Anda di halaman 1dari 638

-------. r--r~- -._- - -.

- L
)
...:.I

PRINCIPLES . OF
ELECTROMECHANICAL· ­
ENERGY CONVERSION
CONTENTS

l
preface ix

Chapter 1. Basic Coordinates, Lumped Elements, and Energy.state


Functions
1-1 The Electrical-capacitance Element 2

1·2 The Electrical·inductance Element 7

1-3 The Electrical Dissipative Element 13

1-4 Basic Mechanical-Translational Coordinates 15

1-5 Basic Mechanical·Rotational Coordinates 20

1-6 Summary of the Basic Coordinate>! and the

Lumped-element Relationships 23

1-7 Equilibrium Equations, State Variables, and

State Functions 24

1-8 The Capacitance State Function 26

1-9 Energy-state Function for a System of Capacitances 29

1-10 Partial Derivatives of the Electric·field State

Functions 32

1-11 Energy-;;tate Functions for a System of Inductances 34

1-12 Translational Mechanical-energy·state Functions 37

1-13 Summary of the Conservative Element Energy-

and Coenergy-state Functions 41

'probkms 44

Chapter 2. Equilibrium Equations from Energy-state Functions:

Lagrange's Equation 48

2-1 Generalized Forces from Energy-state Functions 48

2-2 Mechanical-equilibrium Equations from Energy-

state Functions 50

2-3 A Restricted Form of Lagrange's Equation 53

2-4 Degrees of Freedom and Generalized Coordinates 59

2-5 A Complete Formulation of Lagrange's Equation

for Conservative Mechanical Systems 60

2-6 Loop Equations from Energy.state Functions 69

2·7 Nodal Equations from Energy-state Functions 74

2-8 Lagrange's Equation for Conservative Mechanical

and Electrical Systems 78

2-9' A Brief Look at Electromechanical Coupling 79

xiii

l .._ - .. _--._- ~---~-


xiv Contents

2-10 The Rayleigh Dissipation Function 81

2-11 Summary 91

Problems 92

Chapter 3. Formulation of Equilibrium Equations for Electro­


mechanical Systems 99

3-1 A Two-port Magnetic-field Electromechanical

Lumped System 100

3-2 A Multiport Magnetic-field Electromechanical

System 106

3-3 A Two-port Electric-field Electromechanical

Lumped System 109

3-4 A Multiport Electric-field Electromechanical

System 113

3-5 Mechanical Forces of Electrical Origin from the

Principle of Virtual Work 115

3-6 The Force of Electrical Origin Using the Coenergy­


state Function 119

3-7 The Principle of Virtual Work for a Multiport

Electric-field System 121

3-8 Forces of Electrical Origin for a M ultiport

Magnetic-field System 124

3-9 Interaction Forces from Lagrange's Equation 125

3-10 Interaction Forces from Two Forms of the State

Functions 133

3-11 Lagrange's Equation with the Two State-function

Forms 139

3-12 Summary 141

Problemil 142

Chapter 4. The Analysis of linear Systems 148

4-1 Characteristic Properties of the Linear System 149

4-2 The Linear Differential Operator 153

4-3 The Constant-coefficient Linear Differential

Operator 156

4-4 Block-diagram Notation 159

4--5 D-C Steady-state Response 166

4--6 Sinusoidal Steady-state Response 168

4-7 Frequency-response Characteristics 175

4-8 The Logarithmic Form of the Frequency Response 178

4-9 Laplace Transform Techniques 190

4-10 The Partial-fraction Expansion: Poles and Zeros 200

4-11 Time-response Characteristics from Pole-zero

Locations 202

------_..._­
Contents

4-12 The Routh Criterion 204

4-13 Summary 208

Problems 210

Chapter 5. Response Characteristics of Electromechanical Systems 214

5-1 Static Operating Points 215

5-2 Linear Incremental Differential Equilibrium Equa­


tions Valid about an Operating Point 221

5-3 Stable and Unstable Operating Points 224

5-4 Incremental Operation as a Mechanical-to-

Electrical Transducer 227

5-5 Magnetic Fields from Permanent l\iagnets 234

5-6 A Permanent-magnet Mechanical-to-Electrical

Transducer 238

5-7 The Moving-coil Audio Loudspeaker 243

5-8 An Electromechanical-accelerometer Transducl'r 248

5-9 Summary 257

Problems 258

Chapter 6. Construction of the Primitive Machine 266

6-1 Magnetic Structure of the Primitive Machine 267

6-2 Construction of the Stator Circuits 270

6-3 Stator Magnetic Field 272

6-4 Commutator-and-Brush Rotor Winding 276

6-5 Rotor Magnetic Fields 280

6-6 Primitive-machine Stationary Inductance

Parameters 283

6-7 Primitive-machine Rotational Inductances 288

6-8 Multipole Primitive-machine Inductance

Parameters 294

6-9 Primitive-machine Resistance Parameters 301

6-10 Identical Rotor Distribution Factors 302

6-11 Summary 304

Problems 306

Chapter 7. Equilibrium Equations for the Primitive Machine 312

7-1 Reference Directions 312

7-2 Electrical-equilibrium Equations 314

7-3 Mechanical-equilibrium Equations 319

7-4 Multiwinding Ideal Transformer 322

7-5 Primitive-machine Equivalent Circuit 324

7-6 D-C Steady-state and Incremental Operation 326

7-7 Summary 328

Problems 330

xvi Contents

Chapter 8. Direct-current Commutator Machines (I)


334
8-1 Two-winding Commutator Machine 335

8-2 D-C Generator Constraints 340

8-3 Static Operation of the D-C Generator 342

8-4 Trf!.nsient D-C Generator Operation 345

8-5 D-C Motor Constraints 350

8-6 Static Operation of the D-C Motor 352

8-7 Transient D-C :\lotor Operation 360

8-8 Series-connected D;C Motor Constraints 366

8-9 Steady-state Operation of the Series Motor 368

8-10 Transient Operation of the Series Motor 370

8-11 Summary 375

ProblemB 376

C~apter 9. Direct-current Commutator Machines (II) 382

9-1 Ward Leonard Speed-control System 383

9-2 Incremental Transient Response of the

Ward Leonard System 387

9-3 Compound D-C Generator Steady-state Operation 392

9-4 Compound-generator Transient Response 395

9-5 The Amplidyne Rotating Amplifier 398

9-6 The Metadyne Constant-current Amplifier 405

9-7 Nonlinear Characteristics: Saturation Effects 406

9-8 Self-excited Shunt Generator 408

9-9 Transient Response of the Self-excitcd Generator 412

9-10 Air-gap Fields in Actual D-C Machines 414

9-11 Commutating or Interpole Windings 417

9-12 Compensating Windings 421

9-13 Summary 423

Problems 425

Chapter 10. The A-B Primitive Machine 431

10-1 The Slip-ring Winding 432

10-2 Rotating D-Q Rotor Surface-current Distributions 435

10-3 Equivalent A-B and D-Q Rotor Surface-Current

Distributions 436

10-4 A-B Machine Equations and Parameters 440

10-5 Lagrangian Formulation of the A-B Machine 445

10-6 Power Invariance 449

10-7 Three-phase to Two-phase Transformations 451

10-8 Impedance Matrix Transformations 457


v

10-9 Summary 461

ProblemIJ 463

Content. It

r.:
Chapter 11. The Synchronous Machine 467

11-1 An Elementary View of the Synchronous Machine 468

11-2 A Primitiv<,-machine .Model 470

11-3 Constraint Equations for Steady-state Operation

on a Balanced Infinite Bus 474

11-4 Synchronous Torque of Electrical Origin 479

~
11-5 Sinor Diagram of the Synchronous Machine 485

11-6 Constant-power Operation 492

11-7 Per-phase Equivalent Circuit for the Nonsalient­


pole ~lachine 502

lI-R Phy:-ical Significance of the Torque Angle 504

11-9 J)t'termination of Synchronous-machine Parameters 508

11-10 Transient Torqu<;-angle Characteristics 513

\l-ll Electrical Respons(' to a Three-phase Short Circuit 520

11-12 Summary 529

Problems 530

Chapter 12. The Induction Machine 536

12-1 An Elementary View of the Induction :\lachine 531

12-2 Primitive-machine Model of the Three-phase

Induction Machine 541

12-3 Constraint Equations for Steady-state Operation

on a Balanced Infinite Bus 547

12-4 Two-phase Rymmetrical-component Transformation 550

12-5 Rinusoidal Steady-state Equivalent Circuit 553

12-6 Torque Relation8 557

12-7 Per-phase Quantities and the Per-phase

Equivalent Circuit 558

12-8 The Leakage-Mutual Equivalent Circuit 565

12-9 The Approximate Equivalent Circuit and the

Experimental Determination of Machine

Parameters 569

12-10 The Circle Diagram for the Induction Machine 575

12-11 Steady-state Performance Characteristics 584

12-12 Transient Performance Characteristics 591

12-13 The Two-pha..<;e Servomotor 596

12-14 The Single-phase Induction Machine 603

12-15 Summary 609

Problems 612

Appendix Conversion Factors and Physical Constants 621

Bibliography 623

Answers to Problems
627

Index 633

BASIC COORDINATES, LUMPED


ELEMENTS, AND ENERGY-STATE
FUNCTIONS

The mathematical description of physical systems is


very often accomplished through the specification of
certain dependent coordinates as functions of the
independent coordinate time. For example, sup­
pose we are interested in a mechanical system com­
posed of a mass suspended from some rigid support
by a coil spring. A reasonable description of the
motion of this simple system could be the specifica­
tion of the position of the mass with respect to the
r rigid support for each instant of time. The depe~d­
ent coordinate, in this case, is the position of the
mass, and the independent coordinate is the time of
this measurement. If the motion of this mass were
not simply up and down, we might require two or
maybe even three coordinates to completely specify
the position of the mass at each instant in time.
If these dependent position coordinates are
known functions of time, then we know exactly what
the complete motion of the system will be. How~
ever, this is not the only manner in which system
motion could be described. As an alternative, the
velocities of the mass could have been used as
dependent coordinates. This second choice of coor~
dinates is not independent of the first, since it can
he ohtained hy simply taking the first time deriva­
tive of the position coordinates. In fact, there are
\ really an infinite number of coordinate schemes for
specifying the motion of this very simple mechanical
system.
In this first chapter certain common coordi~
nates for describing lumped elements in electrica.l
systems and in mechanical systems in translational
and rotational motion will be introduced. The con-
2 Principle. of e/KtromKhClnicClI·energy conyemon

cept of a lumped element mathematically separates physical effects which


are actually occurring together. For example, an inductance coil is often
approximated by a lumped-resistance element in series with a lumped­
inductance element. Thus the ohmic losses of the physical inductance
coil are taken into account by the resistance element and the magnetic­
field effects are accounted for by the inductance element. Such a separa­
tion of physical effects is possible in systems involving only static or really
quasi-static time variations. In all cases our coordinate description of
these lumped elements will be sufficiently general to include nonlinearities.
The physical world we are trying to describe is nonlinear, and concepts
must include this analytic difficulty if they are to be of any practical use.
Let us begin this discussion of fundamental coordinates with electrical
systelfls. In cireuit analysis the usual description of the operation of a
network is in terms of current8 and voltages. The lumped elements of
resistance, capacitance, and inductance imply relationships between the
current and voltage variables at the terminals of the devices. Thus for
electrical systems, voltage and current as functions of the independent
coordinate time might be considered as good choices for our basic depend­
ent coordinates. The main reason for using these quantities in our work
is t hat they are easily measured. There are many deviees for observing
and measuring the time-dependent functions representing voltages and
currents. One major drawback to this choice is the fact that only linear­
capacitance and -inductance elements can be described conveniently by
voltage and current coordinates. What we must do is go to the basic
electromagnetic-field concepts for ideas which bridge the gap between
physical deviees and lumped-eireuit theory. From this starting point
sufficiently basic fundamental coordinates will emerge.

1-1 THE ELECTRICAL-CAPACITANCE ELEMENT

The capacitance element is a good starting point. In linear-circuit theory,


the symbol for the capacitance shown in Fig. 1-1 is described by the differ­
ential relationship

i(t) = C dv(t) (1-1)


dt

which states that the current through the capacitance element is propor­
tional to the time rate of change of the voltage across the element. For a
given element this proportionality constant ha.s units of farads if the cur­
rent i is in amperes, the potential rise v is in volts, and the time t is in
seconds. These are all consistent mks unit's. Equation (1-1) can be
Coordinate., element., and Itate func:tioM 3

c
Fig. 1-1 The IUII'\ped-eapa.citanee symbol.

written in an equivalent form as

v(t) = bf,: i(t) dt + v(t 1) (1-2)

simply by solving, or really manipulating, Eq. (1-1). Equation (1-2)


gives the voltage across the capacitance element at some particular time
t as being equal to the voltage at some previous time tt plus the integral
of the current from time h to time t multiplied, or weighted, by the recip­
rocal of the capacitance coefficient C.
Suppose we arbitrarily select time h to be far enough in the past that
v(t 1) is equal to zero. If tl were equal to - co, then we could be sure
that the voltage across the capacitor was zero at that time. In fact, the
capacitor probably was not even manufactured at time - co. Thus Eq.
(1-2) becomes

v(t) = bf~. i(t) dt (1-3)

The current function i(t) is defined as the time rate of flow of positive
charge, or

i(t) = dq(t) . (1-4)


dt

where q(t) is charge as a function of time expressed in units of coulombs;


thus one coulomb per second equals one ampere. Now integrating Eq.
(1-4) gives

q(t) - q( - co) = f~. i(t) dt (1-5)

Taking q( - co) equal to zero, since our element was not even constructed
at that time, Eq. (1-5) gives the integral in Eq. (1-3) as being some
quantity of charge at time t. Physically, this charge q(t} is found to be
accumulated on the capacitor plates, with positive charge on the top plate
in Fig. 1-1 and a negative charge of equal magnitude on the bottom plate
for positive i(t). The current is defined as the rate of flow of positive
4 Principle, of electromechanical-energy canverr;on

Fig. 1-2 The defining-characteristic curve for


y the Jumped-capacitance element.

charge; therefore, for positive current, positive charge is being deposited


on the top plate.
'Jlhus for the linear-capacitance element, Eq. (1-3) becomes
1
vet) = C q(t) (1-6)

which says that the voltage across the capacitor is linearly related, or
simply proportional, to the charge on the capacitor plates. From a physi­
cal standpoint the capacitance element has the ability to accumulate
charge, and in the process a voltage or potential difference appears across
its terminals. The charge appears as a fundamental coordinate in this
mechanism, and it is chosen as the first of our basic coordinat~. The
first time derivative of the charge, dq(t)/dt, or more simply, q(t) , is the
current through the capacitor, namely, i(t). We have taken i11e charge
as the basic coordinate, with its first time -derivative, current, being one of
the two common circuital variables. In succeeding discussions involving
the inductance parameter the second fundamental electrical coordinate,
called flux linkage, is introduced. The first time derivative of the flux
linkages turns out to be voltage. Thus the usual network variables, cur­
rent and voltage, are to be obtained as the first time derivatives of our
basic coordinates charge and flux linkage.
Equation (1-6) gives a general experimental scheme for describing
this capacitance element. For a physical device havi'ng the ability to
accumulate charge, an expression relating the voltage across the termi­
nals of the device to the charge it accumulates is required. If the charge
is plotted as a function of the voltage, then the capacitanceC(v) is the
slope of this curve at any point, as shown in Fig. 1-2. If the plot were a
straight line, then the capacitance parameter would be constant and we
would then have a linear capacitor. However, if the plot is not a straight
line, as shown in Fig. 1-2, then the slope changes as a function of the
charge, or the voltage, and the element is nonlinear. "Linea.r" or IInon­
linear" simply means the charge vs. voltage plot is straight or curved,
respectively. We shall call a curve such as Fig. 1-2 the defining-charac­
Coordinate., element., and "ote function. S

Fig. 1-3 Double-valued characteristic curve


exhibiting hysteresis.

teristic curve for the lumped-capacitance element, since this curve com­
pletely defines the element.
All of the characteristic curves will be restricted to single-valued func­
tions. Only one value of charge q in Fig:i-2 corresponds to a 'specific
terminal voltage v. Certain dielectric materials exhibit a double-valued
characteristic as sfiown in Fig. 1-3. The value of q for any specified value
of v depends on the past history of the voltage v. -This behavior whereby
1!.he value of q 'for increasing v does not correspond to the value qf q for
decreasing v is termed hysteresis. A large number of hysteresis loops, in
the form of Fig. 1-3, are actually needed to describe the device completely.
As an approximation for such characteristics a single-valued curve is
drawn through the centers of the hysteresis loops. The energy loss ass0­
ciated with cycling around the hysteresis loop is then taken into aCcount
by a lumped-resistance element.
I

Example 1-1
The characteristic curve for a parallel-plate capacitor with a Rochelle-ealt dielectric
has the form shown in Fig. 1-4. The lack of symmetry results from dielectric polari­
zation giving rise to reSidual charge analogous to residual magnetism for ferromagnetic
materials. For voltage values between v. and I'" the characteristic curve can be

Fig. 1-4 Parallel-plate capacitor with a Ro­


chelle-tlalt dielectric.
6 Principle, of e/ectromechanical·energy conversion

A spherical capacitor.

approximated by the expression

v "'I ao + a,q + a,.q' + a,ql


Determine the capacitance valid over the interval v. < v < Vb.
SolutWn: The capacitance for any value of voltage or charge is defined &8 the
slope of the characteristic curve. Therefore. if q is expressed &8 a function of v, we have

C(v) - dq(,')
do

However, the given expression has vasa function of q; thus the capacitance is found
more conveniently from
1 dv(q)

C(q) - (j(J - a, + 2a.q + 3a aq'

which can be put into the form


Co

C(q) - 1 + Aq + Bq'

where Co. A, and B are appropriate constants.


By using a polynomial of high enough order, any single-valued characteristic
curve can be expressed analytically. Over the limited region v.. < v < Vb a third-order
polynomial in q W&8 judged to be sufficient. For a more restricted voltage variation a
straight-line approximation is often used, and over such a limited range the capaci­
tance is linear.

Example J·2
Let us now determine the capacitance of a physical configuration consisting of two
perfectly conducting concentric spheres. The radius of the inner sphere is R and
that of the outer sphere, R.. The space between the two spheres is half filled" with
polystyrene having a relative dielectric constant of 2.7. The remainder of the space is
filled with air, &8 shown in Fig. 1-5.
Solution: Let us begin by connecting two leads to the inner and outer spheres, &8
shown in Fig. 1·5. Terminal variables q and v are arbitrarily assigned to these two
leads, also &8 shown in the figure. The only restriction on the assignment of these
terminal variables is that energy flows into the device if the two variables have like
signs. These reference conditions correspond to the original conditions in Fig. 1-1.
Coordinat••, .I.ment., and .fat. function. 7

Now if we construct a. concentric spherica.l Gaussi80n surface between the two spheres,
the electric displacement vector isl

D_.-L ..
4orT.
where q is the eh80rge on the inner sphere, with an equal and opposite charge existing
on the outer sphere, r is the arbitrary radius of the Gaussian surface, and a, is the unit
ra.dial vector.
The respective electric fields in the polystyrene and air are given by
D q
Es,017. - - - -
2.7'0 4'1'1'1(2.7'0)
a.
and

The potential of the inner sphere with respect to the outer sphere is designated by v in
Fig. 1-5. By definition, this potential difference is given as

v - - !cE'dl
where for the contour C we shall choose a radial path from the outer sphere to the inner
sphere, thus making
dl - -dr a,
On solving for v by substituting the electric fields into the contour integral, we have
q ( ~ (R.+ R.)f2 --+
v--
4....
R.
dr
2.7r'
f R.
(R.+R.)f2
­dr)
r'
which reduces to
q [ 1 + 3.4 1]
v - 4.... 2.7R. 2.7(R. + R.) - R.
The capacitance, ILl! seen at the two terminals, is given by the slope of the q versus v
characteristic curve. For fixed values of R. and R. this characteristic is a stra.ight
line, and therefore
c_ 4..••
farads
1/2.7R. + 3.4/2.7(R. + R.) - l/R.
By using fa - 8.854 X lO-u'farad/m, and by expressing R. and R. in meters, we
would have the capacitance C in farads.

1·2 THE ELECTRICAL-INDUCTANCE ELEMENT


For the inductance parameter the discussion follows steps similar to those
for the capacitance element. The symbol for the linear inductance,
I For 80n excellent review of electrostatic fields in material bodies see R. Plonsey
and R. E. Collin, "Principles and Applications of Electromagnetic Fields," chap. 3,
McGraw-Hill Book Company, New York, 1961.
8 Principles of electromechclflical·energy conversion

shown in Fig. 1-6, is described by the differential equation


v(t) = L di(t) (1-7)
dt
which states that the voltage across the inductance element is proportional
to the time rate of change of the current through the element. If the cur­
rent is expressed in amperes, the time in seconds, and the potential rise in
volts, the proportionality constant has units of henrys. Integrating Eq.
(1-7) with respect to time gives us

i(t) = ~ f. v(t) dt (1-8)

assulling that i at time 00 equals zero.

In order to interpret the integral in Eq. (1-8), let us review Faraday's


law for the voltage induced in a single closed circuit or coil. In terms of
the magnetic-flux-density vector, this basic law clln be written as

eind = - ~ /s B . dS (1-9)

With refererwe to Fig. 1-7, B is the magnetic-flux-density vector, in webers


per square meter, and dS is the vector increment of surface area directed
outward as shown in Fig. 1-7. In order to have a magnetic-flux-density
vector B in the direction shown in Fig. 1-7, a current must be inserted in
side 2 of the coil c. Therefore, comparing reference conditions in Figs.
1-6 and 1-7, we must have
v (1-10)
in order to have the positive side of the terminal voltage on the input cur­
rent line. Substituting Eq. (1-10) into Eq. (1-9) and integrating with
respect to time gives us

(1-11)

The surface integral of the magnetic-flux-density vector over a surface


enclosing, or capping, the coil is defined as the magnetic flux going through,
or linking, the coil. If the coil has more than one turn, then the sum of the

Fig. 1-6 The lumped-inducta.nce symbol.


Coordinat••, .lements, and stot. functions 9

Fig. 1-7
law.
Illustration of Faraday's
-----2~~J i

surface integrals taken for each turn will be involved in Eq. (1-11). The
total of these surface integrals represents the total flux linkages of the
entire multiturn coil. Therefore, symbolically
Mt) - X( - 00) (1-12)
where the flux linkages X have units of webers. The unit weber-turns is
also used; but the two are really equivalent, since turns are dimensionless.
The flux linkages in the coil at time - 00 is taken to be zero~ thus
Eq. (1-8) becomes

i(t) = i Mt) (1-13)

upon substituting Eq. (1-11) and then Eq. (1-12) into Eq. (1-8). For the
linear-inductaQce parameter the flux linkages are proportional to the
applied current. An inductance has the ability to accumulate flux link­
ages, and in the process a current must somewhere exist.! A description
of this element is given by plotting the flux linkages as a function of the
current setting up this linking field. The defining-characteristic curve
for the lumped-inductance element is shown in Fig. 1-8. The slope of
this curve at any point is L(i), the inductance. As before with the capaci­
tance element, a straight-line plot of X versus i gives a constant value for
L and the element is linear, whereas a curved plot of X versus i makes L a
function of either Xor i and the element is nonlinear. "Linear" or "non­
linear" simply states that the X versus i plot is straight or curved, respec­
tively. The fundamental coordinate is that of flux linkage. The first
1 Notice that the statement does not say that this current must exist in the same
inductance element. In our work with electrical circuits, the mutual-inductance
parameter is formulated from flux linkages in one element due to current in another
element.
10 Principl•• af elecfromecmlnieal•.,..,.gy conver.ion

Fig. 1-8 The defining-eharacteristic curve for


i the lumped-inductance element.

time derivative of this coordinate gives the common circuital variable,


volta~e.

Example 1·3
Figure 1-9 shows a cross-sectional view of a cylindrical iron frame. On the round
center post of the frame a coil of N turns is wound. The diameter of this cylindrical
center post is d meters, and the height is I meters, as shown in Fig. 1-9. A88uming
the magnetic-intensity vector to be very nearly zero everywhere except in the center
post, determine the inductance of the winding. Figure 1-10 gives a B-B hysteresis
loop for the iron.
Solutwn: The inductance can easily be determined once we have the proper
characteristic curve for the device. This involves a. plot of the coil flux linkages as a
function of the coil current.
Ampere'" circuital law is very helpful in determining magnetic fields in this type
of system. This basic law, in integral form, is given by

¢cH • dl - Is J . dS
where H is the magnetic-intensity vector, in ampere-turns per meter, and J is vector­
current density, in amperes per square meter. The law tells us that the line integral
of the magnetic-intensity vector around any closed contour is equal to the surface
integral of the vector-eurrent density taken over any surface whose boundaries are the

- d

r-- .....---,
I I
+
~ lax
,..-Contour

I - Ii I
I
I
I
I :{ ~
~
N turns
I

•I
L ____ - - J+
.0--­

'­- Fig. 1-9 ero_tiona! view of


a cylindrical magnetic device.
8

Third-order opproximotion

Normol magnetization curve

Fig. 1-10 Normal magnetization curve and hysteresis loop for the iron core. A
third-order approximation to the magnetization curve is also shown.

chosen contour. Once a direction has been chosen for walking around the contour,
thus defining the vector dl, the incremental-surface vector dS has its direction estab­
lished by the right-hand rule as follows: If the fingers of the right hand are curled in
the direction of the contour, then the thumb of the right hand defines the direction
for dS. This is a very important physical law, and the reader should spend sufficient
time to review itS meaning.
By applying Amp~re's circuital law to the contour shown in Fig. 1-9, we have

H.l - -Nt) (1-14)


where H. ill the magnetic-intensity vector in the poIIitive::r; direction in the center poIIt.
The minlUlsign l'eCIults from the fact that the total current crosses every surface cappinl
the contour in the direction out of the paper, while the right-hand'thumb points into
the paper for the chosen contour direction. In the remainder of the cylindrical
structure we are 8.88uming H to be zero beca\Ule the path reluctance is much less than
that of the center post.
One hysteresis loop for the iron is shown in Fig. 1-10. The normal magnetization
curve joins the tips of the hysteresis loops and is a single-valued function approximatinl
the actual B-H relationship. Any normal magnetization curve can be written analyt­
ically in the form of a power series or polynomial expansion. Owing to the first- and
third-quadrant symmetry shown in Fig. 1-10, the power-series expansion contains
only odd powers. Expanding H as a function of B gives

The dashed curve in Fig. 1-10 terminates this series after the cubic term. On sub­
stituting the third-order approximation for H in Eq. (1-14), we have

(l-U5)
12 Pr;rK;;p'e. of eredromechoni«l'-energy connrs;on

By fonning the surface integral of B over each of the N turns, the total flux
linkages can be expressed &8
N

). - ~ /s> B,,· dS.


1:-1

where the symbolism S. is used to mean the surface over the kth turn. The vector
direction of each dS" must be chosen to correspond to the assumed current direction in
the kth turn and the right-hand rule. Thus, referring to Fig. 1-9, wrapping the
fingers of the right hand around the ooil in the direction of the current, the thumb
points up; hence
dS. - -dS"..
Now since B is constant in the center P08t and therefore is constant over the area of
each t1frn, the total flux linkages are given by
). __ NB....d"
4

where ...d"/4 is the area of the center post. Multiplying this area by B. gives the total
flux in the center post linking all N turn8, so by multiplying by N we have the total
flux linkages. Upon solving for B" in terms of ). and substituting into Eq. (1-15), we
have
(1-16)

where
and

By using Eq. (1-16), the )...q characteristic curve can be drawn. The incremental
inductance or slope of the characteristic curve is given by

henrys

In this example the inductance is expressed &8 a nonlinear function of the flux linkages
rather than the current. In theory, Eq. (1-16) oould be solved for). &8 a function of q
and the inductance could then be expressed &8 a function of q.

Example 1-4
U8ing the magnetic a.ssembly shown in Fig. 1-9, con8ider cutting an air gap of g meters
in the center post &8 shown in Fig. 1-11. Con8idering the permeability of the iron to be
much greater than that of air, and neglecting fringing fields at the air gap, now deter­
mine the inductanee of the winding.
Solu.tion: Since the penneability of the iron is many orders of magnitude greater
than that of air, it is reasonable to assume that the magnetie-vector intensity in all
of the iron, includip.g the center post, is very nearly zero compared with that in the air
gap. Again using Amp~e's eireuital law with the contour shown in Fig. 1-11, we
have simply

H"g - -Nq
Ir,:~,::::"
..
I
t Contour

I'
r
-q
f
I

Fig. 1-11 Cylindrical magnetic


•I
device with an air gap cut in the
IL ____ - - . JI
center post.

where H. is the component of the magnetic-intensity vector in the air gap in the direc­
tion of the unit vector a.. Thus, the magnetic-intenaity vector in the air gap is ,iven
by
H - - !!ja.
9
and for the vector-magnetic-flux density we have

B - "oil - - "oNq a.
9
where ". - 4r X 10-' hIm is the permeability of air.
Since fringing fields in the air gap are being neglected, the magnetic-f1ux-ciensity
vector in the central iron cylinder must be equal to that in the air gap. By forming
the surface integral of B over each of the N turns, the total flux linkages can be deter­
mined to be
N
A- 2: Is. B•. dS. - "oNl(~/4) q
I-I '.\
The plot of Aversus q is now a straight line with constant slope. The inductance,
which is the slope of this ).. vel'8us q plot, is therefore
L _ ~ _ (rNd)1 X 10-' henrys
dq 9
upon substitution of the value for ,... Notice that cutting a small air gap in the iron
path results in a linear inductance eliminating the nonlinear characteristics of the iron.
The magnitude of the inductance is greatly reduced by the air gap.

1·3 THE ELECTRICAL DISSIPATIVE ELEMENT

The lumped-cape.eita.nce and -inductance elements will be shown to store


energy only in electric and magnetic fields, respectively. All electrical
14 PrillCiple, of electromechanical· energy converlion

Fig. 1·12 An incremental tube of the conducting


1:.5 medium.

systems dissipate some energy per unit time in the form of heat. The
term "dissipate" is used because the heat energy is usually lost by the
Ielectrical system to its physical surroundings. A third element is required
to account for this energy dissipation. In a conducting medium, the
current~density vector J is related to the electric-field vector E by 'an
expression of the form
J = O'E (1-17)
In this equation, 0' is the conductivity of the medium, in mhos per meter.
:For a linear medium, 0' is constant; but, in general, q may be a function
of either J or E.
Figure 1-12 shows an incremental tube of the conducting medium of
cross-sectional area I:J.S and length I:J.l. The tube is so selected that J is
normal to the surface I:J.S and is relatively constant over the surface.
Also, E does not change over the length I:J.l. The two ends of the tube are
equipotential surfaces, and any current entering one end exits through
the other end. No current Bows through the side walls. The surface
vector I:J.S is collinear with J and E; so by forming the scalar product of
I:J.S with Eq. (1~17), we have

J I:J.S =q I:J.S E I:J.l (1-18)

after multiplying and dividing by I:J.l. The quantity J I:J.S is the total cur­
rent in the incremental tube, and E I:J.l is the potential difference from one
end of the tube to the other end.
The incremental tube can be modeled by the lumped-resistance el~
ment whose symbol is shown in Fig. 1-13a. By using q for current and >.
for potential difference, Eq. (1-18) can be written as
q = G(>')>' (1-19)
where the admittance G = q I:J.S/l:J.l has units of mhos and I/G = R is the
resistance with units of ohms. The defining-characteristic curve for the
resistance or admittance element, as shown in Fig. I-13b, is a plot of q as a
Coordinot••, ./ement., and .tate function. 15

1
i
Gmhos Slope = NIH =GO,)•
Rohms

(a) (b)
Fig. 1-13 (4) The lumped-resistance symbol. (b) The defining-charaeteristio
curve.

function of X. If the conductivity of the medium 0' is constant, then G is


also constant and the q-X characteristic is a straight line, thus defining
a linear dissipative element. However, if 0' is a function of q or X, then
the admittance G is also a function of q or Xand we have a nonlinear dis­
sipative element. In that case, G(X) is the slope of the q-X characteris­
tic curve at any point.

1-4 BASIC MECHANICAL-TRANSLATIONAL COORDINATES


In our future work two types of mechanical systems are encountered.
These are mechanical systems in
1. Translational motion
2. Rotational motion
A particular system may actually be undergoing both types of motion
simultaneously. Once the concepts are formulated for either of these
mechanical modes, the other will become intuitively obvious. Let us
begin with translational motion.
The lumped-linear-mass element shown in Fig. 1-14 can be described
by the following differential relationship:
f(t) = M dv(t) (1-20)
dt

Fig. 1-14 The symbol for the lumped-m&88 element.


16 Principle, of e/ectromKoonic:al·energy c:onverrion

The force of f newtons, Fig. 1-14, is positive when directed downward.


The brace is used to indicate that the velocity of v m/sec is measured
with respect to a stationary reference frame. Both sides of the mass are
moving with the same velocity, so it would be trivial to draw v as the
velocity of one side of the mass with respect to the other side. The small
arrow shown on the brace indicates that v is taken positive when directed
downward. Newton's second law for the linear case states that the net
applied force on the mass is proportional to the time rate of change of its
velocity, with the proportionality constant M expressed in units of kilo­
grams for the mks units of force and velocity.
Now integrating Eq. (1-20) gives us

v(t) = M1 J'_ .. f(t) dt (1-21)

taking the velocity at time - ac to be zero. The integral term in Eq.


(1-21) is defined as the momentum p; therefore, we can write

p(t) - p( - ac) = r.. f(t) dt (1-22)


With the momentum at time - 00 equal to zero, Eq. (1-22) becomes
1
v(t) = M p(t) (1-23)

with the momentum and velocity simply proportional in this linear case.
The defining characteristic for the lumped-mass element in general is
a plot of its momentum as a function of its velocity as shown in Fig. 1-15.
The slope of the momentum vs. velocity plot at any point is the mass of
the element. If the plot is a straight line, then the mass IS constant and
we have a linear-mass element. If the plot is curved, the element is non­
linear with the mass being a function of the momentum, or of the velocity.
As with all of the electrical elements, "linear" or "nonlinear" is deter­
mined by the defining characteristic being straight or curved.
Momentum is chosen as a basic mechanical coordinate for a transla­
tional system. Taking the first time derivative of the momentum gives
us the force applied to the mass. As before with the electrical coordinates,

Fig. 1-15 The defining characteristic for the


v lumped-mass element.
Coordinates, element., c",d state function. 17

K m/newton
(complioncel

Fig. 1-16 The symbol for the lumped-spring element.

the first time derivative yields a variable with which we have had more
previous association, in this case a mechanical force. Since the velocity
v has been used in the description of the mass element, the second basic
mechanical coordinate should be the integral with respect to the time of
the velocity, or simply position.
The lumped-linear-spring element, shown symbolically in Fig. 1-16,
can be described by an equation of the following form:

vet) =K df(t) (1-24)


dt
where f is the force, in newtons, applied to the spring, v is the velocity, in
meters per second, of one side of the spring with respect to the other side,
,and K is a proportionality constant of the spring, in meters per newton.
The constant K is termed the compliance of the spring. By manipulating
Eq. (1-24), a more common expression for the spring parameter can be
obtained. IJ}tegmting both sides gives us

f(t} I'
= K1 _.. vet) dt (1-25)

taking the force at time - 00, f( - 00), to be equal to zero.


The integral in Eq. (1-25) is simply the position of the top of the
spring, Fig. 1-16, with respect to the bottom. Symbolically, this is
expressed by

x(l) - x( - 00) = f . vet) dt (1-26)

Taking the position of the top measured with respect to the bottom at
time - 00 to be zero gives us
1
f(t) = K x(t) (1-27)

as the defining equation for a linear spring. Usually in mechanics the


spring constant, which is the reciprocal of the spring compliance, is used
in an equation with the form of (1-27). However, in order to maintain
complete symmetry with the electrical elements, compliance will always
be used.
18 Principle. of electromechanit:Ol-energy convenion

Slope =K(fl
Fig. 1·17 The defining characteristic for the
f
lumped-spring element.

A plot of position vs. applied force, as shown in Fig. 1-17, defines the
general lumped-spring element. The slope of the characteristic at any
poibt is the compliance of the spring. A straight-line plot gives a constant
compliance K and a linear spring. Similarly, a curved plot yields K as a
function of the position x or the force f and the spring is nonlinear. As in
all previous cases, "linear" or "nonlinear" is a description of the defining
characteristic.
The fundamental mechanical coordinates are momentum and posi­
tion. The first time derivatives of these coordinates are force and velocity,
respectively. 1 The dissipative, or viscous, damping element shows the
use of these first time derivatives. Figure 1-1& pictorially shows this
element. The force applied to the element is shown as p, and the velocity
t represents the velocity of the top side with respect to the bottom side,
as indicated by the brace and arrow in Fig. 1-18a. The defining rela­
tionship for a linear-damping element is given by
p = Dt (1-28)
where D is in units of newton-seconds per meter, using consistent mks
units. A plot of p versus X, shown in Fig. 1-18b, gives the general defining
relationship for the damping element. The slope of the plot at any point
is the damping coefficient D(x). Again a straight-line plot represents a
linear element and a curved plot represents a nonlinear element.
I Very often in beginning mechanics courses the scheme of selecting force and
position as basic variables is used. In view of the discussion in this chapter, this
certainly appears to be a rather hybrid selection. In beginning electrical-circuit
theory a more consistent choice of current and voltage is used.

Slope =f) Ii )
Fig. 1-18 (a) The sym­
bol for the lumped-viacous
damper. (b) The defining­
(al (bl characteristic curve.
Coordinates, elem8lll$, ond stote functions 19

Example 1-5
A common electromagnetic audio loudspeaker is shown in Fig. 1-19. The voice coil is
wound on a cylindrical member attached to the apex of a right~ircular paper cone.
The edges of the paper cone are rippled and attached to a rigid-eircular frame. A
constant radial magnetic field (not shown in Fig. 1-19) generates a force on the voice
coil proportional to the coil current. This mechanical force of electrical origin is given
by !(t) in Fig. 1-19. As the apex of the cone moves horizontally in response to the
force !(I), the front and back surfaces of the cone push and pull on the surrounding air
and generate sound waves. Determine a lumped-element model for horizontal
translational motion of the voice coil.
Solution: The mass M of the moving system consists of the voice coil, the cylin­
drical coil form upon which it is wound, and some portion of the paper cone. Owing
to the rippled edges of the paper cone, a spring compliance will tend to restore the
moving system to some particular equilibrium point in the absence of a force I<t).
The air surrounding the paper cone contributes a damping force opposing the motion
of the system. In Fig. 1-19 the variable i; is used to define the velocity of the voice coil
with respect to the stationary frame of the speaker. The small arrow on the brace
indicates the velocity to be positive when the voice coil moves to the right.
Figure 1-20 shows a lumped-element model for the system assuming only transla­
tional motion. The model system in Fig. 1-20 is so drawn that the motion is in the
vertical direction with i; defined positive when the mass M is moving downwa,d.
The force !(I), when positive, is trying to make i; negative; thus!(t) is shown upward.
The compliance of the paper cone and the damping effects of the air are given by the
lumped elements K and D.
The compliance coefficient K could easily be determined experimentally by plac­
ing a static force on the voice coil and measuring the static displacement. By plotting
displacement VB. force, a eurve similar to Fig. 1-17 is obtained. The slope of the
curve at any point is the compliance. The effective mass M and damping coefficient
I D are not quite so easy to determine. The defining~haracteristic curve for the m8.88

\,
~

Fig. 1-19 The moving system of an


electromagnetic audio loudspeaker.

\\
20 Principles of e/ectromechanical·energy conversion

.J
,l
Fig. }·20 A lumped-element model of the loud­
speaker.

element, given by Fig. 1-15, requires a plot of the momentum of the moving systems as
a fun~tion of its velocity. Both momentum and velocity are difficult quantities to
measure. Simply weighing the elemente in the moving system would give a reasonable
approximation to the effective mass. Similarly, the damping coefficient is defined as
shown in Fig. 1-180, where we must plot force VB. velocity. Such measurements are
not easily made. Certain indirect tests, such as step- and frequency-response
measurements, are generally used for determining these parameters. These methods
are discussed in detail in later chapters.
The force or gravity is not included in the model system shown in Fig. 1-20. We
are not hanging a mass M by a spring and damper from a rigid support in a gravita­
tional field. Figure 1-20 is a lumped-element representation of the actual system
shown in Fig. 1-19. The lumped-element model experiences translational motion in
only one direction, and for convenience we have chosen this direction to be vertical.
The equilibrium equations for the model have a correspondence to the actual system
only over a limited operating range for which the model parameters are valid. Usually,
the most difficult part of any analysis is the determination of the model and the evalua­
tion of the parameters.

