, there are many situations in which the
loss mechanism is known well enough that it can be idealized away. The “thermodynamic”
arguments for force density take advantage of this and employ a “conservative” or lossless energy
conversion system. 3.1 Energy Approach to Electromagnetic Forces: f + Magnetic Field v x - System
Figure 4: Conservative Magnetic Field System To start, consider some electromechanical system
which has two sets of “terminals”, electrical and mechanical, as shown in Figure 4. If the system
stores energy in magnetic fields, the energy stored depends on the state of the system, defined by (in
this case) two of the identifiable variables: flux (λ), current (i) and mechanical position (x). In fact,
with only a little reflection, you should be able to convince yourself that this state is a single-valued
function of two variables and that the energy stored is independent of how the system was brought
to this state. Now, all electromechanical converters have loss mechanisms and so are not themselves
conser vative. However, the magnetic field system that produces force is, in principle, conservative
in the 4 sense that its state and stored energy can be described by only two variables. The “history”
of the system is not important. It is possible to chose the variables in such a way that electrical power
into this conservative system is: P e dλ = vi = i dt Similarly, mechanical power out of the system is: P
m = f e dx dt The difference between these two is the rate of change of energy stored in the system:
dWm = P e dt − P m It is then possible to compute the change in energy required to take the system
from one state to another by: a Wm(a) − Wm(b) = Z idλ b − f e dx where the two states of the system
are described by a = (λa, xa) and b = (λb, xb) If the energy stored in the system is described by two
state variables, λ and x, the total differential of stored energy is: ∂Wm ∂Wm dWm = dλ + dx ∂λ ∂x
and it is also: dWm = idλ − f e dx So that we can make a direct equivalence between the derivatives
and: ∂ f e W = − m ∂x In the case of rotary, as opposed to linear, motion, torque T e takes the place of
force f e and angular displacement θ takes the place of linear displacement x. Note that the product
of torque and angle has the same units as the product of force and distance (both have units of work,
which in the International System of units is Newton-meters or Joules. In many cases we might
consider a system which is electricaly linear, in which case inductance is a function only of the
mechanical position x. λ(x) = L(x)i In this case, assuming that the energy integral is carried out from λ
= 0 (so that the part of the integral carried out over x is zero), Wm = Z λ 1 1 λ 2 λdλ = 0 L(x) 2 L(x) This
makes f e 1 = − λ 2 ∂ 1 2 ∂x L(x) 5 Note that this is numerically equivalent to f e 1 = − i 2 ∂ L(x) 2 ∂x
This is true only in the case of a linear system. Note that substituting L(x)i = λ too early in the
derivation produces erroneous results: in the case of a linear system it produces a sign error, but in
the case of a nonlinear system it is just wrong. 3.1.1 Example: simple solenoid Consider the magnetic
actuator shown in cartoon form in Figure 5. The actuator consists of a circular rod of ferromagnetic
material (very highly permeable) that can move axially (the x direction) inside of a stationary piece,
also made of highly permeable material. A coil of N turns carries a current I. The rod has a radius R
and spacing from the flat end of the stator is the variable dimension x. At the other end there is a
radial clearance between the rod and the stator g. Assume g ≪ R. If the axial length of the radial gaps
is ℓ = R/2, the area of the radial gaps is the same as the area of the gap between the rod and the
stator at the variable gap. µ CL x R g N turns R/2 µ Figure 5: Solenoid Actuator The permeances of the
variable width gap is: µ P 0πR2 1 = x and the permeance of the radial clearance gap is, if the gap
dimension is small compared with the radius: 2µ0πRℓ µ π P 0 R2 2 = = g g The inductance of the coil
system is: N2 µ 2 2 1 2 0πR N L = = N 2 P P = R1 + R2 P2 + P2 x + g Magnetic energy is: 6 λ Wm = Z 0 1
λ 2 λ 2 idλ = = 0 x + g 2 L(x) 2 µ πR2N2 0 0 And then, of course, force of electric origin is: f e ∂W = − m
λ 2 = 0 d 1 ∂x − 2 dx L(x) Here that is easy to carry out: d 1 1 = dx L µ 2 2 0πR N So that the force is: f e
λ 2 (x) = − 0 1 2 µ 2 0πR N2 Given that the system is to be excited by a current, we may at this point
substitute for flux: µ0πR2N i λ = L(x)i = x + g and then total force may be seen to be: f e µ = − 0πR2N2
i 2 (x + g) 2 2 The force is ‘negative’ in the sense that it tends to reduce x, or to close the gap. 3.1.2
Multiply Excited Systems There may be (and in most electric machine applications there will be)
more than one source of electrical excitation (more than one coil). In such systems we may write the
conservation of energy expression as: dWm = Xikdλk k − f e dx which simply suggests that electrical
input to the magnetic field energy storage is the sum (in this case over the index k) of inputs from
each of the coils. To find the total energy stored in the system it is necessary to integrate over all of
the coils (which may and in general will have mutual inductance). Wm = Z i · dλ Of course, if the
system is conservative, Wm(λ1, λ2, . . . , x) is uniquely specified and so the actual path taken in
carrying out this integral will not affect the value of the resulting energy. 7 3.1.3 Coenergy We often
will describe systems in terms of inductance rather than its reciprocal, so that current, rather than
flux, appears to be the relevant variable. It is convenient to derive a new energy variable, which we
will call co-energy, by: W′ m = Xλiii − Wm i and in this case it is quite easy to show that the energy
differential is (for a single mechanical variable) simply: dW′ e m = Xλkdik + f dx k so that force
produced is: ∂W′ fe = m ∂x 3.2 Example: Synchronous Machine µ µ F F F’ F’ A’ A A’ B B’ B B’ C C’ C C’
A θ Rotor Stator Gap Figure 6: Cartoon of Synchronous Machine Consider a simple electric machine
as pictured in Figure 6 in which there is a single winding on a rotor (call it the field winding and a
polyphase armature with three identical coils spaced at uniform locations about the periphery. We
can describe the flux linkages as: λa = Laia + Labib + Labic + M cos(pθ)if 2π λb = Labia + Laib + Labic +
M cos(pθ − )if 3 2π λc = Labia + Labib + Laic + M cos(pθ + )if 3 2π 2π λf = M cos(pθ)ia + M cos(pθ − )ib
+ M cos(pθ + ) + Lf if 3 3 It is assumed that the flux linkages are sinusoidal functions of rotor position.