1-5 BASIC MECHANICAL-ROTATIONAL COORDINATES

All the lumped-rotational elements are defined in an analogous fashion.


The fundamental coordinates for rotational motion are angular momentum
land angular'position 8. The use of these rotational coordinates can
readily be demonstrated by their application to the three lumped elements
of inertia, rotational spring compliance, and rotational viscous damping.
The inertia element is pictorially shown as the rotating disk in Fig. 1-21.
The angular veloc~ty of the disk with respect to some reference frame is
indicated by w. T~e small curved arrow on the brace is used to indicate
the positive sense of w with respect to the stationary reference. The torque
applied to the inertia is shown as the symbol T. Notice that when wand
T have like signs, they are in the same direction. For a linear element
these two variables are related by the equation

T(t) = J dw(t) (1-29)


dt

Coordinates, e'ements, and dote Function. 21

:{ ----+---if ~/t----
Fig. 1-21 The symbol for a lumped-inertia

element. 'WffH/rHHH/ffHrRHWffff///J.

where w is in units of radians per second, the applied torque is in newton~


meters, and the inertia has consistent units of kilogram-meters squared.
Integrating Eq. (1-29) gives the angular velocity as

w(t) = 1/1
J _00 T(t) dt (1-30)

with w( - 00) taken to be equal to j/;ero. The integral in Eq. (1-30) is


defined as the angular momentum, and thus

l(t) - l( - 00) = f~oo T(t) dt (1-31)

The units for angular momentum 1 are newton-meter-seconds, according


to Eq. (1-31). When l( - 00) is taken equal to zero, Eq. (1-30) becomes
simply

w(t) = J1 l(t) (1-32)

"\ In general, the description of the inertia element is accomplished by


plotting the angular momentum l as a function of the angular velocity, as
shown in Fig. 1-22. The slope at any point is the value of the inertia,
with a straight-line plot giving a constant inertia value. If the plot is

Ir curved, then the element is nonlinear and the inertia is a function of the
angular momentum or the angular velocity. Angular momentum is the
basic coordinate being used. Notice that the first time derivative of this
coordinate 1 gives the common torque variable. The angular velocity w
is used in the defining characteristic of Fig. 1-22. The second basic coor­
dinate is angular position 8. The first time derivative of this coordinate,
namely, 9, is the angular velocity w.

Fig. 1-22 The defining characteristic of the


lumped-inertia element.
22 Principle. 0' eledromechoniCQI.energy conversion

Fig. 1-23 (a) The sym.


bol for the lumped-tor­
sional compliance element.
(b) The defining charac­
(al (b) teristic.

The lumped-torsional compliance element, shown pictorially in Fig.


1-23a, is defined by the characteristic of Fig. 1-23b. For a linear charac­
teri,tic, the defining relationship can be given as

T = i. 8 (1-33)

with the compliance K. in units of radians per newton-meter. For a


nonlinear-torsional-spring element a more complicated expression relating
the twist to the applied torque must be used in order to describe the curve
of Fig. 1-23b.
The lumped-rotational viscous-damping element in Fig. 1-24a is
defined by the characteristic of Fig. 1-24b. The angular velocity II indi­
cates the speed of one side of the damper with respect to the other side.
The applied torque 1is shown positive in the direction of positive angular
velocity. For a linear element Fig. 1-24b is a straight line and can be
described by
l "= D,II (1-34)
where the rotational-viscous coefficient D, has units of newton-meter­
seconds per radian.
The similarity between the mechanical-translational and -rotational
systems is obviously very close. In our later work we shall often refer
to simply the mechanical-position coordinate to indicate either transla­

o
~

--1].. . - ­
-:)I-,-i
Fig. 1-24 (a) The
symbol for the lumped­
rotational viscous-damp­
ing element. (b) The
(al (b) defining characteristic.
CoordilKlte., elemenfl, and Itate function' 23

tional position x or rotational position (J. As an abbreviation the symbol


for translational position is retained in general formulations, with the
understanding that the rotational position coordinate (J should be inserted
if the mechanical portion of the system is actually rotational, and not
translational. The same argument applies to the usage of the momentum
coordinates p and l.

1-6 SUMMARY OF THE BASIC COORDINATES AND THE


LUMPED-ELEMENT RELATIONSHIPS
Table 1-1 summarizes the basic coordinates for the three systems to be
used in our electromechanical studies. Of the six basic coordinates in
these three types of systems, only the mechanical position x and rotational
position (J are commonly used as basic variables. Most often the first
time derivatives such as voltage and current are used as fundamental
variables. The reason for this choice rests in our measuring abilities and
devices. Voltage is far easier to measure than flux linkage, and current
can be measured more readily than charge. Similar arguments apply to
force and torque compared with their respective momenta. Angular and
translational position are sometimes easier to observe than velocities.
Quite often, therefore, the position coordinates x and (J are used as vari­
ables. Thus in the case of four of the six coordinates the first time
derivative yields some phenomenon which is readily measurable, and for
this reason these variables are more common in our beginning work.
However, in order to show the physical processes involved in the electro­
mechanical devices we shall study, the six more basic coordinates are to
be used.

TABLE 1-1· Summary of the basic coordinates and their first time derivatives

Coordinatea Fi1'llt time derivativea


System Descrip­
Sym­ 8ym­
Description Units Units
hoi hoi tion

Electrical q Charge coul q-i Current amp
>. Flux linkage webe1'll ~ - v Voltage volts
Mechanical p Momeqtum newton-sec P -I Fqrce newtons
(translational) ~ Position mcters ;i; - v Velocity m/sec
Mechanical l Angular newton-m­ 1 - T Torque newton-m
(rotational)
IJ
momentum
Angular
poIIition
soo
rad , .... Angular rad/sec
velocity
24 Principles of electromechanical-energy conver.ion

TABLE 7·2 Summary of the lumped-element symbols

Capocitance Inductonce Conductonce

Electrical
~
J C(i)=d~aA ] "il.d~ all ] alii. :i
Mass Spring Damper

]"lil'~ JKIPI'~ jo",.~


Mechonical
(transiationol)

Inertia Torsional spring Rotationol damper

Mechanical
(rotationol) ~ ~

. a9
~
. ai
J(Q)=~ K,(l) '" (if 0,(91 = d8

For each of the three physical systems, three lumped elements have
been introduced. Table 1-2 summarizes these nine elements, showing
their schematic symbol and their individual defining differential relation­
ship. For example, the lumped-capacitance element is defined by plotting
the charge q on the plates of the capacitor as a function of the applied
terminal voltage >.. Thus the capacitance for any particular charge or
voltage is the slope of this plotted curve as given by dq/dX. A Iinear­
capacitance element has a straight-line plot, whereas a nonlinear element
has a curved characteristic. However, in either case the capacitance is
still clearly defined as the slope of the characteristic curve. Notice in
Table 1-2 that all other elements are defined in a simila'l: fashion.

1-7 EQUILIBRIUM EQUATIONS, STATE VARIABLES, AND


STATE FUNCTIONS
Assume that we wish to describe the instantaneous operation of a system
composed of the lumped-electrical and -mechanical elements discussed in
the preceding sections. This description could be adequately given by
expressing both basic coordinates for each element as functions of time.
For example, consider a single capacitance element. If the instanta.neous
charge on the capacitor, in addition to its defining-characteristic curve,
were known, then the complete operation for each time instant would be
Coordinates, element" and sfate functions • 25

known. Here only one basic coordinate has to be specified, since the
characteristic curve provides the second coordinate. Consider, as a sec­
ond case, two mutually coupled inductances. The flux linkages in either
coil are a function of both the current in the same coil and the current in
the magnetically coupled coil. In order to specify the instantaneous
operation of the first coil, both its total flux linkages and coil current would
have to be known. The characteristic curve for the element would give
the portion of the flux linkages due to the current in the same coil. A
second characteristic curve is required to provide the flux linkages in the
first coil due to current in the second coil. For a complete description of
the first coil, currents in both coils are required with the characteristic
curves in order to specify the total flux linkages and thereby describe the
operation.
The general procedure for determining the system coordinates as
functions of time generally involves solving a set of simultaneous differ­
ential relationships known as equilibrium equations. These equations
dynamically balance specific terminal variables for the lumped elements
according to the interconnections of the elements in the system. For
electrical systems, the equilibrium equations are formulated by means of
Kirchhoff's current and potential laws. These two laws generate formal­
ized procedures yielding loop equatians or nodal equations, depending on
how the coordinates are chosen. Equilibrium equations for mechanical
systems are generally formulated by means of Newton's second law or
D'Alembert's principle. involving a summation of forces or torques. For
noninteracting electrical or mechanical systems these procedures are
adequate to yield the system equilibrium equations.
The variables used in the formulation of the equilibrium equations
are often called slate variables. The term "state variable" is used because
the equations specify the state of each element and therefore the state of
the system. For example. consider a capacitance element having a char­
acteristic curve as shown in Fig. 1-25. At a given time instant the charge
on the capacitor has a value qQ' The state of the capacitor at this time
instant is therefore defined by point P in Fig. 1-25. The quantities qa

Fig. 1-25 The state point for a lumped-capaci­


tance element.
26 Principle. of elecfromKhanic:ol·enerQY conversion

and the corresponding ~.. are the values for the state variables q and ~
at the specified time instant. The point P on the characteristic curve is
termed a 8tate point. As time progresses, the charge on the capacitor will
vary, and equivalently the state point will move along the characteristic
curve. Thus the motion of the state point as a function of time also
describes the operation of the element.
For electromechanical systems, involving an interaction between the
electrical and mechanical parts of the system, the formulation of the
equilibrium equations directly using Kirchhoff's laws and D'Alembert's
principle is often very difficult. To aid in the formulation problem, an
energy technique will be used. This technique requires the formulation
of energy-8tate functions. These functions of the state variables at any
sp~cific time instant have values determined by the values of the state
variables at that time instant. Therefore, the values of the state func­
tions at a specified time depend only on the state points (or equivalently
on the state variables) at that time instant and in no way on the past
history describing how the state points were reached. As the state func­
tions are formulated in the following sections, these ideas will be discussed
further. The energy formulation gives the same equilibrium equations
as obtained by Kirchhoff's laws or Newton's law. The main justification
for this energy method rests in the routine procedures involved, par­
ticularly in the case of electromechanical systems.
The energy analysis will first be carried out for con8ervative electrical
and mechanical systems. The term "conservative" means that the system
has only the ability to store energy; energy cannot in any way be con­
sumed or dissipated. Thus, for electrical systems, only capacitance and
inductance elements are to be included; initially, no resistance elements
will be considered. Similarly, for mechanical systems, damping elements
will be omitted. However, bear in mind that suitable procedures will
finally be discussed to include these dissipative elements. Also, we shall
limit our considerations to electrical and mechanical systems which are
noninteracting. These are the types of systems considered in previous
courses in circuits and mechanics, respectively. This will afford an oppor­
tunity to check and compare the energy formulation with formulations
that are familiar to us. In the next chapter the broader class of elec­
tromechanical systems involving electrical and mechanical interactions
will be discussed in detail.

1-8 THE CAPACITANCE STATE FUNCTION


Let us begin by developing certain ideas concerned with energy storage
in the capacitance element. Referring to Fig. 1-26, the instantaneous
Coordinates, element., ond dote function. 27

Fig. 1-26 A single lumped-eapacitanee element.

power into a single capacitor is given by


P.(t) = }.(t)q(t) (1-35)
where >'(t) and q(t) are respectively the terminal voltage and current vari­
ables expressed as functions of time and the subscript e indicates electric­
field storage. The energy per unit time flowing into the capacitance
element is being stored in an electric field. The increment of energy sup­
plied in a time dt is given by

dW.(t) = P.(t) dt = }.(t) dq(t) dt (1-36)


dt
The total energy supplied from some initial time to to some later time
t is obtained by integrating Eq. (1-36), giving us
W.(t) - W.(t o) = f,: }.(t) d~~t) dt
W.(q) - W.(qo) = }qO
('I >"(q') dq' (1-37)

where q is the charge at time t and qo is the charge at time to. Notice how
the functional dependence of the voltage >" changes from t to q' as the
integral is manipulated to an equivalent form. In Eq. (1-37) the primes
are used to depote the variables involved in the integration. The
unprimed variables are the limits of the integral, and in the case of q they
define the state of the element at any time t.
Usually we shall be interested in the total energy stored in the electric
field of this capacitance element. If time to is so chosen that qo = 0, then
the energy stored at this time is zero, meaning
W.(t o) = W.(qo) = 0 (1-3S)
Then Eq. (1-37) reduces to

W.(q) = lol }.'(q'} dq' (1-39)

Equation (1-39) is the total electric-field energy stored in the capacitance


element at any time to Figure 1-27, showing the defining characteristic
for the capacitance element, gives a simple interpretation for this integral.
The total energy stored at any time t is the shaded area above the defining­
28 Principle, of electromechanical-energy c;onve,.ion

q'

Characteristic curve

Fig. 1·27 A graphical inter·


pretation of the electric-field en­
ergy- and ooenergy-state functions.

characteristic curve. Thus the electric-field energy storage can easily be


determined graphically by simply counting the squares contained in this
area.. For each instant in time, q( t) has a different value, and thus this
area above the curve, whi ch is the energy stored, changes in a fashion
determined by the characteristic curve.
An energy term in the form of Eq. 0-39) is a state function. Up to
this point we have described the state of a capacitance element by specify­
ing either its charge q or voltage X. If the present state of the capacitor
whose characteristic is shown in Fig. 1-27 is at point P, then specifying
either q or X would give the state of the element. This same state point
could also be defined by giving the shaded area above the curve, with only
one point P corresponding to a certain specified amount of area. Thus,
the function defining this area also defines the state of the system.
The state function depends only on the final state of the element.
The final value of the charge, at time t, is all that determines the energy
stored for a given defining characteristic. With a single-valued charac­
teristic curve, the energy stored is completely independent of the manner
in which the final state, at time t, is reached. In other words, between
some initial time to and the final time t the charge on the capacitor could
assume a wide range of values, but the energy stored depends only on the
final state of the capacitor at this time t. If the characteristic curve is
not single-valued, then the energy stored depends on the manner in which
the final state is reached. If such hysteresis phenomena actually exist,
a single-valued characteristic is used as a first approximation. Calcu­
lations involving double-valued characteristics are analytically very
difficult.
A second energy-state function has importance in our future work.
Referring to Fig. 1-27, the shaded area below the characteristic curve can
Coordioote., element" and .tote function. 29

be given by

W~(A) = 10). q'(X') A' (1-40)

The quantity W~(A) is known as the coenergy stored in the capacitance,


and it is denoted by the prime. This quantity is an energy-state function
in the same way that the energy stored is one. If we specify the area
below the characteristic curve, then the state of the element is uniquely
defined. Only one state point P could exist with a certain given area
below the curve. Thus the coenergy-state function is every bit as good
a description as the energy function given by Eq. (1-39).
For a specific final state of the capacitance element at time t, given
by q(t) and A(t), the energy and coenergy are related by
W.(q) + W~(A) = qA (1-41).
Equation (1-41) simply states that the total area in the rectangle with
sides q and A is equal to the energy, which is the shaded area above the
characteristic curve in Fig. 1-27, plus the coenergy, which is the shaded
area below the curve. For a given characteristic curve the energy and
coenergy functions are not independent of each other.l In the special
case of a linear element the characteristic curve is a straight line and the
energy and coenergy are each equal to half the area of the q by Arectangle.
However, with nonlinear elements these two functions are, in general, not
equal and care must be used to a void interchanging the two state functions.

1-9 ENERGY-STATE FUNCTION FOR A SYSTEM OF CAPACITANCES

So far the electric-field energy- and coenergy-state functions have been


developed for only a single element. These ideas can readily be general­
ized to the systeJ!! of capacitance elements shown in Fig. 1-28. In a very

I Since each of these functions specifies the state of the element and the element has

one unique state, the sta.te functions quite obviously cannot be independent.

Fig. 1-28 A system


composed of n lumped
capacitances.
30 Principle. of electromechanical-energy conversion

general case the voltage across anyone of the capacitors could be a func­
tion of the charge on all n capacitors. For the usual capacitance element
the voltage can almost always be expressed as a function of the charge
on that capacitor alone. However, let us keep open the possibility of the
mutual-capacitance parameter. The total energy stored in the system
of Fig. 1-28 can be given by

(1-42)

witH the final energy stored being a function of the charges on all n capaci­
tors. The prime on the coordinates is used to indicate the general charge
or voltage on the capacitance elements, and the unprimed coordinates refer
specifically to the state-point coordinates. Therefore, the unprimed coor­
dinates will specify the actual state of the elements and, ultimately, the
state of a system of interconnected elements. This added complication of
having primed and unprimed variables arises because we have to perform
integrations with respect to a certain coordinate from one value of the
coordinate to another value of the coordinate. We must distinguish
between the variable over which we are integrating and the limits on this
integration. Hence the primed and unprimed variables.
In order to evaluate Eq. (1-42), we must give some consideration to
how the final state of the system is reached. Suppose, for simplicity, that
the total final charge on each capacitor is deposited in sequence. Thus,
with q;, q~, . . . ,q~ all equal to zero, q~ is raised from its initial value of
zero to its final value, at time t, indicated as simply q" without the prime.
The first integral on the right side of Eq. (1-42) becomes
(q~ ~ I I , " ( ql . , , ,
Jo }.\(qltq:, . . • ,q,,) dql -+ Jo }.\(qltO, . . . ,0) dq\

Similarly, the second integral is evaluated from

where q~ is a constant held at its final value ql throughout the integration


over q;. All other charges q~, q~, . . . , q~ are zero in this integration.
The remaining integrals are all evaluated in a similar fashion. The value
of the state function is independent of the manner in which the final state
is reached. We have evaluated the total energy stored by increasing the
various charges, from zero to their final value, in sequence. The final
answer would be the same if all of the n capacitors had been charged
simultaneously. Only the final state is important, and none of the ele­
CoordifIGte" e'ement" and ,tate function' 31

ments care how they got there. 1 As an abbreviation, Eq. (1-42) is written
as
.
=
.-12: fo'l' X:(q~,q;, . . . ,q~) dq:
W.(ql,q" . . . ,q..) (1-43)

However, remember that the evaluation of Eq. (1-43) involves holding


(n - 1) of the charges at some value as the integration over the ith charge
is performed.
The coenergy-state function can be developed in a similar fashion.
By symmetry, from Eq. (1-43) the total coenergy is given by
.
W:(X1,X" ... ,X,,) = 2: fo~' q~{X~,X~, ..• ,X:) dX;
1-1
(1-44)

The discussion of the evaluation of the energy equation (1-43) applies to


the coenergy equation (1-44). Here all voltages but the ith one are fixed
as the integration proceeds over the ith element. As with the single
element, the energy and coenergy are constrained by some relationship.
In the general case, the relationship takes the form .

W.{qhq2, . . . ,q..) + W:(XI,X2, ... ,X..) = L" q,X, (1-45)


i-I

Example ,·6
Three nonlinear capadtors have the following defining characteristk-s:
ql - 3(kd~ q. - 4(k.)~i q. - 7{k.)~i
Determine the energy- and coenergy-state functions for this system of capacitors.
Also, show that the sum of these two state functions satisfies Eq. (l-45).
Solution: Let' us start with the coenergy-state function. If we increase the
three voltages successively from zero to their final-state values kIt k., k., then Eq.
(1-44) for the coenergy becomes
,
W.(k"k.,k.) -
r~' "
10 r A"I A.
ql(k"O,O) dA, + 10 q.(k"k.,O) dA. + 10 q.(""k.,k.) U".
f , ( I. I J{ I

By substituting the defining relations for the three capacitors, we have

W'(k I, k 2, ka) -
Al 3(k')~' dA' + hAt hAl
0 4(k')~i dA' + 0 7(k')~ dA'

h0 1 t 2: , J •

which reduces to give us


W;(k"k.,k.) - lk," + _'l;~k.~i + 51lk.'i (1-46)
as the electric-coenergy-state function.
I This is similar to a golf game in which, after poorly hitting three shots on a par 4
hole, the golfer hits a 20()..yard approach into the cup. He scores a par as hill final
ate, and tha.t is all that is recorded on the score card.
32 Principle. of electromechanical-energy conversion

Similarly, Eq. (1-43) for the electric-energY-6tate function can be written as


W ( ) ( o. ( I , I
• q"q.,q. - }o ",(quO,O) dq, + }oO. "i(q"q.,O)
( ( I , (0., "
dqi + }o ~I(q"q.,q.) dq,
,

Solving the defining relationships for the voltages in terms of the charges and sub­
stituting lead to

W(
,qt,q.,q. ) (fl.' (1j"q,'
-}o (q'(l1;i'.' ) *
') dq,• +}o (Il' (tfJ.)'
dq.' +}o l ' dq.•

This equation reduces to


1 2 1
W.(q"q.,qa) - 108 qt' + 5(4)U q.* + 6(7)' q.' (1-47)

8.8the energy-etate function for the system.


Let U8 now evaluate the sum of the energy- and coenergY-6tate functions. Taking
the fust terms in Eqs. (1-46) and (1-47), we have

(~~,H + 1~ q,,) - (~~,~, + 1~ ~I~i)


where we have used the defining relationship between the state points ql and ~I.
These two terms combine to give
3~,~s _ ql~'

thU8 verifying the first term in the summation of Eq. (1-45). In a similar fashion the
remaining two terms of this summation can be checked for these two state functions.
Just for practice, complete this verification.

1.10 PARTIAL DERIVATIVES OF THE ELECTRIC-FIELD STATE FUNCTIONS


In the formation of equilibrium equations using the energy-state functions
we shall be required to calculate certain partial derivatives. The manner
by which the equilibrium equations result from the state functions can be
shown in the following way, using the lumped electric-field elements as an
example. By formulating the partial derivative of the electric-field
energy stored in our system of n capacitors with respect to the charge on
the kth capacitor, we have
OW.(ql,q2, . . . ,qk, . ,q..)
.""---~=
Oqk

In performing a partial differentiation of a function of many va.ria.bles, all


q's except qk are held constant. Thus, from Eq. (1-43) we have only
OW.(qJ,q2, ..• ,qk, ,q.. )

Oqk

O~1r. '?l /001 ;':(q~,q~, . . . ,q;, ••• ,q~) dq;


Coordinotes, e/emenls, ond dote functions 33

which reduces tol


aW.(qJ,q2, . . . ,qk, . • • ,q,,) _ ' ( ) (1 A8)
" - /\/1; qhqs, . . . ,q/l;, •.. ,q" '""*
uq/l;
By taking the change in the energy stored in the system with respect to an
incremental change in the charge on one of the capacitors and holding the
charge on all of the other capacitors fixed, we obtain the voltage on that
capacitor. Kirchhoff's loop equations are formulated by summing volt­
ages around certain loops in an electric circuit. Equation (1-48) gives us
a scheme for obtaining the voltage across a capacitance element from a
particular energy-state function.
Notice that these partial derivatives have been taken with respect to
a particular charge qlc. Voltage terms resulted from using charge as the
variable of interest. If loop equations are desired, currents or q's are
selected as variables of interest. Then all our manipulations with the
energy-state functions will involve taking partial derivatives with respect
to q's or 4's. If node equations are used, rather than loop equations, then
node voltages are the variables of interest. In that case partial deriva­
tives should be taken with respect to particular X's or X's. Taking the
partial derivative of the electric-field coenergy-state function with respect
to a particular voltage Xk gives us
aW~(Xl,X2' . . . ,X", . . . ,X,,)

'aX/I;

= at i !o~' q;(X:,x~,
4: i - I
. . . ,X~, . . . ,>.~) dX~ (1-49)

referring to Eq. (1-44). Equation (1-49) reduces to


aW~(XhX2" •.. ,XIc, ••• ,X..) (" , ;.) (1 50)
aXk = q/l; /\1,/\2, • • • ,/\k, • • • ,/\.. ­

Thus the change in the coenergy-state function with respect to the voltage
on the kth capacitor, when all other voltages are held fixed, gives the
charge on the kth capacitor. Taking the total first time derivative of both
sides of Eq. (1-50) would yield the current through this kth capacitor.
Nodal equations are formulated by summing currents at each node.
Therefore, selecting X's or X's as coordinates of interest will give rise to
nodal voltage-equilibrium equationlf when partial derivatives are formed
with respect to these coordinates. After the energy and coenergy func­
tions for all of the remaining electrical and mechanical elements are
formulated, the procedures whereby equilibrium equations are obtained
from state functions will be systematically formalized.
I See W. Kaplan, "Advanced Caleulus," p. 220, Addison-Wesley Publishing Company,

Inc., Reading M888., 1956. AlllO see Prob. 1-15.


34 Principle. of electromechanical-energy conveTl;on

In developing the voltage across the kth capacitor in Eq. (1-48),


suppose that we wished to work with the coenergy function instead of
the energy-state function. Using Eq. (1-45) the same partial derivative
with respect to qt becomes
aW.(ql,q2, . . . ,qt, . . . ,q,,)
oqt

(1...51)

where all >"'s in the last two terms are available as functions of the q's.
Expanding Eq. (1-51) we have

(I-52)

Notice how the chain rule is used in taking OW;;Oqk. Using the result
given by (I-50), the summation is seen to be equal to zero, and thus
Eq. (1-52) gives the same result as (1-48). Equation (1-51) clearly
shows that a partial derivative of an energy function is not equivalent to
the same partial derivative of the corresponding coenergy function. The
result given by Eq. (1-50) can also be obtained in a similar fashion from
Eq. (1-45) using the energy-state function instead of the coenergy
function.

1.11 ENERGY-STATE FUNCTIONS FOR A SYSTEM OF INDUCTANCES


Figure 1-29 shows a system of n lumped inductances. The flux linkages
X and current q in each inductance element are appropriately indicated.
The total instantaneous power supplied to the n inductances is given by
(1...53)
where the SUbscript m on the power P indicates that magnetic-field energy
is being stored. The total magnetic-field energy stored in an increment
of time dt is

dW",(t) = liI{t) dX~it) dt + q2(t) dX;~t) dt + + q... (t) dA,,(t)


dt
dt

(1-54)
Coordinat••, ./emenl., and .tal. funclion. 35

Fig. 1-29 A system com­

posed of n lumped induct­

ancel.
~-----}::
If the energy stored at t = 0 is taken to be zero, then the total energy
stored from time 0 to time t is given by

= 10' (Mt) d'A~it) dt + 10' q2(t) ~% dt


"
W .. (t)

+ ... + 10' q..(t) d'A;,?) dt (1-55)

which can be reduced to

(1-56)

where X" X2, • • • • X". without the primes, are the respective flux linkages
at the time t. By using the symbolic notation of the preceding section,
.
Eq. (1-56) can be condensed to
.
W...(X 1,X 2 • • • • ,X..) = 2 Io~' q~(X~,X;,
,.1 .
• • • ,x:) dX; (1-57)

Notice the complete duality between Eq. (1-57) for the magnetic-field
energy storage in a system of n inductances and Eq. (1-43) for the electric­
field energy storage in a system of n capacitances. In the magnetic-field
case, the coordinates X and q are involved, as opposed to q and Xfor the
electric-field case. This is not surprising, since the original defining char­
acteristics of Figs. 1-28 and 1-29 have the same symmetry.
The coenergy-state function for the system of n inductances is given
by

(1-58)
36 Principl•• of ./ectromechanico/-en.rgy conversion

Also, the energy and coenergy are related through an expression similar
to Eq. (1-45) as follows:
.
W...(X I ,X 2, ••• ,X,,) + W~(ql,q2' ... ,q,,) = L: X;q;
i-I
(1-59)

When the partial derivative of the magnetic-field energy taken with


respect to X is formed, terms required for a nodal formulation should
result. Performing this differentiation leads to
oW...(X h X2 , ••• ,Xk, ... ,X,,)

aXk

(1-60)
I
which reduces to
aw.. (XItX" •.• ,Xk, .•. ,X,,)
aX =
. ('\ '\
qk Al,A2, •••
'\ '\ )
,Ak, • • • ,A"
k
(1-61)
Thus, taking the change in the total magnetic energy stored with respect
to a change in the flux linkages of the kth element, when all other flux
linkages are held constant, gives the current in the kth inductor. Again
observe how the selection of X as the variable of interest, for a nodal
formulation, leads to a current term. Performing a similar operation
with Eq. (1-59) and the coenergy-state function gives the same result.
The partial derivative of the coenergy function with respect to a
particular current leads to
aW~(qlJq2, ... ,qk, . . . ,q,,)

Oqk

.. ,q~, . . . ,q~) dq: (1-62)

from Eq. (1-58). This expression becomes


OW~(ql,q" ... ,qt, . . . ,q,,) _ '\ (. .
,.. - Ak ql,q" ,qt, . . . ,q,,) (1-63)
uqt

which, when the first total time derivative of both sides is taken, gives us
the voltage across the kth inductance. Loop equations involve currents
as variables; thus, taking a derivative with respect to a current gives us a
voltage term which we can then sum in the loop formulation of equilibrium
equations.
Thus far we have considered the applicable state functIons for the
capacitance and inductance elements. The admittance or resistance ele­
ment has a different type of state function. In the capacitance and
inductance elements, energy is stored in electric and magnetic fields,
Coordinates, e'ements, and stote function. 37

respectively. The energy-storage terms are the state functions we have


been discussing. With the dissipative element, no such energy storage
occurs. All energy flowing into the element is converted into heat, and it
is lost to the electromechanical system. For this reason the required
state functions for the dissipative elements will be formulated after the dis­
cussion of all conservative or nondissipative elements is completed.

1·12 TRANSLAnONAl MECHANICAl-ENERGY-STATE FUNCTIONS


The inertial and spring elements in the translational- and rotational­
mechanical systems are conservative elements. Energy storage in a mass
or inertia element is in the form of kinetic energy. The total instantaneous
power supplied to a system of n lumped-mass elements is given by
(1-64)

where :t.(t) is the velocity of the ith mass and p.(t) is the force applied to
the ith mass element. In an increment of time dt the kinetic energy
received by the system is given by

dT(t) = XI(t) dvJ ?) dt + X2(t) dv ;?) dt

+ ... + x.. (t) dV..dt(t) dt (1-65)

where T is the symbol for kinetic energy, the Xi are the respective veloci­
ties, and the Vi are the respective momenta. Taking the kinetic energy
at time tOto be zero, the total kinetic energy stored at some later time
t is given by

T(t) = 10 XI(t) dvJit) dt + fot X2(t) dV;?) dt


1

+ ... + 10(I x.. (t) dV,,(t)


dt
dt (1-66)

which becomes

,v..) = 10('" ./(VlIV2'


Xl
, , ... ,V..') dVI'
+ 10 XI VI,V2,
( "• • /( I I
. • . ,v..
') dV2'

+ . : . + 10 ( "• • /(' I
x.. Vl,V2, • • . (1-67)

Using our abbreviated notation, the total stored kinetic energy is given by
..
T(Vl,VS, •.• ,V..) = L fo"; x~(v:,v~, ... ,V:) dv~
i-I
(1-68)
38 PrillC;pl•• of .lectromechonicol-energy conversion

p'

Characteristic curve

p(t}

~ a

Fig. 1-30 A graphical in­


terpretation of the kinetic­
energy- and -coenergy-state

.; (tl x functions.

Equation (1-68) gives the energy stored in an element as the area


above a characteristic curve. The integral in Eq. (1-68) has the form
Jop x' dp', which is the shaded area above the characteristic shown in Fig.
1-30. Equation (1-68) is sufficiently general to allow the velocity of any
one of the n mass elements to be a function of the momentum of all the n
elements in the system. Practically speaking, we do not have a mutual­
mass parameter, but should one come along, our formulation is sufficiently
general for its inclusion.
The kinetic-coenergy function is the shaded area below the charac­
teristic curve in Fig. 1-30. For the system of n inertias the kinetic­
coenergy-state function has the form
..
T'(Xl,X2, ••. ,x,,) = l
i-I
Jot. p~(.t~,x;, ... ,x~) d:t; (1-69)

These two state functions are related by


.
T(Pl,P2, ..• ,p,,) + T'(Xl,X2• ••. ,x,,) =
.-1.r PiX. (1-70)

which is a general statement that the total rectangular area for a given
state point is the sum of the respective energy and coenergy functions.
Equilibrium equations for mechanical systems are most often formu­
lated on the basis of Newton's second law or D'Alembert's principle for
both translational and rotational systems. Equations in this form are
summations of forces or torques, respectively. With these mechanical
systems, the dual formulation corresponding to loop equations for elec­
trical systems exists, but it is not commonly used. We sum forces from
free-body diagrams, but very seldom do we sum velocity-type terms.
CoOl'Clinotee, e1_eb, and date fund/one 39

The reason for this preference lies in the fact that velocity terms would
involve the momentum as coordinates, and our interest is not usually
centered around the system's momenta. The force formulation, on the
other hand, involves position as coordinates, and most often the sys­
tem's configuration is of prime interest.
With the position coordinates of major interest, all partial derivatives
of state functions should be taken with respect to z's and x's. Thus
aT' (XI,X" • • • ,:tk. . . . ,:tIl)

aXk

a ~
= ax ( ~, ,( ., .,
~ Jo Pi Xl,X"
.,
• • • ,X/o, • • •
.') "'...,
,x" ai i (1-71)
"-1
which gives us simply
aT'(x"x" ... ,x/o, ... ,x,,) (.. . . ) (1-72)
a:t,. = Pit XI,X" . . . ,X/o, . . . .x"

Taking the first total time derivative of both sides of Eq. (1-72) gives us
the required force on the kth inertia element as a function of the n veloci­
ties'in the system. With all of the mass elements that we shall consider,
the momentum will be a function of only the velocity of that element.
Thus in Eq. (1-72) Pic will be only a function of XII. The formulation of
the kinetic-energy- and -coenergy-state functions is presented in a com­
pletely general way in order to parallel the development of the correspond­
ing functions for the electrical elements.
For the translational spring, energy is stored in the form of potential
energy. A system of n springs, in an increment of time dt, could change
its potential-energy storage by an amount
dV(t) = PI~t) d Xl(t) dt + p,(t) dz,(t) dt

dt dt

+ ... + P.. (t) dx,,(t)


dt
dt
,
(1-73)

where the symbol V is used for potential-energy storage. Equation (1-73)


is obtained by multiplying the total power (energy per unit time) input
to the n springs by an increment of time dt. The Pi are the forces on the
springs, and the dx./dt terms are the resulting velocities; thus Pi dx./dt is
the power into each spring element. Selecting time t = 0 such that the
potential energy is zero [that is, V (0) = 0], the total stored potential
energy in the n springs at some later time t is given by
V(t) = (' . (t) dXI(t) dt
Jo PI dt
+ )0(' P2. (t) dx,(t)
dt
dt

+ ... + )0(' P..(t) dx..(t)


dt
dt (1-74)
40 Prin<;iple, of electromechanical· energy converaion

which is equivalent to

(1-75)

With the abbreviated summation notation used with the previous ele­
ments we can write Eq. (1-75) as
..
V(Xl,X2, •.. ,x..) .... 1 /0'" P:(X~,X~,
_-I
(1-76)
I
Notice the complete symmetry of Eq. (1-76) for the potential-energy
storage in a system of n springs and the kinetic-energy storage in a system
of n inertias, given by (1-68), with the interchange of p's and x's between
the two state functions.
For completeness, we shall also formulate the potential-coenergy
functions as being
.
" {Pi '( , .,
1., )0 Xi Pl,P2' . . . ,P..') d Pi
I
(1-77)
i-I

and the relationship between these two state functions is


n
+ (1-7S)
V(Xl,X2, . . . ,x..)
.-1
V'(Pl,P2, . . . ,P..) = LXiV.

As discussed with the mass element, all our mechanical-equilibrium equa­


tions are usually written as summation of forces, with the x's and .t's as
the coordinates of interest. Therefore, partial derivatives are to be taken
with respect to x's or .t's, and we use the potential-energy function.
Taking the partial derivative of Eq. (1-76) with respect to Xk, we have
aV(Xt,X2, . . . ,Xk, .•. ,x..)

aXk

= i
a~·,_1 /0'" pHxl'x~, ... ,x~, ... ,x~) dx; (1-79)

Equation (1-79) reduces to


aV(XhXt, . . • ,x.. ,x,,) .( )
--'-=-..:.:....--a=--X""'k'-----'--"- = Pk XI>X2, . . . ,X.\:, .•• ,x.. (I-SO)

which is the force across the kth spring as a function of the position
coordinates.
Coordinal••, .I.menl" and .101. fllndion. 41

Similar energy and coenergy functions can be developed for the


rotational-inertia and spring elements. The development is absolutely
identical with that for the translational mass and spring elements with
the substitution of the rotational momentum l for translational momentum
'P and rotational position IJ for translational position x.

1.13 SUMMARY OF THE CONSERVATIVE ELEMENT ENERGY- AND


COENERGY-STATE FUNCTIONS
Table 1-3 gives a complete summary of the energy- and coenergy-state
functions for the six nondissipative or conservative element~. The energy
and coenergy functions are not independent of each other. For a set of
state points represented by ak and Pk we can always write J

Energy (a1,a:, . . . ,an)


It

+ coenergy (81,82. .tJ..) L aiPi (1-81)


i-I

Table 1-2, which indicates the defining relationships for all the elements,
shows that the conservative elements are always defined as a relationship
between one basic coordinate and the first time derivative of the other
coordinate. Thus, for example, the rotational inertia element has a" = lk
and Pk = ~".
In evaluating the functional relationship of the form
n

l fo"" P~(al,a2, . . . ,a.) da;


,-I
(1.82)

remember that the system can arrive at its final-state points in any fashion
with relationship (1-82) still maintaining the same value. This means
that with all the other a'S equal to zero, a~ could be increased from zero to
its final value al. Then by holding a~ constant at this final value aI, a~
could be increased from zero to its final value a2. In turn, each of the
other a~ could be increased to its final value as the summing index i
proceeded from 1 to n. Any otlier scheme leading to the same final-state
points ai, a2, . . . , an would give the same value for (1-82). For
example, each of the a;could be increased in turn from zero to half its
a;
final value. Then, with each at half its final value, each could in turn
be increased to the final value. The result would be the same as with the
total increase previously described. Very often, the actual system will
have all its coordinates changing simultaneously from zero to their final
values.
t;

TABLE 1·3 Summary of the conservative element energy· and coenergy-state functions

Element Energy-state functions Type of energy function

Capacitance
W.(ql,qt, . . . ,q.) "" L" f'~' ,
i-I
0 ,(q\>q"",
1 " dq,
,q.) Electric-field energy

W.(~I'~"
1
. . . • ~.) "" L" ~~' , 0 q,(~u~t.
11 . . ,~~) d>.~ Electric-field coenergy
i-I

Inductance
W...(>'I,>'., . . . ,>'.) '" L
n
~X' ,
0 q'(>'I'>"" , d>.,f
.. ,>'.) Magnetic-field energy
i-I

W''"(q l,qt,
. * ••
q) =
J" Lf'
i-I
n

0 >.'(qf
i uq,,···
.f .f) dq'i
,q" Magnetic-field coenergy

.
Mass (translational)
T(pl,PI, ..• ,p.) "" Lt
i-I
0
Pi
Xi(PIIP",
" 0 0 1 dp,t
,P.) Kinetic energy (translational)

..
T
too
(XI,XI, • 0 • ,x.) -
0 L t;t, 0
flo' of
p,(tux . . . ,x.) dx,
"
0'
Kinetic coenergy (translational)
i-I
"'----_..........

TABLE 1·3 Summary of the conservative element energy· and coenergy·state functions (continued)

Element Energy-state functions Type of energy function

Spring (tran8Iational)
. Lt'"'
" , ,,
V(x\,Xt, ••• ,x.) - 0 p,(x\,x, • •.. ,x.) dx. " Potential energy (translational)
i-I

,. . . L" t
V (PI,P" ••. ,P.) -
P '.'.'
0 ' x,(PltP,•.•. . ' dp.
,P.) .' Potential coenergy (trane1ational)
.-1
Inertia (rotational) "
T(ll.l" ••• •1.) - Lf"'11
.-1 ('
8.(lI,l" . . • , ) dl. Kinetic energy (rotational)

T'(8\,8, • .•.• 8.) - L" t i , I ,


0 , l,(8 1,8" ... ,I..) d i
, (,
Kinetic coenergy (rotational)
.-1
Spring (rotational) .
V( 111.11" ••• ,II.) - Lt"I'
i-I
0 "
,(111.11,•••• ,II.) dll, Potential energy (rotational)

,I.) ­ L 10;' II:(/;,t~,


"
V'OI.t" . • . • • • ,1;.) dl; Potential coenergy (rotational)
.-1 ~

t
44 Prirn:iple$ of electromechanical-energy conversion

PROBLEMS I

1.1 For values of applied voltage between zero and v.. a certain capacitor can be
dellCribed by the characteristic equation

q - q.. sin' .!!. for 0 S v S v..