As it turns out, many electrical machines work best (by many criteria such as smoothness of torque
production) 8 if this is the case, so that techniques have been developed to make those flux linkages
very nearly sinusoidal. We will see some of these techniques in later chapters of these notes. For the
moment, we will simply assume these dependencies. In addition, we assume that the rotor is
magnetically ’round’, which means the stator self inductances and the stator phase to phase mutual
inductances are not functions of rotor position. Note that if the phase windings are identical (except
for their angular position), they will have identical self inductances. If there are three uniformly
spaced windings the phase-phase mutual inductances will all be the same. Now, this system can be
simply described in terms of coenergy. With multiple excitation it is important to exercise some care
in taking the coenergy integral (to ensure that it is taken over a valid path in the multi-dimensional
space). In our case there are actually five dimensions, but only four are important since we can
position the rotor with all currents at zero so there is no contribution to coenergy from setting rotor
position. Suppose the rotor is at some angle θ and that the four currents have values ia0, ib0, ic0 and
if0. One of many correct path integrals to take would be: W′ m = Z ia0 Laiadia 0 + Z ib0 (Labia0 + Laib)
dib 0 + Z ic0 (Labia0 + Labib0 + Laic) dic 0 + Z if0 2π 2π M cos(pθ)ia0 + M cos(pθ − )ib0 + M cos(pθ +
)ic0 + Lf if 0 3 3 dif The result is: W′ 1 = L i 2 + i 2 2 m a a0 b0 + ico + Lab (iaoib0 + iaoic0 + icoib0)
2 +M if0 2π 2π 1 ia0 cos(pθ) + i 2 b0 cos(pθ − ) + ic0 cos(pθ + ) 3 3 + Lf i 2 f0 Since there are no
variations of the stator inductances with rotor position θ, torque is easily given by: ∂W′ Te = m 2 π =
−pM if0 π 2 ia0 sin(pθ) + ib0 sin(pθ − ) + ico sin(pθ + ) ∂θ 3 3 3.2.1 Current Driven Synchronous
Machine Now assume that we can drive this thing with currents: ia0 = Ia cos ωt ib0 = Ia cos ω 2π t −
3 2π ic0 = Ia cos ω t + 3 if0 = If 9 and assume the rotor is turning at synchronous speed: pθ = ωt
+ δi Noting that cos x sin y = 1 sin(x − y) + 1 sin(x + y), we find the torque expression above to be: 2 2
1 1 Te = −pMIaIf sin δi + sin (2ωt + δi) 2 2 1 1 4 + sin δi + sin 2 2 π 2ωt + δi − 3 + 1 1 4π sin
δi + sin 2ωt + δi + 2 2 3 The sine functions on the left add and the ones on the right cancel,
leaving: 3 Te = − pMIaIf sin δi 2 And this is indeed one way of looking at a synchronous machine,
which produces steady torque if the rotor speed and currents all agree on frequency. Torque is
related to the current torque angle δi . As it turns out such machines are not generally run against
current sources, but we will take up actual operation of such machines later. 4 Field Descriptions:
Continuous Media While a basic understanding of electromechanical devices is possible using the
lumped parameter approach and the principle of virtual work as described in the previous section,
many phenomena in electric machines require a more detailed understanding which is afforded by a
continuum approach. In this section we consider a fields-based approach to energy flow using
Poynting’s Theorem and then a fields based description of forces using the Maxwell Stress Tensor.
These techniques will both be useful in further analysis of what happens in electric machines. 4.1
Field Description of Energy Flow: Poyting’s Theorem Start with Faraday’s Law: ∂B~ ∇ × E~ = − ∂t and
Ampere’s Law: ∇ × H~ ~ = J Multiplying the first of these by H~ ~ and the second by E and taking the
difference: ∂B~ H~ · ∇ × E~ − E~ · ∇ × H~ = ∇ · E~ ~ × H = −H~ · E J ~ ~ dt − · On the left of this
expression is the divergence of electromagnetic energy flow: S~ ~ = E × H~ 10 Here, S~ is the
celebrated Poynting flow which describes power in an electromagnetic field sysstem. (The units of
this quantity is watts per square meter in the International System). On ~ ~ the right hand side are
two terms: H · ∂B dt is rate of change of magnetic stored energy. The second ~ term, E · J~ looks a lot
like power dissipation. We will discuss each of these in more detail. For the moment, however, note
that the divergence theorem of vector calculus yields: Z S vo ume ∇ · ~dv = l Z Z S~ · ~nda that is, the
volume integral of the divergence of the Poynting energy flow is the same as the Poynting energy
flow over the surface of the volume in question. This integral becomes: ZZ Z ∂B~ ~ · − ~ · S ~nda = E J~
~ + H volume · ∂t ! dv which is simply a realization that the total energy flow into a region of space is
the same as the volume integral over that region of the rate of change of energy stored plus the term
that looks like dissipation. Before we close this, note that, if there is motion of any material within
the system, we can use the empirical expression for transformation of electric field between
observers moving with respect to each other. Here the ’primed’ frame is moving with respeect to the
’unprimed’ frame with the velocity ~v E~ ′ = E~ + ~v × B~ This transformation describes, for example,
the motion of a charged particle such as an electron under the influence of both electric and
magnetic fields. Now, if we assume that there is material motion in the system we are observing and
if we assign ~v to be the velocity of that material, so that E~ ′ is measured in a frame in which thre is
no material motion (that is the frame of the material itself), the product of electric field and current
density becomes: E~ · J~ = E~ ′ − ~v × B~ · J~ ~ = E ′ · J~ − ~v × B~ · J~ ~ = E ′ · J~ + ~v · J~ × B~
In the last step we used the fact that in a scalar triple product the order of the scalar (dot) and
vector (cross) products can be interchanged and that reversing the order of terms in a vector (cross)
product simply changes the sign of that product. Now we have a ready interpretation for what we
have calculated: If the ’primed’ coordinate system is actually the frame of material motion, E~ ′ 1 · J~
= σ |J~| 2 which is easily seen to be dissipation and is positive definite if material conductivity σ is
positive. The last term is obviously conversion of energy from electromagnetic to mechanical form:
~v · J~ × B~ = ~v · F~ where we have now identified force density to be: F~ ~ = J × B~ 11 This is the
Lorentz Force Law, which describes the interaction of current with magnetic field to produce force. It
is not, however, the complete story of force production in electromechanical systems. As we learned
earlier, changes in geometry which affect magnetic stored energy can also produce force.
Fortunately, a complete description of electromechanical force is possible using only magnetic fields
and that is the topic of our next section. 4.2 Field Description of Forces: Maxwell Stress Tensor Forces
of electromagnetic origin, because they are transferred by electric and magnetic fields, are the result
of those fields and may be calculated once the fields are known. In fact, if a surface can be
established that fully encases a material body, the force on that body can be shown to be the integral
of force density, or traction over that surface. The traction τ derived by taking the cross product of
surface current density and flux density on the air-gap surface of a machine (above) actually makes
sense in view of the empirically derived Lorentz Force Law: Given a (vector) current density and a
(vector) flux density. This is actually enough to describe the forces we see in many machines, but
since electric machines have permeable magnetic material and since magnetic fields produce forces
on permeable material even in the absence of macroscopic currents it is necessary to observe how
force appears on such material. A suitable empirical expression for force density is: 1 F~ ~ = J × B~ − 2
H~ · H~ ∇µ where H~ is the magnetic field intensity and µ is the permeability. Now, note that
current density is the curl of magnetic field intensity, so that: F~ = ∇ × H~ 1 × µH~ − H H µ ~ ~ 2 1
· ∇ = µ ∇ × H~ × H~ − H~ · H~ ∇µ 2 And, since: 1 ∇ × H~ × ~ = H H~ · ∇ f s y H~ − 2 ∇
orce den it is: H~ · H~ F~ = µ 1 H~ · 1 ∇ H~ − µ∇ H~ 2 · ~ H − ~ 2 H · H~ ∇µ = µ H~ · ∇
H~ − ∇ 1 µ 2 H~ · H~ This expression can be written by components: the component of force in
the i’th dimension is: Fi = µ X ∂ ∂ 1 H 2 k Hi µ H ∂xk − ∂xi 2 k X k k ! The first term can be written
as: µ X ∂ ∂ ∂ Hk Hi = µHkHi Hi µHk ∂xk ∂ k X xk k − X ∂xk k 12 The last term in this expression is
easily shown to be divergence of magnetic flux density, which is zero: ∇ · B~ = X ∂ µHk = 0 ∂xk k Using
this, we can write force density in a more compact form as: ∂ Fk = µ µH δ 2 iHk − ik H ∂xi 2 X n n !