2v..
where q.. is a constant. For values of applied voltage greater than v.. the character­
istic equation is simply

q - q.. for v Z v..


a. Is this a linear or nonlinear capacitor? Explain your answer.
b. Sketch the characteristic curve for positive values of the applied voltage.
c. Determine the capacitance of the device for all positive values of the applied
voltage. Sketch the capacitance on the same plot with part b.
d. What is the peak capacitance value and at what voltage does it occur?
1-2 Determine the capacitance per unit length between two very long perfectly
conducting cylindrical tubes. The radius of the inner cylinder is Rit and the radius
of the outer <'ylindl'r is R,. The inner half of the space between the cylinders is filled
with polYHtyrcne having a relative dielcetric constant of 2.7. The remainder of the
space iH fillC!1 with air. Is this eylindrical capacitor linear?
1.3 The inductance of a magnetic-field device is as shown in Fig. PI-3.
a. Is the device linear or nonlinear? Explain your answer.
b. Sketch the characteristic curve for the device, and label all slopes and important
axis values.

L",/2 ----J------,

i Fig. Pl-3

1-4 Two windings are placed on the same iron core. The first winding has 800
turns, and the second winding has only 500 turns. With a current in the 800-turn
winding the following data are taken:
Current in 800 turns, amp o 0.1 0.3 0.4 0.5 0.6 0.7 0.8 0.0 1.0
Flux linking 800 turns, kilolines o 13 38 49 56 60 64 67 69 71
Flux linking 500 turns, kilolines o 13 38 47 53 57 59 61 63 64
Carefully draw the following curves on one sheet of graph paper.
a. Plot the characteristic curve for the 8oo-turn coil in mks units.
b. Plot the flux linkages in the 500-turn coil as a. function of the current in the
800-turn coil.
I Answers to odd-numbered problems are listed at the end of the book.
Coordinate" elements, and ,tate functions 45

;
l c. From part!! G and b determine the self-inductance of the 800-turn coil and the
[ mutual inductance between the two coils. Plot these inductances as a function of the

t
!"
current in the 800-turn coil.
1-5 The radio telescope shown in Fig. Pl-5 consists of a large antenna attached
by means of a tower-like structure to a large base. The telescope can be turned by
means of a motor connected through appropriate gearing to the base. Let the tor­
sional compliance of the tower be K, rad/newton-m. The base and antenna inertias
are J Band J .. kg-m', respectively. The torque applied to the base by the motor is T ..
newton-m. Assume that the wjndage torque on the antenna is proportional to the
angular velocity with a coefficient of D, newton-m-ilec/rad.
G. Draw a lumped-element model for the radio telescope. Indicate all sources
and the angular velocity of the base and antenna on this model system.
b. In tcrms of WB and "'A write expressions Cor the kinetic-cnergy- and -eoenergy­
state functions T(lA,l,,) and" T' (WA''''8) or T'( dA,b,,).
c. Write expressions Cor the potential-energy- and -eoenergY-iltate functions
V(8.. ,IJs) and v'(iA,IB).

I
t
II Fig. Pl-5
I
1-6 The charge VB. voltage characteristic Cor a particular electric-field device can
t be approximated by

v - aq' + bq
where and b are constant!!.
G
G. Determine the capacitance of the device as a Cunction oC the charge.
b. Find the clcctric-encrgy- and coencrgy-statc fun('tions. Evaluate these CURC­
tions Cor q - 3 coul.
c. Check Eq. (1-48) from the results oC part b.

I 1-7 The slope of the characteristic curve for a certain capacitance element is
given by
I ~ - aq./l-.'
f•
I where G and q:
are constant!!. When the voltage v acrO!!ll the capacitor is zero, the

I
cbarge II is al80 zero.
Coordinate" element" and dote function' 47

1-11 The average length of each turn on the inductor in Prob. 1-10 is 7.0 in.
Il.Find the resistance of the winding if copper wire w.ith a diameter of 0.028 in. is
used. (The conductivity of copper is listed in the table of physical constants.
Appendix.)
b. Sketch the characteristic curve describing this lumped-resistance element.
c. For a certain temperature increase the conductivity decreases by 10 per cent.
Does this temperature change affect the curve plotted in part b? If it does. sketch
the new curve.
1-12 When a step-current function is applied to a certain lossy capacitor, the
voltage reaches 63 per cent of its final value in 2.5 sec. The capacitor can be modeled
by a linear lossless capacitor in parallel with a linear resistor. With a d-c voltage
source across the lossy capacitor the power dissipated is 85 watts.
Il. Determine the electric-field energy stored in the system, with the voltage
source connected.
b. Also find the electric-field coenergy under the conditions of part Il.
1-13 According to Einstein's special theory of relativity. the mass m of a particle,
instead of being a constant. is given by
m.
m - -:r:==='F'F7.
i VI - (vic)·
I
where m. is the rest mass (or the m&88 at zero velocity), c is the velocity of light, and v
!

is the velocity of the mass m.
Il. Draw the characteristic curve for the relativistic mass element by using the
relativistic velocity vic as the abscissa. Remember that II cannot be greater than c.

t b. Determine the kinetic-coenergy-state function.


c. Show that for II «c the state function in part b reduces to the usual moV'/2
, for the nonrelativistic mass element.
1-14 For the nondissipative elements the area above the characteristic curve in
the rectangle correspondinll: to a lI:iven state point represents energy storage by the ele­
ment. For examples, see Figs. 1-27 and 1-30. ' What does the corresponding area for

the three dissipative elements represent?


1-15 Show that'the manipulation leading to Eq. (1-48) is correct. This same
proof applies to reducing Eq. (1-49) to the form of (1-50). and also to the partial
derivatives of the energy- and coenergy-state functions for the remaining lumped
elements.
!
Hint: By usinll: the reference cited in the note on page 33 show that

() !cll.•,".(q"q•• . . . ,q.), , ",,(q••q••


• . . . ,q.)
dq" ""
(lq. ()

Since the system can reach its final state in any manner, assume that the kth element
is the last to be assembled. Therefore all terms in the summation will not contain q.
except
46 Principle. of electromechanical-energy cOlIYersion

a. Ie this element linear or nonlinear? Justify your answer.


b. Find an equation giving the characteristic curve for the element.
c. Find the coenergy-state function W:(v).
d. Use Eq. (l-41) to find the energy-state function W.(v).
e. Write an expre88ion for the capacitance C of the element. Is the energy
stored equal to -lCv'? Explain.
1-8 Three magnetically coupled coils have the following characteriatic rela­
tionahips:

", - 6i , - 3i. + 4i. ". - 4i, + i. + 7i.


where" and i are the respective flux linkages and currents.
a. Find the coenergy-state function for the aystem W~(i ..i.,i.) by raiRing the
currents from zero to their state values in numerical order.
I b. Repeat part a raiaing i. first, then i" and then finally i. to their respective atate
values. Why muat thia answer be identical with that to part a?
c. What feature of the characteristic equations makes part a equal to part b?
d. Invert the characteristic equations, giving the currents as functions of the flux
linkages, and find the magnetic-energy-state function W .. (",,"""')'
e. Is the energy-state function equivalent to the coenergy-state function for any
particular state point?
f. Verify Eq. (l-59). Also show that 0-61) and (1-63) are valid for "I and i l .
1-9 The magnetization curve for a WOO-turn iron-core inductor is as follows:

Coil current, rna 5.4 8.6 13.5 22.5 27.4 38.2 54.0 65.2

Core flux, kilolines 66 100 133 160 167 173 180 183

a. Plot the characteristic curve for this inductance element. Plot flux linkages,
in weber-turns, as a function of the coil current, in amperes.
b. For a coil current of 30 ma graphically determine the magnitude of the energy­
and coenergY-fltatc function.
c. Graphically determine the inductance, in henrys, as a function of the coil
current. Plot the inductance VB. coil current on the same graph as used in part a.
d. If the inductor has a resistance of 100 ohms, determine the approximate
terminal voltage for a coil current of (30 + 2 sin 377t) milliamperes. Note: 1 weber ­
10· lines ~ 10' kilolines.
1-10 An inductor is constructed by winding 10,000 turns uniformly around a
toroidal iron core with a rectangular cross section. The toroid has an inside diameter
of 4.0 in. and an outside diameter of 6.0 in. The height of the toroid ia 2.0 in. The
magnetization curve for the iron core is as follows:

Flux density, kilogauss 0 2 4 6 8 10 12 14 16 18


Magnetic intensity, oersteds 0 18 30 37 43 48 54 63 88 156

a. Determine the characteristic curve for the element in mb units.


b. At what value of direct coil current does the maximum inductance occur?
c. What ia the value of the maximum inductance?
d. For the current determined in part b find the magnitude of the energy- and
coenergy-state functions.
Hint: By uRing the conversion table in the Appendix, convert all units to mks
units. Use 5 in. as a mean diameter in calculating the path length.
chaptn

EQUILIBRIUM EQUATIONS FROM

2 ENERGY -STATE FUNCTIONS:

LAGRANGE'S EQUATION

In the preceding chapter we have developed energy­


and coenergy-statc functions for the six con!lcrva­
tive (Le., nondis8ipative) lumped elements in elec­
trical, translational-, and rotational-mechanical
systems. Table 1-3 gives a complete summary of
these functions, in each case completely showing
functional dependence. The state functions shown
in Table 1-3 arc much more general, in many cases,
than most practical clements require. Particularly
with the translational-mass and rotational-inertia
elements, the momentum of anyone clement is usu­
ally a function of only the velocity of that one single
element. The capacitance clement also, most gen­
erally, has a similar property. The voltage Oil a
single capacitor is usually not a function of the
charge on any other capacitors in the system. At
this point in the development, the state functions
for the non conservative clements have not been con­
sidered. Energy storage does not exist with this
type of element, and so the scheme used with the
conservative clements does not apply. After con­
sidering how these state functions aid in the formula­
tion of equilibrium equations, we shall add these
nonconservative elements into the formulation.

2-1 GENERALIZED FORCES FROM ENERGY-STATE


FUNCTIONS
As these state functions were being developed in·
Chap. 1 certain partial derivatives were arbitrarily
taken. Essentially, we found that the voltage or
48 current for the electrical elements, force or velocity
Lagrange's equation 49

TABLE 2-1 Summary of the state·function partial derivatives

Coord i-
Element nates or Proper partial derivative Type of analysis
interest

Capacitance ilW·~·-.:_.:-,I!.'!l ".(q" Potential summa-


q, q = ,q.) tion (loop)
aq. :
• aw:(~. ,k.} Current summation
>., " - q.(k" ,>...} (nodal)
ak.
Inductance aW..(>. .. ... ,>'.) Current summation
>., k -.--.----~-
.. q.(>.., ,>'.) (nodal)
a>..
I aW~(q" ,q.) Potential summa-
q, q = >'.(ql, ,q.) tion (loop)
aq.

Mass (trans. p ) Velocity Bummation


lationa\) p,p
t • ~ ~ , It
.. ;t.(p" . .. ,p.)
aplt
aT'(x" .. · ,x.)
Xt i
I

ax.
=­ P.(x•• .. . ,x.) Force summation

Spring (trans­ aV(x., .. • IX..) Force summation .


lational) x,x .. P.(x, •. .. ,x..)
itx.
aV'(p" ... ,P.)
p,p .--~.,.-.--.
.. x.(p., ... ,P.)
~.--
Velocity summation
apt

Inertia (rota­ itT(I" . .... ,I..) Angular-velocity


tional) I, i .----- -~.-
.. b.(l" .. ,I.) summation
al.
aT'(b" .. · ,b.)
8, b, .. Mb" .. ,b.) Torque summation
abo
Spring (rota­ aV(8" .. ,8.) Torque Bummation
tional) 8, b --­ - i.(I1" . • ,II..)
a8k
I av'a.. .. · ,i..)
I,
ai.
.. 11.0" .. . ,1.) Angular-velocity
summation

for the translational-mechanical elemelits, and torque or angular velocity


for the rotational-mechanical elements could be recovered from the state
functions by means of these partial derivatives. The variables of voltage,
current, force, velocity, torque, and angular velocity are the common
quantities used in formulating equilibrium equations, and they are often
termed gencralized forccs. The formation of equilibrium equations by
means of these partial derivatives of the state functions should certainly
50 Principles of electromechanical-energy conversion

seem reasonable, since the generalized forces can be recovered by this


process. The problem of selecting the proper partial derivative of the
proper state function becomes quite simple once the general scheme is
noticed. If we are interested in a set of equilibrium equations involving
the a and a coordinates, then all partial derivatives must be taken with
respect to a or a. Table 2-1 gives a complete summary of all such
partial derivatives for the six lumped conservative eleme~s.

2-2 MECHANICAL-EQUILIBRIUM EQUATIONS


FROM ENERGY-STATE FUNCTIONS

,the mechanical-equilibrium equations are generally formulated as func­


tions of x, X, 9, and 8, or, in other words, as functions of the position and
velocity coordinates. Equilibrium equations in this form involve summa­
tion of forces for translational systems and summation of torques for
rotational-mechanical systems. As previously pointed out, we seldom
are interested in the momentum coordinates for mechanical systems.
Most often, the position coordinates are of prime interest. From Table
2-1 we can see that for the mass and inertia elements the kinetic-coenergy
functions, rather than the kinetic-energy functions, should be used, and for
the two spring elements the potential-energy functions, and not the poten­
tial-coenergy functions, are the proper choice.
The system shown in Fig. 2-1 is composed of a lumped-spring element
and a lumped-mass element free to move only in the vertical direction.
The braces in Fig. 2-1 are used to designate the velocity of each element
with respect to the stationary reference. Thus the brace drawn across
the spring element shows the velocity of the bottom end with respect to
the top end to be x. m/sec. The small arrow on the brace shows XII to
be positive when the bottom end is moving downward from the top end,
which is held stationary. Similarly for the mass element, the brace
designates the velocity of the mass to be Xb m/sec with respect to the
stationary reference, taken positive when the mass is moving downward.
The forces PII and Pb have their reference directions so chosen that energy

Spring

lPa

I>!:" I Fig. 2-1


elements.
Unattached lumped-m8.1!8 and -spring
Lagrange', equafion 51

flows into the element when both terminal variables have the same signs.
Thus for both x" and Po positive, potential energy is being stored in the
spring. Also for both Xb and Pb positive, kinetic energy is being stored by
the mass element.
The spring element is described by a characteristic plot of the dis­
placement x" as a function of the applied force po. For simplicity assume
that the spring is at its free length (that is, p" is zero) when x" is equal to
zero. The mass element is described by a characteristic plot of its
momentum Pb as a function of its velocity Xb. From the characteristic
plot the potential-energy-state function for the spring element is given by

(2-1)

and the kinetic-coenergy-state function for the mass element is given by

(2-2)

The primes are used to indicate the variables over which the integration
takes place. The unprimed variables are the final state variables defining
the state point on the characteristic curves.
Taking the partial derivative of both sides of Eq. (2-1) with respect
to x" leads to • ,

(2-3)

which is the external force applied to the spring element. Similarly, the
external force applied to the mass element can be recovered from the
kinetic-coenergy-state function. By forming the total time derivative of
the partial derivative with respect to Xb of both sides of Eq. (2-2), we have

-d
dt
[fJT'(Xb)]
- - - - - = -d [ - 8
8Xh dt 8Xb
!c:b '(") d"] =
0
Pb xI> XI> . ('Xb )
Pb (2-4)

which is the external force applied to the mass element.


Let us now attach the mass element to the bottom end of the spring
as shown in Fig. 2-2. By this connection of the two lumped elements,
the velocity coordinates in I<'ig. 2-1 are related by
(2-5)

where x is the single velocity coordinate shown in Fig. 2-2. If xo , x., and
x are so arranged that the spring is at its free length when theya.re equal
to zero, then Eq. (2-.;) can be integrated to give
Xo=Xb=X (2-6)
52 Principle. 01 e/ectromed!onical-energy conversion

I(

Fig. 2-2 Assembled mass-spring system.

Equation (2-6) is formally known as a hOWnomic equation of constraint.


The term "constraint" is used because the connected system requires or
qonstrains the coordinates to vary in a certain fashion. Once the mass is
attached to the spring, the motion of the spring is constrained to follow
the motion of the mass. The original equation of constraint, given by
(2-5), is a differential relationship between the position coordinates. We
were able to integrate Eq. (2-5) to obtain Eq. (2-6); therefore, Eq. (2-5)
is an integrable constraint equation, and it is termed a holonomic equation of
constraint. The term "holonomic" means that the differential relation­
ship between the coordinates can be integrated to give an algebraic equa­
tion (not a differential equation) between the coordinates.
On substitution of Eq. (2-6) into Eqs. (2-1) and (2-2) for the state
functions the potential energy and kinetic coenergy become functions of
the single coordinate x. The external force applied to the spring is now
given by oV(x)/ox, and the external force applied to the mass becomes
d[aT' (x) / ox]/ dt. If a force f(t) is applied to the bottom of the mass, as
shown in Fig. 2-2, then D'Alembert's principle of dynamic equilibrium
requires that f(t) be balanced by the force applied to the mass and the
force applied to the spring. Therefore, using the energy-state functions,
we have
!i [aT'~x)] + aV(x) = f(t) (2-7)
dt ox ox
For linear mass and spring elements, the state functions for the sys­
tem shown in Fig. 2-2 would be
x%
T'(x) = iMx 2 V(x) = 2K (2-8)

where K is the spring compliance and M is the mass. By substituting


Eqs. (2-8) into Eq. (2-7), we obtain .

Mi + ~K = J(t) (2-9)

which is the equilibrium equation for the system.


lagrange's equation S3

2-3 A RESTRICTED FORM OF LAGRANGE'S EQUATION

Now let us see if we can extend these ideas to a system composed of many
lumped mass and spring elements. Before these elements are attached
together, individual force and velocity terminal variables are assigned.,
For the mass elements a kinetic-coenergy-state function can be written
in the general form

T '("Xo,Xb, • • •
~ Jo
) = L (s. p"'(""
X ..,Xb, • ••
) d"XI< (2-10)
Ie

Similarly, for all spring elements a potential-energy-state function can be


written in the general form
~ ("'., I , ,
V(x.. ,X/I, • • •) = L Jo P.(x..,x(J, . . .) dx. (2-11 )

The mass elements are designated by lowercase English-letter subscripts,


and the spring elements are denoted by lowercase Greek-letter subscripts.
Assume that the mass and spring elements are now joined together to
form a conservative translational-mechanical system. The junction
points for the lumped elements are termed mechanical nodes. Assume
that the elements are joined in some fashion at n nodes. We can now
write a set of constraint equations between the velocities defined at each
of the n nodes and the original element velocities as follows:
Xo = ftJ(XI'X~, ••• IX,,)
Xo = f.(XhX2, •.• ,x,,)
(2-12)
x.. = f,,(XI,X2, ..• ,x,,)
XII f/l(Xl'X~,:" ,x,,)

If the constraint equations given by (2-12) are holonomic, then they can
be integrated to yield relationships between the original element coordi­
nates and the new node coordinates. Substituting the integrated equa­
tions of constraint into the state functions, given by (2-10) and (2-11),
yields
(2-13)
and
f2-14)

In the next section we shall examine cases where the state functions'have
a form more general than is proposed by Eqs. (2-13) and (2-14).
S4 Principle.: of electromechanical-energy conversion

Taking the total time derivative of aT'/az" gives us


!!:.. [aT'(Zl,;i;2' . . .
dt ax"
,X,,)] -_ p"'ma.. ( . •
Xl,X2,.·. ,x..
• )
(2-15)

as is seen by examining Table 2-1. Equation (2-15) is the total external


force applied to all mass elements connected to the kth mechanical node.
Similarly, forming the av lax" leads to
aV(XI,X2, •.. ,x,,) _ ","prino: ( )
ax" - I'k XI,X2, • • • ,x.. (2-16)

which is the total external force applied to all spring elements connected
to the kth mechanical node. If Q" is a total of all forces externally applied
io the kth mechanical node of the system, then D'Alembert's principle of
dynamic equilibrium requires Q" to be balanced by the mass and spring
forces at the kth node as given by Eqs. (2-15) and (2-16). Thus we have
!!:.. [aT'(X1,X2, .' .. ,X,,)] + aV(XI,X2, .•. ,x,,) = Q"
dt aXI; aXk
fork = 1,2, .. _ ,n (2-17)

By writing one equation as k assumes the values 1, 2, . . . ,n, a set


of n equilibrium equations for the system is obtained. A solution of these
n equilibrium equations provides the motion of the system. Equation
(2-17) is a re8trictedjorm oj Lagrange's equation. An additional term must
be added to Eq. (2-17) to account for holonomic constraint equations
where the original velocities are functions of both the new velocities and
the new position coordinates. For these more complicated constraint
equations T' becomes a function of both;i; and x and Eq. (2-17) must be
extended as shown in Sees. 2-4 and 2-5.

Example 2·J
Write the equilibrium equations for the linear m8.88-8pring system shown in Fig. 2·3.
Assume that springs K K t , and K. are stretched distances a, b, and c, respectively,
when x. and Xb are equal" to zero.
Solution: In Fig. 2-4 the system is broken apart to show each lumped element
separately. Individual terminal variables are defined for each element. For
example, compliance K I has the velocity of its bottom end with respect to its top end
given by x.. The small arrow on the x, brace means that the velocity is taken posi·
tive when the bottom moves away from the top. The force pIon the compliance K, is
defined as positive when it causes a positive velocity, or when it also is directed down·
ward. The terminal variables on the mass elements have their velocities defined with
respect to the stationary support. Both velocities x. and x. are positive when directed
downward. The terminal forces applied to these m&lJSe8, namely, p. And P., Are alflo
positive in the downward direction and are applied by pushing from the stationary
support.
Lagrange', equation 55

t
t'
i
f }:.

I
t.
;"

Fig. 2-3 A simple m8.ll8-llpring Fig. 2-4 The maBII-llpring system


translational-mechanical system. di8a8llembled.

Referring to Table 1-3, we can formulate the kinetic-coenergy function by


expanding
I)

T'(x.,x.) 2: 10Z1 p:{x~,x~) d.t:


i-4

The relationships betwecn the momenta and the velocities for the linear-maBII elements
&regiven by

and 80 the kinetic-coenergy function becomes

T'(x.,:t.) - }M IX:' + iM ti.'


The potenti&l function from Table 1-3 is given by
3
V(XI,X"XI) -
i-1
2: 10%1 p:(x:,x~,x~) dx:
For the three linear springs we have

• XI
PI - K.
­

where Zit X., x. are the amounts each,spring is stretched. When X., X., and x. are zero,
the forces exerted by the springs are respectively zero. The potential-energy function
therefore expands to give
56 Principles of electromec:nanic:al.ener9Y conversion

By connecting the elements together, as shown in Fig. 2-3, constraints are placed upon
the individual terminal variables. We can write the relationships between the
coordinates x. and i. and the original coordinate variables as follows:
Xl .., x.
x, - -:to + Xb
xa =- :4
By integrating the first three of these equations, we have
XI - +a
x.
x. - +
- x. x. + b
x. - x. + c
with a, b, and c as the respective stretched positions when x. and Xb are equal to zero.
The kinetic..j)oenergY-l!tate and potential~nergy-8tate functions are now given by
T'(:t.,X!.) - !M,x.· + !M.x.' (2-18)
V(X.,Xb) _ Lx,,-±~~~ + ( -x. + Xb + b)' + (Xb + e)1 (2-19)
2K, 2K, 2K.
upon substitution of the new coordinates. The external forces acting on the mechani·
cal nodes are seen by inspection to be
Q. - -j,(t)

with a negative sign indicating opposition to the assumed positive sense for the
coordinate and a positive sign indicating that the force is in the same direction as the
positive sense for the coordinate.
Equation (2-17) for. the first coordinate becomes

By expanding these partial derivatives with the use of Eqs. (2-18) and (2-19), we have

(2-20)

Similarly, Eq. (2-17) for the second coordinate becomes

~ [ilTJ~.•,Xb)] + ilV(X.,Xb) _ Qb

cit ilx. ilx.

which expands to give us

i (M,x.) + -x. + x. + b + x. +~ _ j.(t) '(2-21 )


dt K, K.

Equations (2-20) and (2-21) are the two required equilibrium equations. A simple
free-body diagram of each of the masses shows these equations to be valid force summa­
tions according to Newton's laws or D' Alembert's principle. The purpose of this
example is t~ illustrate the detailed procedures involved in the use of the energy
formulation suggested by Eq. (2-17). For a problem this simple, such an involved
method is certainly not warranted. However, (or more complex systems the same
general routine will yield equilibrium equations when not all of the forcet! are quite 80
obvious.
Lagrange'_ equation 51

Fig. 2-5 A lumped-element model of


a motor driving a load torque.

Example 2·2
Figure 2-5 represents a motor driv.ing a rotating loed. The motor is represented by
the rotating inertia J ,. The torque developed by the motor is some explicit function of
time TI(t). The loed, represented by the inertia J., is coupled to the motor by means
of a drive shaft with compliance K,. In addition, the load contains a resisting torque,
labeled T.(t), which is some function of time. The combination of an automobile
engine, drive shaft, and rear wheels, in its most basie form, eould be represented by
the lumped system of Fig. 2-5. For this simple rotational system determine the
equilibrium equations by using the energy formulation.
Solution: By breaking the system apart, as shown in Fig. 2-6, terminal variables
are defined for eaeh of the elements. For the motor inertia J I , the angular velocity

'I.
with respect to a stationary reference is designated as II. and the torque applied to the
inertia is Similarly, the angular velocity of the left side of the shaft compliance
with respect to the right side is I, and the torque applied is I,. In the same manner,
terminal variables are defined for the loed inertia J •.
The kinetic-coenergy function for the system is formulated by referring to Table
1-3. For this example, we have
2
T'(II,.I.) - 2: !c/' t,(I~,I~) dl;

.-1

The angular momenta are related to the respective angular velOcities by the linear
relationships

and thus the kinetic ooenergy is simply

Similarly, from Table 1-3, the potential energy is given for this system by

V(e,) - 10
( " I' I
,,(8,) dB,,

Fig. 2-6 The lumped-element model disa8Bembled.


58 Prirn:iple£ of electromechanical-energy converl;on

For the linear shaft-compliance element, the torque /, is related to the angular twist by
1
1. - -II.
K,
and so the potential function becomes

When these three lumped elements are connected together into the system shown
in Fig. 2-5, constraints are placed on the individual element variables as follows:

giving the original coordinates as functions of the new "all-hooked-up" coordinates.


Inl terms of the new coordinates the kinetic-coenergy- and potential-energy-state
functions are given respectively by
T'(b.,bb) - iJ,b.' + iJ.b." (2-22)
V(II.,II.) '" (II. 2~8I1b)1 (2-23)

The assumption has been made that, when II• ... II., the shaft compliance is in its free,
or untwisted, position.
The external forces for a rotational system are in the form of torques. From
Fig. 2-5, the externally applied torque on the a mechanical node is given by

and on the b mechanical node by


Q.... -T.(t)
The sign in each ease is selected by observing whether the external torque is in the
same direction as or the direction opposite to the assumed coordinate direction. Thus
T,(t) is in the same direction as b. and Q. has a positive sign, while T.(t) is opposite to
b. and Qb has a negative sign.
Equation (2-17) for the first coordinate becomes

~ [aT'(fJ.,b.)] + aV(II.,II~ ... Q.


dI. afJ. all.
Upon expanding the partial derivatives by using Eqs. (2-22) and (2-23), we have

d 0. - II.
dt (J.fJ.) + -1[6 ... T.(t) (2-24)

which is one of the equilibrium equations. In a similar fashion, Eq. (2-17) for the
second coordinate becomes

which reduces to

(2-25)
Lagrvnge', equation S9

This is the second equilibrium equation obtained after substituting Eqs. (2-22) and
(2-23).
Obtaining the equilibrium equations by the energy method seems to be a lengthy
job. For a problem as simple as this one, the equilibrium equations could be written
by inspection. The major use of the energy formulation is for systems involving
certain interaction force terms. These force terms are often difficult to postulate by
inspection, and the energy formulation becomes a more useful tool. Consider this
example and the one preceding it to be illustrations of a new technique and not a
suggestion for complicating a problem you could have solved quite simply by using
your previous experience.

2-4 DEGREES Of fREEDOM AND GENERALIZED COORDINATES


Before generalizing the restricted form of Lagrange's equation, as given
by Eq. (2-17), we must carefully examine certain basic coordinate require­
ments. For a mechanical system composed of n lumped mass and spring
elements completely unattached, a total of n coordinates would be required
to completely specify the configuration of each element. For example, a
set of n position coordinates XI(t), x:(t), . . . ,x..(t) could be used for the
n elements. If the defining characteristic curve for each element were
available, the associated momentum and forces could be determined. The
velocities for the elements are obtained by differentiating these position
coordinates.
If the n elements are now connected together in a particular configura.­
tion, the original n coordinates are no longer free to vary independently.
Connecting the lumped elements together establishes equations of con­
straint between the coordinates. If the constraint equations involve only
relationships between the coordinates, then these are called holonomic
constraints. If the constraint equations involve nonintegrable relation­
ships between coordinate differentials, then the system has nonholonomic
constraints. The number of the degrees of freedom possessed by a system
is defined as the number of coordinates minus the number of constraints
between the coordinates. For a system having n coordinates and m con­
straint equations, there are (n - m) degrees of freedom.
Repeating the restricted form of Lagrange's equation given by Eq.
(2-17), we have

\.+ OV(XI,X:,.
i !i IOT'(XI,X2,
dt

OXt
,x,,)
ox"
• ,x,,)
=
Q
t
for k = 1, 2, . . . ,n (2-26)

It Upon expanding Eq. (2-26) we would have a total of n equilibrium equa­


tions describing the motion of the system. Now suppose the system con­
nections between the lumped elements impose m additional constraint
I
t
60 Principle. of electromechanical-energy conven;on

equations. We would then have a total of (n + m) equations with only


n unknowns. All of the (n + m) equations cannot be independent.
In order for Lagrange's equation to give a complete set of equilibrium
equations which are sufficient to describe the motion of the system, the
coordinates used must be mutually independent. Therefore, the coordi~
nates must be so selected that no constraint equations involving the coor­
dinates can be written. Such coordinates are known as generalized coor­
dinates. 1 The number of generalized coordinates required to describe the
configuration of a system is always equal to the number of degrees of
freedom possessed by the system. If we have n generalized coordinates,
then the system has n degrees of freedom, since the number of constraint
eq~ations between the generalized coordinates must be zero.
The state functions can be formulated in any convenient coordinate
system. However, before these state functions are used in Lagrange's
equation, a set of generalized coordinates must be substituted for the
original set of coordinates. Examples 2~1 and 2-2 in the preceding section
illustrate this procedure.
Suppose we have a system with ,OJ degrees of freedom, but the par­
ticular coordinate system we have chosen describes the system with a
total of six coordinates. Therefore, one constraint equation between the
six coordinates must exist. If the constraint equation were holonomic
(i.e., a relationship between the coordinates and not their differentials),
then the constraint equation could be solved for one of the coordinates
in terms of the other five. By use of this relationship, the selected coor­
dinate could be eliminated in all of the state functions. The state func­
tions can now be used in Lagrange's equation, since the five remaining
coordinates form a generalized set.
If the constraint equations were nonholonomic, then relationships
between the coordinates could not be obtained because the differential
constraint equations could not be integrated. The state functions would
therefore continue to be functions of the six nongeneralized coordinates
and Lagrange's equation could not be used to obtain the equilibrium equa­
tions for the connected system. In the next section we examine the effect
on Lagrange's equation of a general transformation of coordinates.

2-5 A COMPLETE FORMULATION OF LAGRANGE'S EQUATION

FOR CONSERVATIVE MECHANICAL SYSTEMS

The formulation of the energy-state functions required for Lagrange's


equation is very often more convenient in terms of a set of nongeneralized
I The term "generalized coordinates" is somewhat of a misnomer, since the generalized

set of coordinates is in reality a very specific set.


Lagrange's equation 61

coordinates, or coordinates for which constraint equations can be written.


Examples 2-1 and 2-2 have energy-state functions formulated in this man­
ner. In Sec. 2-4, however, we found that the coordinate set used in the
state functions must first be transformed to a generalized set of coordi­
nates before the state functions can be used in Lagrange's equation. A
transformation from a sct of nongeneralized coordinates to a set of gen­
eralized coordinates can usually be accomplished only when the constraints
are holonomic (Le., the equations of constraint are relationships between
the coordinates and not their differentials). In this section we shall
propose a vcry general coordinate transformation and study its effect upon
the restricted form of Lagrange's equation formulated in Sec. 2-3.
Suppose that we decide to change from the original XI, X2, • • • , Xn
coordinates used in Eq. (2-17) to a new set of coordinates denoted by

XI = XI(~I,~2, • • • ,~m,t)
X2 X2(~I,~2, • • • ,~ ... ,t)
(2-27)

where ~ is the symbol used for these m new coordinates and the possibility
of an explicit dependence on the time t is included. I The total time deriva­
tive of anyone of the old X coordinatcs is given, in general, by

(2-28)

Equation (2-28) shows that the old velocities Xt are functions of not only
the new velocities E; but also the new position coordinates ~;, where i is
equal to any integer, from 1 to m. 011 taking the partial derivative of t,he
old velocities Xt with respect to one of the new velocities Ei, we have

(2-29)

This equation reduces simply to

(2-30)

since E. appears only in the term with aXk!a~i as a coefficient.


Taking the restricted form of Lagrange's equation as given by (2-17)
or (2-26) and mUltiplying through by aXk!aEi, which is equivalent to

I The letter ~ is a general coordinate which, for our electromechanical systems, could
be charge, flux linkage, position, or momentum, both angular and translational.
62 Principles of electromechanical-energy conversion

aXk/iH;" gives us
iJXk ~ [aT'(XI,X2, .~ .. ,X,,)] + aXk [aV(XI'X2" . ,X,,)] = aFt Qk
a~, dt aXk a~, ax" a~,
for k 1, 2, . . . ,n (2-31)

The total differential of the product of two functions can be expanded as


follows:

(2-32)
I
Also, the total time derivative of the last term in Eq. (2-32) is equivalent
to

(2-33)

If the old velocities X/o are not a function of the new position coordinates
~i, then the last term in Eq. (2-32) will reduce to zero.
Observe that the first term on the right-hand side of Eq. (2-32) is
identical with the first term in the equilibrium equation (2-31). Substi­
tuting Eq. (2-32) into Eq. (2-31) with the identity of Eq. (2-33) yields
!
for k = 1,2, .. ,n
(2-34)
II
For each value of k one equation similar to Eq. (2-34) can be written.
Adding all of these n equations together gives us as a result j
!
I
i
i
!
Lagronge', equation 63

By using Eqs. (2-27) and (2-28), the state functions T' and V can be
expressed in terms of the m new coordinates and the time t. After making
the substitution of coordinates, the following identities can be written:

,x,,) aXk
a~.
(2-36)
aT'al'~2' ... ,~... ,tl,t2' ,~ ... ,t)
at.
n
" aT' (XI,X2. . . . ,xn ) aXk (2-37)
L,
• -1
aXk at•
aV(~I,t2' ... ,tm,t) aV(XI,X2, .. ,Xn ) aXk
(2-38)
at. aXk at.

Also let us define

(2-39)

The djdt and summation operations in the first term of Eq. (2-35) are
mutually exclusive and can be interchanged. Therefore, after substitut­
ing Eqs. (2-36) to (2-39) into Eq. (2-35), we have

~ [aT'al'~2' ... ,~... ,the2' ... ,tm,t)]


dt a~.
aT'<EI'~2' I ,~ ... ,tJ,t2' ,t....t) + aV(ehe2, ... ,em,t) = Q.
ae; at.'
for i 1, 2, . . . ,m (2-40)

One equation can be written for each value of i = 1, 2, . . . ,m, giving


a total of m equilibrium equations in terms of the m new coordinates.
If these m coordinates are selected to be a generalized set, Eq. (2-40)
represents a complete form of Lagrange's equation.
Notice in Eq. (2-40) that the kinetic coenergy has become a function
of not only the new velodties but also the position coordinates and pos­
siblyeven the time explicitly. The potential function, on the other hand,
is never a function of the velocities, being only a function of the coordinates
and the time. If the kinetic-coenergy function were not actually depend­
ent on the position coordinates, then Eq. (2-40) would reduce to the
original Eq. (2-17) with the exception of the explicit time dependence in
T' and V.
64 Principle. of electromechanical-energy converdon

In order to further systematize this energy formulation, let us define


a new function
.eal'~2, ... '~""~h~2, ... ,~... ,t)
= T'al,~2, ... ,~... ,~l'~" ... ,~... ,t) - V(tl,h, . . . ,~ ...,t) (2-41)
as the difference between the kinetic coenergy and the potential energy.
This new function .e is called the LOIJrangian, and, in general, it is a func­
tion of the ~'s, the ~'s, and t. In terms of the Lagrangian function, Eq.
(2-40) becomes simply
!!: [a.ea,~,t)]
dt a~i
_ a.ea,l;,t)
a~i
= Qt' f
or t
.
=
1 2
, , ... ,m
(2-42)

whete the abbreviated notation .e(tl;,t) is used to indicate the complete


functional dependence of the Lagrangian as given by (2-41). Equation
(2-42) is equivalent to (2-40), since
a.ea,I;,Q = aT'(~,I;,t) _ 0 (2-43)
a~i a~i
and
(2-44)

as seen from the definition of the Lagrangian.


The set of equations given by (2-42) is the usual form of Lagrange's
equation, and it represents a set of m simultaneous differential equations
in terms of the ~ coordinates and the time t. These are the equilibrium
equations for the system, the solution of which yields all of the m coordi­
nates as functions of time and thus describes the complete operation of the
system. Keep in mind that these equilibrium equations are identical
with those obtained from a usual Newtonian force formulation. If these
equilibrium equations can be formulated directly, then Lagrange's equa­
tion offers no further information. Fo'r many systems, however, these
operating equations are difficult to formulate by summing forces. In
these cases the proper energy-state functions, and thus the Lagrangian,
can often be developed. Then by a straightforward application of
Lagrange's equation, the desired equilibrium equations are obtained.
The purpose of this entire Lagrangian development is to aid in the analytic
description of lumped physical systems. The problems in modeling the
actual system and solving the reSUlting equilibrium equations are in no
way simplified. Only the formulation is greatly systematized.

Example 2·3
A classical problem illustrating the real power of this method is the double pendulum
shown in Fig. 2-7. A mass M, is hung by means of a massless rod of length r, from &
Lagrangci', equation 65.