where we have used the Kroneker delta δik = 1 if i = k, 0 otherwise. Note that this force density is in
the form of the divergence of a tensor: ∂ Fk = Tik ∂xi or F~ = ∇ · T In this case, force on some object
that can be surrounded by a closed surface can be found by using the divergence theorem: ~f = Z F
dv ~ = vol Z T vol ∇ · dv = Z Z T · ~nda or, if we note surface traction to be τi = P k Tiknk , where n is
the surface normal vector, then the total force in direction i is just: ~f = I τida = s I XTiknkda k The
interpretation of all of this is less difficult than the notation suggests. This field description of forces
gives us a simple picture of surface traction, the force per unit area on a surface. If we just integrate
this traction over the area of some body we get the whole force on the body. Note one more thing
about this notation. Sometimes when subscripts are repeated as they are here the summation
symbol is omitted. Thus we would write τi = P k Tiknk = Tiknk. 4.3 Example: Linear Induction Machine
Figure 7 shows a highly simplified picture of a single sided linear induction motor. This is not how
most linear induction machines are actually built, but it is possible to show through symmetry
arguments that the analysis we can carry out here is actually valid for other machines of this class.
This machine consists of a stator (the upper surface) which is represented as a surface current on the
surface of a highly permeable region. The moving element consists of a thin layer of conducting
material on the surface of a highly permeable region. The moving element (or ’shuttle’) has a velocity
u with respect to the stator and that motion is in the x direction. The stator surface current density is
assumed to be: Kz = Re n K ej(ωt−kx) z o 13 Figure 7: Simple Model of single sided linear induction
machine Note that we are ignoring some important effects, such as those arising from finite length of
the stator and of the shuttle. Such effects can be quite important, but we will leave those until later,
as they are what make linear motors interesting. Viewed from the shuttle for which the dimension in
the direction of motion is x ′ − x − ut′ , the relative frequency is: ωt − kx = (ω − ku)t − kx′ = ωst − kx′
Now, since the shuttle surface can support a surface current and is excited by magnetic fields which
are in turn excited by the stator currents, it is reasonable to assume that the form of rotor current is
the same as that of the stator: K = Re n K ej(ωst− ′ kx ) s s Ampere’s Law is, in this situation: o ∂Hy g =
Kz + Ks ∂x which is, in complex amplitudes: K Hy = z + Ks −jkg The y- component of Faraday’s Law is,
assuming the problem is uniform in the z- direction: −jω B ′ s y = jkEz or E ′ ω z = − s µ0H k y A bit of
algebraic manipulation yields expressions for the complex amplitudes of rotor surface current and
gap magnetic field: σ Ks −j µ0ωs s k 2g = 1 + j mu0ωsσs Kz k 2g j K Hy = z kg 1 + j mu0ωsσs k 2g
14Massachusetts Institute of Technology Department of Electrical Engineering and Computer Science
6.685 Electric Machines Class Notes 1: Electromagnetic Forces c 2003 James L. Kirtley Jr. 1
Introduction Stator Stator Conductors Rotor Rotor Conductors Bearings Shaft End Windings Air Gap
Figure 1: Form of Electric Machine This section of notes discusses some of the fundamental
processes involved in electric machinery. In the section on energy conversion processes we examine
the two major ways of estimating elec tromagnetic forces: those involving thermodynamic
arguments (conservation of energy) and field methods (Maxwell’s Stress Tensor). But first it is
appropriate to introduce the topic by describing a notional rotating electric machine. Electric
machinery comes in many different types and a strikingly broad range of sizes, from those little
machines that cause cell ’phones and pagers to vibrate (yes, those are rotating electric machines) to
turbine generators with ratings upwards of a Gigawatt. Most of the machines with which we are
familiar are rotating, but linear electric motors are widely used, from shuttle drives in weaving
machines to equipment handling and amusement park rides. Currently under development are large
linear induction machines to be used to launch aircraft. It is our purpose in this subject to develop an
analytical basis for understanding how all of these different machines work. We start, however, with
a picture of perhaps the most common of electric machines. 2 Electric Machine Description: Figure 1
is a cartoon drawing of a conventional induction motor. This is a very common type of electric
machine and will serve as a reference point. Most other electric machines operate in 1 a fashion
which is the same as the induction machine or which differ in ways which are easy to reference to
the induction machine. Most (but not all!) machines we will be studying have essentially this
morphology. The rotor of the machine is mounted on a shaft which is supported on some sort of
bearing(s). Usually, but not always, the rotor is inside. I have drawn a rotor which is round, but this
does not need to be the case. I have also indicated rotor conductors, but sometimes the rotor has
permanent magnets either fastened to it or inside, and sometimes (as in Variable Reluctance
Machines) it is just an oddly shaped piece of steel. The stator is, in this drawing, on the outside and
has windings. With most of the machines we will be dealing with, the stator winding is the armature,
or electrical power input element. (In DC and Universal motors this is reversed, with the armature
contained on the rotor: we will deal with these later). In most electrical machines the rotor and the
stator are made of highly magnetically permeable materials: steel or magnetic iron. In many common
machines such as induction motors the rotor and stator are both made up of thin sheets of silicon
steel. Punched into those sheets are slots which contain the rotor and stator conductors. Figure 2 is a
picture of part of an induction machine distorted so that the air-gap is straightened out (as if the
machine had infinite radius). This is actually a convenient way of drawing the machine and, we will
find, leads to useful methods of analysis. Stator Core Stator Conductors In Slots Rotor Conductors In
Slots Air Gap Figure 2: Windings in Slots What is important to note for now is that the machine has
an air gap g which is relatively small (that is, the gap dimension is much less than the machine radius
r). The air-gap also has a physical length l. The electric machine works by producing a shear stress in
the air-gap (with of course side effects such as production of “back voltage”). It is possible to define
the average air gap shear stress, which we will refer to as τ . Total developed torque is force over the
surface area times moment (which is rotor radius): T = 2πr2 ℓ < τ > Power transferred by this device
is just torque times speed, which is the same as force times 2 surface velocity, since surface velocity
is u = rΩ: Pm = ΩT = 2πrℓ < τ > u If we note that active rotor volume is , the ratio of torque to volume
is just: T = 2 < τ > Vr Now, determining what can be done in a volume of machine involves two things.
First, it is clear that the volume we have calculated here is not the whole machine volume, since it
does not include the stator. The actual estimate of total machine volume from the rotor volume is
actually quite complex and detailed and we will leave that one for later. Second, we need to estimate
the value of the useful average shear stress. Suppose both the radial flux density Br and the stator
surface current density Kz are sinusoidal flux waves of the form: Br = √ 2B0 cos (pθ − ωt) Kz = √ 2K0
cos (pθ − ωt) Note that this assumes these two quantities are exactly in phase, or oriented to ideally
produce torque, so we are going to get an “optimistic” bound here. Then the average value of
surface traction is: 1 2π < τ >= Z BrKzdθ = B0K0 2π 0 The magnetic flux density that can be developed
is limited by the characteristics of the magnetic materials (iron) used. Current densities are a function
of technology and are typically limited by how much effort can be put into cooling and the
temperature limits of insulating materials. In practice, the range of shear stress encountered in
electric machinery technology is not terribly broad: ranging from a few kPa in smaller machines to
about 100 kPa in very large, well cooled machines. It is usually said that electric machines are torque
producing devices, meaning tht they are defined by this shear stress mechanism and by physical
dimensions. Since power is torque times rotational speed, high power density machines necessarily
will have high shaft speeds. Of course there are limits on rotational speed as well, arising from
centrifugal forces which limit tip velocity. Our first step in understanding how electric machinery
works is to understand the mechanisms which produce forces of electromagnetic origin. 3 Energy
Conversion Process: In a motor the energy conversion process can be thought of in simple terms. In
“steady state”, electric power input to the machine is just the sum of electric power inputs to the
different phase terminals: Pe = Xviii i Mechanical power is torque times speed: Pm = TΩ 3 Mechanical
Electro Converter Electric Power In Mechanical Power Out Losses: Heat, Noise, Windage,... Figure 3:
Energy Conversion Process And the sum of the losses is the difference: Pd = Pe − Pm It will
sometimes be convenient to employ the fact that, in most machines, dissipation is small enough to
approximate mechanical power with electrical power. In fact, there are many situations in which the
loss mechanism is known well enough that it can be idealized away. The “thermodynamic”
arguments for force density take advantage of this and employ a “conservative” or lossless energy
conversion system. 3.1 Energy Approach to Electromagnetic Forces: f + Magnetic Field v x - System
Figure 4: Conservative Magnetic Field System To start, consider some electromechanical system
which has two sets of “terminals”, electrical and mechanical, as shown in Figure 4. If the system
stores energy in magnetic fields, the energy stored depends on the state of the system, defined by (in
this case) two of the identifiable variables: flux (λ), current (i) and mechanical position (x). In fact,
with only a little reflection, you should be able to convince yourself that this state is a single-valued
function of two variables and that the energy stored is independent of how the system was brought
to this state. Now, all electromechanical converters have loss mechanisms and so are not themselves
conser vative. However, the magnetic field system that produces force is, in principle, conservative
in the 4 sense that its state and stored energy can be described by only two variables. The “history”
of the system is not important. It is possible to chose the variables in such a way that electrical power
into this conservative system is: P e dλ = vi = i dt Similarly, mechanical power out of the system is: P
m = f e dx dt The difference between these two is the rate of change of energy stored in the system:
dWm = P e dt − P m It is then possible to compute the change in energy required to take the system
from one state to another by: a Wm(a) − Wm(b) = Z idλ b − f e dx where the two states of the system
are described by a = (λa, xa) and b = (λb, xb) If the energy stored in the system is described by two
state variables, λ and x, the total differential of stored energy is: ∂Wm ∂Wm dWm = dλ + dx ∂λ ∂x
and it is also: dWm = idλ − f e dx So that we can make a direct equivalence between the derivatives
and: ∂ f e W = − m ∂x In the case of rotary, as opposed to linear, motion, torque T e takes the place of