The double pendulum. f

rigid support. A second mass M. is hung by a similar massless rod of length r. from
the first mass M ,. Find suitable equilibrium equations for this system. Assume both
pivot points to be frictionless.
Solution: A simple xy cartesian-coordinate system is set up with its origin at
point 0 in Fig. 2-7. The position of each mass at any instant in time is given by
XI, 1/, for mass MI and X•• ,¥. for mass M.. Now the Lagrangian for this system is
given by iii
.e(x,,:t••/iI./i.,x.,X•• II ..!I') ... T'(x.,:i:,./i,,/i,) - V(x••x,.!I',1I.)
However, with this system, no spring elements are involved; thus, the potential func­
tion V is equal to zero, and all we must find is the kinetic-coenergy function T'. t
Since all these masses are linear elements, the kinetic-coenergy function is readily
derived to be
T'(x.,:i:"y,.y.) ....!M.x,· + !M,y.' + !M,x.' + !M.Y.'
and with V ... 0 the Lagrangian is simply
.e - T' ... !M.(x,· + !i,') + !M.(x,' + Ii.') (2-45)
The external forces acting on this system arise from gravitational forces acting in
the 11 direction on both masses. On mass lIf. the external force of gravity is simply
Mig, and on mass M. the gravitational force is M,g. Therefore, the external forces
for the cartesian set of coordinates are 68 follows:
Q" - M.g Q., - 0 (2-46)
•Q., ... M,g Q., ... 0
Positive signs are used, since the forces are tending to increase the cooroinate.
t Many authors, in dealing with tllis classic problem. place the gravitational
effects into the potential function with the concept tha~ change in elevation causes
changes,in this form of potential energy. Ho\~ever. in our treatment we- shall restrict
V to containing only spring (orees and consider gravitational forces only 68 external
forces. For an alternative treatment sec 8. W. McCuskey, "Advanced Dynamics,"
p.54, Addison-Wesley Publishing Company, Inc., Reading, Mass., 1959.
66 Principle, of elec:fromec:honicol-energy conversion

The cartesian set of coordinates do not represent a generalized set suitable for
Lagrange's equation. The pendulum rods impose holonomic constraint equations
which can be expressed as
and
These are holonomic constraints because they are relationships between the coordinates.
The angles /I, and /I, shown in Fig. 2-7 represent a set of generalized coordinates.
Either of these coordinates can be varied independently of variations in the other
coordinate. The system has only 2 degrees of freedom, and the two coordinates 8,
and /I, specify the complete configuration.
The original cartesian coordinates can be transformed to the new /I coordinates
by the relationships
z, - r, sin /I, z, - r, sin /I, + r, sin /I, (2-47)
~, - r. cos /I. y, - r, cos /I. + rt cos /I,
These transformation equations automatically satisfy the two constraint equations.
Notice also how these equations correspond to the general transformations of Eqs_
(2-27).
The first total time derivatives of the old coordinates are derived to be
x. - r.iJ. cos /I, i, os r.iJ. cos /II + r,iJ, cos /I,
li. - -r.iJ. sin 8, iI. - -rliJ. sin 8. - r,iJ, sin 8,
On substituting these relationships into the Lagrangian given by (2-45), we have
£(8,,8,,/1,,/1.) .. !M.r.'8. 1 + j-M.[r,'8I' + r,'8,'
+ 2rlr,818. cos (/I, - 8,)1 (2-48)
using the trigonometric identity for the cosine of the difference of two angles. The old
velocities are functions of the new velocities 8. and 8, and also the new coordinates 81
and /I.. Thus the Lagrangian becomes a function of both the new COOrdinaU3 and the
velocities, even though the potential function V is equal to zero.
The generalized forces in the new generalized coordinate system can be evaluated
by using Eq. (2-39). For the /II coordinate we have
oy. Q oz. Q oy, Q oZ,
Q,," Q" 08. + 'I 0/1. + " 081 + ", 08.
By use of Eqs. (2-46) for the original forces and the transformation equation (2-47),
the first generalized force becomes
(2-49)

Similarly. the second generalized force is evaluated from

which reduces to simply

Q" - M ,g( - r, sin 8.) (2-50)

Since the new Lagrangian, given by Eq. (2-48), is a function or angular coordinates
rather than translational coordinates, the external forces must be in the form of torques
rather than lineal forces; therefore the form of Eqs. (2-49) and (2-50) should seem
Lagrange'll equation 61

reasonable. The minus signs in (2-49) and (2-50) indicate that these torques are
tending to decrease the coordinates BI and B,.
Lagrange's equation given by (2-42) can now be used with the Lagrangian func­
tion (2-48) and the generalized external forces given by Eqs. (2-49) and (2-50). The
procedure is as fol1ows:

For i - 1:

Taking the total time derivative of ;U!/a6 1 gives us


_d [a,c( 81,8I,BhBI)] - (M I + MI)TI I'. + MarlT1'1 cos (BI - B,)
dt a6.

Substituting these quantities into Lagrange's equation, using Eq. (2-49) for Q." gives
us finally

(M I + MI)Tllillir MITITI'I COB (BI - BI ) + M,TIT,6I ' sin (BI - BI)


- - (M I + M ,)grl sin BI (2-51)
as one of the two equilibrium equations.

For i - 2:

In a similar fashion', by using the expressions for the Lagrangian and Q,., we obtain
M,r,I" + M,rlr./Il cos (B. - B,l M,rlr,6 1'sin (B. - B,)
- - M t(/TI sin B. (2-52)
as the second equilibrium equation.
These two equilibrium equations are very complicated nonlinear differential equa­
tions in the two coordinates BI and B,. An attempt to formulate these equations
directly from Newton's second law or D' A1embert's principle is certainly a very diffi­
cult job, at the least. By using Lagrange's equation, we have acquired these equations
in a very systematic fashion.

Exampl.2-4
The pivotsl point for the ideal pendulum shown in Fig. 2-8 is constrained to move at a
constant angular velocity of w rad/sec around a circle of radius a 8.8 shown. At time
, - 0 8.8I!Iume that the frictionless pivot point is located at the bottom of its circular
path. Find the equilibrium equation for this system.
68 Principle. of e/edromechallical·ellergy cOllversioll

,
x

Fig. 2-8 An ideal pendulum with a rotat­


ing pivot point.

Solutwn: In terms of the x'y' cartesian-coordinate system the kinetic coenergy of


the m8.8l! M, located at the general state points (x,y), is given by
T'(x,y) - iMx' + iMy'
Since no springs are involved in this problem, the potential-energy function is taken
equal to zero. The system's Lagrangian is therefore given by
,c(x,y) - T' - V - !MX' + !My'
The xy cartesian coordinates are not suitable generalized coordinates, since they are
related by the holonomic constraint equation
(x - a sin ",t)' + (y - a cos wt}l - rl

We have two coordinates and one equation of constraint, so the system has only
1 degree of freedom. The 11 coordinate shown in Fig. 2-8 is a suitable generalized coor­
dinate. The transformation equations from the original cartesian coordinates are

x = a sin ...t + r sin 11 y-acoswt+rcosll

thus expressing the old coordinates in terms of the new coordinate. Notice that both
x and yare not only functions of (J but also explicit functions of the time t. For the
first time derivatives we have

x- a... cos wt + r8 cos fI y = - a", sin ",t - r8 sin (J

Substituting these relationships into the Lagrangian gives us

.c( 8,8,t) - !M!a.... ' + r 8' + a...r8 cos (wt


1 - 11)1 (2-53)

Again observe the explicit time dependence of the Lagrangian.


The only external force on the system shown in Fig. 2-8 is the gravitational force.
Therefore, the original forces in the cartesian system are given by

Q. - Mg Qz - 0
Lagrunge's equatjon 69

These forces can be transformed to the generalized coordinate system by using Eq.
(2-39), which expands to the form
a'll ax
Q. - Q. a(l + Q. a(l
thus giving us
Q. - Mg( -r sin (I)
The minus sign indicates that the torque is opposing an increase in the generalized
coordinate angle (I.
Lagrange's equation for this single coordinate is simply
~ [a.e(6,(I,t>] _ a.e(6,(I,t) _ Q.

tit a6 a(l

Therefore, from Eq. (2-53) we obtain


a.e( 6,(I,t) 1 .
~ - Mr'6 + ~Ma,wr cos (Ull - (I)

and, upon taking the total time derivative, we have

;ud [a.e(6,(I,t)]
~ - Mr" + ~Mawr[
1 •
-w sm (wt - (I) + 6 sm
.
(wt - (I)l

Also by using (2-53), we have


a.e( 6,(1,0 1M? A ' (
a(l - - ~ a!r"sm wt - (I
)
--

Upon substituting these results into Lagrange's equation, after some simplification,
we fin&lly have
a,w'
, - Trsin (wt (I) - - ~ sin (I (2-54)
r

The second term on the left side of the equilibrium equation (2-54) results from the fact
that the kinetic-coenergy function, in terms of the new polar coordinates, is a function
of not only 6 but also e and t. Thus, we have the additional term -aT'(6,(I,t)/ae
introduced into the equilibrium equation. This represents a mechanical inter­
action of the forces of constraint on the motion of the system.

2-6 LOOP EQUATIONS FROM ENERGY-STATE FUNCTIONS

The formulation of equilibrium equations for mechanical systems usually


proceeds in terms of the position and velocity coordinates as the variables
of interest. Table 2-1 shows that formulations with x, X, 6, and 9 as the
coordinates of interest lead to force or torque summations. These equilib­
rium arrangements are commonly called Newtonian mechanics, since
equations in this form are based on Newton's second law. With electrical
~ systems, the formation of equilibrium equations is accomplished by using
;t,
~~
~~.
~;;:
-"::"
70 Pril'lCip/e$ of electromechanical-energy conversion

tJ ") ) tJ
tJ ") ") tJ
~ tJ tJ tJ Fig. 2-9 Loop-current variablee as generalized
coordinates.

Kirchhoff's two laws involving summations of potentials around closed


loops and currents at nodes. These two laws lead us to the use of two
schemes for the formation of electrical-equilibrium equations known as
loop epuatians and Mdal equatians, respectively. With a loop formulation
we equate the algebraic sum of the instantaneous voltages around all
closed loops in the electrical system or network to zero. Either the charge
coordinate q or its first time derivative, the current variable q, is used as
the coordinate of interest. The nodal formulation equates the instan­
taneous algebraic sum of currents entering or leaving each node in the net­
work to zero. All currents are expressed in terms of the flux-linkage
coordinate or its first time derivative, the voltage variable. Nodal equa­
tions, therefore, involve hand}. as the coordinates of interest.
Whe~ we formulate the equilibrium equations for an electromechani­
cal system, we must decide between the loop and nodal schemes for the
electrical part of the system. The mechanical part of the system will
always be formulated with position and velocity as the coordinates of
interest. Referring to Table 2-1, a loop formulatian requires magnetic­
coenergy- and electric-energy-state functiom. These state functions can be
formulated in terms of voltage and current variables defined at the termi­
nals of each electrical lumped element. Since the magnetic-coenergy and
electric-energy functions are being used, the element current or charge
variables are involved in the state functions. However, upon connection
of the elements in a network, constraint equations will exist between the
element current coordinates. These constraint equations are Kirchhoff's
current law applied at each connection point or node of the network.
A set of generalized coordinates which always satisfy the current
law at every node in the network are the loop-current variables shown
in Fig. 2-9. If a loop current enters a node, it also leaves the node and
thus automatically satisfies the current law. Figure 2-9 has selected the
interior loop currents as an independent set of generalized coordinates.
The subject of network topology deals with techniques for determining
other sets of suitable generalized loop currents. l
I See, for example, F. M. Reza and S. Seely, "Modern Network Analysis," chap. 4,

McGraw-Hili Book Company, New York, 1959.


Lagrunge'. equation 71

With the magnetic-eoenergy- and electric-energy-state functions


expressed in terms of a set of generalized loop variables, the equilibrium
equations for the network are given by

!! [aW~(til,tit' ... ,q,,)] + 8W.(q.,qt, ..• ,q,,) = Q",


ell aq", 8q",
for k = I, 2, . . . , 11. (2-55)

where til, q2, . . . , qlt are an independent set of loop currents in the sys­
tem. Taking the partial derivative of the magnetic coenergy with respect
to the kth loop current gives us the flux linkages in the inductors in the
kth loop, as seen from Table 2-1. Taking the total time derivative of
these flux linkages would yield the voltage across these inductances. Thus
the first term on the left side of Eq. (2-55) represents the voltage across
the inductors in the kth loop of the network as a function of all 11. loop
currents in the n-Ioop network.
Also from Table 2-1 we see that, by taking a suitable partial deriva.­
tive of the electrUt.f!eld energy, the voltages across the capacitors are
obtained. The genmlized external-force term Ql on the right-hand side
of Eq. (2-55) involves the algebraic sum of voltage sources in the kth loop,
taken positive if directed aiding the assumed loop-current direction and
negative if opposing that direction. These voltage sources are generally
not functions of the coordinates, but are usually explicit functions of time.
Remember that a constant or d-e voltage source is just a very special
function of time. Equation (2-55), in terms of our energy-state functions,
tells us simply that the sum of the voltages across the inductors and capaci­
tors in the kth loop must equal the algebraic sum of the voltage sources in
the kth loop. This is simply a statement of Kirchhoff's potential law.
Notice, however, thAt the state functions in Eq. (2-55) are formulated as
functions of the 11. loop currents til, ti2, . . . , qlt and their integrals
ql, q2, . . . ,q,.. To be used in Eq. (2-55), these variables must represent
a generalized set of coordinates. In general, we might find it more con­
venient to formulate these functions in terms of the branch currents in the
network and then relate the branch currents to the loop currents.

Example 2-5
By using an energy formulation, determine the equilibrium equations, on a loop basis,
for the linear electrical circuit shown in Fig. 2-10.
Solution: First let us define voltage and current terminal variables for each of the
lumped-linear inductance and capacitance elements in the system. These variables

-- are designated by the a, b, c, and d subscripts in Fig. 2-11. Notice in each CMe that the
reference directions for the terminal voltage across each element and the current
through the element are 80 chOl!en that, if both are of like sign, energy flow is into the
element.
72 Principles of eledromechonicol-_rgy conversion

Fig. 2-10 A linear network.

For a loop analysis, we must formulate the electric-energy-state function and the
majpletic-eoenergy-state function. From Table 1-3, the electric-field storage is given
by
W.(q.,q4) - )0
( q• • "
A.(q.,O) dq.
,
-r )0
( q• • ' ,
Ad(q.,qd) dqd
,

since this form of en~rgy storage takes place only in the two capacitors. The capaci­
tors are linear. Therefore we have

and the electric-field energy storage becomes


q.t qdt

W.(q.,qd) - 2C. + 2C
t
as a function of the element coordinates.
Also from Table 1-3, the magnetic-eoenergy function for this network is given by
, .( q. I I I (q. I I ,
W..(q.,q.) .. jo h.(q.,o) dq. +)0 h.(q.,q.)dq.

with the magnetic-field IItorage taking place in the inductance elements. From FiV;.
2-11 we have

The minus sign precedms the mutual term in these two expressions because the mutual­
field contribution to the flux linkages is in opposition to the self-induced flux linkages

Fig. 2-11 The network disconnected.


Lagrange', equation 73

in each of the two inductances. The current ~b in inductor L. is 8.88umed to flow into
the dotted side of this element. This current sets up flux linkages in a certain direction
in the inductance L,. Current q. is 8.88umed to flow into the nondotted side of induct­
ance L" and therefore, according to the dot convention, it sets up flux linkages opposite
to those set up by the current ~b. Substituting these flux-linkage (unctions into the
integral expressions gives us .
,rq· ,rq· I
10 (L.~.-O)d~6+ 10 (-M~.+L~.)d~.
W.,(4b.~.) -
I I

with ~. in the second integral acting as a constant. Finally, the magnetic coenergy as
a function of these element coordinates becomes
W~(~.,~.) - tL'~b' - MM. + j-Lt4.'
In order to obtain the desired loop formulation, the loop-current coordinates must
be substituted for the element coordinates as follows:
q.(t) - ~,(t) q.(1) - ~.(t) - ~.(t)

~(t) - ~,(t) ~4(t) - ~,(t)

These four equations represent holonomic constraints which can easily be integrated.
The only charge coordintiles appearing in the state functions are q. and q4. By
integrating the first and lail. constraint equations, we have

for all time after the network is connected together. The energy-state functions in
terms of the loop coordinates now become
q.' q.'
W.(q.,q,) - 2C, + 2C. (2.56)

W~(~l'~') - tL,~,· - M~I(~l - ~.) + tL.(~1 - ~.). (2-57)


simply by substitution of the loop coordinates into the state-function equations. The
external forces based on the loop coordinates are given by

QI - 1110) - IIt(t)· Q. - tI.(t)

The equation for Q. is made up of all voltage sources in loop 1 taken positive if aiding
~, and negative if opposing ~.. Similarly, for Q, we have all sources in loop 2 taken
positive if aiding ~. and negative if opposing ~•.
By using Eq. (2-55), we have
Fork-I:

~[aW~(~I'~')] + aW.(q"q.) _ Q.

dt a~. aq,

From Eq. (2-56)


aW.(q"q.) q.

aq, - C.

and from Eq. (2-57)


aW~(~I'~')
a~,
- LI~' - M(~I - ~.) - M~, + L.(~l - ~.)
14 Principl•• 0' ./ectromechanica/-energy conversion

Thus from Eq. (2-55) with k - 1 we have

i [Llql - M(ql - q,) - Mq. + L,(ql - q.)] + ~'. - fI,(e) - fI,(t)

using the expression for Q,. Taking the total time derivative and rearranging terms
gives us, finally,
1
(L , - 2M + L.}ij. + C I q. + (-L. + M)lb - fll(t) - lI,(t) (2-58)

8.8 the firat-loop equilibrium equation. For k .. 2, Eq. (2-55) becomes

~ [aw~(q ..q.)] + aW.(q.,q,) _ Q,

tit aq. aqt

By lusing Eqs. (2-56) and (2-57) for the state functions and the expression fev Qt, we
have •

which becomes

(-L. + M)/b + (L.)ij. + ~ q. - fI,(t) (2-59)

when we take the total time derivative.


Equations (2-58) and (2-59) are readily seen to be the correct equilibrium equa­
tions for this network.

2·1 NODAL EQUATIONS FROM ENERGY-STATE FUNCTIONS


Nodal-equilibrium equations are formulated on the basis of Kirchhoff's
current law. From Table 2-1 we see that a nodal formulation requires the
electric-coenergy- and magnetic-energy-state functions. The voltage and
current variables defined at the terminals of each lumped element can be
used in the formulation of these state functions. Since the electric­
coenergy and magnetic-energy functions are required, the flux-linkage or
voltage coordinates are involved in the state functions. Upon connection
of the lumped elements in a network, constraint equations will exist
between the voltage coordinates. These constraint equations are Kirch­
hoff's potential law applied around each closed loop of the network.
A set of generalized potential coordinates which automatically satisfy
the potential law around every closed loop are the nodal variables shown
in Fig. 2-12. At each node the potentials Xl, X2 , • • • , X8 are defined with
respect to the datum node. We can see how the potential law is satisfied
around each closed loop by considering the sample loop shown in Fig. 2-12.
The respective voltages between nodes are defined in terms of the nodal
Lagrange', equation 75

.:{ ~$
i
'----v---"
63 +

Fig. 2-12 Nodal variables as gen­


eralized coordinates.
xe
-b- Datum
node

coordinates as follows:

The sum of potentials around this particular closed loop is


el + e, + e, + e. ='0 (2-61)
as seen by substituting Eqs. (2-60). In a similar fashion the nodal coor­
dinates satisfy the potential law for every closed loop of the network.
Since no constraint equations can be written for the nodal coordinates,
these represent a suitable generalized set. Alternative generalized sets
can be formed by selecting other nodes as datum.
With the electric-coenergy- and magnetic-energy-state functions
expressed in terms of a set of generalized nodal variables the equilibrium
equations for the network are given by
!i [aW:(Ah . '.. ,A,,)] + aw..(X 1, ••• ,x..) = Q.
• a~ a~
for k = 1,2, . . . ,n (2-62)
The A's are the respective node voltages inthe network. Therefore, the
partial derivative of the electric-field coenergy with respect to the kth­
node voltage gives the total charge on all capacitors touching the kth node.
The first term on the left side of Eq. (2-62) takes the total time derivative
of this charge, and thus it represents currents out of the·kth node through
capacitors. Similarly, the second term on the left side of Eq. (2-62)
represents currents leaving the' kth node through inductances. The gen­
eralized force term Q. on the right side includes all current sources touch­
ing the kth node. In this summation of current sources, a positive sign
is used if the source is directed into the node, trying to make the node
variable A. positive. If the source is oppositely oriented, then a negative
sign is used.
76 Principle. of e1edl'Omechonieol-energy convenion

Equation (2-62) is simply Kirchhoff's current law at each node in the


network. As with the previous formulations we have considered, quite
often it is more convenient to formulate these state functions in terms of
terminal coordinates of each element. These coordinates are then related
to the generalized nodal variables. This substitution of variables essen­
tially gives the physical hookup of all the individual lumped elements.
We tear our system apart to form the state functions and then, through
a coordinate transformation, connect it back together. l

Example 2·6
Determine the nodal equilibrium equations, by using an energy formulation, for the
net"'ork shown in Fig. 2-13.
8olutwn: First the terminal variables for each of the lumped elements are desig­
nated as shown in Fig. 2·13. The only rule on their selection is that energy flow into
the element if both the respective;f.. and q are positive. The original formulation of our
state functions, leading to Table 1-3, is based on this arrangement of the terminal
variables for each of the lumped elements. Observe the way elements LI and C 1 have
their terminal variables defined in an opposite fashion.
For a nodal formulation, we must find the electrie-field-coenergy and magnetic­
field-energy storage. By use of Table 1-3, and the fact that all elements are linear,
these funetions are given respectively by
W~(4;f..,,) - iCI~' + iCt:f..,,'
)..1 ~.
W -().•• ~) - 2LI + 2L.
for the system shown in Fig. 2-13.
The generalized nodal coordinates must be substituted for these individual terminal
coordinates in order to use the nodal energy equation (2-62). The eoordinate rela.­

lOne is reminded of Humpty-Dumpty if the king's horses and men had been more
successful.

c,

.",. DQtum
Fig. 2-13 A linear network.
Logronge', eqIJOtion 77

tionships, obtained by inspection of Fig. 2-13, are 118 follows:


)...(t) - )...(0 )...(0 - )...(t) - )..,(t)

4(0 - -)..,(t) )..,,(t) - )...(t)

These constraint equations are holonomic and can eMily be integrated. The magnetic­
energy-state function requires only A. and >.e, so by integrating the first and third rela­
tionships we have
A.(t) - A,(t) ,
By substituting th, relationships into the state ~unctions, we have
W~()",,)..,) .. -tCI).." + -tC.)..,' (2-63)
W ..( 'A',.
A) - A.'
2L,
+ (A. 2L,
-- A,)' (2-64)

The external forces for this network are readily seen to be

with current sources into the respective nodes taken 118 positive and out of the nodes
118 negative.
By using Eq. (2-62), tie have
For 11: - 1:
d[OW:()",,)...>] (lW..(A.,A.) ... Q

dt (1)..1 + (lA, I

From Eqs. (2-63) and (2-64)


oW:()..,,)...) •
-~-'-...;. ... C,I\,

0)...

oW...(A.,A.) ... ~ A, -- A,
+
(lA. L . · L,

A, A, AI
.
Thus, Eq. (2-62) with 11: ... 1 gives us, finally,
..
C,A, + 1:; + L. -- L. ... i,(l) -- i.(t)

after we take the total time derivative and substitute for Q,. Thus we have the
correct nodal equation for node 1.
For 11: ... 2:

Equation (2-62), with 11: ... 2, gives the second nodal equation 118

C.X. + ~: -- ~ - i.(I) + i.(t)


which is again obeerved to be correct.
78 Principle$ of electromedtanical-energy cOt'lVer$ion

2·8 LAGRANGE'S EQUATION FOR CONSERVAnVE MECHANICAL


AND ELECTRICAL SYSTEMS
For conservative (i.e., nondissipative) mechanical syst~s, we found it
convenient to define a new state function known 8.8 the L4;rangian. The
Lagrangian function, defined by Eq. (2-41) for mechanical systems, is the
system's kinetic-coenergy-state function minus the system's potential­
energy-state function. In general, the Lagrangian for a mechanical sys­
tem is a function of position and velocity coordinates and possibly even the
time explicitly. In terms of the Lagrangian function, the system's equilib­
rium equations are given by Lagrange's equation in the form of Newton's
law or D'Alembert's principle. Now we would like to incorporate con­
serlative electrical systems into the same routine procedure.
Since we have two different formulations in the case of electrical sys­
tems, a choice in the formulation of the Lagrangian is provided. For a
loop formulation the Lagrangian should be taken as

£(ql, ... ,q",ql, ... ,q..) = W:(ql, .. ,q,,)


I - W.(ql, . . . ,q..) (2-65)

where the /j's are the loop currents in the network. Substituting this
Lagrangian into Lagrange's equation (2-42) gives us Eq. (2-55). This
latter equation is the proper energy-state formulation yielding loop equilib­
rium equations.
For a nodal formulation the proper Lagrangian is given by

£(X 1, : •• ,X"':>'I> . . . •:>.,,) = W:(X h •• ,X,,)


A - W",(:>'1t ••. ,:>...) (2-66)
" ,
where the X's are the node voltages. Again, substituting the Lagrangian
of (2-66) into Lagrange's equation would yield Eq. (2-62), which is the
energy-state-function formulation giving nodal equilibrium equations.
In Eq. (2-65), which gives the proper Lagrangian for a loop formula­
tion, and Eq. (2-66), which gives the Lagrangian for a nodal formulation,
notice that a certain energy-state function is being subtracted from a
selected coenergy-state function. In our later work we shall be dealing
with systems having both electrical and mechanical parts. The Lagran­
gian for such an electromechanical system will contain both mechanical­
and electrical-coenergy-state functions minus mechanical- and electrical­
energy-state functions. The coenergy portion of the general Lagrangian
is to be caIIed the total coenergy-state funetion. Similar1y, the energy por­
tion of the Lagrangian is termed the total energy-state function. Therefore,
in summary, the total system coenergy-state function indicated by the
Lagrange'. equation 79

script 3' is given by

", (.x,x,)..,
"
q't) -- T'('X,x, t) W~(q)
+ W;()..) loop
(2-67)
nodal
where the upper line indicates a loop formulation on the electrical part of
the system and the lower line is for a nodal formulation. The total sys­
tem energy-state function indicated by script 'l) is given by
loop
(2-68)
nodal
The total system Lagrangian is defined by

.c (x ,x')..,A,
4,Q't) = (x ' x')..,q't) - (x 'A,q't)
3' 'l)
loop
nodal
(2-69)

again using the upper set of symbols for a loop formulation and the lower
set for a nodal formulation. The equilibrium equations for the system
are then obtained by using Lagrange's equation, which, in a general form,
is given by

for k 1,2, . . . ,n (2-70)

using ~ to represent the position and charge, or position and flux-linkage,


coordinates.

2-9 A BRIEF LOOK AT ELECTROMECHANICAL COUPLING


A vast extension of the scope of this entire formulation is available to us
simply by carefully examining the functional dependence of the total sys­
tem coenergy-state function and energy-state function given respectively
by Eqs. (2-67) and (2-68). Notice that 3' is composed of the mechanical­
kinetic coenergy T' and either the magnetic- or electric-field coenergies
W: W;.
or The mechanical kinetic coenergy can be a function of not
only velocity but also position coordinates and the independent time coor­
dinate. This function T' was originally formulated in Table 1-3 as a func­
tion of only the mechanical velocity coordinates, or the x's. It becomes
a function of velocity, position, and time most often because of a coordinate.
transformation to a set of generalized mechanical coordinates.
- Let us suppose that the electrical coenergies W: or W~, in a certain
coordinate system, were functions of more than just the 4's or X's. For
.~"
;
example, assume that W:
were a function of all the loop currents, meaning
all the 4's, and also certain of the mechanical coordinates in an electro­
§f
t
80 Principle. of e/ec:tromechonicol.energy conversion

mechanical system. In other words, suppose


W: = W:(q,x) (2-71)
where we use our shorthand functional notation to indicate that W:
is dependent on all the q's and x coordinates. Since the magnetic coen­
ergy is a function of the mechanical position coordinates, the term
-aw:(q,X)/aXk would be included in the equation for the kth mechanical
coordinate. Let us carefully examine how this comes about and what it
means.
Taking a loop formulation for the electrical portion of the system,
the total coenergy-state function is given by
::J'(x,x,q,t) = T'(x,x,t) + W:(q,x) (2-72)
Using the assumption of Eq. (2-71), the magnetic coenergy depends on the
mechanical position coordinates. Similarly, the total energy-state func­
tion is given by
'll(x,q,t) = V(x,t) + W.(Q).-.L (2-73)
From Eq. (2-69), the Lagrangian is given by
.SK.
£(x,x,q,q,t) = T'(x,x,t) + W:(q,x) - V(x,t) - W.(q) (2-74)
For the kth mechanical coordinate, Lagrange's equation (2-70) gives us the
following equilibrium equation
-!!:. [aT'(X,x,t)] _ aT'(x,x,t) _ aW;'(q,x) + aV(x,t) = Q~ (2-75)
dt axlc. aXk iJXk iJXIc
Carefully verify that Eq. (2-75) is correct by using the Lagrangian func­
tion of (2-74). Notice the importance of keeping the precise functional
dependence of each of the state functions.
The terms in the equilibrium equation (2-75) have the following inter­
pretations. On the left-hand side, the first two terms are inertial forces.
The fourth term, involving the mechanical potential energy, represents
the spring forces. The QIe on the right-hand side includes all the exter­
nal forces applied to the kth mechanical node. The remaining term,
-iJW:(q,x)/iJ1:le , seems to be something new. This term is a mechanical
force in one of the mechanical-equilibrium equations, which arises because
of an electrical-coenergy storage, and therefore is called a mechanical force
of electrical origin. Lagrange's equation is predicting a mechanical force
due to the electrical portion of the system, provided the electrical state
functions, for some physical configuration, are dependent on the mechani­
cal coordinates in addition to the electrical coordinates. This is the basis
for all electromechanical devices. In the following chapters we shall study
wide classes of electromechanical devices. The fundamental principle
lagrange's equotion 81

underlying their operation rests on the functional dependence of their state


functions. Thus, the operation of loudspeakers, microphones, relays,
instrumentation transducers of all types, etc., can be analyzed by the
same general procedure stemming from the development of proper state
functions and the use of Lagrange's equation. The interaction forces
between the electrical and mechanical portions of the system will be
included in a very natural way by using Lagrange's equation. After
extending Lagrange's equation to include dissipative elements, we shall
study these interacting electromechanical systems in great detail.

2-10 THE RAYLEIGH DISSIPATION FUNCTION

The general form of Lagrange's equation for conservative electromechani­


cal systems is given by Eq. (2-70). For convenience this equation is
repeated as follows:
!i [~.ca,~,t)] for k = 1, 2, . . . , n (2-76)
dt ate
where the general coordinate ~ represents translational position x, rota­
tional position 0, and electrical charge q for a loop formulation or electrical
flux linkage A for a nodal formulation. Each of the terms in Lagrange's
equation (2-76) represents certain mechanical forces or electrical voltages
or currents. Thus, the entire equation gives us the mechanical-equilib­
rium equations identical with Newton's second law and electrical-equilib­
rium equations identical with either of Kirchhoff's two laws.
The first term on the left-hand side of (2-76) includes all inertial
forces and inductive voltages for a loop formulation or capacitive currents
for a nodal formulation. The second term includes all spring forces and
capacitive voltages 'or inductive currents for a loop or nodal formulation,
respectively. In addition, certain electromechanical interaction terms
will stem from these two terms, as is briefly shown in Sec. 2-9. Thus, the
terms stemming from the Lagrangian .c(t,t,t) include all of the effects
of the six conservative lumped elements. Extemal forces, explicitly
dependent on the time t, are grouped in Qk and are placed on the right­
hand side of Lagrange's equation.
The three lumped dissipative elements are introduced in Chap. 1 and
are summarized in Table 1-2. All physical systems contain some dissipa­
tion, and consequently any practical analysis using Lagrange's equation
would not be complete without some provision for including these ele­
ments. There are generally two ways in which dissipation can be handled.
The forces due to the dissipative elements can be suitably incorporated
with'The external forceS-Q;:' For example, consider the mechanical-trans­
lational damping element shown in Fig. 2-14a. The characteristic curve
82 Principle. of e/ectromechanical·energy conversion

Fig. 2·14 (a) A lumped


translational damping ele­
i ment. (b) Its defining-char­
(a) (II) acteristic curve.

for this element, shown in Fig. 2-14b, relates the restraining force p to the
velocity i of one side of the damper relative to the other side. An expres­
si~m for p as a function of i can be written, and the force as a function of
velocity can then be included in the equilibrium equations. Such a
procedure is familiar to us from previous work in mechanics. For the
electrical resistance or admittance element the same idea can always be
used. For a loop analysis the relationship expressing the terminal poten­
tial across the dissipative elements as a function of the loop currents is
needed. This potential term is then properly inserted into the appropriate
loop equations. For a nodal analysis the current through the dissipa­
tive elements as a function of the node voltages is placed into the proper
nodal equations.
A second approach can also be used to introduce the dissipative terms
into the equilibrium equations. The defining-characteristic curves for
the dissipative elements show the first time derivative of one basic coordi­
nate as a function of the first time derivative of the other basic coordinate.
Thus for the translational damping element shown in Fig. 2-14b the force
p is plotted as a function of the terminal velocity i. For the electrical
admittance or resistance element the current q is plotted as a function of
the terminal potential X, as shown in Fig. 2-15. The state point P defines
a particular state of the element where the current is q and the terminal
potential is }.. (both without the primes).
For the electrical element, the Rayleigh dissipation function is defined
as the shaded area above the characteristic curve in Fig. 2-15. Thus for
the state point P the electrical Rayleigh function is given by
(2-77)

Similarly, the electrical co-Rayleigh diss~pationfunction is the shaded area.


below the characteristic curve, and therefore it is given by

F:(X) = /0). q'(X') dJ..1


where the prime denotes the cofunction.
Lagrange', equafion 83

Fig. 2-15 The Hayleigh and co-Hay­


leigh di!l8ipation functions for the lumped
electrical 1088 element.

For a system of n electrical lumped dissipative elements the elec­


trical Rayleigh and (~o-Rayleigh functions are defined as follows:

(2-79)

(2-80)

In both Eqs. (2-79) and (2-80) the possibility of having a lllutual resistance
element is included. For example, Eq. (2-79) has the potential across the
ith element as a function of all the n terminal currents. Normally, the
voltage across a resistance element is only a function of the current in that
single element.
In a similar fashion Rayleigh and co-Rayleigh dissipation functions
can be formulated for mechanical systems in translational and rotational
motion. For a system of n lumped viscous damping elements the mechan­
ical Rayleigh and co-Rayleigh functions are defined as follows:

. ) " (in.,(.t .,
\'
,1'" = "")0 Xi 1'hP%, (2-81)
..
i-I

F'"' (.Xl,X2,
. • • •
.)
,X,. = \' )(t.
"" "(" .,
0 Pi Xl,X%,
.,) .> ...'
• . • ,X" (.I.;l;i (2-82)
i-I

The possibility of mutual viscous damping is also included in these two


functions.
The usefulness of the Rayleigh and co-Rayleigh dissipation functions
can readily be appreciated by considering certain partial derivatives.
84 Principle. of electromechanical·energy conversion

From Eqs. (2-79) and (2-80) we can show that


8F.(ih,Q2, . . . ,q.. ) }. (. . h )
8q. = • Ql,Q2, • • • ,'I" (2-83)

and

(2-84)

Equation (2-83) shows that the partial derivative of the electrical


Rayleigh function with respect to the current in the kth element gives the
voltage across the kth resistance element. If the electrical Rayleigh func­
tion is formulated as a function of a set of generalized loop currents, then
}.At represents the voltage across all of the resistances in the kth loop.
Similarly, Eq. (2-84) shows that the partial derivative of the co-Rayleigh
function with respect to the potential across the kth element gives the
current through the kth resistance element. Using a generalized set of
nodal voltages in the formulation of the electrical co-Rayleigh functions
would make Ijk be the sum of the currents leaving the kth node through
resistance elements.
For the mechanical Rayleigh and co-Rayleigh functions given by
(2-81) and (2-82) appropriate partial derivatives give the velocity across
the kth damping element and the force transmitted by the kth element as
follows:

(2-85)

(2-86)

For an electromechanical system an overall Rayleigh dissipation


function can be formulated. The mechanical-equilibrium equations are
usually formulated as a summation of forces expressed as functions of the
velocity or position coordinates. Therefore, according to Eq. (2-86), the
mechanical co-Rayleigh function should be used. The proper electrical
dissipation function depends on whether a loop or nodal formulation is
used for the electrical-equilibrium equations. Thus, in general, the total
Rayleigh dissipation function is given by
loop (2-87)
nodal
where the upper line of symbols is used for a loop formulation and the
lower line for a nodal formulation.
If the system Rayleigh dissipation function given by Eq. (2-87) is
expressed in terms of a set of generalized coordinates, then it can be
Lagrange', equation 85

incorporated in Lagrange'~ equation as follows:


~ [c£(e.~,t)] _ c£(~.~,t)
dt C~k C~k
+ cffW = QI; for k = 1, 2, . . . ,n (2-88)
c~.t
Notice that ff is a function only of t which could represent translational
or angular velocity, voltage, or current. Equation (2-88) represents a
very general form of Lagrange's equation including dissipative elements.
This form is of sufficient generality to readily include all of the electro­
mechanical systems we shaH study.

Example 2·7
A P-N junction diode is shown in Fig. 2-16a. The terminal characteristics for the
dioue arc shown in Fig. 2-16b. The terminal voltage is assumed to be limited such
that reverse breakdown in the Zener region does not occur. The characteristic shown
in Fig. 2-160 can be described by the equation

where 1. is the so-called saturation current, q. is the charge of an electron, k is Boltz­


mann's constant, and T is the absolute temperature in degrees Kelvin.' Determine
an expression for the co-Rayleigh dissipation function.
Solutwn: The electrical co-Rayleigh dissipation function is defined by Eq. (2-80).
For the simple case of a single element, (2-80) reduces to the form of Eq. (2-78).
ThUll, we have

1See E. J. Angelo, "Electronic Circuits," 2d ed., chap. 3, McGraw-Hill Book


Company, New York, 1964.

r 1 N

i.
P

+
Zener
breakdown

(al (bl
Fig. 2-16 (a) A P-N junction. (b) Its terminal ehara.eterietic.
86 Principle, of electromechanical-energy conversion

it ~ -----i'" ~
i~Pl {i ip",
m, mechanical dampers
(both tranSlational and

j { 10 , 10", j
rotational)

n, eleCfrical admi!!ances

t
Fig. 2-17 A system of linear meehani('ul and electrical dissipative elements.

On substitution of the ~haraeteri8tic equation the.eo-Rayleigh function becomes


,
F.(k) - 10fA I.
[
exp
(q.k')
kT

which reduces to

Taking the ilF;/akwould give q, the current through the diode.


For the I'lpecial case of linear dil'lsipative clements a simpler form for
the Rayleigh function is very usefuL Figure 2-17 shows the dissipative
portion of a system containing m linear mechanicardampers and n linear
electrical admittances. We therefore have the following relationships for
these lOss elements:
i = 1,2, . . . , m (2-89)
and
or i = 1,2, . . . ,n (2-90)

where the ith resistance R. is the reciprocal of the ith admittance, Gi •


Substituting Eqs. (2-89) and (2-90) into the definition of the Rayleigh
dissipation function of Eq. (2-87) gives us
loop
(2-91)
nodal

Notice that the Rayleigh dissipation function is a quadratic function of


the velocities (both translational and rotational) and loop currents or
node voltages. For the linear loss element, Eq. (2-91) shows the Rayleigh
Lagrange'. equation 87

function to be simply one-half the instantaneous power supplied to each


element.

Example 2·8
For the mechanical system shown in Fig. 2-18 write the equilibrium equations by
using Lagrange's equation. Include the force of gravity on each of the two m&88e8.
The coordinates x. and x, are 80 designated that, when x. - x, - 0, spring K. is
stretehed a distance a, spring K, is compressed a distance b, and spring K, is stretched
a distance c. Motion is constrained to be in only the vertical direction.
Solution: Because of the simplicity o{ this problem, let us not define terminal
variables for each element. In other words, let us just formulate the state functions
directly in terms of x" Xl, x., and t" which are clearly a set of generalized coordinates.
The coenergy (unction for the system is given by
3'(x,.x,t) - j-M,x,' + iM:i:.'
and is actually only a function of the x's.
For the total energy (unction we have

1 1 1
'll(x.t) - 2K, (x. + a)' + 2K. (x, - x, - Ill' + 2Ka (x, + c)'

which depends on only x for this case. Each of the terms in parentheses represents
the displacement of one end of the spring with respect to the other. Combining
the coenergy- and energy-state fl!nctions, the Lagrangian for the system is given by

£(x,x,t) - ~ Md,' + ~ M,x." - 2~' (x, + a)1 - 2~' (x. - x, - b)1


__1_ (XI + e)1 (2-92)
2K.
The Rayleigh dissipation function for the linear damping elements is given by

S'(i) - iD,i,' + tD,(i, - i,p + iD,x,'

Fig. 2-18 A lumped-linear


mechanical-translational sys­
tem. l Gravity f(f}
88 Principle. of eledromechanicol-energy conversion

Notice that each term in the Rayleigh function is simply one-half the damping
coefficient D times the square of the velocity of one side of the damper with respect to
the other side. This represents one-half the power dissipated by each damping
element.
The external forces are as follows:
Q. - M.g -- f(t)
where g is the acceleration of gravity.
Now, using Eq. (2-88) with k - I, we need
(J~ _ M,x,
(Jz,
ii£ 1 1
(Jz, - -- K, (z, + a) + K. (z, -- z, -- b)
a~ _ Dd, __ D,(:t. - x,)
(Jz,
which combine, according to (2-88), to give

M,i, + i.. (z, + a) + ~. (x, + b z,) + D,x, + D,(xl -- x,) - M,g (2-93)

after some slight rearrangement of signs and terms.