force f e and angular displacement θ takes the place of linear displacement x. Note that the product
of torque and angle has the same units as the product of force and distance (both have units of work,
which in the International System of units is Newton-meters or Joules. In many cases we might
consider a system which is electricaly linear, in which case inductance is a function only of the
mechanical position x. λ(x) = L(x)i In this case, assuming that the energy integral is carried out from λ
= 0 (so that the part of the integral carried out over x is zero), Wm = Z λ 1 1 λ 2 λdλ = 0 L(x) 2 L(x) This
makes f e 1 = − λ 2 ∂ 1 2 ∂x L(x) 5 Note that this is numerically equivalent to f e 1 = − i 2 ∂ L(x) 2 ∂x
This is true only in the case of a linear system. Note that substituting L(x)i = λ too early in the
derivation produces erroneous results: in the case of a linear system it produces a sign error, but in
the case of a nonlinear system it is just wrong. 3.1.1 Example: simple solenoid Consider the magnetic
actuator shown in cartoon form in Figure 5. The actuator consists of a circular rod of ferromagnetic
material (very highly permeable) that can move axially (the x direction) inside of a stationary piece,
also made of highly permeable material. A coil of N turns carries a current I. The rod has a radius R
and spacing from the flat end of the stator is the variable dimension x. At the other end there is a
radial clearance between the rod and the stator g. Assume g ≪ R. If the axial length of the radial gaps
is ℓ = R/2, the area of the radial gaps is the same as the area of the gap between the rod and the
stator at the variable gap. µ CL x R g N turns R/2 µ Figure 5: Solenoid Actuator The permeances of the
variable width gap is: µ P 0πR2 1 = x and the permeance of the radial clearance gap is, if the gap
dimension is small compared with the radius: 2µ0πRℓ µ π P 0 R2 2 = = g g The inductance of the coil
system is: N2 µ 2 2 1 2 0πR N L = = N 2 P P = R1 + R2 P2 + P2 x + g Magnetic energy is: 6 λ Wm = Z 0 1
λ 2 λ 2 idλ = = 0 x + g 2 L(x) 2 µ πR2N2 0 0 And then, of course, force of electric origin is: f e ∂W = − m
λ 2 = 0 d 1 ∂x − 2 dx L(x) Here that is easy to carry out: d 1 1 = dx L µ 2 2 0πR N So that the force is: f e
λ 2 (x) = − 0 1 2 µ 2 0πR N2 Given that the system is to be excited by a current, we may at this point
substitute for flux: µ0πR2N i λ = L(x)i = x + g and then total force may be seen to be: f e µ = − 0πR2N2
i 2 (x + g) 2 2 The force is ‘negative’ in the sense that it tends to reduce x, or to close the gap. 3.1.2
Multiply Excited Systems There may be (and in most electric machine applications there will be)
more than one source of electrical excitation (more than one coil). In such systems we may write the
conservation of energy expression as: dWm = Xikdλk k − f e dx which simply suggests that electrical
input to the magnetic field energy storage is the sum (in this case over the index k) of inputs from
each of the coils. To find the total energy stored in the system it is necessary to integrate over all of
the coils (which may and in general will have mutual inductance). Wm = Z i · dλ Of course, if the
system is conservative, Wm(λ1, λ2, . . . , x) is uniquely specified and so the actual path taken in
carrying out this integral will not affect the value of the resulting energy. 7 3.1.3 Coenergy We often
will describe systems in terms of inductance rather than its reciprocal, so that current, rather than
flux, appears to be the relevant variable. It is convenient to derive a new energy variable, which we
will call co-energy, by: W′ m = Xλiii − Wm i and in this case it is quite easy to show that the energy
differential is (for a single mechanical variable) simply: dW′ e m = Xλkdik + f dx k so that force
produced is: ∂W′ fe = m ∂x 3.2 Example: Synchronous Machine µ µ F F F’ F’ A’ A A’ B B’ B B’ C C’ C C’
A θ Rotor Stator Gap Figure 6: Cartoon of Synchronous Machine Consider a simple electric machine
as pictured in Figure 6 in which there is a single winding on a rotor (call it the field winding and a
polyphase armature with three identical coils spaced at uniform locations about the periphery. We
can describe the flux linkages as: λa = Laia + Labib + Labic + M cos(pθ)if 2π λb = Labia + Laib + Labic +
M cos(pθ − )if 3 2π λc = Labia + Labib + Laic + M cos(pθ + )if 3 2π 2π λf = M cos(pθ)ia + M cos(pθ − )ib
+ M cos(pθ + ) + Lf if 3 3 It is assumed that the flux linkages are sinusoidal functions of rotor position.