Similarly, with k - 2, we require

which combine, according to (2-88), to givc

M.i, + ~. (z: - Z, -- b) + ~. (z, + c) + D,(:t, - :tl) + D.:t, - Mt(1 -- 1m


(2-94)
F..quations (2-93) and (2-94) are the required equilibrium equa.tions for the lumped­
linear mechanical system. Be sure to check these equations by using the conventional
Newtonian formulation.
The stea.dy-state or static positions of the two masses, taking the external driving
function 1(0 to be zero, can be readily determined by letting

x, - X, and z, - X,
where X, and X, are constant!!. With the forces of gravity as the only external fornell,.
.
(2-95)

it is certainly reasonable to assume that the system with damping will reach some ,
constant, non-time-varying configuration. Substituting Eqs. (2-95) into the equi~"
librium equations (2-93) and (2-94) gives us . ~

1
K. (Xl + a) + K,1 (Xl + b - X,) - MIg
1 l '

K, (X. -- XI -- b) + K. (XI + c) - Mt(1

Logrange', equotion 89

+
+
vlttlr:

"Ql
Fig. 2-19 A lumped-lin­
ear electrical network.

with all time derivatives of constants equal to zero. If the spring compliances, the
two m8.8EleS, and the constants a, b, and c are known, the static configuration of the
system can be determined.

Example 2-9
I·'or the lumped-linear electrical network shown in Fig. 2-19, determine the equilibrium
equations on a loop basis by using Lag.range's equation.
Solution: By using the three generalized loop current coordinates. shown in Fig.
2-19, the following total state functions are determined.
Total coenergy function:

Total energy function:

Total Rayleigh dissipation function:

The external forces. or voltage sources for a loop formulation, are given by

Q. - -Vt(t) + v.(l)
The total Lagrangian for this system is
90 Principle. of e/edromechanical·energy conversion

Using Lagrange's equation, given by (2-88), we need the following terms for
k - 1:

Combining these terms with the external source Ql according to Eq. (2-88) gives us

L,(lb - ii.) - M(ij, - ij.) + -(~ ql + -c~- (q. - q,) + R.(q. - q,) + RMI - q,)
J' a
- VI(t) - v,(t) (2-96)

aftet we take the total time derivative d/dt required by the equation. Equation (2-96)
is readily seen to be the correct loop equation for the first loop.
By using the Lagrangian, the Hayleigh dissipation function, and the external
sources, derive the remaining two equilibrium loop equations. At this point in the
development of these energy techniques, it is important that you be able to manipulate
these functions properly, particularly in regard to partial and total derivatives.
Therefore, take flufficient time to complete this example carefully.
Notice in this example that we have written the state functions :I', 'll, and !f
directly in terms of the loop variahle8 without defining a preliminary set at the terminals
of each element. With a small amount of experience in settihg up these functions for
simple linear systemll, a great deal of time can he saved hy rememhering the form of
these functions. Their original definitions, given by TaMe 1-3, are important when a
new, unfamiliar situation is presented.

Example 2-10
For the lumped-linear network shown in Fig. 2-20 determine the equilibrium equa­
tions on a nodal basis by using the Lagrange formulation.

Fig. 2-20 A lumped-linear electrical network.


Lograng.'s equotion 91

Solution: For the nodal variables shown in Fig. 2-20, the Lagrangian and Ray­
eigh dissipation functions are given respectively by
1 1 1 1 1
,£, - 2 CIX,' + 2 C.(Xs - X,), - 2L. Xl' - 2L. (X, - X.), - 2L. XI'
!f -4Gl~'lS + j-Gt(X, - X,), + jGaX.' + jG.(X, - X,), + jG,(X, - Xa)2 + lG.x,'
The external forces, or current sources for a nodal formulation, are
Qt ... it(t) Q, == it(t)
The first nodal equation is obtained by letting k = 1 in Eq. (2-88). For k =- 1
we require the following terms:

t
I By substituting into Lagrange's equation, we have

I
I
l
including the external for('e Q.. This equation is readily seen to be the correct nodal
equation for the first node.
Just for practice, complete the example by obtaining the remaining two nodal
equations.

2·11 SUMMARY

i Let us see where ~e now stand. Our objective has been to develop a
systematic procedure for determining a consistent set of equilibrium equa­

Il tions for electromechanical systems. Previous courses in circuits and


mechanics have taught us procedures which are actually just the writing
of these equations by simple inspection of the system. Obviously, if we
! can write the equations immediately, by inspection, then this is the best
method, provided that we arc neither nearsighted nor farsighted so that
terms arc erroneously included or omitted. In this chapter an energy
technique which generates these equilibrium equations has been developed.
The formulation of certain energy-state functions follows a very sys­
tematic, well-defined procedure. Lumped elements, which are not neces­
sa.rily linear elements, are definable in terms of characteristic curves relat­
ing particular basic coordinates and possibly time derivatives of these
coordinates. For conservative elements, energy-state functions depend
only on the physical construction of the elements and on the final state
92 Principles of e/edromechanical-energy conversion

of the system in terms of the coordinates of the characteristic curves. For


diasipative elements the Rayleigh and co-Rayleigh dissipation functions
are defined. Lagrange's equation (2-88) gives the recipe for manipulating
these state fundions to ohtain the required system equilihrium equations.
Lagrange's equation requires that all state functions be expressed in
terms of a set of generalized coordinates. A consistent set of generalized
coordinates describe the configuration of the system and are completely
independent from one another. Therefore, no equations of constraint
can be written between generalized coordinates, and thus the number of
generalized coordinates is equal to the degrees of freedom possessed by
the system. When expressed in terms of generalized coordinates, the
total coenergy-state function depends not only on the time derivatives of
cobrdinates (that is, X, 8, q, ).) but also on the actual coordinates and pos­
sibly even the time explicitly. The total energy-state function, however,
can only be a function of all the coordinates and the time, and never a
function of the time derivatives of the coordinates.
Through numerous examples in this chapter we have seen how the
Lagrange formulation generl!-tes equilibrium equations for separate elec­
trical or mechanical systems including both conservative and dissipative
elements. In the next chapter we shall be concerned with electromechani­
cal systems where there is an interaction between the electrical and
mechanical portions of the system. Once we are able to formulate the
energy-state functions for these devices and systems, then Lagrange's
equation will, in a routine fashion, give us the proper equilibrium equa­
tions. Even the forces of interaction will be properly and simply derived
and included in the equilibrium equations. It is for these classes of prob­
lems that this energy formulation is of greatest value.

PROBLEMS

2-1 The rotational-mechanical system shown in Fig. P2-1 consists of two fly­
wheels with inertia.s J\ and J, coupled by a shaft having a. compliance K\. A third

Fig. P2-1
lagrottge'$ equation 93

flywheel with inertia J, is coupled to J. by means of a torsional spring having a com­


pliance K.. The inertia J, is mounted on the shaft by using II. frictionless bearing.
a. Find II. La.grangian for the system in terms of the indicated set of generalized
coordinates.
I,. Find tho externl\l tor'lue stJurcllI; Qt, (h, and Q,.
c. Write the three equilibrium equations {or the system by using Lagrange's
equation.
2-2 The mechanical system shown in Fig. 2-3 is disassembled as shown in
Fig. 2-4. In Fig. 2-4 a total of 10 variables are shown to describe the terminal condi­
tions of each of the five lumped elements. However, in Fig. 2-3 only two velocity
coordina.tes are required to describe the system.
a. How many degrees of freedom does this mechanical system have?
b. For the 10 variables in Fig. 2-4 how many constraint equations do we have
when the system is reassembled?
c. How many constraint equations exist before the system is reassembled?
d. Are these constraint equations holonomic? Explain.
e. List the constraint equations for part b and indicate those that result from
assembling the system.
2-3 A mass !of is attached to a stiff rod of length I suspended below a spring of
compliance K as shown in Fig. P2-3. Neglect the mass of the rod and spring. The
support point at the bottom of the spring is constrained to move only vertically, and
the pendulum can swing only in the plane of the paper. By using the two generalized
coordinatcs y. and fI, write the equilibrium equations for the system.
Him: First formulate the Lagrangian in terms of the xy cartesian coordinates and
then transform these coordinates to the two genere.\ized coordinates.
~~~-------------.x

Fig. P2-3 y

2-4. Formulate the equilibrium equations for the system shown in Fig. P2-3.
Assume that the top of the spring is moved vertically up and down according to the
equation y, - a sin wt, where '" is a constant angular frequency, a is II. constant ampli­
tude, and 1/1 is the distance from the top of the spring to the origin of the X1/ coordinate
system.
2.5 A spherical pendulum, shown in Fig. PZ-5, consists of a mass 11'/. suspended
from a frictionless pivot by a weightless rod of length I. The pendulum is not con- .
strained to swing in a plane. The only constraint is that the mass m rema.in on the
surface of a sphere of radius I.
a. In terms of the xyz coordinates, write the kinetie-coenergy-etate function for
the system.
94 Principles of e/ectromechc:mical.energy conversion

b. Explain why the xyz coordinates are not a suitable set of generalized coordi­
nates. Show that the two spherical-coordinate angles (6,4» are a suitable set of gen­
eralized coordinates.
c. Write the constraint equations between the two sets of coordinates. Are
these equations holonomic? Why?
d. Determine the Lagrangian for the system in terms of the generalized
coordinates.

,/ y
/
/
/
/
"
----y /

l
Jf

m Fig. P2-5

2-6 A flyball speed governor is illustrated in Fig. P2-6. The four rods of length
l are pivoted at the three point masses and at the origin. The rods have negligible
mlU!8. The system rotates about the x axis, causing the two identical masses M. to
move away from the axis and thus pull mlU!8 M 2 toward the origin against the spring
having compliance K. The position of Mt is a function of the angular velocity of the
system; thus, the device can be used as a speed governor. AU pivot points are fric­
tionless, andM. slides along the x axis without friction. When the two maBSe8 M.
are on the x axis, the spring is at its free length.
a. Write a Lagrangian for the system in terms of the three cartesian coordinates
x, y, z. The z axis is out of the paper.
b. Can we use the Lagrangian of part a in Lagrange's equation? Explain your
answer.

Fig. P2-6
l Lagrunge', equation 95

c. Write a. La.gra.ngia.n in terms of 8. a.nd 82 by first writing suita.ble constra.int


equa.tions a.nd then tra.nsForming the Lagrangian found in pa.rt G. (Assume 8.... 0
when the system is in the xy plane.)
d. Determine the equilibrium equa.tions for the system.
e. For II - "'0 rad/sec. find an expression for the steady-sta.te value of 8,.
2-7 For the conserva.tive electrical'network shown in Fig. P2-7:
G. Write the La.gra.ngian For the system on a. loop basis.
b. Write the equilibrium equations for the system by using Lagrange's equation.
c. The equilibrium equa.tions for the double pendulum are given by (2-51) and
(2-52). Show tha.t Eqs. (2-51) a.nd (2-52) for sma.ll a.ngula.r displa.cements of the
double pendulum reducc to the sa.mc form as the equations in part b. Note tha.t the
double pendulum contains mechanica.l mutual coupling.

Fig. P2-7

2-8 Four Iinea.r-lumped conservative electrical elements a.re shown in Fig.


P2-8. Termina.l va.ria.bles are a.lso shown for each of the elements.
G. Find the magnetic-coenergy-state function and the electric-energy-sta.te func­
tion in terms of the current va.riables.
b. Write the Lagra.ngia.n for the system of four isolated elements.

2-9 The four elements shown in Fig. P2-8 a.re joined to form the network shown
in Fig. P2-9.
G. Why a.'re the origina.l termina.l currents q., q~, q" and q4 not suitable genera.lized
coordina.tes? Ca.n loop currents q, and ti. serve as generalized coordina.tes? Why?
b. Write four constraint equa.tions expressing the origina.l currents in terms of the
loop currents til and tho Are these constraint equa.tions holonomic? Expla.in your
a.nswer.
C. Substitute the constra.int equa.tions into the La.gra.ngian found in Prob. 2-8b
a.nd find the loop equilibrium equa.tions by using La.gra.nge's equa.tion.

L,
.~C'

Fig. P2-9
YIlIQ 0
96 Principles of e/ectromeehollieal.enerllY cOllversioll

2-10 Repeat Prob. 2-9 with the four elements connected into the network shown
in Fig. P2-lO.

Fig. P2-10

2-b The four linear-lumped elements shown in Fig. P2-8 are now joined to form
the circuit shown in Fig. P2-11. Assume that the mutual-coupling coefficient Mis
equal to zero.
a. Write 'the electric-coenergy. and magnetic·energy-state functions in terms of
the individual voltage variables he, }.bJ he, and }.•.
b. Write a Lagrangian suitable for a nodal analysis in terms of these four voltage
variable!! of part a.
c. After the elements are connected into the circuit of Fig. P2·11, can the four
variables of part a serve as generalized coordinateS in Lagrange's equation? Why?
d. Write suitable constraint equations between the nodal variables and the
original four variables.
11. Transform the Lagrangian in part b in such a way that it becomes a function of
generalized coordinates, 'and use Lagrange's equation to find the two nodal equations
for the network;

Fig. P2.11

2.12 For the linear mechanical and electrical systems shown in Fig. P2-12 write
an appropriate Lagrangian and Rayleigh dissipation function and determine the
equilibrium equations by using Eq. (2-88). Do you notice any similarity between
these three systems? Hint: Look at the state functions.
2-13 A certain fan can be described by the following torque-speed data:
Speed, rpm o 200 400 600 800 1000 1200 1400 1600 1800

Torque, lb·ft o 0.4 1.0 1.7 2.7 3.7 5.0 6.7 8.5 11.2

a. Draw the characteristic curve for this diBBipative element. Use mks units.
b. For a speed of 1500 rpm find the magnitude of the Rayleigh and co-Rayleigh
dissipation functions.
c. Evaluate the rotational viscous damping coefficient D, at 1500 rpm. Compare
jD",,' with the state functions found in part b.
Lagrange'$ equation 97

Ie)
Fig. P2-12

!
Gravity

Fig. P2-14
98 Principles of electromechanical-energy conversion

2-14 In packaging a fragile piece of equipment a rigid box is to be used. A


spongy packaging material is tightly packed on all sides of the equipment. The
packaging material can be modeled 8.8 a lumped spring and a loss element, 8.8 shown in
Fig. P2-14. A cartesian-coordinate system is placed at the center of the piece of
equipment. With the equipment at the origin each of the six identical springs is
compreased a distance a. If the equipment is moved to any point (x,y,;:), the motion
is sufficiently small that the six springs still remain on their respective axes. Assume
that the damping force of the packaging material is proportional to the velocity.
a. Write the Lagrangian and the Rayleigh function for the system.
b. Determine the three equilibrium equations.
c. From the equations in part b, find the steady-state position of the system.
chapbr

FORMULATION OF EQUILIBRIUM
3
EQUATIONS FOR
ELECTROMECHANICAL
SYSTEMS

In Chaps. 1 and 2 an energy formulation is developed


for obtaining differential equilibrium equations
describing the dynamic operation of lumped elec­
trical and mechanical systems. For both electrilJal
and mechanical systems lumped elements are graph­
ically described in terms of characteristic curves
relating the basic terminal coordinates for each ele­
ment. From these curves the energy- and coenergy­
state functions are defined for all of the conservative
or nondissipative elements. These include induct­
ance and capacitance electrical parameters and
the inertial and spring mechanical parameters. For
dissipative elements the Rayleigh and co-Rayleigh
dissipation fUIlctions are used to include these ele­
ments in the scheme of the energy formulation.
The analytic mechanism which selects and com­
bines these energy-state functions, giving us the
desired equilibrium equations, is known as La­
grange's equation. In order to use Lagrange's equa­
tion, a Lagrangian function for the system has to be
formulated in terms of a set of generalized coor­
dinates. This Lagrangian combines prescribed
energy- and coenergy-state functions for the con­
servative elements in the system. By using the
Lagrangian with the total Rayleigh dissipation func­
tion for the nonconservative elements, Lagrange's
equation provides the recipe for manipulating par­
ticular partial and' total derivatives which yield
Kirchhoff's laws for electrical systems and New­
99 ton's second law for mechanical systems.
100 Principle, of electromechanical· energy conversion

So far we have only used the Lagrange technique on either electrical or


mechanical systems separately. Systems involving interactions between
electrical and mechanical parts have not as yet been considered. In fact,
for most of the systems discussed in Chaps. 1 and 2, you probably could
have formulated the equilibrium equations directly from your background
work in electrical circuits and mechanics without any new-fangled energy
method. Example 2-3 for the double pendulum and Example 2-4 for the
revolving pendulum are the only places where the energy technique
really seems to buy us anything. In Sec. 2-9 we looked briefly at how the
Lagrangian energy formulation could help us with a system involving
electromechanical interactions. It is with this type of problem that the
new method has its greatest value. All of the interaction forces will he
obtaiped in a very systematic fashion without any change in Lagrange's
equation and the formulation we have developed in Chap. 2. The energy­
state functions are somewhat more general for the electromechanical sys­
tem. In this chapter we shall carefully examine this new generality.
Then, by using the basic first law of thermodynamics, which is commonly
called conservation of energy, we shall develop these interaction forces.
We shall then see that by using Lagrange's equation in a natural way,
meaning that all partial derivatives are properly taken, these same forces
can be correctly obtained and placed in the resulting equilibrium equa­
tions. In other words, the formulation of equilibrium equations for a
wide class of electromechanical systems is a very easy job by using the
Lagrangian technique.

3-1 A TWO-PORT MAGNETIC-FIELD ELECTROMECHANICAL


LUMPED SYSTEM
Many physical devices are constructed in such a fashion that they have
both electrical and mechanical parts. l\1icrophones, audio speakers,
relays, solenoid actuators, and instrumentation transducers for measuring
position, velocity, and acceleration are just a few such mechanisms. A
slight generalization of our energy formulation is required to study such
devices. In t}1is section we shall examine the energy- and coenergy-state
functions for these electromechanical elements. The electrical portion of
the de~e can store energy in either an electric or magnetic field. There­
fore, we have two general classes of devices, namely, electric-field elec­
tromechanical devices and magnetic-field electromechanical devices.
Most physical devices contain both electric- and magnetic-field energy
storage, but one effect is usually dominant, so this classification is not too
restrictive from a practical standpoint.
Figure 3-1 shows a particular magnetic-field electromechanical device.
Electromechanical-system equations 101

;
,..---....,........::

l"ig. 3-1 A simple two-port eleetrome­


chanical system with magnetic-field energy
storage.

Notice that the device has a set of electrical terminals in addition to "
movable I1Hl(!hanical member. The term port is generally used to describe
the place in the system where either electrical or mechanical energy can be
supplied or extracted. The system in Fig. 3-1 has a single electrical port
and a single mechanical port. The voltage and current variables for the
electrical port are designated by ). and q, respectively. Similarly, the
velocity and force variables for the single mechanical port are given by
x and p, respectively. The velocity x of the movable member is meas­
ured with respect to the stationary member, and x is taken equal to zero
when the air gap betwecn the stationary and movable members is equal
to zero.
Now we would like to examine the various state functions describing
the magnetic-field system shown in Fig. 3-1. For this purpose, let us
assume that the velocity :i; of the movable member is equal to zero. In
other words, the movable member is held at some fixed position x. As
far as the electrical portion of the system is concerned, we now have a
simple lumped-inductance element. As discussed in Sec. 1-2, this element
is conveniently characterized by a plot of the total flux linkages X as a
function of the coil current q. Figure 3-2 shows typical characteristic
plots for various fixed positions of the movable member. These positions
are designated by x"' Xb, and xc. Thus, with a current qb and with the
movable member held at X6 the state of the system is described by point B
in the X'-Ij' plane of Fig. 3-2.1 Holding the current constant at the
value qb and repositioning the movable member to position x" causes the
state of the system to move to point A with an increase in the flux linkages
to a value of X For any final current q and position x a single point on
G•

the plane of Fig. 3-2 is determined. In fact, specifying two out of the
three variables X, q, and x is sufficient to describe the particular state point
of the system. The plot of Fig. 3-2 gives the third variable once the other
two are specified. Notice that nothing has been said about how the state
1 The primes on the coordinates are used to indicate integration variables. Coordinate
symbols without the primes refer to system state points, or state values of the
coordinates.
\

102 Prim;ipl., of .'.dromechanical·_rgy conversion

Fig. 3-2 Characteristic curves for


various fixed positions of the movable
ti' memher of the magnetic-field systcm.

point ls reached. The only thing that matters in determining the state
point of the system is the particular final values of ~, q, and x, and not how
these values were obtained.
With the mechanical member held at a fixed position x the magnetic­
field energy storage can now be computed in the usual way as being

W... {>',x) /0). q'(>.',x) d'A' (3-1)

in accordance with Sec. 1-11. The only difference here lies in the fact
that the current q is a function of not only the flux linkages>. but also the
fixed position x of the movable member. Thus the magnetic-energy
storage W .. becomes a function of both ~ and x. The integral given by
Eq. (3-1) is readily interpreted as the shaded area above the particular
characteristic curve for the particular final position Xb, as shown in Fig.
3-3. The magnetic energy stored in the system, for a given final state of
the system, is a function of only the final state coordinates. These state
coordinates define a state point, a.nd the shaded area above the curve
going through this state point is the energy stored in the system.

Fig. 3~1 Energy- and coenergy-state


functions for the two-port magnetic-field
ti' system.
Elec'romechc:mical-q..em eqllClfiolll 103

In a similar fashion the coenergy stored in the system, defined as the


shaded area below the curve, is shown in Fig. 3-3. This magnetic-coen­
ergy-state function can be calculated from

W:"(q,x) = /0 X'(q',x) dq'


4
(3-2)

and is a function of both the current q and the position coordinate x ..


The sum of the energy- and coenergy-state functions must be the total
area of the X-q rectangle; thus
(3-3)

Equations (3-1) to (3-3) are not very different from the corresponding
equations for the simple electrical-inductance element. The mechanical
part of the system offers a small additional complication in that the
position coordinate influences the state (unctions. However, in all these
integrations x is treated as a constant and is simply carried along, and it
appears in the final results.

Exomple 3·1
The device pictured in Fig. 3-4 is a common electromechanical actuator, or force gen­
erator. Hupplying a current to the electrical winding will cause the plunger to tnove
upward, thereby pulling the mechanical load to which the plunger might be attached.
The device provides a means of controlling mechanical motion by an electrical signal.
Connecting the plunger to the valve stcm of a hydrl\ulic or pneuml\tic valve gives us a
remotely operated, electrically controlled valve. Connecting 0. pair of electrical
contacts, one to the stator and one to the plunger, makes this device the common
electrical relay. A very large number of rommon devices are founded upon the simple
mechanism shown in Fig. 3-4.
Let us, at this poil\t, find the magnetic--energy- and -eoenergy-state functions for
the electromechanical system shown in Fig. 3-4. Assume the permeability of the iron

r----.,
I
I
I Iron
t stator
I
I
I
Contour • Iron
plunqer
Fig. 3-4 An electromechanical sys­
tem with magnetic-field energy storage.
10" Principlea of electromechanical· energy conversion

ill many orders of magnitude greater than that of air. Also, take the permeability of
the brMII to be equal to that of air. Neglect all fringing fields on the ba.sia that the
air-ga.p length x and thiekness d of the bra.ss are small eompared with the radius of
the plunger.
Soluli<m: The energy- and eoenergy-etate funetions ean be readily computed onee
the eharacteristie eurves for the deviee are known. We require the total flux linkages
X in the eoil a.s a function of the eurrent I} in the eoil and the Position x of the plunger.
The flux link:agee are determined in the following manner.
The magnet ie-intensity vector in the bra.ss and air ean be found by using Amp~re'8
eircuitallaw, whieh can be written in the form

¢c B • d1 - Is J . dS
This vector equation tells us that the line integral of the magnetic-intensity vector H
arounlany e10sed contour C must equal the total current croesing a surface 8 with
this eo tour a.s its periphery. When the contour shown in Fig. 3-4 is ehosen, Amp~re's
eircuital law becomes
H.x + H"d - -Nq
where H.a. and H"a. are the magnetic-intensity vectors in the air and bra.ss, respec­
tively. The unit vector a. is positive in the positive x direction. Notice that we are
traversing this contour in the positive x direction, as shown by the arrows on the
contour in Fig. 3-4. By placing a surface over the contour, we !lee that the total eur­
rent NI} croeses the surface in a direction out of the paper. By the usual right-hand
convention, currents into the paper are positive for the contour direction we are using. 1
Therefore, the total current NI} croesing the surfaee ha.s a negative sign in Amp~re's
circuital law. The magnetic-field intensity in the iron is taken to be zero compared
with that in the bra.ss or air. This approximation is based on the permeability of
iron being many orders of magnitude greater than that of bra.ss or air.
Neglecting fringing, the magnetic-f1ux-density vector in the bra.ss equals that in
the air gap between the plunger and the bra.ss. Therefore, the magnetic-f1ux-denllity
vector is given by
B - ,.,II.a. - ,.,II"..
where the permeability of air and bra.ss is taken to be equal to,.". By substituting
into Amp~re'8 cireuitallaw, we have for the magnetic-f1ux-density vector
B ___ ,.oNl} ..
d+x
The total flux ,., existing in the bra.ss and in the air gap is defined by

,., - IsB'dS
where the surface 8 croeses through either the bra.ss or the air gap. By evaluating the
surface integral, we have
,., _ __ ,.oNAI}
d+x
I Place the fingers of your right ha.nd along the contour pointing in the direction of
the arrows. Your thumb, which points into the paper, gives the correct convention
for positive currents.
EI.ctromechon;cal••yllfem equofiOM 105

Fig. 3-5 Characteristic curves for the


magnetic-field elet'tromechanicai system. q'

where A is the face ares of the plunger or the brass, as shown in Fig. 3-4. Positive
flux is taken in the positive z direction; thus, the negative sign indicates that the flux
is in the negative z direction.
Since all of the flux. links each of the N turns, the total flux linkages are given by

(3-4)

where the negative sign arises from the fact that A, as shown in Fig. 3-4. would become
positive if the flux 'I' were increasing in the upward or negative x direction. In other
words, A equals -N d./dt. since. is positive for Rux in the positive x direction, or
downward.'
Figure 3-5 shows a plot of h as II. function of q for various fixed values of x. Notice
that h is a linear function of q; thus, all these characteristics are straight lines.
From Eq. (3-2) the ma.gnetic-coencrgy-state function is given by

W~(q,z) - (4 h'(q',X) dq'


Jo •
w' (Ii z) _ (4 p.NtA q'dq' (3-5)
.. , Jo d +z
, p.NtAq"
W.. (Ii,x) - 2(d + x)
The magnetic energy stored can be ca.lculated from Eq. (3-1). On rearranging
(3-4). we have

k(h ) _ (d + z».
'I ,x PeNtA

, This discussion of signs is usually quite confusing. An easy check is simply to use
the right-hand rule as follows. Curl the fingers of your right hand around the coil in
the assumed direction for positive current. Your thumb, which now points upward,
gives the corresponding flux direction for positive current. This direction corresponds
to the positive sign on >.; therefore, h should be positive if 9 is positive, as given by
Eq. (3-4).
106 Principle. of electromechanical-energy con"e,.,ion

If this i8 8ub8tituted into Eq. (3-1), the magnetic-energy-state function i8 simply


(d + :l;:)A"

W...(A,:l;:) - 2p..N"A

If the current q(A.:l;:) i8 substituted into the expression for the coenergy-state function
given by Eq. (3-5), we see that

w... - W~
This result is to be expected, since the A versus q characteristics shown in Fig. 3-5 are
all straight lines. For any state point P the energy is equal to the area of triangle
OPQ and the coenergy is equal to the area of triangle OPR. Obviously, area OPQ
equals area OPR, since each equals half of the area of rectangle OQPR. Also, we can
show that

in accordance with Eq. (3-3).

3-2 A MULTIPORT MAGNETIC-FIELD ELECTROMECHANICAL SYSTEM

Let us now generalize these magnetic-energy and coenergy-state functions


to include a magnetic-field electromechanical device having m mechanical
ports and n electrical ports instead of just the simple case of one mechanical
and one electrical port that we have so far been considering. Assume that
the (m + n)-port system arrives at its final state by first fixing all m of the
mechanical coordinates. The n electrical flux linkages are then increased
in sequence from zero to their respective final values. The magnetic
energy stored by arranging the system in this fashion would be

,x"') dX~ (3-6)

following the general form of Eq. (1-56), where the mechanical ports are
not present. Notice in the first integral of Eq. (3-6) that all At coordinates
except A~ are zero. The second integral holds A~ constant at its final
value AI with A~, A~, ..• , A: still equal to zero. The same sequence con­
tinues for each successive integral. Again let us remember that even
though the system does not reach its final state (as represented by AI,
Xl, . • • ,X",XIIX2, • • • ,x"') in such a manner, the energy-state function
calculated this way is still correct, since it depends only on the final state,
and not on how the system got there. When we use the abbreviated
Electromechanical-system equation, 107

notation introduced in Chap. 1, Eq. (3-6) for the magnetic energy stored
becomes

W ... (A,X) I"


i-I
lo~' q~(A/,X) dA: (3-7)

The shorthand notation can be interpreted only by referring to Eq. (3-6).


By duality we can see that the magnetic-coenergy-state function can
be calculated as

W~(ql,q:, .. ,q",XhX" . . . ,x..)


9
= 10 ' A;(q;,O. ' .. ,O,X"X2, . ,x..) dq;

+ )0( q, "2
1 (. .1 °
~ Ql,Q" . . . , ,X"Xt, .•. ,x.. ) d Q,
.1

+ ... + It A:(q"q2, . , . ,q:,x"Xt, , .. ,x...) dq~ (3-8)

Here we have fixed each of the m mechanical coordinates with all currents
equal to zero. Then, in sequence, we increase each current to its final
value and hold it there as we adjust the succeeding current.s. The coen­
ergy expression can be abbreviated simply by
.
lV~(q,x) = I loti' A~(q',X) dq:
i-I
(3-9)

The energy- and coenergy-state functions are still related by


..
W... (A,X) + W~(q,x) = L Aiqi (3-10)
i-I

again using our shorthand notation.

Example 3-2
A particular magnetic-field electromechanical device having two mechanical ports and
three electrical ports has the following characteristic relationships:

>-,{qloq"q.,x"x,) - 5x,x.qI' + i!.


Xl
~ q,
q, + ""2

", >-,(q"q"q,.xlox,) - i!.


Xt
q, + 7xlx,q,' + ~
,xIX:
q, (3-11)

>-.(q"q"q.,x,x,) - .!
X,
ql + ~
XIX,
q, + 9xl'x,"h

Determine the magnetic-coencrgy-state fun('tion for this device.


Solution: The magnctic-coenergy-state function is formulated by using Eq. (3-9),
or in expanded form by (3-8). For the particular device here being considered the
108 P";tK:;ple. of electromechanical-energy conv.,..ion

general formulation for the coenergy becomes


W' (. . ) (q" (" ).1 (q.,... ,
.. ql,q.,q"x.,x. -)0 X. q.,O,O,x.,x. dq. +)0 X.(q.,q.,O,x.,x.) dq.
(q,. .' •
+)0 X,(q"q.,q.,Xl,x.) dq.

We have fixed the mechanical coordinates at some particular values x. and x.. Then,
in succession, we increase the currents q;, q;,
and q;
from zero to their final values,
given by ql, ql, and q., respectively. The state of the device at any time instant is
given by specifying q.. q., q., x., and x.. Analytically we have taken the system to its
final state and computed the coenergy in the process. The resulting expression for the
coenergy is correct even though the system does not actually reach the final state in this
manner. On evaluating the coenergy, we have

As an additional exercise let us evaluate the coenergy by raising the currents from
q;
zero to their final values, taking first, then q;,
and finally q;.
The general coenergy
expression, from Eq. (3-9), now becomes
' . ('I"., I
W .. (q,x) - )0 X.(O,O,q.,xt,x.) d4. + )0(v, X.(O,q.,q.,x.,x.) dq.
I I ,

('ii
+ )0 I I I
X,(q"4'.,q.,x,,x.) dq,

When the defining flux-linkage relationships are substituted, the coenergy is given by

When these three integrals have been evaluated, the resulting coenergy is seen to be
identical with Eq. (3-12).
The original characteristic relationships, given by Eqs. (3-11), have a certain
symmetry. Notice that the contribution to X, from current q. is identical to the con­
tribution to x. from q,. A similar statement can be made for all other mutual terms,
or terms involving flux linkages in one portion of the system due to currents in another
portion. This property of too defining relationships must be true in order for the state
functions to be independent of the path by which the final state is reached.
If the magnetic-energy-state function were required, then Eqs. (3-11) would have
to be solved for the currents in terms of the flux linkages in order to use Eq. (3-7) for
the energy-state function. The defining relationships of (3-11) make this look like a
very tough job. A simpler procedure would be to use Eq. (3-10), which gives the
magnetic-energy-state function for this device as

W.. - q,x. + q.X, + q.x. - W~


Eledromechonkol-qstem equofiofl$ 109

Substituting Eqs. (3-11) for }.I, }.o, and }.I and Eq. (3-12) for W~ would give the mag­
netic-energy-etate function in terms of the currents ql, ql, and q, and mechanical
coordinates :1:1 and :1:,. This is not the usual functional representation of W.., since
normally the flux linkages rather than the currents are used. However, with the
defining relationships (3- \I) 80 difficult to manipulate, we are practically forced to
accept W .. as a function of the q's and x's rather than the }.'s and x's.
This discussion is included to point out a slight difficulty that we have not, as yet,
faced with our energy formulation. The difficulty arises when the energy and coenergy
are correctly formulated but are functions of variables differing from their defining
relationships. As we proceed with the energy formulation, keep this problem in mind
and we shall see what its consequences and the required correction factors are.

3-3 A TWO-PORT ELECTRIC-FIELD ELECTROMECHANICAL


LUMPED SYSTEM

The second classification of electromechanical devices includes those which


store electric energy in the form of an electric field. Figure 3-6 illustrates
the general form of this type of device, with one electrical and one mechan­
ical port. Two conducting plates are arranged in such a manner that if a
voltage>" exists between the plates, then an electric field is established and
electric-field energy storage is present. The bottom plate is movable with
respect to the top plate, and its general velocity and force mechanical port
coordinates are defined by :i; and p, respectively. The velocity :i; is meas­
ured with respect to the stationary upper plate. Similarly, the voltage
and current variables of the electrical port are defined by ). and q,
respectively.
The characteristic curves for an electric-field device are plots of the
accumulated charge q as a function of the potential difference >... These
curves are plotted fot various fixed positions of the bottom plate. Such a
set of curves is shown in Fig. 3-7. The state of this device is, in general,
defined by giving q, >.., and x as functions of time. If the characteristic
curves, shown in Fig. 3-7, are available, then only two of these three coor­
dinates need be given, since the curves will supply the third.
The characteristic curves cal! be determined in many different ways.
For simple geometries the electric fields can be determined and the curves
can be calculated analytically. For more complicated geometries a flux-

Fig. :J-6 A simple two-port electro­


mechanical system with electric-field energy
storage.
110 Principle, of electromechanical· energy convenion

Fig. 3-7 Characteristic curves for various


fixed positions of the movable member of
i.' the electric-field system.

plotting technique can often be used to obtain the characteristic curves. I


If the physical device is available, the characteristic curves can be deter­
mined experimentally. The method chosen to obtain the characteristic
curves largely depends on the complexity of the system being studied and
the required accuracy of the solution.
The energy- and coenergy-state functions for the electric-field device
can now be formulated in a manner exactly analogous to that for the
magnetic-field system in Sec. 3-1. When the mechanical position is held
fixed at the final state value, the charge can be increased from zero to its
final value. The electric-field energy-state function under these condi­
tions is

W.(q,x) = fov X'(q',x) dq' (3-13)

where x is held fixed or constant throughout the integration with respect


to q'. Referring to Fig. 3-8, if the state of the system at some specific
instant of time is at point B, then the coordinates have values q", x", and
x". The electric-field energy-state function, as given by Eq. (3-13), for
the state point B is seen to be the shaded area above the x' = x" curve in
the rectangle with sides ql> and x", as shown in Fig. 3-8.
In a similar fashion, with x held fixed, the electric-field coenergy is
given by

w:(~,x) fo'!.. q'(~/,X) d~' (3-14)

For state point B in Fig. 3-8 the coenergy integral is the area in the
ql> by ~ rectangle below the x' Xo curve. The energy and coenergy
defined by the integrals of Eqs. (3-13) and (3-14) must give the total area

I For an excellent discussion of flux plotting see R. P!onsey and R. E. Collin, "Principles
and Applications of Electromagnetic Fields," chap. 5, McGraw-Hill Book Company,
New York, 1961.
Electromechanicol-,y"em equation, 111

Fig. 3-8 Energy- and coenergy-state


functions for the two-port electric-field
system.

of the rectangle when t.hese state functions are added together. Thus we
can, as usual, write
W.(q,x) + W;(X,x) (3-15)

showing that the energy and (!oenergy are not independent state functions.

fxampl.3-3
An electrostatic voltmeter is shown pictorially in Fig. 3-9. The device consists of a
set of stator plates electrically connected together. A set of rotor plates, al80 con·
nected electrically together, are interspaced between the stator plates as indicated in
the figure. However, the stator and rotor are electrically insulated from each other.
The rotor plates are mounted on a shaft to which are attached a torsional spring and
a pointer. A later example in this chapter shows how the device can serve as a volt-

Fig. 3-9 A top view of an


electrostatic voltmeter.
112 Principle. of electromechanical· energy conver.ion

Pointer

LII Ir----""'!{---:"",-,...,,.:
A
....f-+'--- • + + +
1------';;.;;,-;;;..;;;
..
• + ....

........
.. .. .. +

.... + ..
.. .. + ..
-
q

.. . . .. .. .... +
Stator
plotes ..,..-Pivot
Rotor
plates

Fig. 3-10 An edge view of the electrostatic voltmeter.

meter. For the time being, let us simply determine the electric-energy- and -coen­
ergy-6tate functions for this electric-field electromechanical device. Assume that
there are four stator plates and four interspaced rotor plates. Neglect fringing fields
at the edges of the plates.
Solution: The energy- and coenergy-state functiQns can be readily determined once
the characteristic curves for the device are known. These curves are plots of the
charge on either the stator or rotor as a function of the voltage applied between stator
and rotor for various fixed values ofthe angular position of the rotor. In other words,
we need the function q().,II).
Figure 3-10 shows an edge view of the device. Suppose we place a positive
charge q on the rotor plates by means of the current q indicated at the terminals of the
voltmeter. Then an equal amount of negative charge will be found on the adjacent
IlUrfaces of the stator plates. The charge distribution is shown in Fig. 3-10 by the
respective plus and minus signs. GaU88's flux theorem can be used for determining
the electric field in the space between the stator and rotor plates. This theorem can
be given in integral form by

where D is the electric-displacement vector, 8 is a closed surface, V is the volume con­


tained by the surface 8, and p is the volume charge density. Applying this theorem in
the space labeled A in Fig. 3-10, we would have

D. [
a +
(...R,· - ... R,') -z;- II] - 14q
where D. is the magnitude of the electric-displacement vector in the direction of the
unit vector a.. The bracketed term OD the left side of the equation is simply the
adjacent area between the stator and rotor plates in the space labeled A. Refer to
Fig. 3-9 for the definitions of R" R" and a. The charge q is divided into 14 equal
Electromechanica/.system equotions 113

parts on the rotor plates, as shown in Fig. 3-10.' The electric-field vector in the space
A is therefore given by

E-~- q ....
'0 7' 0(R,' - RI')(OI + e)
where <0 is the permittivity of air, equal to 8.854 X 10- 11 farad/m.
The voltage of the rotor with respect to the stator is designated in Fig. 3-10 as~.
By definition, the potential difference is related to the electric-field vector by the
following line integral:

~ - - Ie E·dl

where the contour C starts at the stator plate and ends at the rotor plate. In the
IIpace A we can choose a simple straight-line contour; thus
d1 - -/h ••
The minus sign is necessary because the selected contour goes from the stator to the
rotor, which in the space A is in the negative x direction. Evaluating~, we have
>- _ _ ~_~_!Lh____

7.0( R.· RI')(OI + e}

where h is the separation between stator and rotor, as shown in Fig. 3-10. The char­
aeteristic curvCl! of the electrostatic voltmetcr arc thcrefore obtainable from

q(>-,e} _ 7••Ul '·h- RI") (01 + e)>- (3-16)

From Eq. (3-14) the electric-iloenergy-state function is given by

W (>-)
•' ,e - 7••(R." - RI") (
2h a + 9)>-. (3-17)

Similarly, using >-(q,9), the electric~nergy-state function according to Eq. (3-13)


would be
.Iuj'
w.(q,e) - 14.o(R,i _ RI')(a + 9) (3-18)

We can easily show, by substituting Eq. (3-16) into Eq. (3-18), that here W. - W:.
The equality of the energy- and coenergy-state functions is expected, since q is a linear
function of >- for fixed values of Ii. Thus the characteristic curves would be straight
lines. We can also verify Eq. (3-15), which shows that the sum of and W.
must W:
equal q>-. This last step is left for you to perform as a brief exercise.

3·4 A MULTIPORT ELECTRIC-FIELD ELECTROMECHANICAL SYSTEM

As in the case of the magnetic-field electromechanical system we shall find


that generalizing these results will be advantageous. Therefore, consider
I The equation is a valid approximation only for a eertain limited range of rotor
angular displacements. We are tacitly ll.88uming that the rotor never lea,ves this
restricted region.
114 Principle. 01 electromechanical· energy conver,ion

an electric-field electromechanical device having m mechanical ports and


n electrical ports. The state functions can be developed in the usual way
by first fixing all m mechanical position coordinates. We completely build
the system with all mechanical elements set at their final state positions.
We then successively raise the n electrical charges from zero to their final
values. The total electric-field energy stored would be

(3-19)
I
This equation for the electric-field energy stored is identical with Eq. (3-6)
if ). and q are replaced in (3-6) by q and X.
The first integral on the right-hand side of Eq. (3-19) has all q' coor­
dinates except q;
equal to zero. Therefore, this first integral evaluates
q;
the energy stored as is raised from zero to its final value ql with all other
q' coordinates equal to zero. All the mechanical coordinates are held at
their final values. The second integral has q;
held at its final value ql
with all other q' coordinates still equal to zero as q; is' increased from zero
to its final state value q2, thus evaluating the contribution to the final
energy stored during the second process. The same scheme then con­
tinues for q~, q~, and on up to q:. Throughout the entire process the
mechanical coordinates are held fixed at their final state values. Since
the energy stored is a state function, the result obtained by Eq. (3-19) is
identical with that obtained by the coordinates reaching their final state
values by means of any other path. Equation (3-19) can be written more
conveniently by using the abbreviated notation introduced in Chap. 1.
Thus, we can write
n

W.(q,x) = I
i-I
q
fo , X:(q',x) dq: (3-20)

Equation (3-19) interprets the abbreviations used in (3-20). This short


form is useful only if you remember how it is to be expanded. On looking
at Eq. (3-20) alone, the expansion is far from obvious.
Since the discussion for the electric-field device exactly parallels that
for the magnetic-field electromechanical system, the electric-field coenergy­
state function should be quite evident. In shortened form, the coenergy
can be given as

W~(X,x) = i
i-I
fo'A' q:(X',x) dX: (3-21)
Elec,romechonica/ •• y.,.m equation. 115

where the coenergy is expressed as a Cunction of all the voltages and


mechanical position coordinates.
Again we have a constraint between the energy and coenergy func­
tions. Their sum must obey the relationship
.
W.(q,x) + W:(}",x) = L qi}", (3-22)
1-1

Thus, if we can calculate either the energy or the coenergy, then we can
find the other state function. If the q' versus }", characteristics are all
straight lines for various constant values for the mechanical coordinates,
then the energy- and coenergy-state functions are equal. However, if the
device is not linear, then these two state functions are not in general
equivalent, or equal, and, as we shall see, care must be exercised to select
the proper function.

3-S MECHANICAL FORCES OF ELECTRICAL ORIGIN I (

FROM THE PRINCIPLE OF VIRTUAL WORK

In this section we shall investigate the mechanical forces in the electro­


mechanical systems which are due to the electrical portions of the device.
Relays, audio speakers, solenoid actuators, electric meters, and many other
common devices have mechanical motions due to certain electrical inputs.
The nature of these forces of electrical origin is now to be analytically
examined. We must be able to completely specify these forces if the
equilibrium equations for the devices are to be formulated. Lagrange's
equation, introduced in Chap. 2, is completely sufficient for formulating
the equilibrium equa,tions. In fact, we shall see that the development of
the Lagrangian function for electromechanical systems is identical with
the procedures for the separate electrical or mechanical systems discussed
in Chaps. 1 and 2. The only difference lies in the fact that the electrical
state functions involve both electrical and mechanical coordinates, as seen
in the preceding sections oC this chapter.
However, instead of using Lagrange's equations simply in a "cook­
book" Cashion, let us investigate the mechanical Corces of electrical origin
from a slightly different standpoint. Figure 3-11 shows, in a general
form, an electromechanical device having the capacity to store electric
energy in an electric field. The device has one electrical port with

Electric
Fig. 3-11 A two-port \0881_ electric- Electromechonicol device ,,'
field electromeehanicalllYlltem.
116 - Principl.s of electromechanical· energy conversion

voltage and current terminal variables defined by X and q, respectively.


The device also has one mechanical port, corresponding to a single movable
mechanical member, with defined velocity :i: and external applied force p.
Let us assume that, at some particular instant in time, energy is being
supplied to the device through both the electrical and mechanical ports.
The rates at which electrical and mechanical energy are being supplied
are, of course, the electrical- and mechanical-port powers. The powers
are given simply by

p. = Xq (3-23)
p ... = xp (3-24)

where p. and p .. are the electrical and mechanical powers, respectively,


and k, q, i, and p are the port variables as defined in Fig. 3-11. In a
differential increment of time dt the electrical and mechanical energy
supplied is found simply by multiplying both sides of Eqs. (3-23) and
(3-24) by the time increment dt.
At this point in the discussion, assume that this electromechanical
device is lossless. Later we shall see that if loss, or energy dissipation
(usually in the form of heat), is present, it can be accounted for by a
lumped dissipative element external to the system. i Now consider where
the electrical and mechanical energy, which is pouring in through the two
ports, goes. Since the device is assumed not to dissipate any of the
energy, where can it go? To be absolutely rigorous, the answer to this
question depends on many factors. The energy could, in part, be radiated
in the form of electromagnetic waves. Part of the energy could also be
converted into mass according to Einstein's law. However, with most
electromechanical devices these two effects are so small as to be negligible
for all practical purposes. Part of the energy could be stored in mechani­
cal kinetic energy, since the mechanical member could, in general, achieve
some finite velocity. Suppose, however, that we are very careful in how
we supply the energy SO that none goes into kinetic storage. This means
that the mechanical member is so constrained that it virtually does not
move or that it moves with a very, very small velocity. Also assume that
the system shown in Fig. 3-11 contains no springsj thus we have no
mechanical potential-energy storage. The major receiver of the incoming
energy is the electric field. The assumption made here is that all of the
supplied electric and mechanical energy goes into electric-field storage.
For a magnetic-field device, the energy will ~o into magnetic-field storage.

1 A pra.ctical inductive element stores magnetic energy but also has some dissipa.­
tion. UsUAlly we account for this dissipative component by placing a resistance in
series or pa.rallel with a lossless inducta.nce. The II!.me procedure is used with a
practica.l capacitive element.
Electromechanical''Y.tem equation. 117

Formally, the statement that all the energy supplied must be


accounted for, meaning that no energy is created or destroyed, is the con­
servation of energy law. ThiR statement is also known as the first law of
thermodynamics, where the u.,:ual phrasing says that the energy supplied
to a system goes into internal storage and output work.- Since output
work is simply an algebraic sign change on energy flow in through the
mechanical port, this is equivalent to the previous statement. Therefore,
in an increment of time dt, the first law applied to the system of Fig. 3-11
can be written as
i. dq + P dx = dW.(q,x} (3-25)
where X dq and p dx are the electric and mechanical energies flowing in
through their respective ports in the differential time increment dt.
These results are obtained by multiplying Eqs. (3-23) and (3-24) by dt.
Weare using i. dq and p dx rather than q fA and :t dp simply for conven­
ience. The term dW.(q,x} in Eq. (3-25) represents the change in the
electric-field storage in this time increment dt. Expanding the total dif­
ferential of the electric-field storage gives us

dW (q x) = aW.{q,x) dq
• , aq
+ a~-"(q,~
ax
dx (3-26)

Remember that in taking the partial derivative with respect to q, ~he


variable x is held constant, and in taking the partial derivative with respect
to X, q is held constant.
Using Eq. (3-13), the first partial derivative in {3-26} can be manipu­
lated as follows:

{3-27}

The result of Eq. (3-27) has a simple graphical interpretation, as shown in

Fig. 3-12. If the system is operating at state point A, then the energy

Fig. 3-12 Graphical interpretation of


8W./Bqi z held constant. i'
118 Pr;fIC;p'•• of eleetromeehanica'·.tJftr9Y con.,en;on

stored is given by area OAD. Changing q by the increment CD, while


holding x fixed, changes the energy stored by an amount proportional to
the area ABCD. Thus the change in W., specified by aW., for incremental
changes in q is very nearly equal to }" times the change in q. Forsmall
enough changes in q, }" does not change appreciably between points A and
B. Thus the change in W. divided by the change in q, while x is held
fixed, equals }", the voltage at this state point.
Substituting (3-27) and (3-26) into Eq. (3-25) gives us
}" dq + P dx = }" dq + aW.(q,x)
ax
dx (3-28)

which reduces to
I. aW.(q,x) I

P= ax (3-29)

Equation (3-29) tells us that the externally applied mechanical force p is


equal to the change in the electric-field energy storage with respect to an
incremental displacement in the position of the mechanical member while
the charge q is held fixed. Figure 3-13 shows graphically what this partial
derivative means. The two characteristic curves are drawn for the
mechanical member first held at position x and then at position (x + ax),
where ax is a small increment in this position. The change in the energy
stored while the charge is held fixed is proportional to area OAB. There­
fore, the aW.(q,x)/ax is the ratio of a.rea OAB to ax 8.8 the increment ~x
becomes smaller and smaller and approaches zero 8.8 a limit.
Let us now backtrack a little bit and review a few things about the
externally applied force p. We did not allow any energy storage in the
form of mechanical kinetic energy. This would mean that the velocity
of the mechanical member is constrained to be very close to zero. The
only way we have to impose such a constraint is through this external
force p. Suppose that the electrical parts of the system were generating
a force which was being exerted on the mechanical member. Then in

q'

Fig. 3-13 Graphical interpretation


i.' of aw./ax; q held constant.
Electromechanical.system eqtIOtionJl 119

Fig. 3-14 The force of electrical origin


is equal and opposite to the externally
applied force.

order for the member to be held at zero velocity, the external force p
would have to be equal and opposite to the force of electrical origin.
Figure 3-14 illustrates this situation, withf. as this force of electrical origin.
The force of electrical origin can be determined by using Eq. (3-29) as
being

(3-30)

since p = f.. Notice in Fig. 3-14 that the force of electrical origin is
positive if it is in a direction opposing the assumed positive direction for x.
In other words, positive f. tries to make the velocity of the mechanical
member negative according to the sign convention being used.
The procedure we have used to determine the force of electrical origin
bears the imposing title of the principle of virtual work. Since the velocity
of the mechanical member is held at zero, meaning x = there can be no 0:
displacement or change in x. Therefore, no work can actually be done,
or no energy can be supplied to the system through the mechanical port.
However, in order to achieve the result of Eq. 3-29, we assume a virtual
or infinitesimal displacement in x as p struggles to hold x as near to zero
as possible. We are always sufficiently successful in this manipulation of
p that the mechanical kinetic energy is negligible, but not quite successful
enough to have dx equal to zero, causing a question to be raised concerning
its cancellation in going from Eq. (3-28) to Eq. (3-29).

3·6 THE FORCE OF ELECTRICAL ORIGIN USING


THE COENERGY-STATE FUNCTION
The force of electrical origin can also be calculated from the coenergy­
state function. Equation (3-15) gives a relationship between the energy­
and the coenergy-state functions, which for convenience is repeated as
W.(q,X) + W~(~,x) = q~ (3-31)
This relationship simply means that for any state point, defined by q, ~,
and x, th$:l total area of the q by ~ rectangle is equal to the sum of the
energy- and the coenergy-state functions. In Fig. 3-13, with the system
120 Principle, of electromechanical-energy conversion

at state point A, the energy, area OAC, plus the coenergy, area OAD, sum
to give the rectangular area ODAC.
The total differential of the energy-state function can be expanded
as follows:
dW.(q,x) = d[q>.. - W;(>",x)]

= q dX + Xdq - aw:('J..,x) dX _ aW;('J..,x) dx


(3-32)
ax ax
From Eq. (3-14), we have
aW;(>..,x)
a>.. = q (3-33)

Therrfore, the total differential of the energy-state function becomes

dW.(q,x) Xdq - aw~~x,x) dx (3-34)

upon simplification of (3-32) by using (3-33). Upon substituting Eq.


(3-34) into the first-law equation (3-25), we have

>.. dq + P dx >.. dq - aW;(>"f) dx


ax
(3-35)

which reduces to
. aW;(>--,x)
p = - ax (3-36)

A graphical interpretation of Eq. (3-36) is shown in Fig. 3-15. As


x is changed by an increment ~x, while the voltage is held constant, the
state of the system moves from point A to point B. The change in the
coenergy is given by the shaded area OAB. Equation (3-36) is the ratio
of the area OAB to ~x in the limit as ~x goes to zero, the voltage being
held constant at its state value X. Since the force of electrical origin is

q'

Fig. 3-15 Graphical interpretation of the


i' aw:/Ox; ~ held constant.
Electromechanical.,ystem equation, 121

equal and opposite to this externally applied force p, we have, finally,

f. = _ aw~(x,x)
(3-37)
ax
as the force of electrical origin from the coenergy-state function.
Notice the negative sign in Eq. (3-37). Figure 3-15 shows the coen­
ergy increases as x is increased by an amount ax. On the other hand,
Fig. 3-13 shows that the energy-state function decreases (i.e., the change
in the energy is negative) as x goes to (x + ax). Therefore, a negative
sign must appear in Eq. (3-37) for f. in terms of the coenergy if it does not
appear in Eq. (3-30) for the force of electrical origin in terms of the energy­
state function.
There is one further word of caution concel1ling these results.
Observe that the energy-state function in Eq. (3-30) is a function of q
and x. The aW.(q,x)/ax requires that q be held constant. Therefore W.
must be a function of only q and x. Expressing W. as a function of X
and x would not, in general, give the correct force of electrical origin.
The sanJ(lll'9;rgument holds for expressing the coenergy W~ in Eq. (3-37) as
a function of only Xand x, not q and x. In Sec. 3-10, we shall reformulate
the force of electrical origin by using the energy as a function of X and x
and the coenergy as a function of q and x. Sometimes these functions can
be calculated only in this form, and in these cases it is important to be able
to find f. from the reversed functional dependences. For the time being
let us work with the more straightforward forms given by Eqs. (3-30)
and (3-37).

3·7 THE PRINCIPLE OF VIRTUAL WORK FOR A MULTIPORT


ELECTRIC-FIELD SYSTEM

Let us now generalize the results of Secs. 3-5 and 3-6 for an electric-field
system having m mechanical ports and n electrical ports. Applying the
first law of therodynamics to such a system gives us

n '"
L Xi dqi + L Pi dXi = dW.(q),q2,
i-I i-I
. . . ,q.. ,XI,XS, ••• ,x...) (3-38)

again assuming that all energy supplied to the electrical and mechanical
ports goes into electric-field storage. The first term on the left-hand Side
of Eq. (3-38) represents alfUteeileTgy pouring in through the n electrical
ports. The second term is the energy through the m mechanical ports.
The right-hand side is the change in the electric-field energy storage.
122 Prin<;ip/e. of eledromeehoniccr/-energy conversion

Expanding the total differential gives us

dW.(q,x) ,., ~ oW;(~,x) dq. + ~o~;~~,X) dx; (3-39)


i~l q. i~l 1

where we use our shorthand functional notation to indicate that W. is a


function of all the q's and all the x's, as shown in Eq. (3-38). We can
show that
oW.(q,x) = >-. (3-40)
oq.
by referring to Eq. (3-19) and remembering that the partial derivative
with respect to qi means that all other q's are held fixed.
Substituting Eqs. (3-40) and (3-39) into Eq. (3-38) gives us
'" '"
\' . -d - - \' oW.(q,x) d _
.....
'- p. x. - '- ox- X.
(3-41)
i_] i-I t

after canceling the identical summations appearing On both sides of the


equation. Equation (3-41) can be manipulated into the form

~
'-
[ .
Pi -
oW.(q,x)] d
ox- Xi
0 (3-42)
i-I '

The only way to satisfy this equation for any arbitrary set of virtual dis­