As it turns out, many electrical machines work best (by many criteria such as smoothness of torque
production) 8 if this is the case, so that techniques have been developed to make those flux linkages
very nearly sinusoidal. We will see some of these techniques in later chapters of these notes. For the
moment, we will simply assume these dependencies. In addition, we assume that the rotor is
magnetically ’round’, which means the stator self inductances and the stator phase to phase mutual
inductances are not functions of rotor position. Note that if the phase windings are identical (except
for their angular position), they will have identical self inductances. If there are three uniformly
spaced windings the phase-phase mutual inductances will all be the same. Now, this system can be
simply described in terms of coenergy. With multiple excitation it is important to exercise some care
in taking the coenergy integral (to ensure that it is taken over a valid path in the multi-dimensional
space). In our case there are actually five dimensions, but only four are important since we can
position the rotor with all currents at zero so there is no contribution to coenergy from setting rotor
position. Suppose the rotor is at some angle θ and that the four currents have values ia0, ib0, ic0 and
if0. One of many correct path integrals to take would be: W′ m = Z ia0 Laiadia 0 + Z ib0 (Labia0 + Laib)
dib 0 + Z ic0 (Labia0 + Labib0 + Laic) dic 0 + Z if0 2π 2π M cos(pθ)ia0 + M cos(pθ − )ib0 + M cos(pθ +
)ic0 + Lf if 0 3 3 dif The result is: W′ 1 = L i 2 + i 2 2 m a a0 b0 + ico + Lab (iaoib0 + iaoic0 + icoib0)
2 +M if0 2π 2π 1 ia0 cos(pθ) + i 2 b0 cos(pθ − ) + ic0 cos(pθ + ) 3 3 + Lf i 2 f0 Since there are no
variations of the stator inductances with rotor position θ, torque is easily given by: ∂W′ Te = m 2 π =
−pM if0 π 2 ia0 sin(pθ) + ib0 sin(pθ − ) + ico sin(pθ + ) ∂θ 3 3 3.2.1 Current Driven Synchronous
Machine Now assume that we can drive this thing with currents: ia0 = Ia cos ωt ib0 = Ia cos ω 2π t −
3 2π ic0 = Ia cos ω t + 3 if0 = If 9 and assume the rotor is turning at synchronous speed: pθ = ωt
+ δi Noting that cos x sin y = 1 sin(x − y) + 1 sin(x + y), we find the torque expression above to be: 2 2
1 1 Te = −pMIaIf sin δi + sin (2ωt + δi) 2 2 1 1 4 + sin δi + sin 2 2 π 2ωt + δi − 3 + 1 1 4π sin
δi + sin 2ωt + δi + 2 2 3 The sine functions on the left add and the ones on the right cancel,
leaving: 3 Te = − pMIaIf sin δi 2 And this is indeed one way of looking at a synchronous machine,
which produces steady torque if the rotor speed and currents all agree on frequency. Torque is
related to the current torque angle δi . As it turns out such machines are not generally run against
current sources, but we will take up actual operation of such machines later. 4 Field Descriptions:
Continuous Media While a basic understanding of electromechanical devices is possible using the
lumped parameter approach and the principle of virtual work as described in the previous section,
many phenomena in electric machines require a more detailed understanding which is afforded by a
continuum approach. In this section we consider a fields-based approach to energy flow using
Poynting’s Theorem and then a fields based description of forces using the Maxwell Stress Tensor.
These techniques will both be useful in further analysis of what happens in electric machines. 4.1
Field Description of Energy Flow: Poyting’s Theorem Start with Faraday’s Law: ∂B~ ∇ × E~ = − ∂t and
Ampere’s Law: ∇ × H~ ~ = J Multiplying the first of these by H~ ~ and the second by E and taking the
difference: ∂B~ H~ · ∇ × E~ − E~ · ∇ × H~ = ∇ · E~ ~ × H = −H~ · E J ~ ~ dt − · On the left of this
expression is the divergence of electromagnetic energy flow: S~ ~ = E × H~ 10 Here, S~ is the
celebrated Poynting flow which describes power in an electromagnetic field sysstem. (The units of
this quantity is watts per square meter in the International System). On ~ ~ the right hand side are
two terms: H · ∂B dt is rate of change of magnetic stored energy. The second ~ term, E · J~ looks a lot
like power dissipation. We will discuss each of these in more detail. For the moment, however, note
that the divergence theorem of vector calculus yields: Z S vo ume ∇ · ~dv = l Z Z S~ · ~nda that is, the
volume integral of the divergence of the Poynting energy flow is the same as the Poynting energy
flow over the surface of the volume in question. This integral becomes: ZZ Z ∂B~ ~ · − ~ · S ~nda = E J~
~ + H volume · ∂t ! dv which is simply a realization that the total energy flow into a region of space is
the same as the volume integral over that region of the rate of change of energy stored plus the term
that looks like dissipation. Before we close this, note that, if there is motion of any material within
the system, we can use the empirical expression for transformation of electric field between
observers moving with respect to each other. Here the ’primed’ frame is moving with respeect to the
’unprimed’ frame with the velocity ~v E~ ′ = E~ + ~v × B~ This transformation describes, for example,
the motion of a charged particle such as an electron under the influence of both electric and
magnetic fields. Now, if we assume that there is material motion in the system we are observing and
if we assign ~v to be the velocity of that material, so that E~ ′ is measured in a frame in which thre is
no material motion (that is the frame of the material itself), the product of electric field and current
density becomes: E~ · J~ = E~ ′ − ~v × B~ · J~ ~ = E ′ · J~ − ~v × B~ · J~ ~ = E ′ · J~ + ~v · J~ × B~
In the last step we used the fact that in a scalar triple product the order of the scalar (dot) and
vector (cross) products can be interchanged and that reversing the order of terms in a vector (cross)
product simply changes the sign of that product. Now we have a ready interpretation for what we
have calculated: If the ’primed’ coordinate system is actually the frame of material motion, E~ ′ 1 · J~
= σ |J~| 2 which is easily seen to be dissipation and is positive definite if material conductivity σ is
positive. The last term is obviously conversion of energy from electromagnetic to mechanical form:
~v · J~ × B~ = ~v · F~ where we have now identified force density to be: F~ ~ = J × B~ 11 This is the
Lorentz Force Law, which describes the interaction of current with magnetic field to produce force. It
is not, however, the complete story of force production in electromechanical systems. As we learned
earlier, changes in geometry which affect magnetic stored energy can also produce force.
Fortunately, a complete description of electromechanical force is possible using only magnetic fields
and that is the topic of our next section. 4.2 Field Description of Forces: Maxwell Stress Tensor Forces
of electromagnetic origin, because they are transferred by electric and magnetic fields, are the result
of those fields and may be calculated once the fields are known. In fact, if a surface can be
established that fully encases a material body, the force on that body can be shown to be the integral
of force density, or traction over that surface. The traction τ derived by taking the cross product of
surface current density and flux density on the air-gap surface of a machine (above) actually makes
sense in view of the empirically derived Lorentz Force Law: Given a (vector) current density and a
(vector) flux density. This is actually enough to describe the forces we see in many machines, but
since electric machines have permeable magnetic material and since magnetic fields produce forces
on permeable material even in the absence of macroscopic currents it is necessary to observe how
force appears on such material. A suitable empirical expression for force density is: 1 F~ ~ = J × B~ − 2
H~ · H~ ∇µ where H~ is the magnetic field intensity and µ is the permeability. Now, note that
current density is the curl of magnetic field intensity, so that: F~ = ∇ × H~ 1 × µH~ − H H µ ~ ~ 2 1
· ∇ = µ ∇ × H~ × H~ − H~ · H~ ∇µ 2 And, since: 1 ∇ × H~ × ~ = H H~ · ∇ f s y H~ − 2 ∇
orce den it is: H~ · H~ F~ = µ 1 H~ · 1 ∇ H~ − µ∇ H~ 2 · ~ H − ~ 2 H · H~ ∇µ = µ H~ · ∇
H~ − ∇ 1 µ 2 H~ · H~ This expression can be written by components: the component of force in
the i’th dimension is: Fi = µ X ∂ ∂ 1 H 2 k Hi µ H ∂xk − ∂xi 2 k X k k ! The first term can be written
as: µ X ∂ ∂ ∂ Hk Hi = µHkHi Hi µHk ∂xk ∂ k X xk k − X ∂xk k 12 The last term in this expression is
easily shown to be divergence of magnetic flux density, which is zero: ∇ · B~ = X ∂ µHk = 0 ∂xk k Using
this, we can write force density in a more compact form as: ∂ Fk = µ µH δ 2 iHk − ik H ∂xi 2 X n n !
where we have used the Kroneker delta δik = 1 if i = k, 0 otherwise. Note that this force density is in
the form of the divergence of a tensor: ∂ Fk = Tik ∂xi or F~ = ∇ · T In this case, force on some object
that can be surrounded by a closed surface can be found by using the divergence theorem: ~f = Z F
dv ~ = vol Z T vol ∇ · dv = Z Z T · ~nda or, if we note surface traction to be τi = P k Tiknk , where n is
the surface normal vector, then the total force in direction i is just: ~f = I τida = s I XTiknkda k The
interpretation of all of this is less difficult than the notation suggests. This field description of forces
gives us a simple picture of surface traction, the force per unit area on a surface. If we just integrate
this traction over the area of some body we get the whole force on the body. Note one more thing
about this notation. Sometimes when subscripts are repeated as they are here the summation
symbol is omitted. Thus we would write τi = P k Tiknk = Tiknk. 4.3 Example: Linear Induction Machine
Figure 7 shows a highly simplified picture of a single sided linear induction motor. This is not how
most linear induction machines are actually built, but it is possible to show through symmetry
arguments that the analysis we can carry out here is actually valid for other machines of this class.
This machine consists of a stator (the upper surface) which is represented as a surface current on the
surface of a highly permeable region. The moving element consists of a thin layer of conducting
material on the surface of a highly permeable region. The moving element (or ’shuttle’) has a velocity
u with respect to the stator and that motion is in the x direction. The stator surface current density is
assumed to be: Kz = Re n K ej(ωt−kx) z o 13 Figure 7: Simple Model of single sided linear induction
machine Note that we are ignoring some important effects, such as those arising from finite length of
the stator and of the shuttle. Such effects can be quite important, but we will leave those until later,
as they are what make linear motors interesting. Viewed from the shuttle for which the dimension in
the direction of motion is x ′ − x − ut′ , the relative frequency is: ωt − kx = (ω − ku)t − kx′ = ωst − kx′
Now, since the shuttle surface can support a surface current and is excited by magnetic fields which
are in turn excited by the stator currents, it is reasonable to assume that the form of rotor current is
the same as that of the stator: K = Re n K ej(ωst− ′ kx ) s s Ampere’s Law is, in this situation: o ∂Hy g =
Kz + Ks ∂x which is, in complex amplitudes: K Hy = z + Ks −jkg The y- component of Faraday’s Law is,
assuming the problem is uniform in the z- direction: −jω B ′ s y = jkEz or E ′ ω z = − s µ0H k y A bit of
algebraic manipulation yields expressions for the complex amplitudes of rotor surface current and
gap magnetic field: σ Ks −j µ0ωs s k 2g = 1 + j mu0ωsσs Kz k 2g j K Hy = z kg 1 + j mu0ωsσs k 2g 14

You might also like