placements dXi is to have the bracketed term equal to zero for each i.
Therefore, we have

Pi - oW.(q,x) =0 i 1,2, . . . , m (3-43)


OXi
By using arguments absolutely identical with those used in discussing the
two-port device, we can show that the force of electrical origin on the ith
mechanical member equals the ith externally applied force P., which is
used to maintain the virtual displacement constraint. Therefore, the
forces of electrical origin on the m mechanical members are given by

<I.). = oW.(q,x) i = 1,2, . . . ,m (3-44)


ox,

We can also calculate the forces of electrical origin by using the coen­
ergy-state function instead of the energy-state function. Equation (3-22)
gives the following relationship between the energy and coenergy:
n
W.(q,x) + W:O"x) = L q,'''i
,-1
(3-45)
I • ••

f/ecttomechanical.ay.tem equation. 123

The total differential of the electric-field energy storage is, therefore,

dW.(q,x} =d[ L" q.>.. - W;(X,x}] (3-46)


i-I

which can be expanded to give

Be sure that you understand this expansion of the total differential of Eq.
(3-46) to give Eq. (3-47). This is not trivial, and it is a very important
step.
By using Eq. (3-21), we can show that
aW;(X,x}
~~--- = q. (3-48)
a>.;
Now we substitute Eq. (3-48) into the third term on the right side of Eq.
(3-47) and ·cancel the first and third terms. Substituting the remaining
expression for dW. into the first-law formulation of Eq. (3-38) gives us,
finally,

I [pi + ~w;~>.,x}]
i ... l ...
dx. = 0 (3-49)

after canceling like sup1mations on both sides of the equation and slightly
manipulating the result. As before, in order to satisfy Eq. (3-49) for all
dx. virtual displacements, the bracketed term must itself be equal to zero.
Therefore, since (f.). = Pi, the force of electrical origin on the ith mechani­
cal member is given by
. (I') = _ aW;(X,x) ~ = 1,2, . . . ,m (3-50)
j J.... ax,
Notice again the negative sign when the coenergy-state function is used.
Remember that the force of electrical origin is positive if its sense oppo8e8
the assumed positive direction for Xi.
Equation (3-44) gives us the forces of electrical origin calculated from
the energy-state function for an electric-field device having n electrical
ports and m mechanical ports. Similarly, Eq. (3-50) gives us these same
forces calculated from the coenergy-state function. The results from
these two equations are identical except for one small point. Equation
124 Principle. of e/ectromechanical·energy conversion

(3-44) gives all forces of electrical origin as functions of all the q's and x's.
This would be suitable for a loop formulation on the electrical part of the
system where the q's or q's are the coordinates of interest. In Sec. 2-6 we
showed that the electric-field energy W. is always used when loop equa.­
tions are desired. Similarly, Eq. (3-50) gives the forces of electrical origin
as functions of the X's and x's, which is proper for a nodal analysis on the
electrical portion of the system. In Sec. 2-7 we demonstrated that the
el~ric-field coenergy-state function is used when a nodal analysis is
desired.

3·8 FORCES OF ELECTRICAL ORIGIN FOR A MULTIPORT.


I MAGNETIC-FIELD SYSTEM

The pieces of this puzzle should be starting to fall into place. Notice
that no new state functions have to be introduced when electromechanical
interactions are present in the system. The only thing new is that the
electrical state functions are dependent on both electrical and mechanical
coordinates. Then partial derivatives of these electrical state functions,
taken with respect to mechanical coordinates, yield the mechanical-inter­
action forces. All of the derivations in the preceding section have been
for a multiport electric-field system. Since a complete duality exists
between the electric- and magnetic-field state functions, we can predict
the magnetic-field results directly. The forces of electrical origin on the
ith mechanical member of a magnetic-field electromechanical system are
given by

(3-51)

in terms of the magnetic-field energy-state functions expressed in terms


of all the X's and x's. Using the magnetic-field coenergy-state function
we have

(3-52)

where the 4's and x's are involved. As a review of the principle of virtual
work, carefully develop the magnetic-field interaction forces, given by
Eqs. (3-51) and (3-52), for a general n-electrical-port m-mechanical-port
system starting from the first law of thermodynamics. Also give some
personal thought to the virtual displacement concept as you perform this
derivation.
• :l!1

E'.ctromechonico'·,yst.m equation. 125

3-9 INTERACTION FORCES FROM LAGRANGE'S EQUATION


We have seen in Chap., 2 how Lagrange's equation could be used to yield
equilibrium equations for electrical or mechanical systems systematically.
All the systems were composed of lumped elements with definable terminal
coordinates. The state of these systems could be expressed in terms of
energy- and coenergy-state functions involving the terminal coordinates.
The complete form of Lagrange's equation, including the Rayleigh dis­
sipation function and nonconservative external forces, is given by Eq.
(2-88). For con venience this equation is repeated as
!}.. [a.ca,~,t)] _ a.ca,~,t) + a(fW = Qk
dt a~k a~k a~k
k = 1, 2, . . . ,n (3-53)
The symbol ~ represents a generalized set of coordinates, both mechanical
and electrical. The Lagrangian .c(t~,t) can be a function of all the coor­
dinates, all their first time derivatives, and in some cases even the time t
explicitly. The Rayleigh dissipation function (fW is a function of all the
first time derivatives of the coordinates only. QIr. represents the sum of
the external forces acting on the kth node, or loop, of the system.
In Sec. 2-8 we carefully formulated the Lagrangian function in such
a manner that system equilibrium equations would 'result when this func­
tion was substituted into Eq. (3-53). Let us briefly review the formula­
tion of the Lagrangian function. The Lagrangian is defined as the
difference between two general state functions known as the total system
coenergy and the total system energy functions. The total system coen­
-ergy function consists of the sum of the mechanical kinetic coenergies
and certain electric-c,oenergy-state functions, as follows:

" (.x,x'X,(j't) = T'('x,x,


"
t) + W:(X,x)
W:(q,x) loop
nodal
(3-54)

When a loop formulation for the electrical part of the system is desired,
the upper line of symbols is used; and when a nodal formulation is
required, the lower line of symbols is used. The total system coenergy­
state function is designated by 3'; the mechanical-coenergy function is T';
and the magnetic- and electric-coenergy-state functions are given by W:
and W:.respectively. Remember that the mechanical position coordinate
x stands for both the translational- and the rotational-mechanical position
coordinates.
Equation (3-54) is identical with the original formulation of the total
system coenergy function for separate electrical or mechanical systems,
as given by Eq. (2-67). except for one small point. In Eq. (3-54) we are
including the possibility that the electrical-coenergy functions and W:
126 p,.inciple. of electromechanical·energy conve,.,;on

W: could involve both electrical coordinates and mechanical coordinates.


In other words, we have included in the formulation of the total coenergy
function the electromechanical lumped elements discussed ill the preceding
sections of this chapter.
The second total system state function involved in the formulation
of the Lagrangian is the total system energy function. The total system
energy function contains the mechanical potential-energy functions plus
the proper electric-energy-state functions as follows:

V(x t) + W.(q,x) loop


(3-55)
, W .. (X,x) nodal
where the upper set of symbols is used for an electrical loop formulation
and tile lower set is used for a nodal arrangement. The symbol 'U repre­
sents the total system energy function; V is the mechanical potential
energy; and W. and W ... are the electric- and magnetic-energy-state func­
tions, respectively. Equation (3-55) is similar to Eq. (2-68), which
applies to systems composed only of lumped mechanical and electrical
elements. The important extension lieR in the fact that Eq. (3-.')5)
includes lumped electromechanical elements, since the electric-energy­
state functions are shown to be dependent on both the electrical coordi­
nates and the mechanical position coordinates.
Let us pause for a moment and review what we are trying to do here.
The objective of this section is to investigate how Lagrange's equation
can be used for systems containing lumped electrical, mechanical, and
even electromechanical elements. Equations (3-54) and (3-55) have
reviewed the original formulation of the total system coenergy and energy
functions which we know are proper for systems containing only electrical
and mechanical elements. As a reasonable guess the electromechanical
elements have been included in the electric-energy- and -coenergy-state
functions in the original scheme. Now we must check this more complete
formulation and see if proper equilibrium equations are stilI generated by
Lagrange's equation given by (3-53).
By definition, the Lagrangian function is formulated as

.c (x,x,t:t) = :I' (x,x,t;t) - 'U ( x,~:t) ~::al (3-56)

where :I' and 'U are defined by Eqs. (3-54) and (3-5.1), Now suppose we
have an electromechanical system with electric-energy storage only in
magnetic fields. In addition, we have decided to formulate our electrical­
equilibrium equations on a nodal basis. For this particular case the
Lagrangian would be given by
.c(x,x,X,t) = [T'(x,x,t)] - [V(x,t) + W..(X,x)] (3-57)
f/ec'romechonic:o/·,y.'em equalion. 177

Notice that none of the electric-field terllls are present and that the mag­
netic energy, as a function of>.. and x, is used, since a nodal formulation is
desired. Substituting the Lagrangian of (3-57) into F.q. (3-Sa) gives us,
for the kth mechanical node,
~ [aT'<x,x,t)] _ ~T'(x,x,t) + ~V(x,t) + ~W...(>..,x)
dt aXil: aXil: ax" aXil:
+ as:(~,).) = Q/t (3-58)
aXil:
The first three terms in Eq. (3-58) are the inertial forces and the
spring forces acting on the kth mechanical node. The as:/ax" adds the
dissipation forces, and QII: includes all non conservative external forces on
the kth mechanical node. The remaining term is the aW .. (>",x)/ax•.
Equation (3-51) tells us that this term represents the force of electrical
origin on the kth mechanical node. In a very natural way, it arises in the
mechanical-equilibrium equation in the proper place with the proper sign.
For the same magnetic-field electromechanical system, suppose we
desired a loop formulation of the ('lcctrical-equilibriul1l equations. The
Lagrangian would therefore he given hy
.£(t,x,q,t) = [T'(x,x,t) + W~(q,x)J - (V(x,t)] (3-59)
here using the magnetic-coenergy-state function W~. t The equilibrium
equation for the kth mechanical node, from Lagrange's equation (3-53),
would now hecome
!i [aT'(X,x,t)] _ ~T'(x,x,t) _ aW~(q,x) + ~V(x,t)
dt aXil: aXil: a:tk ax.
+ as:(~,q) = Q. (3-60)
ax.
This equilibrium equation is identical with Eq. (3-58) except for the force
of electrical origin. However, Eq. (3-52) has shown us that -aw:/ax/t
is the correct force of electrical origin when the coenergy-, rather than the
energy-, state function is used.
A similar exercise for a system containing an electric-field electro­
mechanical Jumped element would again show that the forces of electrical
origin are correctly genel'ated and inserted into the equilibrium equations
by the Lagrangian method. The beauty of this whole scheme is that we
really did not have to alter Lagrange's equation in any significant way to
include the interaction forces. The electrical state functions became
slightly more involved for electromechanical elements, but the Lagrangian
formulation did not really change. The fact that the equilibrium equa- .
t Carefully examine Eq. (3-59) and be sure you understand each and every deta.il of
the terms and their functional dependance.
128 Principle. of electromechanical-energy convel"'ion

tiona for electromechanical systems are generated in such a neat, systematic


fashion gives Lagrange's equation its greatest appeal.

Example 3-4
For the electrostatic voltmeter shown in Figs. 3-9 and 3-10 determine the system
equilibrium equations by \l8ing a nodal basis for the electrical portion of the system.
Include a viscous damping torque at the rotor pivot points. Also include the possibility
of some ohmic loss due to a nonzero conductivity in the medium between the stator and
rotor members. Use Lagrange's equation to aid in this formulation.
Solution: The electrostatic voltmeter is shown schematically in Fig. 3-16. A
source, whose voltage is to be determined by the device, is connected to the electrical
tenninals. The ohmic loss due to the nonzero conductivity of the dielectric is repre­
sented by the lumped linear admittance G, with units of mhos. The total viscous
damping baused by the pivots has a net value of D newton-m-sec/rad. The total
inertia of the rotor and pointer about the pivoting axis is J kg-mt, and the compliance
of the torsional restraining spring is K rad/newton-m. The spring exerts zero torque
when B equals zero.
In order to \l8e Lagrange's equation, we m\l8t first formulate the system's La­
grangian function. The Lagrangian is composed of the difference between the total
system coenergy-state function and the total system energy-state function, as indi­
cated by Eq. (3-56). Equation (3-54) tells us how to formulate the total coenergy
function. Since we are using a nodal formulation on the electrical portion of the sys­
tem, the total system coenergy function becomes

The mechanical kinetic coenergy for the linear inertial element is simply
T'(~) - !J~.

In Example 3-3, Eq. (3-17) gives the electric-field coenergy for the system 88

W;O\,9) - a(". + 9)}.·


where
7. o(R t l - R,l)
a - 2h

Refer to Fig. 3-9 for an explanation of the physical dimensions of the device.

Fig. 3-16 A schematic diagram of the


Datum electrosta.tic voltmeter.
E/ectromechanica/-,y"em equation. 129

The total system energy function for a nodal formulation is

'!J(II,>-) - V(II) + W ..(>-,II)

as seen from Eq. (3-55). However, since no magnetic-field storage is present, the
system energy function reduces simply to

III
'!J(8) - V(II) - 2K

which contains only the one linear spring element. Thus the Lagrangian for this
system is giv~n by
, 1 III

.e(II,e,}..) - :2 Jill + a(a + e)}..1 - 2K

upon subtraction of '!J(/I) from :J'(II,II,}..).


The Rayleigh dissipation function for the system contains the two viscous loss
elements and is given by

The only remaining terms required for the Lagrangian formulation are the externally
applied forces. In the mechanical portion of the system, these forces take the form of
applied torques. Since there are no externally applied torques,

Q, - 0

where the symhol~. _"ed for the externally applied torques influencing the (I
coordinate.
The externally applied forces on the electrical part of the system take the form of
current sources when a nodal formulation is used. However, in this system we have
only an externally applied voltage source. If the voltage source had some series
impedance, we might tran.sform to an equivalent current source and shunt impedance.
With no series element present, we have an even simpler situation. The node voltage
" in Fig. 3-16 is not really an unknown generalized coordinate, since the external
voltage source v specifies this coordinate for every instant in time. Thus, in reality
we do not even have to write a nodal electrical-equilibrium equation, and Q"" the
external electrical force, is not required. If we had decided to use a loop formulation,
then the one mechanical equation plus two electrical equations would have been
required to describe the system. The nodal formulation is far simpler, requiring only
the one mechanical equation.
Lagrange's equation for the mechanical coordinate is simply

By using the Lagrangian and the Rayleigh function, the required terms are evaluated
as follows:

a.c(II,II,}..) _ JII !!.[a.c(II,(I,,,)] -JB


and
ai tit . all
130 Prineip/e. of e/ectromechonical-energy convenion

Similarly, we have
it£(11,/I,1.) _ tV.' _ !
0/1 K
and for the dissipation term

Now substituting these results into Lagrange's equation, we have as the desired
equilibrium equation

J6 + K/I + D6 - tV.' - 0 (3-61 )

after 80me slight rearrangement of terms.


Thelfinlt three terms on the left-hand side of Eq. (3-61) are respectively the
inertial torque, the spring restraining torque, and the viscous friction torque. The
term Olll is the torque 01 electrical origin causing the system to have some deflection.
Since the voltage 80urce v equals lI, the driving torque is actually a specified function of
time, and thus we have one equation and one unknown coordinate 8. At this point
our interest lies strictly in formulating equilihrium equations. In later chapters
we shall work on techniques for obtaining their solutions, and thus the actual system
response. However, one point concerning the operation of this electrostatic volt­
meter can still be made. If a constant or doC voltage were applied to the electrical
terminals, then the torque of electrical origin would al80 be a constant proportional
to the square of the applied voltage. Under steady-state conditions the deflection of
the pointer would be of such magnitude that the spring resisting torque would exactly
cancel or oppose this doC driving torque of electrical origin. Thus the device could
serve as a doC voltmeter if the scale were properly calibrated.

Example 3·5
Determine the equilibrium equations for the electromechanical actuator discussed in
Example 3-1 and shown in Fig. 3·4. A schematic representation of the device is shown
in Fig. 3-17. The resistance of the N-turn coil is shown as a lumped-linear resistor R•.
To the terminals of this winding a voltage source v(t), with internal resistance R .. is
connected as an electrical driving function. The mechanical plunger is spring-loaded,
80 that when the position coordinate x equals a distance b, the spring exerts no force on
the plunger. The equilibrium equations are to be determined by using a loop basis for
the electrical portion of the system.
Solution: Since a loop formulation for the electrical part of the system is to be used,
the total system coenergy function, according to Eq. (3-54), becomes

:l'(:I:,x,q) - iM:/;1 + W~(q,x)


The magnetic-coenergY-l!tate function W~(q,x) for this device has been determined in
Example 3-1 and is given by Eq. (3-5) as being
aql

w:,(q,x) - d +x

where a - pJ{IA/2. Refer to Fig. 3-4 for a definition of these parameters.


Eledromedtanical.system equations 131

Mg 9rovifofionol
force
Fig. 3-17 A schematic
diagram of the mag­
netic-field electromechani­
cal actuator.

The system energy function, referring to Eq. (3-55), is simply

with the electric-energy-state function equal to zero because there is no electric­


energy storage in the system. Notice that, when x - b, the energy stored in the
spring equals zero. Thus the Lagrangian for the system becomes
"('
"" X,x,,,
") _ !M" +~
2 x d +x
_ (x 2K
- b)'

upon gathering terms from the expressions for 3' &nd 'U.
For the Rayleigh dissipation function we have

IJ(i,q) - 1-Di' + i(R. + R.W


The external forces complete the bookkeeping for the Lagrange formulation. These
are given by
Q. - Mg
88 the gravitational force acting on the x coordinate, and by
Q. - v(1)
as the voltage source acting on the q coordinate.
The mech&nical-equilibrium equation is determined from Lagr&nge's equation
by using
~ [as!(i,z,q)] _ as!(:t,x,q) + aIJ(t,q) _ Q
rlt at ax ax •
132 Principle. of electromechankal·_rgy conversion

These partial derivativea are evaluated as follows:


~ [''J,c(:~X,q)] _ ~ (Mx) _ Mi:

iJ,cC i:,x,q) aq" X - b

ax - - Cd + X)I - -r

iJ;J(x,q) _ Dx

iJx

The mechanical-equilibrium equation i8 therefore given by

M:t + Dx + x ~ b + (d ~"X)" - Mg (3-62)

Notice again how naturally the force of electrical origin has entered into the equilibrium
equation.
The eleetrical-equilibrium equation is obtained in a similar fashion from
!! [~,c(:t'X,q)] _ iJ'c(i:,x,q) + iJil'(x,q) .. Q

dt iJq iJq iJq •

For these derivatives, we have


.!! [il.c(X,X q)] _ t d ( 2aq ) _ 2a.. 2aqx t

dt iJq - di d + x - d +
x q - (d x)!
+
iU!(:j;,x,q) _ 0

ilq

~Jilt,ll) - (R. + R,)q


Thus the electrical-equilibrium equation is

d~x Ii + (R. + R,)q (3-63)

after some 8light rearrangement of terms.


Let us take a moment to examine each of the terms in Eq. (3-63). The first term
represents the inductive voltage drop in the solenoid winding due to the loop current
changing with time. The second term on the left-hand side of (3-63) is the voltage
across the two resistors in the loop. Now, the third term is something new. This
term is formally known as a voltage of mechanical origin. The voltage is induced in the
winding owing to the motion or velocity x of the iron plunger. Notice that the
voltage term depends on the current q and in general is not zero even if the coil current
should remain constant. The sum of these three voltage terms is equal to the applied
voltage v, thus completing the equilibrium equation. Again notice how easily and
naturally this equation, including the electromechanical interaction effects, has been
obtained.
Equations (3-62) and (3-63) are the two required system equilibrium equations
whose solution would yield x and q, or if desired, fIt as functions of time. The solution
of these equations, in general, is, however, quite difficult, since they represent a
simultaneous set of nonlinear differential equations. Our efforts in the following
chapters will be devoted to techniques which scrape a' great deal of information con­
cerning system operation from such a set of nonlinear equations.

t Carefully verify that this result is correct. Remember that dldt is a total. not a
partial, derivative.
Electromechanical-system equation, 133

3-10 INTERACTION FORCES FROM TWO FORMS


OF THE STATE FUNCTIONS
In Example 3-2, a set of relationships specifying the flux linkages in a
magnetic-field electromechanical device as a function of the current and
mechanical position coordinates are given. In general, for a magnetic­
field electromechanical device, defining relationships can usually be written
in the form
Xl = Xl(ch,Q2, ,Qn,XI,X2, . . . ,x...)
X2 = X2(QhQ~, ,ti",XI,X2, . • . ,x..)
(3-64)

assuming the device to have n electrical ports and m mechanical ports.


The device of Example 3-2 has only three electrical ports and two mechani­
cal ports, but it includes the general functional dependence, indicated by
Eqs. (3-64), whereby each of the flux linkages can depend on all of the
current and position coordinates.
By using expressions in the form of Eqs. (3-64), the magnetic-field
coenergy-state function can readily be found from its defining relationship
given by

(3-65)

Notice in this formulation of the magnetic coenergy that it is a function


of the n electrical currents and the m mechanical position coordinates. In
Sec. 3-8 the force of electrical origin was determined from the magnetic­
coenergy-state function, and, according to Eq. (3-52), it is given by
(f.), = - ~W~(Q,x) (3-66)
ox,
for the ith mechanical node in the system.
Suppose we wish to evaluate the force of electrical origin from the
energy-state function rather than the coenergy-state function. Equation
(3-51) shows how this can be accomplished provided the energy-state
function is dependent on the flux linkage and the position coordinates.
The defining expression for the magnetic-energy-state function is given by

W ..(X,x) (3-67)

Thus, quite normally, the energy-state function turns out to be dependent


on the Xand x coordinates. However, Eq. (3-67) tacitly assumes that the
134 Principle, 01 electromechanical-energy conversion

relationships given by (3-64) can be solved for the currents as functions


of the flux linkage and the position coordinates. In other words, in order
to use Eq. (3-67), we must obtain relationships in the form
Ih = Ih(AJ,A 2, • • • ,A~,Xl,X2' ••. ,x...)

q2 q2(AI,A2, . . . ,A~,XhX2, ••• ,x...)

(3-68)

Considering Example 3-2, this looks like a very difficult job, but the
magnetic-field energy can still be evaluated by knowing the coenergy
because the two state functions are related by
"
If{... + W~ = L
10-1
A"q" (3-69)

Therefore, in general, the magnetic-field energy-state function is given by

W ... (q,x) L" qltAk(q,x)


1;-1
W~(q,x) (3-70)

Notice, however, that by use of the relationship of Eq. (3-70) the magnetic
energy will turn out to be a function of the current and position coordinates
and does not have its usual A, x dependency predicted by (3-67).
Can we still find the force of electrical origin from W",(4,x)1 The
answer is yes, but we must proceed with some caution. Suppose we take
the a/ax, of both sides of Eq. (3-70) while holding all currents constant
during the partial differentiation. Thus we have

(3-71)

upon interchanging the differentiation and summation operations. The


last term in Eq. (3-71) is recognized as the desired force of electrical origin,
referring to (3-66). Therefore we have

(3-72)

as the interaction force determined from the energy-state function


expressed in terms of the current and position coordinates.
In Sec. 3-8 the force of electrical origin "from the magnetic-energy­
state function is given by Eq. (3-51) as being

(J.)I = aW",(A ,x) (3-73)


ax.
Eledro",ee#tanical-S:Y$te", fICJIIOtioM 135

with W... as a function of A and x. Equation (3-72) shows us that the


electrical force when evaluated from W...(q,x) involves the oW.../ox; with
a second correction term included. Omitting the second term in (3-72)
will give an erroneous value for the interaction force.

Example 3·6
To illustrate these ideas, consider a. m8ognetic-field electromech8onic8ol system with the
following cha.r8octeristic rel8otionshipll:

(3-74)

The device has two electric80l ports 80nd two mech80nical parts. Determine the force
of electrica.l origin on the first mechanical node from the coenergy-state function Bnd
then from the energy-st8ote function.
Solution: Expanding Eq. (3-65), the coenergy-state function can be found by
ev80luating

Upon substituting the defining flux-linkage relations given by (3-74), we obtain for the
magnetic coenergy

The force of electrica.l origin on the first mechanical node, according to Eq. (3-66), is
given by

(/.). - (3-75)

using the expression for the magnetic coenergy as a function of q and z.


Now let us see how this same force of electrical origin can be determined from the
m8ognetic-energy-state function. If we attempt to evaluate this energy-state func-'
tion by using the defining relationship of Eq. (3-67), then we must be 80ble to expre88
q. and q. as functions of X., X., ZI, and z.. Theoretically, we can do this by manipul8ot­
ing Eqs. (3-74) and solving for the two currents in terms of the flux-linkage and posi­
tion coordinates. A casual glance at (3-74) tells us that such a manipulation is quite
a difficult job. However, by using Eq. (3-69), the energy-state function can be found
from

Notice that when we use this relationship between the two state functions, the energy
is a function of the current and position coordinates rather than the more usual flux­
linkage and position coordinates.
136 Principles of electromechanical· energy conversion

Upon substitution of Eqs. (3-74) and the magnetic eoenergy, tbe energy-state
fundion becomes

We must now use F.£!. (;J·72) to evaluate tbe forre of electrical origin. Simply taking
the aW../ax, will not give the correct (I.), because of the unusual functional repre­
sentation of W... On expanding Eq. (3-72) we have

whieh is evaluated as follows:

Hubstituting th{,lIe quantities into the expression for (f,)" after som{' lIimplification of
terms, gives the for('e of electrical origin on the first mechanical node as
. 12x,'q.q. 7x,'q,'
(j,), "" -x,X,I[" - -~ - -'--4­

which is identi('al with Eq. (:1-75) calculated from the magnetic-coenergy-state


function. To be sure that you understand these manipulations clearly, ('·aicuiate (f.),
from the ('oenergy- and energy-state functions.

Suppose that we know the current coordinates as functions of the


flux-linkage and position coordinates for a magnetic-field electromechani­
cal system. This means that relationships in the form of Eqs. (3-68) arc
available. Therefore, we could readily evaluate the energy-state func­
tion in terms of the X and x coordinates by using Eq. (3-67). The force
of electrical origin on the ith mechanical node is then given by
aw",(X,x)
(3-76)
ax;

The coenergy-state function, in general, could reasonably be obtained


only from the relationship
,.
l: XklMX,x)
k-l
- W",(X,x) (3-77)

where W:is a function of X and x. Taking the partial derivative of


Eq. (3-77) with respect to Xi, we have

(3-78)
Electromechanical-system eqwtions f37

using Eq. (3-76). Again we see that expressing the state function in
terms of the alternative electrical coordinate requires a second term in the
expression for the interaction force.
In the case of an electric-field electromechanical device similar argu­
ments apply. Assume that we were able to obtain defining relationships
expressing the charges as functions of the voltage and position coordinates.
This means that a set of expressions in the form
ql = ql(X h X2, ,Xn,XIoX2, . . • ,x..)
q2 = Q2(}.I,X 2, • • • ,X",XI,X2, . . . IX..)
(3-79)

are available. By using these functional relationships, the electric-coen­


ergy-state function can be calculated from its defining relationship given
by

(3-80)

Notice that the electric eoenergy is a function of the voltage and position
coordinates. From this coenergy expression, the force of electrical origin
on the ith mechanical node is given by

(3-81)

Suppose that we wish to calculate the interaction force from the


electric-field-energy function rather than froln the coenergy-state func­
tion. In general, the relations given by (3-79) could be sufficiently
involved that it would be quite difficult to solve for the voltages as
functions of the charge and position coordinates. Therefore, the defin­
ing relation for the electric-field energy-state function, as given by
It

W.(q,x} = l
i-I
10
9
' X:(q~,q~, ... ,q~,XIpX2' .•• ,X.. ) dq: (3-82)

cannot be used, since the x..(q,x) functions cannot be obtained.


However, by using the relationship that exists between the energy­
and coenergy-state functions, the electric-field energy-state function can
be obtained as follows:
n
W.(}',x) = L X,tlJk(X,X) -
4'-1
W:(X,x} (3-83)

When Eq. (3-83) is used, the energy-state function depends on the}. and x
coordinates and not on the q a.nd x coordinates as in its defining relation­
138 Prim;iples of electromechanical-energy conver$ion

ship (3-82). Taking the partial derivative of both sides of Eq. (3-83)
with respect to x, gives us

(f.). (3-84)

as the force of electrical origin on the ith mechanical node as determined


from the energy-state function expressed in terms of the voltage and posi­
tion coordinates. As in the magnetic-field case, notice that expressing
the state function in terms of the opposite electrical coordinate [here we
have W.{>.,x) rather than W.(q,x)] requires a correction term in the expres­
sion for the force of electrical origin. If W. is a function of q and x, then
this ~ame force is given by

(I.), = aW.(q,x) (3-85)


ax.

Instead of starting with relations in the form of Eqs. (3-79), we could


have postulated relationships expressing the voltage coordinates as func­
tions of the charge and position coordinates. Therefore, using Eq. (3-82),
the electric-field energy-state function could be obtained in the form
W.(q,x). Equation (3-85) could then be used to find (I.)., the force of
electrical origin on the ith mechanical node. Now, however, the coenergy
could be obtained only as a function of the q a.nd x coordinates by using
the relationship
.
W~(q,x) = L qk";,k(q,X) -
.\;-1
W.(q,x) (3-86)

TABLE 3-1 Mechanical forces of electrical origin from


energy- and coenergy-state functions
Ana-Iyei, Energy...tate funotiona Coenergy...tate function.

Mrt!i!:Ca~~a.,:: (/.); _ o.w ~(A,.,)


az. (/.)0 -
_ .,W~'(lI.x)
iJz:,
+
n

L .-­
), ali.(A,X}
az,

L.
.1:-1
iJWM'(ti,..,}
Loop analyaia (/.), - ~W ~(q,Xl - 4. oA.(li,.,) (f.); -
aZt lb • ~
.1:-1

Eleetric storage: _ oW.'(q,.,) +


.. ol...Cq,,,)
Loop onalyli.
(j.), _ oW.(q.z)
ax, '{J.}, -
ax!. L fl'~

. .1:-1

Nodal analy.i.
(J.l< _ ow.cA,.,) _
oz. 2:
.1:-1
.-­
A aq.{A•..,)
OZi
(J.); -
aw.'{~,.,)
- --,;;­
ElflCtromechanical-sysfem equations 139

When the a/ax,. of (3-86) is taken, the force of electrical origin is given as

(f.), = aW.(q,x) = _ aW:(q,X) + ~ q" aX.(q,x) (3-87)


ax-.. ax-.. k
4-1 ax­..

Compare Eqs. (3-81) and (3-87), and notice the correction or additional
term that needs to be included if we have W~(q,x) rather than W:(X,x)
for the coenergy-state function.
Table 3-1 summarizes all these relationships for the force of electrical
origin and shows how they can be obtained from 'the various functional
dependencies of the state functions.

3-11 LAGRANGE'S EQUATION WITH THE TWO


STATE-FUNCTION FORMS
If Lagrange's equation is used to find the system equilibrium equations,
then the forces of electrical origin will be determined correctly in a very
routine fashion, provided the state functions are grouped according to
Eqs. (3-54) and (3-5,1'». For convenience, let us repeat these two equa­
tions and review their basic formulation. The total system coenergy­
state function is defined as

3' (x, 'X, t)


x q, = T'(:i: x
' ,
t) + W:(X,x)
W:(q,x) loop
nodal
(3-88)

where the top line of symbols is used for a loop analysis on the electrical
part of the system and the bottom line is used for an electrical nodal
analysis. The total.system energy-state function is similarly given by

'U (x, q,
h,
t) = V(x,t) + W.(q,x)
W .. (h,x)
loop
nodal
(3-89)

where the same scheme is again used for the upper and lower sets of
symbols. The Lagrangian state function is defined as (3' - 'U) according
to Eq. (3-56). Then by using Lagrange's equation (3-53), the system
equilibrium equations can be found.
Notice that Eq. (3-88) defines the system coenergy function to be a
sum of only coenergy-state Junctions. The coenergy function 3' excludes
energy-state Junctions in its definition. Similarly, the system energy-state
function 'U contains only energy-8tate Junctions, excluding the coenergy
Junctions. Thus, if a particular system has magnetic-field-energy storage
and a loop formulation for the electrical part of the system is desired, then
the magnetic coenergy W~ goes into the system coenergy function 3'. On
1M) Principle. of electromechanical-energy conversion

the other hand, if a nodal analysis were desired, then the magnetic-energy
function W", should be inserted into the system energy function 'U.
Also observe that Eqs. (3-88) and (3-89) carefully detail the proper
functional dependence of each state function. The W~ that is substituted
into the J' function must be expressed in terms of the <j's and the x's.
Similarly, the magnetic energy W ... that goes into 'U is a function of the >.'8
and x's. Suppose that in a given problem we require a loop analysis for
the electrical portion of a system containing magnetic-field elements.
And further suppose that we have available only the magnetic-energy
fUllction in terms of the <j's and the x's. Since a loop analysis is desired,
having the q's in the state funetion seems to be a step in the right direction.
Equations (3-88) and (3-89), however, do not have a space for W",(<j,x).
Norm..Uy, for a loop formulation we would use the magnetic-coenergy
function W~(q,x), and indeed that is what we must substitute. By using
the relation
n

L qk>'k -
.\;-1
W",(q,x) (3-90)

we are readily able t.o find the reqllired magnetic-coenergy function.


Thus, in general, the state function in terms of t.he variables required by
Eqs. (3-88) and (3-89) can always be found by using the relation that
exists between the coenergy and energy functions.
By substitution of Eq. (3-90) into the system coenergy function the
corred mechanical force of eleetrical origin is still obtained. On the
ith meehanical node, the force of electrical origin is given at least in part
by
oj'
(3-91)
ax.
as can be seen by 'expanding Lagrange's equation for the ith mechanical
node. Since::;t contains W:' as given by Eq. (3-90), we have

oW~(q,x)
(3-92)
ox.

This checks with the corresponding result in Table 3-I.


All these manipulations of the electrical energy- and coenergy-state
functions st.ill do not allow us to choose freely between a loop and nodal
formulation. For example, if equations describing magnetic-field ele­
ments express the flux linkages in terms of the current and position coordi­
nates in the form of Eqs. (3-64), we are forced into a loop formulation
unless these equations can be inverted into the form of Eqs. (3-68)
whereby the currents are functions of the flux-linkage and mechanical
coordinates. Even though the magnetic-field energy and coenergy func­
Electromechanical-system equations 141

tions can be found, these are functions of the q's and x's, and the magnetic
coenergy or its equivalent in terms of the magnetic-energy-state function
must be used in Lagrange's equation, thus resulting in a loop formulation.
If the defining relationships for the electromechanical elements cannot
be inverted, then the form of these relationships dictates our choice of
a loop or nodal formulation.

3-12 SUMMARY
The energy- and (~oenergy-8tate functions introduced in Chap. 1 are gen­
eralized to include electromechanical lumped elements. With the elec­
trical coordinates all equal to zero the mechanical coordinates are set at
their state values. While the mechanical coordinates are held fixed at the
state values, the electrical coordinates are then taken from zero to their
final state values and the energy- and coenergy-state functions are cal­
culated. Since the state functions depend only on the final state point,
and not on the manner or path by which the system reached the final
point, the energy and cocncrgy funetions calculated in t.he described man­
ner are valid. The state functions for electromechanical elements are
similar to the simple lumped elements discussed in Chaps. 1 and 2 with a
dependence on the mechanical coordinates in addition to an electrical
coordinate dependence.
By using the principle of virtual work, an experiment is analytically
conducted to determine the mechanical forces of electrical origin. The
experiment requires that an external force be applied in such a fashion as
to be equal and opposite to the force of electrical origin. By applying
the first law of thermodynamics with the virtual work principle, the exter- .
nal force is obtained III terms of partial derivatives of the state functions
describing the electromechanical elements.
A study of Lagrange's equation shows that the mechanical forces of
electrical origin are correctly inserted into the mechanical-equilibrium
equations simply by using the Lagrange formulation in the systematic
manner described in Chap. 2. For electromechanical systems the elec­
trical state functions used in the Lagrangian are dependent on both
mechanical and electrical coordinates. However, the same state func­
tions are required for the Lagrangian, and the formulation introduced in
Chap. 2 for separate electrical or mechanical systems is applicable for the
electromechanical systems. For the electrical-:-equilibriumJlQ.uatioIls,
Lagrange's equation also gives the electrical voltages or currents of
mechanical origin.
The state functions for the lumped electromechanical elements, in
general, can have two functional forms. The mechanical forces of elec­
142 Prim;;p/es of electromechanical-energy conversion

trical origin can be found from the electrical state functions expressed in
terms of either of two electrical coordinates. Table 3-1 summarizes these
forces for all of the possible functional dependencies of the energy- and
coenergy-state functions. Since the Lagrangian formulation requires
prescribed state functions with prescribed functional dependencies in
synthesizing the two functions 3' and 'lJ, the interaction forces given in
Table 3-1 are calculated in a routine fashion by using Lagrange's equation.

PROBLEMS

3,l In the derivation of the energy- and coenergy-l'ltate functions for lumped
elemen'ts with both electrical and mechanical ports, the mechanical coordinates are
analytically fixed at their state-point values. The state functions are then calculated
by raising the electrical coordinates to their final values. Discuss the energy storage
during each step of this process, and explain why we follow such a procedure.
3-2 A parallel-plate capacitor is constructed from two circular conducting plates
each having a radius of 5 cm. A lO-volt doC 8upply is connected across the two plates.
With a spacing between the plates of 0.5 em:
a. Calculate the magnitude of the energy- and coenergy-l'ltate functions.
b. Calculate the force, in newtons, tending to pull the plates together.
c. If the spacing is slowly reduced to 0.01 em, calculate the new force tending to
pull the plates together and the work done by an external agent in repositioning t.he
plates to this new spacing.
3-3 The magnetic-field transducer shown in Fig. P3-3 has an N-turn coil wound
on an iron core. The iron rotor has a cylindrical surface and is axially mounted in the

N turns

.,(l) Fig. P3-3


ElfKtromfKoon;cal-,y,tem equation. 143

air gap with 181 S 01/2. A torsional spring of compliance K, restrains the motion' of
the rotor. When 8 - .,/8, the spring is in its free position. The rotor has inertia J,
and its pivot points otTer a viscous restraining torque with a coefficient D,. The gap
length g is much less than the rotor radius r, so fringing fields can be neglected.
Assume that the permeability of the iron is much greater than that of the air.
a. Determine the coil flux linkages as a function of the' rotor angle /} and the coil
current i.
b. Find the inductance of the coil. Is the inductance linear?
c. Determine the magnetic coenergy W~(i,8) and tbe magnetic energy W ..(>.,8).
3-4 a. For the magnetic-field transducer discussed in Prob. 3-3 find the torque
of electrical origin exerted on the rotor if the coil current is held constant at i-i.,
Use the appropriate coenergy-state functions.
b. Repeat part a, but use tbe appropriate energy-state function bolding the flux
linkages constant at >. - >'0.
c. Compare the answers obtained in parts a and b.
d. Graphically illustrate the significance of the partial derivatives in parts a and b.
3-5 For the magnetic-field transducer discussed in Prob. 3-3:
a. Write a suitable Lagrangian for a loop formulation on the electrical portion of
the system.
b. Find a suitable Rayleigh dissipation function.
c. Determine the system's equilibrium equations by using Lagrange's equation.
3-6 ti-he electric-field transducer shown in Fig. P3-6 consists of a cylindrical
copper t~be of radius a with a solid copper plunger of radius b held axially within the
tube bYI an inSUlating bushing. The plunger moves only in the .:I: direction and is
restrain4jd by a spring with compliance K and a linear damper D. When x - 0, the
plunger~ halfway intO the tube and the spring is at its free length. If a is only slightly
larger th~n b. we can neglect all fringing fields.
a. Dl\termine tbe charge on tbe plunger as a function of the position of the
plunger and., tbe voltage between the plunger and the tube.
b. Find"the capacitance of the transducer. Is the capacitor linear?
c. Determine the electric coenergy W.(q,x) and tbe electric energy W:(~,.:I:).

_x
K
t­ "I
T f(tl
2a 211 AI
L

G
~
Insulating 0
bushing

Fig. P3-6 -
iltl
1.44 Principles of electromechanical-energy conversion

3-7 Q. For t.he electric-field transducer discussed in Prob. 3-6 find the force of
electrical origin exerted on the plunger if the capacitor voltage is held constant. Use
the appropriate coenergy-state function.
b. Repeat part a, hut use the proper energy-state function holding the charge
fixed.
c. Compare the answers obtained in parts Q and b.
d. Graphically illustrate the significance of the partial derivatives in parts a and b.

3-8 For the electric-field transducer discussed in Prob. 3-6:


Write a suitable Lagrangian for a nodal formulation on the electrical portion
Q.

o(the system.
- b. Find a suitable Rayleigh dissipation function.
c. Determine the system's equilihrium equations by using Lagrange's equation.
3-9 For the two parallel iron pole faces shown in Fig. P3-!l, determine the
attract\ve force hetween the poles in terms of the air-gap flux density n, in webers per
square meter, and the pole-face area A, in square meters. Assume that the iron has
infinite permeahility compared with air and neglect fringing around the air gap.
Hinl: Remember that the energy stored in a volume V is

W .. - Iv H~B·dV
where B - "oH in air and H - 0 in the iron.

Fig. P3-9

3-10 A eylindrical iron rod of radius Q is inserted into the cylindrical iron frame,
shown in Fig. P3-lO, through two cylindrical hrass hushings of thickness t. The flux
density established by the coil wound around the iron rod is B webers/m' as shown.

Bross
--B
bushing ..L:::.-._,.....j,.,.,.,,---''--------=,.,.,.,'''----'
00000000
0000000
0000000

Fig. P3-l0
Electromechanical-system equations 145

Assume the iron has infinite penneability compared with the brass. Take the
permeability of the brass to be equal to that of air, ",.. Neglect fringing on the basis
that t « a and assume the field in the brass to be radial.
a. Find the total magnetic-field energy- and coenergy-state functions for the
system in terms of the flux density B and the physical dimensions.
b. Find the force of electrical origin from the energy-state function.
3.1 I For the electromechanical system shown in Fig. P3-11 the capacitance
element C is described by the following relationship:
q - A)"(x. - x,)-'
where q is the charge on the plates and A is a constant. All other elements, both
mechanical and electrical, arc linear.
a. Find the Rayleigh dissipation function 5().,x.,x,) for the entire system.
b. Find the total ele"tric-cocnergy storage W;().,T"T,).
c. Find a suitable potential function 'U().,x"x,) for the entire system.
d. Formulate the system's Lagrangian.
e. Derive the equilibrium equations for the system. Expand all total derivatives_

Fig. P3-11

3.12 An clectrodyn,amometer movement used in both a-e and d-e indicating


instruments is shown schematically in Fig. pa-12. The movable coil is connected to a
pointer. Its rotation acts against a torsional spring of compliance K,. Assume a
viscous damping coefficient D, and I!. total inertia J to also influence the pointer rota­
tion. The damping is due to air vanes connected to the moving-coil axis. The

'Flexible
connection
Fig. P3-12
146 Principle6 of e/ecfromechonico/-energy (;On.,er6;on

mutual coupling M between the fixed and movable coils is a function of the rotational
coordinate 8.
a. Find the functional relationship for M such that the torque of electrical origin
is independent of 8.
b. Show how the device can be used as a true rms current instrument.
c. By using Lagrange's equation and the functional relationship of part a,
determine the equilibrium equations for the system.
S-IS The electric-field force transducer in the electromechanical system shown
in Fig. P3-13 can be described by the following characteristic equation relating the
charge to the terminal voltage and plunger position;
q - aA~~ + b},.{x - C)4

When x - c, the plunger is centered between the two plates and the spring is at its
free length.
a. Write a Buitable Lagrangian for the system.
b. Write the Rayleigh diBBipatiol) function and the external generalized forces.
c. Find the system equilibrium equations.
d. From the equations in part c write the force of electrical origin and the gen­
erated current of mechanical origin. Indicate the reference sense for these two
interaction terms.

ir------lI------Jt------1

t
i(tl

Fig. P3-13

3-14 For the electromechanical system shown in Fig. P3-14 equilibrium equa­
tions which formulate the circuit containing the electric-field device on a nodal buill
and the circuit containing the magnetic-field device' on a loop basis are required_
The characteristic equations for the coil and capacitor are lioii folloWJI:
Coil flux linkages - ). - Q2,(A + R1:21)
Capacitive charge - q - AI' D .;E:r:1
Electromechanica/·system equations 147

a. Write a suitable LagrllJlgian for the system. Take the spring force to be zero
when:l:, - :l:J.
o. Determine the Rayleigh dissipation function and the external BOUrces.
c. By u8ing Lagrange's equation, find the four equilibrium equations.

X,
A

Fig. P3-14

3-1S A certain electric· field transducer having two electrical ports and two
mechanical ports haa the following characteri8tic equation8:

a. Determine the energy 8tOred in the electric field W.(A"At,:I:,,:I:.). Notice the
functional dependence of W •.
o. Determine the force of electrical origin on the two mechanical nodes from the
energy-atate function found in part a.
c. Repeat part 0, but use a 8uitable coenergy-atate function.
d. Write a Lagrangian for the 8ystem using the energy-8tate function found in
part a.
e. Find the currents bf mechanical origin flowing oul of the two electrical nodel.
3-16 The characteristic equation for a magnetic-field electromechanical device
is given by
X - X.(l - e-·b )
where X. and a are constants.
a. Show that the magnetic-field stored energy i8

W ..(q,:I:) - ~ (1 - e-·'-) - X.qe-·"


o. Find the force of electrical origin from W ..(q,:I:).
c. Repeat part 0, but use a suitable coenergy-atate function.
d. Write a Lagrangian for the 8ystem.
chapter

THE ANALYSIS OF

4
LINEAR SYSTEMS

The preceding three chapters have established gen­


eral techniques for the formulation of the equilib­
rium equations for a wide class of electromechanical
systems. Energy-state functions for the lumped
elements of the system are first formulated. By
operating on these state functions according to La­
grange's equation, the system equilibrium equations
are obtained. In general these equations are non­
linear differential equations with both constant and
time-varying coefficients. The purpose of this
chapter is to develop general techniques for obtain­
ing information concerning the response of these.
systems to a broad class of forcing functions.·
Notice that we are not proposing to find general
solutions for these equilibrium equations. Such
solutions would certainly be desired, but in general
we are unable to obtain complete analytic solutions
for nonlinear differential equations. Therefore; we
shall be looking for techniques which yield, for many
cases, system-response informatio:n sufficient for our
needs. The idea is to learn enough about the
approximate operation of the system in specific
.instances to be able to predict the operation in many
other situations reasonably well:
Linear systems, or systems described by linear
constant-coefficient differential equilibrium equa­
tions, form the starting point for this study. Pre­
vious courses in circuit theory and mechanics have,
for the most part, . been concerned with lumped­
linear systems. We know how to obtain the com­
plete response for linear systems for a wide range of
practical input forcing functions. Certain so-called
classical techniques can beqsed for the solution of
the equilibrium equations. Also of great import­
148 ance are the transform,particularly the Laplace
The analysi, of linear systems l49

transform, methqtls for obtaining complete solutions giving the desired


system response.' We shall first review the general techniques available
for the study of linear constant-coefficient equilibrium equations. Then,
by usin~Jechnique known formally as Liapunov's first llleth,!<!, iE.eSe
lineaiideas can--be~extended to-ourhoiilriiearequationst~- yield a great
deaforiiiforiiiatloii-conc-errung their operatiou:----------------"
~e-vei.y'impoItantprailtical consideration should not be overlooked
at the outset of such a study. The parameters of any particular physical
system which might be of interest to us are known only to a certain degree
of accuracy. Therefore, even if a complete analytic solution to the '
equilibrium equations formulated in terms of these coefficients could be
found, we would still be in some doubt about how closely the solution
would correspond to the actual system response. Thus we should always
keep in mind that the approximate techniques that we shall use on system
equilibrium equations might certainly yield answers which are just as
helpful to us as the complete analytic solution to these equations if the
solution were available. The ultimate test, of course, is to see whether
the actual physical system does what we expect it to do. This discussion
points out the great importance of the laboratory for verification and, in
many cases, for subsequent alteration of our system to obtain a desired
mode of operation.
". ,

\
.4-1 CHARActERISTIC PROPERTIES OF THE LINEAR SYSTEMl

Let us begin by reviewing some of the important characteristics of a linear


system.. Figure. 4-1 shows a general block-diagram representation of a
li~fla:r syste:r1i with a single input or forcing function represented by /3(t)
and a singl~vesponse function x(t).. 'In general, for such a system we can
always write a differential equation in the following form:
dnx dn-1x dx
a..(t) dtn + a..-l(t) dt,,-l + ... + al(t) dt + ao(t)x
d~/3 m 1
= bm(t)df:.. + bm-l(t) ddtm-
- /3 d/3
1 + ... + b1(t) dt + bo(t}/3 (4-1)

The input function (3(t) is usually some known function of time, and thus
the complete right-hand side Of Eq. (4-J-) can be determined as some

1 For additional information see D. K. Cheng, "Analysis of Linear Systems," chap.


1, Addison-Wesley Publis4ing Company, Inc., Reading, Ma8Iil., 1959.

'. Input
PI t)
Response
x It)
Fig. 4-1 , A generallumped-ltnear system.
150 Principles of electromechanical-energy conversion

explicit time function which we could represent simply as Q(t). The


response x(t) of the system is unknown, and it is to be obtained by solving
the differential equation.
Notice that all of the an, an-I, . . . , ao coefficients are represented
as functions of the'independent variable time. Since these coefficients
represent physical parameters and dimensions in the system, they can,
quite reasonably, be assumed to remain constant with respect to time.
For the purposes of this discussion, we shall leave the coefficients in their
time-dependent form in demonstrating the important properties of a linear
system.
Equation (4-1) can be simplified to the form
dnx dn-Ix dx
an(t) dtn + an_let) dtn-l + ... + a,(t) dt + ao(t)x = Q(t) (4-2)

with Q(t) representing the known time function, which is the right-hand
side of Eq. (4-1). Suppose that a specific forcing function Q,(t) drives the
system represented by Eq. (4-2). The response to this forcing function
could, in general, be represented by some function X,(t); thus, we can write

+ a,(t) d~, + aO(t)x, = Q,(t)


(4-3)

Figure 4-2a represents this situation. I


Figure 4-2b represents a second case where the input is a different
function Q2(t) and the response is X2(t). Thus we have

1 In Fig. 4-1 the linear system modified the input function (jet) according to the

right-hand side of Eq. (4-1), thus giving a general Q(t), In Fig. 4-2 we are assuming
that this modification of the input does not take place, since Q(t) is actually being
inserted.

t:i,{t) X, (t)

(a) } O.ItI'O,1t1 x,{t) + x21!)

x2{t) (e)

(b)

Fig. 4-2 Superposition as a property of the linear system.


The analysis of linear systems 151

By adding Eqs. (4-3) and (4-4), we can simplify the resulting sum into the
form

an (
t) d"(xl
dt"
+ X2) + a,,-l(t) d,,-l(XI + X2)
dtn 1
+. + al(t) d(Xl: X2) ao(t)(Xl + X2) = Ql(t) + Q2(t)
(4-5)
which is represented by Fig. 4-2c. We can see how (4-5) results by taking
the first term as an example.

(4-6)

If the coefficient an were some function of the response Xl for the first
excitation and X2 for the second excitation, then this combination of terms
as indicated by Eq. (4-6) would not be valid. The fact that an is a function
of t, the independent variable, does not invalidate the combination of
terms.
Equation (4-5) and Fig. 4-2 point out the first important property of
a linear system, namely, that of superposition. Input Ql(t) gives rise to
response Xl(t), and input Qz(t) gives rise to response xz(t). Now inserting
both Ql(t) and Q2(t) simultaneously gives us as a response the sum of the
original two responses Xl(t) + X2(t). In other words, the responses to the
two inputs superimpose themselves one upon the other. The linear sys­
tem handles each individually in such a way that the second input in no
way affects what happens to the first input, and similarly the first input
does not alter the response to the second input. Actually, no physical
system can successfully meet the superposition criteria for all ranges of
its terminal variables.
Take, for example, the simple case of a lumped-linear resistance
described by the relationship
(4-7)
Consider the current ti to be the input and the voltage Xto be the response.
As a first input function let
a constant current
Then the response voltage is

Xl = RIo (4-8)

As a second input let


ti2 = I sin wt
152 Principles of electromechanical-energy conversion

giving a response
X2 = RI sin wt (4-9)
Now inserting a current equal to
(it + (12 = 10 +I sin wt
should give a response
Xl +X 2 R(lo +I sin wt) (4-10)
according to Eq. (4-7). The total response given by Eq. (4-10) is seen to
be the sum of the two responses given by Eqs. (4-8) and (4-9).
Suppose we allowed the direct current 10 to become very very large.
A physical resistor might in fact overheat sufficiently to actually disinte­
grate and make an added sinusoidal response impossible. Thus super­
position does not hold, since we can always postulate a large enough
current to bum up any physical resistor. When an equation in the form
of (4-7) is written, we are always implying that there is a certain allowable
range for the variables over which the relationship is a reasonable analytic
approximation for the physical device. The same statement should be
made in connection with all of the elements we have discussed in the
preceding three chapters, namely, that relationships between coordinates
are valid only for a limited range of coordinate values. The statement is
true even for nonlinear relationships.
The linear system, which can always be described by an equation in
the form of (4-2), has a second important property illustrated by Fig. 4-3.
If an input kQ(t) is applied to the system, where k is any constant, then
the response must be kx(t). The response x(t) is assumed to be the
response to Q(t) acting alone. This property, whereby a linear system
with an input k times as large will simply yield a response amplified by the
same factor k, is known as homogeneity. We can show that Eq. (4-2)
obeys the homogeneity property by simply multiplying both sides by the
constant k, thus setting
d"x d,,-IX dX]
k [ an(t) + an-let) dt,L-l + ... + al(t) dt + ao(t)x
= kQ(t) (4-11)
Since k is a constant, we can rearrange Eq. (4-11) into the form

+ al(t) d~t)
+ao{tHkx) = kQ{t) (4-12)
clearly showing that the response to an input kQ(t) is equal to kx(t).

k Q(tl kx(tl Fig. 4-3 Illustration of the homogeneity prop­


erty for a linear system.
The analysis of linear systems 153

Perfect low-pass filter Nonlinear system


fe = 1,000 cps No,l

OUI xU)

Nonlinear system
NO.2
Fig, 4-4 A system possessing superposition but not homogeneity.

In order for a system to be classed as linear, it must possess both the


property of superposition and the property of homogeneity. Figure 4-4
shows an example of a system which possesses the superposition property
but not the homogeneity property. An input Ql(t) containing component
frequencies all less than 1000 cycles per second (cps) goes through the top
line and nonlinear system 1, giving as a response Xl(t). A second input
Qz(t) containing frequency components all above 1000 cps is carried
through the bottom line in Fig. 4-4 and non.linear system 2, giving as a
response X2(t). Putting in Ql(t) and Q2(t) simultaneously results in an
output Xl(t) plus xz(t), thus verifying the superposition property. If, how­
ever, Ql(t) is made k times its original size and inserted alone, the response
will not be kXl(t). The signal is still filtered through the upper line, since
kQl(t) contains only frequencies below 1000 cps, but the nonlinear system
1 does not give a response of simply kXl(t), or k times the original response.
Thus, the property of homogeneity is not obeyed. Remember that a
linear system possesses both the properties of homogeneity and superposi­
tion, and such a system can be described by a differential equation in the
form of (4-1) or, equivalently, by Eq. (4-2).

4-2 THE LINEAR DIFFERENTIAL OPERATOR

The work we shall do with linear differential equations can be greatly


systematized through the use of the differential operator notation. For
this purpose, let us define the differential operator

d
p (4-13)
dt

The symbol p is an operator, and standing alone it has no meaning. The


operator p must have something upon which to operate. 1 Thus px is
another way to write dx/dt or, equivalently,:t. The second time deriva­
154 Principles of electromechanical-energy conversion

tive is given by

. ~:~ = :t (~~) = p(px) (4-14)

and in a similar fashion

(4-15)

showing that the operator raised to the nth power indicates that the nth
derivative shall be taken.
A general function of the differential operator can be formed as
follows:
(4-16)
where the a", a ..-l, . . . , ao coefficients can, in general, be functions of the
independent variable time. By using this notation, the linear differential
equation (4-2) can be written as simply
L(p)x = Q(t) (4-17)
where x is the unknown coordinate and Q(t) is the forcing function.
The linear differential operator defined by Eq. (4-16) obeys certain
common algebraic rules. Given three distinct linear differential operators
L1(p), L 2(p), and L3(P), the commutative and associative laws of addition
are valid. This means that
Commutative:
(4-18)
Associative:
(4-19)
Also, in general, the distributive and associative laws of multiplication are
valid, as follows:
Distributive:
(4-20)

Associative:

(4-21)
However, the very important commutative law of multiplication is not
generally valid, except under certain important circumstances to be dis­
cussed later. This law is given by
The analysis of linear systems 155

Commutative:
(4-22)
and is not true for the general linear differential operator.
An example will serve to illustrate what these two additive and three
multiplicative laws mean.

Example 4-1
For the three linear differential operators

L ,(p) == 5p + 2t L.(p) == 3 sin tp' + p L 3(p) =p


test the five laws given by Eqs. (4-18) to (4-22).
Solution: The commutative law of addition, Eq. (4-18), requires the evaluation of

[L,(P) + L.(p)]x = [(5p + 2t) + (3 sin t p' + p)]x

== (3 sin t p' + 6p + 2t)x

We must also evaluate

[L.(p) + L,{p)]x == [(3 sin t p' + p) + (5p + 2t)]x

= (3 sin t p' + 6p + 2t)x

thus illWltra.ti~g this commutative law of addition.


The associative law of addition, Eq. (4-19), requires that we show that adding L,
to the sum of L. and L. gives an operator identical to that obtained by adding L, to the
sum of L, and L.. Thus we have

L,(p) + [L.(p) + L.(p)] "" (5p + 2t) + (3 sin t p' + 2p) == 7p + 2t + 3 sin t p'
[L,(p) + L.(p)] +
La(P) = (6p + 2t + 3 sin t pO) + p = 7p + 2t + 3 sin t p'

The distributive law of multiplication, Eq' (4-20), is checked by Wling L, and L.


lIS follows:

[L,(p) + L.(p»)x = [(5p + 2t) + (3 sin t p' + p)]x

+
== (3 sin t p' 6p 2t)x
+
dx
= 3sint d'x
at l
+6
at
+2tx

Also, we have

L,(p)x + L.(p)x = (5p + 2t)x + (3 sin t pO + p)x

dx . dlx dx

= 5 at +2tx +3smt at' + at

thus verifying the distributive law for this example case.


The associative law, Eq. (4-21), is checked lIS follows:

,IL,(p)[L.(p)L.(p)Jlx == [(5p + 2t}(3 sin t p. + p'}]x


- (15 cos t p' + 15 sin t p. + 5p a + 6t sin t pI + 2tpl)X
156 Principles of electromechanical-energy conversion
Let us now attempt to check the commutative law of multiplication, given by
Eq. (4-22). First we have, for the left-hand side of (4-22),
L,(p)[L2(p)xl (5p + 2t)[(3 sin t p' + p)xl
(5p + 2t) (3 sin t ~; + ~:)
d~x .'d 3x d'x . d 2x dx
15 cos t dt 2 + 15 Sln t dt S + 5 dt' + 61 sm t dt' + 2t dt
Now for the right-hand side of (4-22) we have
L~(p)[L,(p)xJ (3 sin t p' + p)[(5p + 2t)x]
(3 sin I p' + p) (5 ~: + 2tX)
. dSx . d 2x . dx d'x dx
15 sm t (Tt3 + 6t sm t dt' + 12 sm t dt + 5 dt' + 2x + 2t dt
Quite obviously, these two results are not equal. We have shown that the commuta­
tive law of multiplication is not valid in this case.
In summary, this example has verified the commutative and associative laws of
addition and the distributive and associative laws of multiplication for the linear
difIerential operator. The commutative law of multiplication for the linear differ­
ential operator was found to be invalid for this example.

4-3 THE CONSTANT-COEffiCIENT LINEAR DiffERENTIAL OPERATOR

The linear differential operator will obey the commutative law for multi­
plication provided the an, a,,-l, . . . ,ao coefficients are not explicit func­
tions of time but are simply constants. This would mean that Eq. (4-16)
defining the operator becomes
(4-23)
with constant coefficients. Having the operator in this form, the four
laws, Eqs. (4-18) to (4-21), are still valid, and the commutative law for
multiplication, Eq. (4-22), is also valid. Equation (4-17) uses the linear
differential operator in a general description of a linear system. By use of
the constant-coefficient operator, given by (4-23), the description reduces
to a linear differential equation with constant coefficients: the type of
differential equation we can readily solve. The techniques used to obtain
solutions essentially stem from the fact that the constant-coefficient linear
differential operator obeys all of the algebraic laws of ordinary coeffi­
cients, including the commutative law of multiplication. The operators
The analysis of linear systems 157

can then be treated as ordinary constant coefficients and the differential


equations can be manipulated as if they were algebraic equations.
Suppose for a particular linear system the following set of n constant­
coefficient differential equilibrium equations is obtained:
L ll (p)Xl + L 12 (P)X2 + ... + Ll"(p)x,, = Ql(t)

L 21 (P)Xl + L 22 (p)X2 + ... + L 2n (p)x n = Q2(t)

(4-24)

Since all of the L n , L u , . . . ,Lnn constant-coefficient linear differential


operators obey all of the laws followed by algebraic coefficients, we are free
to use Cramer's rule for solving Eqs. (4-24) for any particular response
coordinate. Thus for the kth coordinate we have
Lll La Ql(t) LIn
L21 L22 Q2(t) L 2"
~(p)Xk = (4-25)
LuI Lu2 Q,,(t) Lnn
Lu L12 LIn
L2l L22 L 2n
~(p) (4-26)
Lnl Ln2 L"n
In the determinant of Eq. (4-25), Ql, Q2, . . . ,Qn are placed in the kth
column. Notice in (4-25) that ~ operates upon Xk. The cofactors along
the kth column in the determinant on the right-hand side of Eq. (4-25)
should operate on the corresponding source term. Therefore, Eq. (4-25)
can be expanded into the form
(4-27)
where ~ik represents the i-Ie cofactor (Le., the signed i-k minor of the
determinant). Equation (4-27) represents a linear constant-coefficient
differential equation with one unknown coordinate Xk. The right side of
(4-27) is an explicit function of time, since all of the Q(t) forcing functions
are known time functions. Our interest will now focus on specific tech­
niques for obtaining solutions for an equation in the form of (4-27).
Particularly, the d-c and sinusoidal steady-state responses are of interest.
Laplacian techniques will also be considered to include general response
characteristics.

Example 4-2
For the circuit shown in Fig. 4-5 write out the nodal equilibrium equations and obtain
a single equation involving only the first node variable.
158 Prim:iple. of eledromedJanica/.en8rflY conversion

T Datum
Fig. 4-5 A lumped-linear electrical network.

Solution: In terms of the nodal variables ),1 and ),1 (the integrals of the respective
node voltages) the equilibrium equations are
1 J (c • (iii l
[(C1 + C•) dl),l G d),1
dt s + di + L ),1 -
1
d ),.)
= •
-h(t) + • (
$. t)
1

- (Ct d;~l) + (C. d~~1 + G. ~. + L),s) = -i.(l) + ia{tY


In operator notation, these two simultaneous constankoefficient differential equa.­
tionsbecome

[ (C 1 + C.)p· + GIP + LJ ),1 - (C.pS»,. = -il(l) + i.(t)


- (C.P·»,l + ( C.p· + G.P + L) ),. = -i.(t) + i,(t)
By treating the linear differential operators 80S ordinary algebraic coefficients, we
ca.n use Cramer's rule to obtain a single differential equation involving only ),i­
According to Eqs. (4-25) and (4-26), we have

I
\

A(P»,l = I-~l +~. C -cG.p' 1


-$. + ., .p. + .p + Lt

which reduces to

A(P»,l == - ( C.p· + G.p + L) i l + (G.P + L) i. + C.pli, (4-28)

The characteristic determinant A expands to become

A(P) = ClC.P' + (GIC. + G,C + G.Cs)p· + (f: + GIG. + C C.) pi


1
1
i.
+ (G.
Ll
+ L.
Gl) p + _1
, LILI
Equatio~ (4-28) represents 'a fourth-order differential equation. Assuming that the
three current sources are known time fUIlctions, the right side of (4-28) is completely
specified once a.ll the required derivatives are taken. In the following sections of this
chapter we sha.ll review solutions for constankoefficient linear differential equations.
Dividing both sides of Eq. (4-28) by the characteristic determinant A would
appear to yield an eXplicit solution for ),1. This is not the ca.se, however~ because the
The anolysis of linear sy.riems 15?

linear differeiltiaJ. operator, in this case A(p), is defined only when ithaa something
upon which to opera.te.

4-4 BLOCK-DIAGRAM' NOTATION

A pictorial techD;ique for representing a set of differential equations will


aid in formulating equilibrium equations for electromechanical systems,
These block diagrams, as they are commonly called, will greatly systems.- ,
tize the reduction of a large set of simultaneous equations to a single
input-output relationship between any single forcing function and any
particula.r desired resp9nse. For linear systems, block diagrams are con­
structed from the three symbols shown in Fig. 4-6a to c. The block symbol
in Fig. 4-& has meaning in the following way. The variable x(t) is
inserted into the block as shown by the arrow on the left side of the box.
This variable is then operated upon by whatever is written inside the box,
in this case L(p), thus giving the output Q(t). The symbolism of Fig.
4-00 represents the equation .

L(P)x(t) = Q(t) (4-29)

or· the input to the block x(t) is operated upon by L(p), giving the output
Q(t). .
The 8'Ummation symbol, in Fig. 4-6b, adds (or subtracts) the two
variables Xl and X2. Plus and minus signs must be placed on each of the

~ L(p) ~ == L(p)x=Qlt)

Operation (or multiplication1


(al

Summation (summing point)


.. (bl

Fig. 4-6 Basic block-dia.gram sym­ -


x

Distribution (splitting point)


Ie)
::
boIs.
inputs to the summing circle to indicate whether that input should be
added or subtracted, as shown in Fig. 4-6b. The distribution symbol,
shown in Fig. 4-6c, is also known as a splitting point or pick-off point.
It simply distributes a variable, allowing placement into a number of
different positions in the overall diagram.
Let us take a few simple equations and see how they can be repre­
sented by using the. block-diagram notation. The series R-C circuit
shown in Fig. 4-7 has the following loop equilibrium equation,

( Rp + ~) q = v(t) (4-30)

where q is used as the integral of the loop current i. Figure 4-8a shows one
block-diagram representation of Eq. (4-30). An entirely equivalent repre­
sentation is shown in Fig. 4-8b, where the forcing function v(t) is shown
as the input to the system block and q(t) is shown as the output or response.
Figure 4-8b seems like a more natural representation, since the input
to the block is known and it is the output which we desire. If we want
the current i, rather than its integral q, a tandem block can be joined to
Fig. 4-8b, as shown in Fig. 4-8c, with p operating on q to yield i.

~I
qU) vlt)
Rp + lfc •

(a)

I' (t) 1

·1 Rp+1Jc

(b)

vlt) J 1 q(t) r:I i (tI_


! Rp+Vc
f--'---.~ Fig. 4-8 Equivalent block-diagram
representations for the R-C series
Ie) circuit.
The analysis of linear systems 161

(Rp)q

Fig. 4-9 A feedback representation of


the R-C series circuit.

Still another way to represent Eq. (4-30) is shown in Fig. 4-9. The
output of the summing point is given by
1
C q =.v(t) - (Rp)q (4-31)

which is identical with Eq. (4-30). Upon mUltiplying (l/C)q byC, we


have simply q. By use of a distribution symbol, or splitting point, q is
then fed back to block Rp, whose output (Rp)q is available for subtrac­
tion from vet). Figure 4-9 is more complicated than Fig. 4-8. However,
only one lumped element is represented by each of the two blocks: the
capacitor by the upper block and the resistor by the feedback or lower
block.
Notice that a slightly different diagram can be drawn for each manip­
ulation of the original equation. Just as all the manipulated equations
are really equivalent, so all these diagrams of Figs. 4-8 and 4-9 are
equivalent. The rules of algebra tell us how to manipulate equations.
There is a similar set of rules for manipulating these block diagrams, allow­
ing for their reduction [into generally simpler, more convenient forms.
Blocks in cascade, as shown in Fig. 4-10, can be simply combined into
a single block whose operation is the multiple of the operators of the
individual blocks. The cascade combination can be proved as follows:

X2 = L1(P)Xl
X3 = L 2 (p)X2 = L 1L 2 (P)Xl
(4-32)
x, = La(P)X3 LIL2La(P)XI

All the combinations given by Eqs. (4-32) are valid, because all the
constant-coefficient linear differential operators obey the associative and
comrnutll,tivelaws of mUltiplication given by Eqs. (4-21fand (4~22).
Summing points may be separated, reversed in order, and combined
as illustrated in Fig. 4-11. The manipulation of the summing point

Fig. 4-10
-
Reduction of blocks in cascade.
162 Principlef of electromechanical-energy cooverfion

"'2
Fig. 4-11 Manipulations involving summing points.

illustrates the commutative and associative laws of addition. The three


representations in Fig. 4-11 are respectively
X4 = (Xl - X2) - Xa
X4 = (-X2 - X3) + Xl (4-33)
Other combinations, in addition to those given by (4-33), are possible,
but the ones mentioned in connection with Fig. 4-11 should give the
general idea concerning the manipulation of summing points.
Figure 4-12a and b illustrates how a block can be moved past a
splitting point. An easy way to check the equivalence of these manipu­
lations is to see that the transmission from any input to any output
remains the same. In going from Xl to X2, Fig. 4-12a, one has to pass
through a block L(p) in either of the two diagrams. Similarly, in Fig.
4-12b, the transmission from Xl to X2 goes through L(p), and from input
Xl to output Xl the transmission equals L(p) times l/L(p), or simply unity.
Figure 4-13a and b shows us how to move a block past a summing
point. The equivalent diagrams in Fig. 4-13a represent the following
equations, respectively,

X3 = L(P)Xl - X2 X3 = L(p) [Xl - Ltp) X2 ] (4-34)



1--_....-_"'2

- 1--_~"'2

(a)

(0)
Fig. 4-12 Interchanging the positions of blocks and splitting points.
I
I
The analysis of linear systems 163
J

-
(al.

-
Fig. 4-13 Interchanging the positions of blocks and summing points.

which are obviously equivalent equations. Figure 4-13b stands for the
equations
Xa = L(p )XI - L(p )X2 (4-35)
which again are identical equations. N otice th~t the transmissions
from Xl and X2 to Xa in each of the diagrams in Fig. 4-13a and b are kept
identical.
The only additional manipulation we shall need involves the feed­
back configuration shown in Fig. 4-14a. For this common diagrarn, in
terms of the operators G(p) and H(p), the following equations cu¥ be
written:
Xa = G(P)(XI =1= X2)
Therefore, we have
Xa = G(p)Ixl =1= H(p)xaJ
which reduces to
G(p) ] (4-36)
Xa = [1 ± G(p)H(p) Xl

Thus, according to Eq. (4-36), the feedback configuration can be reduced


to the single block shown in Fig. 4-14b. We now can see how Figs. 4-8

_xf'--_..../ G(p) X3 ..
• __1_1_G_(p_I_H_~_I~r--~

(ill
(al
Fill:. 4-14 Reduction of the feedback eonfil!:Uration.
164 Principles of electromechanical-energy conversion

(a) (b)

~ KI
~ (el
O+Jp
~

Fig. 4-15 Block-diagram representation of the three equilibrium equations.

and 4-9 for the R-C circuit are equivalent by simply applying the feed­
back reduction given by (4-36).

Example 4-3

A certain electromechanical system can be described over a particular operating region


by the following set of simultaneous linear differential equations:

(R, + L,p)i1 = v,(t)

-K,il +(R. +
L.p)i. +
K.8 = v.(t) (4-37)

-K1i• .+ (D +
Jp)8 = T(t)

All the forcing functions VI, V2, and T are assumed to be known functions of time.
The unknown dependent variables are ii, i., and 8. All coefficients are constants, and
the linear differential operator p equals dldt. Draw the block diagram for the sys­
tem and reduce to a form which clearly shows the relationships between the response 8
and the three forcing functions V" Vii and T.
Solution: The overall block diagram of the system represented by Eqs. (4-37) is
constructed by combining diagrams for each of the equations. Figure 4-15a, b, and c
represents each of th~ three system equations. Notice in Fig. 4-15b how the second
equation has been manipulated to have i 1 as the input to the diagram and i. as the
output. Similarly, Fig. 4-15c has i. as the input and 8 as the output. Therefore,
quite naturally, the three parts of the figure can be combined to give the complete
block diagram shown in Fig. 4-100. The output of the diagram is the response of
interest, in this case 8. The input to the left-hand summer K.8 is obtained by the
feedback path as shown in the figure. Also notice how the three forcing functions
appear as explicit inputs to the diagram.
Figure 4-16a can be reduced to a more compact form. Combining the blocks in
cascade and moving the block Kd(R. + L.p) past the right-hand summer gives us
Fig. 4-16b. Upon closing the feedback loop according to Eq. (4-36) or Fig. 4-14, we
have the final result shown in Fig. 4-1&. From this figure the transfer relationships, or
transfer functions, between the response 8 and the three forcing functions can be
The analysis of linear systems 165

The diagram of Fig. 4-1& gives the same relationship in a neater and more compact
form. The mathematics involved in manipulating the diagram from Fig. 4-16alto b
and finally to c is simpler than evaluating the determinants necessary when using
Cramer's rule on the original equations (4-37).

Cal

(lI)

+
+
+

(e)
Fig. 4-16 Combination and reduction of the three separate diagrams.
A linear system can always be represented by an equation of the form
L(p)x(t) = Q(t) (4-39)
where the linear" differential operator L(p) is defined by Eq. (4-16), Q(t)
is a specified time-dependent forcing function, and x(t) represents a
\ general unknown system coordinate. In Sec. 4-3 we have seen that a
sy'ktem represented by a simultaneous set of differential equations
involving all of the coordinates can always be reduced to a single equa­
tion in the form of (4-39) provided the L(p) operators contain only con­
stant coefficients. Thus for a set of constant-coefficient linear differential
equations, a single constant-coefficient equation involving only one
coordinate can always be obtained.
Let us now examine the solutions of Eq. (4-39) for certain particular
Q(t) forcing functions. Suppose Q(t) were simply a constant rather
than a general time-varying function. Further, let us assume that the
constant forcing function, which we shall represent by Qo, is applied to
the system for a time of sufficient duration to allow the system to reach
its steady-state response. If the linear system represents a physical
system, then certain loss elements, such as resistors and dampers, must
be present. The steady-state response of such a linear system to a
constant d-c input must also be a constant. At the time instant at
which Q(t) = Qo is applied, the system will usually respond with some
time-varying motion. However, after a sufficient time duration, the
transient motion decays to zero, leaving only a steady-state d-c response.
Analytically we can express these concepts in the following way.
The general coordinate x(t) in Eq. (4-39) must be a constant. Therefore
we have
L(P)xo = Qo (4-40)
where Xo is the constant d-c response to the constant input Qo. Equation
(4-40) assumes the response x in (4-39) is equal to a constant Xo in the
steady state. Since the operator p = djdt operating on a constant gives
zero as a result, Eq. (4-40) can be simplified to give

L(O)xo = Qo
and thus
Qo
Xo = L(O) (4-41)

The result given by (4-41) is obtained by setting p = 0 in all the original


simultaneous differential equations. This leaves a set of algebraic equa­
The analysis of linear systems 167

;1 ­

Fig. 4-17 A linellJ' translational system with 3 degrees of freedom.

tions involving only the various steady-state responses, which can readily
be solved by using determinants and Cramer's rule.

Example 4-4
Determine the differential equilibrium equations describing the linear translationa.!­
mechanical system with 3 degrees of freedom shown in Fig. 4-17. Each of the
rolling masses is assumed to have a viscous friction force on its wheel bearings, as
shown in the figure. The position coordinates are assigned in such a manner that the
three springs lIJ'e at their free lengths when Xl, X" and X3 are zero. Assumingf(t) to be
equal to a constant force fo, find t~e steady-state position of the three masses.
Solution: Let us use Lagrange's equation to formulate the system equilibrium
equations. The Lagrangian for the mechanica.! system, in terms of the coordinates
shown in Fig. 4-17, is given by

.e(x,x) -21 (Mlxl' + M.x,' + Maxa') - -1 [Xli


2
-
Kl
+ ~-=~'"- + (X, K.
x.)2 ]

For the Rayleigh function we have


!f(x) = i(DlXl' + D.x,' + D.x,')
The three required external forces are obtained by inspection to be
Q, = 0 Q. - f<t)
By using Lagrange's equation in the form

~ [a.e(~,x)] _ a.e(x,x) + a!f~x) = Qi for i = 1,2,3


dt ax, ax; ax;
the three equilibrium· equations can readily be obtained. After a small amount of
manipulation, these equations Can be put into the (orm

[ MIP' + Dlp + (~+


KI
~)]
K,
Xl - ~X. = 0
K.
- ~XI
K.
+ [M.p + D.p + (~+
s
K.
~)] x, - ~Xa
K. Ka
0

- ~. x. + (M.p' + Dap + ~a) X. = !(t)


by using our differential operator notation.
168 Principles of electromechanical-energy conversion

Assuming the applied force f(t) to be a constant fo, the steady-state position
response can be found by setting all p's equal to zero. This leaves the following set of
algebraic equations:

The additional subscript zero is being used to indicate the d-c steady-state value for
each of the coordinates. Cramer's rule can now be used to solve these equations for
each of the final positions. ThuB for (XI)O we have

1
0 0

~(XI)O = .l..+.l.. 1
0 - Ka
Kz Ka
1 1
fo - Ka Ka
~(X,)o = l.k a
where ~ l/K,K zK a.
The final steady-state position of the first mass is therefore

Similarly, the final positions of the remaining two masses are given respectively by

and

The d-c stea.dy-sta.te "response is obtained by letting p = dldt go to zero in each of the
original equilibrium equations and then using Cramer's rule to obtain the individual
steady-state responses.

4-6 SINUSOIDAL STEADY-STATE RESPONSE


A second class of steady-state responses is alsQ of great practical impor­
tance. Very often we are interested in the steady-state response to a
sinusoidal forcing function. In many electrical systems sinusoidal volt­
age supplies are probably the most common excitation sources. The
common power transformer can readily alter the impedance level of a
distribution system, thus making transmission over long distances
practical and economical.
A second reason for the importance of the sinusoidal steady state
rests on the fact that physical systems are very often forced by periodic
functions. The generated torque of an internal-combustion engine and
the common household single-phase a-c electric motor are just two
The analysis of lineor systems 169

examples of devices that develop a periodic forcing function. By using


a Fourier sine-cosine series expansion, these periodic functions can be
represented by a sum of sinusoidal functions. The steady-state response
of a linear system to a nonsinusoidal periodic forcing function is equal
to the sum of the sinusoidal steady-state responses to each of the har­
monic terms. Therefore, finding the response to any periodic forcing
function can be reduced to finding a set of sinusoidal responses.
The sinusoidal steady state is important for a third reason. In
testing the capabilities of a physical system, sinusoidal excitations are
often found to be very convenient test forcing functions. These physical
tests are usually called frequency-response tests. If we know the system's
frequency response, we can predict quite accurately the complete response
of the system to a range of inputs widely different from a sinusoid. Thus
the frequency response provides an important tool for studying the com­
plete behavior of a system. Tests of this type, quite often, are easily
performed in the laboratory. In this section we shall review simple
methods for analytically determining frequency-response information.
Again we start with our description of the linear system by an
equation in the form .
L(p}x(t} = Q(t} (4-42)
where the constant-coefficient linear differential operator L(p) is defined
by Eq. (4-23), Q(t) is a time-dependent forcing function, and x(t} repre­
sents a general unknown system coordinate. Let us now assume that
Q(t) is a sinusoidal time function given by
Q(t) = V2 Q sin (wt + a) (4-43)
where Q is the rms magnitude of the sinusoid and 00 and a are constants.
Euler's identity for ei<»1 can be written as
ei<»1 cos wt +j sin wt (4-44)
wherej is the imaginary operator equal to v' -}. By using this identity,
the forcing function Q(t), as given by Eq. (4-43), can be written as
Q(t) = 1m (V2 Qei(.,Ha» (4-45)

where 1m is an abbreviation for the operation of taking the imaginary


part of the term in parentheses. Substituting the forcing function in
the form of Eq. (4-45) into the original system description of (4-42)
gives us
L(p)x(t) = 1m (v'2 QeJ(OJI+a» (4-46)
The response x(t} to a sinusoidal excitation must, in the steady state,
reduce to a sinusoidal form of the same angular frequency w. This
170 Principles of e/ectromechanico/-energy conversion

statement is true provided the system contains some loss element, such
as a resistor or viscous damper. All physical systems have loss elements.
Therefore, the response can, in general, be given by
x(t) = "\1'2 X sin (wt + 1')
= 1m ("\1'2 Xei("'t+-r» (4-47)
where X is the unknown nns magnitUde of the response and l' is an
unknown phase angle. Equation (4-47) gives the complete form of the
steady-state response, with only the magnitude and phase angle as
unknowns.
Before SUbstituting the assumed steady-state response of (4-47)
into Eq. (4-46) let us see how the differential operator affects such a
function. The first time derivative of the steady-state response is
px(t) = p [1m ("\1'2 Xei("'t+-r»]
= 1m [p("\I'2 Xei(",H-r»l (4-48)
upon interchanging the operations of taking the time derivative and
taking the imaginary part. Equation (4-48) further reduces to
px(t) = 1m (jw V2 Xei("'t+-r» (4-49)
when the derivative is actually taken. In a similar fashion we can show
that the nth derivative is given by
p"x(t) 1m [(jw)" "\1'2 Xei{",t+-r)l (4-50)
Now, if the assumed solution (4-47) is substituted into the system
Eq. (4-46), using (4-23) for L(p), we have
1m ([an(jw)n + a _l(jW)n-l + ... + al(jw) + aol "\1'2 Xe;(..t+-r) I
n
= 1m ("\1'2 Qei(",t+a» (4-51)

Notice that the term in square brackets on the left side of (4-51) is
actually just L(jw). Therefore, Eq. (4-51) can be condensed tol
1m [L(jw) "\1'2 Xei("'t+-r)] 1m ("\1'2 Qei(",Ha» (4-52)
Taking the total time derivative of both sides of Eq. (4-52) and
dividing by the constant angular frequency w gives us
(4-53)
Since the imaginary part of j times a complex number is equal to the
real part (Re) of that number, Eq. (4-53) becomes
(4-54)
lOne might be tempted to simply drop the 1m operations on both sides of (4-52).
However, in general, such a manipulation is not valid. For example, 1m (3 + j2) =
1m (5 + j2), but certainly (3 + j2) r6 (5 + j2).
The analysis of linear systems 171

On comparing Eqs. (4-52) and (4-54) we see that both the real parts
of two complex numbers and the imaginary parts of these numbers are
respectively equal. Therefore the numbers themselves are equal, and
we have
(4-55)
upon expanding the exponential. Since e iloll has a constant amplitude
of unity, it may be canceled, along with V2, from both sides of (4-55),
leaving ,
L(jw)(Xeh) = Qe ia (4-56)
The right-hand side of Eq. (4-56), representing the sinusoidal forcing
function, is actually a shorthand notation for the original function of
(4-43). The term Qe ia is a complex number whose magnitude is the
rms amplitude of the sinusoidal time function and whose phase angle ­
is the initial angle of the sine wave. The abbreviation of the original
sinusoidal time function is termed a sinor or sinor quantity, and the
boldface symbol Q is used to denote the sinor. Thus, sinor
Q = Qeia

stands for
Q(t) = y2 Q sin (wt + a)
Notice that we are saying that Q standsjor Q(t), not that Q equals Q(t).
Once w.e know sinor Q and the original frequency w, we can always obtain
the corresponding time function Q(t). In a similar fashion a sinor
quantity for the response can be written as

x = Xeh

standing for

x(t) = V2 X sin (wt + 'Y)


In terms of the sinor quantities the steady-state response can be
obtained from Eq. (4-56) as

Q (4-57)
X = L(jw)
The result is very similar to the d-c steady-state solution given by
Eq. (4-41), except that here p = d/dt becomes jw instead of zero and
sinor quantities instead of d-c quantities are used.
172 Principles of electromechanical-energy conversion

Example 4-5
A particular mechanical system is described by the following constant-coefficient
linear differential equation.

3 :;; + 2:: + 15x = 14.14 sin (2t + i) + 28.28 sin ( 4t - ij)


Find the complete steady-state response of this system.

Solution: In operator notation the describing equation can be written as

(3p! + 2p + 15)x = V2 [ 10 sin ( 2t +~) + 20 sin ( 4t - ij)]


The forcing function for the system is composed of two sinusoidal components. The
superposition principle must be valid for the system, since the equation is linear.
Therefore, we can find the steady-state response to each of the forcing functions acting
separately and then simply add the two responses to obtain the total response to the
two functions acting together.
The sinusoidal steady-state response to the forcing function of 2 rad/sec can be
determined in the following manner. On substituting j2 for p, we have
[3(j2)' + 2(j2) + 15jX = lOe;.. /3
l

where Xl is the sinor response to the first forcing function on the right side of the
equation. Notice that the sinor representation always uses the rms magnitude of the
timc function, and not the peak magnitude. On solving for Xl, we have
lO /60° lO /60°

Xl = (15 _ 12) + j4 = 5 /53.20 = 2 /6.8°

where we use the polar representation of the complex quantities with degrees, rather
than radians, as the angular measurement. The corresponding time response is
therefore
Xl(t) = V2 2 sin (2t + 6.8°)
The second portion of the steady-state response is obtained from
[3(j4)' + 2(j4) + 15jX. = 20e-irls
and therefore we have
20~ 20~
(15 - 48) + j8 34 /166.4°

= 0.59 / -196.4°

The complete steady-state time response is the sum of these two components, which is
x(l) = x,(t) +
x.(t)
x(t) = V2 [2 sin (2t + 6.8°) + 0.59 sin (4t 196.4°)1

Remember to reinsert the V2 when the sinor is changed back into the corresponding
time function.

If our linear system is described by a set of simultaneous constant­


coefficient linear differential equations in the form of Eqs. (4-24), then a
The analysis of linear systems 173

slightly more general procedure should be followed. Again we are


assuming that each of the Ql, Q2, . . . , Qn forcing functions is sinusoidal
or a series of sinusoidal terms and that our interest is in the steady-state
response of the Xl, X2, • • • , Xn coordinates. By a derivation identical
with that for the single equation (4-42), we can show that, for each
f .."mlB~m.~..J9~.~h~ Qjff\'J!1~j;j~l.operator. p ~a.~ be replaced by jw and time
\ --.:-uu J

Lnl(jW)Xl + L n2 (jw)X + ... + Ln"(jw)X,,


2 Q"
Equations (4-58) are complex algebraic equations which can readily "be
solved by the use of Cramer's rule. We must be quite careful to manipu­
late these complex numbers properly. On solving for the kth response,
we have
Lu(jw) L12(jw) QI Lln(jW)
L21(jw) L22(jW) Q2 L2n(jW)
A(jw)X k = (4-59)
. . . . . ...... . ....
Lnl(jW) L,,2(jW) Q" Lnn(jw)
where
Lu(jw) L12(jW) Lln(jW)

L2l(jW) L22(jW) L2n(jW)

A(jw) (4-60)
...... . . . . . 9 . ~ ...
Lnl(jW) Ln2(jW) Lnn(jw)
Expanding Eq. (4-59) and dividing by A(jw) gives us

(4-61J

as the kth response in sinor form. The terms Ai.!; represent the i-k
cofactor of the determinant. The corresponding time response Xk(t)
can easily be recovered from the sinor form of Eq. (4-61).

Example 4·6
A motor driving an inertia load can be modeled by the lumped rotating system shown
in Fig. 4-18. The load has inertia J. kg-m' and a viscous friction, or windage,
component D, newton-m-sec/rad. The shaft coupling between the motor and the
load has a compliance of K rad/newton-m. The motor has a shaft inertia of J, kg-m 2
and windage coefficient of D, newton-m-sec/rad. The developed torque of the motor,
174 Principles of electromechanical-energy conversion

Motor Shaft Lood

Fig. 4-18 A linear model of a motor


driving an inertia load.

given by T(t) newton-m, has a sinusoidal component of the form (V2 T sin wt).
Determine the steady-atate angular-velocity response of the load to this component of
the driving torque.
Solution: In operator notation, the differential equilibrium equations are found to
be
(JIP + Dl + ~p) 01- ~p 0, = T(t)

- ~P 01 + (J2P + D. + ~p) 0, = 0

The division by P has not actually been defined. The terms involving (I/Kp>,O should
be interpreted as being (I/Kp)(pfJ), which reduces to simply (I/K)fJ. Since the
angular velocities of the motor and load are of primary interest, the equilibrium
equations have the th and fJ. coordinates multiplied by p, yielding 01 and 0., respec­
tively. The four coefficients are then divided by P to maintain equality.
Using sinor notation, the steady-state angular velocity of the load is given by

J 1 (jw) + Dl + K(~w) T
1
K(jw)
o
where
1 1
J 1 (jw) + Dl + K(jw)

+ D. + K(~w)
1
- K(jw) J.(j",)

By expanding these two determinants, we have


T
Dl + D. - w'K(D1J. + DJl) + j",(J~ + J. + D D.K 1 ",'KJ1J.)
When parameter values are substituted, we can simplify the equation into the general
form

e =O.h
which has a corresponding time response
W2 = 0, = V2 O. sin (wt + or)
Suppose that the two windage coefficients Dl and D. were small enough to be
negligible. The steady-atate angular velocity of the load, in sinor form, would then
reduce to
The analysis of linear systems 175

If the frequency w of the generated sinusoidal torque happens to be equal to

Wr

then the denominator becomes zero and the velocity of the load oscillates between plus
and minus infinity. This infinite resonant response occurs only if the windage
coefficients are neglected; in practice, they never can be entirely. However, large
torsional oscillations can certainly cause great damage to the physical components in
the system.

4-7 FREQUENCY-RESPONSE CHARACTERISTICS


'\
In the preceding section we have considered the steady-state response
of a linear system to a sinusoidal forcing function. The general character
of the response must be sinusoidal with an angular frequency identical
with that of the forcing function. If the forcing function is composed
of a series of sinusoidal terms, then, by using the principle of superposi­
tion, we know that the response must also be a similar series containing
a harmonic term for each of the components in the forcing function.
Only the magnitude and relative phase displacement of the response
can differ from that of the sinusoidal forcing function. The sinor nota­
tion contains only magnitude and angle information, and thus it pro­
vides a concise tool for sinusoidal excitations.
One of the most pra:ctical techniques for studying the operation of a
physical system is to examine its steady-state response to sinusoidal
inputs over a suitable range of frequencies. For a constant magnitude
of the forcing sine wave, we want to know how the magnitude and rela­
tive phase angle of the output vary as the input frequency is varied.
This information is usually displayed or plotted in three ways:

1. The magnitude ratio, or the ratio of the output sinusoid to the


input sinusoid, and the relative phase angle are plotted as functions
of frequency. This plot is usually called a Bode plot.
2. The magnitude ratio and phase angle can be plotted by using
polar coordinates with the frequency as a parameter of the curve. This
is called a Nyquist or polar plot.
3. The magnitude ratio can also be plotted as a function of the
phase angle on regular cartesian coordinates again having the frequency
parametrically displayed. This is commonly called a NichollS plot.

As a starting point, we shall concentrate on the first representation, or


the Bode plot. This is usually the simplest to determine, and the
ren~aining two representations can easily be drawn from it.
176 Principles of electromechanical-energy conversion

The sinusoidal steady-state solution for a set of simultaneous con­


stant-coefficient linear differential equations is given in sinor notation
by Eq. (4-61), which for convenience can be put into the form

(4-62)

'T'hp kt.h rp.<lnonI'lP in sinor form to a set of 0; sinor forcina: functions can be
A'k
X k = ~ Q. (4-63)

Equation (4-63) is the kth sinor response to the ith sinor forcing function
acting alone. The solution of Eq. (4-63) gives us the sinusoidal steady­
state response to the sinusoidal forcing function Q.(t).
The frequency-response characteristics are plots of the magnitude
and relative phase angle of the response Xk(t) as the frequency of Qi(t)
is varied while the magnitude and phase angle of Q,(t) are held constant.
We desire, therefore, a plot of the magnitude and phase angle of the
transfer relationship or transfer function A.k/ A as a function of frequency.
The term "transfer function" is quite descriptive in the sense that it
gives the transferring recipe whereby the system "cooks up" the response
for a given forcing function.
Let us now consider the most general form that the complex transfer
function A,k/A could have. Both the cofactor and the determinant are
made up of the products of polynomials involving jw to various powers.
Therefore, we should always be able to write Aik/A in the following form:
A,k bm(jw)m + b..._1(jw)m-l + ... + b1(jw) + bo (4-64)
A a,,(jw)n + a.._l(jw),,-l + ... + al(jw) + ao
where m and n are integers. Factoring bo from the numerator and ao
from the denominator changes (4-64) into
Aik =K (bm/bo)(jw)m + (bm_tlbo)(jw)m-l + ... + (b1/bo)(jw) + 1
A (a,./ao)(jw)n + (an_t/ao)(jw)n-l + ... + (at/ao)(jw) + 1
(4-65)

where the coefficient K = bo/ ao and the two polynomials involving jw


are dimensionless. If bo or ao is equal to zero, then jw could be factored
The analysis of linear systems 177

from the numerator or denominator, respectively, and we could divide


through by b i or al. In general, Eq. (4-65) could havejw raised to some
power as a factor in either the numerator or denominator.
A fundamental theorem in algebra tells us that an nth-order poly­
nomial, such as the denominator of (4-65), must have n factors. Further,
since all of the coefficients an, an-I, . , . ,ao are constant real numbers,
we know that these factors can only be made up of the following forms:

1. Real distinct factors such as (j1'W + 1). The term "distinct"


means that the factor appears to the first power, or that it is not repeated.
2. Complex conjugate factors arising from terms of the form "
(jW)2 + 2tl(jW) + 1

W1 2 WI

which factors into the form l

(~~ + tl + j VI - tl2) (~~ + tl - j VI - t12)

3. Repeated factors, which could be real factors or complex conjugate


factors raised to a power greater than unity.

With constant-coefficient polynomials in jw no other types of factors


can exist.
Instead of working with the general form of the complex transfer
function, as given by (4-65), let us consider certain special cases. From
these special examples, we shall be able to extend our thinking to see
what form the most general frequency response might take. First con­
sider a transfer function consisting only of real distinct factors such as
X" ~ik K(j1'lW + 1) (4-66)
Q. = ~ = (j1'2W + 1)(j7'3W + 1)
where 1'1, 1'2, and 7'3 are constants and 7'3 > 7'2 > 7'1. To obtain a Bode
plot, we need the magnitude and phase angle of this complex tralisfer
function as the frequency W is varied. The magnitude of (4-66) is given
by

I~:I (4-67)

and the angle of Eq. (4-66) is similarly given by

I!!:
L....S1i. = tan- 1 nw (4-68)

'Here an wand an w, are being used, where w stands for the frequency, of the
forcing function and w, is a coefficient or parameter of the system. \..
178 Principles of electromechanical-energy conversion

We could plot Eqs. (4-67) and (4-68) as functions of frequency,


which would give us the desired Bode characteristics. However, these
calculations, particularly for Eq. (4-67), are quite involved even for a
simple function. Also, if more factors are present in the numerator
and denominator, the calculation becomes even more difficult. With
a little ingenuity both these difficulties can be surmounted.

4-8 THE LOGARITHMIC FORM OF THE FREQUENCY RESPONSE


A number can be given in terms of decibels (db) by taking the logarithm
to the base 10 of the number and multiplying by 20. Thus a number k
expressed in decibels is given by
k(db) 20 loglO k (4-69)
with the db in parentheses to indicate that k is in decibels. Table 4-1
shows a very interesting relationship between the decibel expression
for a number and the actual value of the number. Notice that as the
value of k is increased by a factor of 10, the value of k, in decibels, increases
by 20 db. If k increases by a factor of 100, which is, of course, 10 times
10, then 40 db is added to the decibel value of k.
By taking the logarithm to the base 10 (loglo) of both sides of Eq.
(4-67) and multiplying by 20, we have

20 loglo I~: I = 20 10glO K + 10 loglo [(-qW)2 + 1]


- 10 loglo [(T2W)2 + 1] - 10 loglo [(TaWP + 1] (4-70)

Remember that the logarithm of a product is the sum of the individual


logarithms and that the logarithm of a quotient is the difference of the
logarithms. Also, the 10glO of a number raised to a power equals that
power times the logarithm of the number.

TABLE 4-1 Decibel


equivalent values
k k(db)

0.001 -60

0.01 -40

0.1 -20

1 0

10 20

100 40
1000 60
The analysis of linear sysfems 179

40
,/'

/'
V
20
/'"
Actr' curve~
--?"
/ - High-frequency
asymptote
+ o Slope = 2.0 db/decode
a
....., LO~-freq~ency
~
asymptote I I
--20 I - ­ I I I I I

Now let us examine the plots of each of these terms as a function


of frequency. The first term, 20 loglo K, does not contain w, so it is
just a constant for all frequencies. The next term, 10 10g1o [(TIW)2 + 1],
can be plotted in a very simple way. Consider the rallge of frequencies
where w is much less than 1/1'1. For these frequencies this first term
can be approximated by
1
10 logio [(TIW)2 + 1] ~ 10 10giO 1 o for w « 1'1
(4-71)

Similarly, for a higher range of frequencies where w is much greater than


l/n, the term is given approximately by
10 10glO [(TIW)2 + I} ~ 10 logio (TIW)2
= 20 10giO nw for -1
w» 1'1 (4-72)

Equations (4-71) and (4-72) are respectively known as the low­


frequency and high-frequency asymptotes of the term. In Fig. 4-19
these straight-line asymptotes are drawn on semilog graph paper. The
low-frequency asymptote, given by (4-71), is zero for all frequencies
where w «l/TI. A horizontal line is drawn at 0 db for all frequencies
up to 1/1'10 The asymptote is a poor approximation for the term near
the so-called cotner frequency where w l/n.
The high-frequency asymptote given by Eq. (4-72) is also a straight
line, as shown in Fig. 4-19, when the frequency w is plotted on a loga­
rithmic scale. The magnitude of the function 20 IOgiO TIW increases by
20 db for each decade in the frequency w. A decade in frequency means
that the frequency has increased by a factor of 10. From w = l/Tl to
180 Principles of electromechanical-energy conversion

w = 10/Tl is a decade in frequency. Notice that for this range the


asymptote has gone from a value of zero to a value of 20 db. From
w = 10ITl to w = 100/Tl the asymptote goes from 20 to 40 db in value.
The high-frequency asymptote is a good approximation to the term for
frequencies where w» 1/71. Therefore, near w = l/Tl this asymptote
also has its maximum error.
By calculating just three points in the vicinity of w I/Tl, using
the low- and high-frequency asymptotes, an exact plot of the magnitude
response contribution of this second term in Eq. (4-70) can be drawn.
At the corner frequency we have w I/TI' and therefore
10 IOglO 2 3 db
At half the corner frequency, w = 0.51Tl and
10 loglo [(TIW)2 + I] = 10 logio 1.25 = I db
At twice the corner frequency, w = 2/Tl and
10 loglo [(TIW)2 + 1] = 10 logio 5 = 7 db
Notice that at twice the corner frequency we have
10 logio 5 10 loglO (1.25 X 4) 10 logio 4 + 10 logIO 1.25
The value of the high-frequency asymptote at w = 2ITI, according to
(4-72), is 10 loglO 4. Thus at both half the corner frequency and twice
the corner frequency we simply add I db to the asymptotes. At the
corner frequency w = 11TI and we must add 3 db to the asymptotes.
By drawing a smooth curve through these three points and along the
asymptotes, we have the actual plot of the term, as shown in Fig. 4-19.

40
I

I I I I
i

20
.0

'0
Low-frequency
/ asymptote Lgh-freqUency
asymptote _
~l~
0
Slope; -20
~
Actua~7 ~
~~ db/decade
~
curve
-20
............
i'--­
-40 I !

O.5/T2 I/T2 2/T2


1.1.1, rod/sec
Fig. 4-20 Actual and approximate magnitude plots for the transfer function
1/(jTf.6J + I).
20 201091

'­ 0__----i----t----1"---..:::--____1,,;;:----t---t
~...
X
-201__~-4--_r----1__-+---·____1,

-40~--~--~----~-~-~---~--~
lIT, 1/T2 1fT3
41, rod/ sec
Fig, 4-21 Total asymptotic and corrected magnitude plots for the transfer function
X./Q. K(jTIW + 1)/0T"" + 1)(jr3w + 1) with T3 > Tz > T"

The third term in Eq. (4-70) can be plotted in exactly the same
manner as the second term, except that the slope of the high-frequency
asymptote is negative 20 db/decade in frequency. The correction
quantities are still 3' db at the corner frequency w = 1/72, and 1 db at
half and twice this frequency. The only difference is that these correc­
tions must be subtracted rather than added to the asymptotes. Figure
4-20 shows a plot of the third term in (4-70). Notice again that the
corner frequency and the asymptotes are all we need to obtain the plot.
The high-frequency slope and corrections are standard for the distinct
factor.
According to Eq. (4-70), the complete plot of 20 logio of the magni­
tude of the transfer function is simply the sum of the individual loga­
rithmic plots. In Fig. 4-21 we begin by drawing the asymptotes of each
of the terms in Eq. (4-70), with 20 logIo K as simply a constant horizontal
line for all frequencies. All the various asymptotic curves can be added
together, giving the total asymptotic curve shown in Fig. 4-21. Just
by using the major corrections of 3 db at each of the three corner'fre­
quencies, the approximate curve can be corrected to give a more exact
plot, as shown also in Fig. 4-21.
The phase angle of the transfer function given by Eq. (4-68) involves
the adding or subtracting of a standard angular characteristic. Figure
4-22 shows a normalized plot of tan- 1 a as a function of the normalized
angle ct, again using semilog graph paper. For values of ex much less
than unity tan- 1 a is approximately zero, and for a much greater than
unity tan- 1 a approaches 90°. From Eq.(4-68) the phase angle of the
transfer function can easily be plotted by using this normalized charac­
teristic. Dividing the normalized angle by the individual time constants
182 Principles of electromechanical-energy conversion

TI, T2, and Ts allows the phase contributions of each term to be read
directly from Fig. 4-22. Adding and subtracting these contributions,
according to whether the factor is a numerator or denominator term in
the transfer function, yields the total phase response as a function of
frequency.
Figure 4-23 shows this resulting phase pharactedstic as given by
Eq. (4-68). Notice that for values of the forcing frequency w much less
than the reciprocal of the time constants the total phase shift of the
transfer function is zero. As w increases, the numerator factor is the
first to become significant, causing the characteristic in Fig. 4-23 to
become initially positive. However, as w continues to increase, the two
denominator factors add negative angular contributions, with the phase
characteristic finally approaching - 90 0 for large values of w. The
numerator factor contributes +90 0 and the two denominator factors
give 1800 total, thus resulting in a net _90 0 at high frequencies.
Having more real distinct factors in the numerator and denominator
than in the transfer function originally proposed by Eq. (4-66) would
simply mean more terms would have to be included in the magnitude
characteristic of Fig. 4-21 and the angle characteristic of Fig. 4-23.
However, both of these plots are formed by adding various quantities,
so the inclusion of more terms does not change the form of the individual
contributions. Adding asymptotes or phase contributions for a more
complicated transfer function is therefore just a simple extension of the
ideas presented for the simple transfer function of Eq. (4-66).
As a second consideration, let us assume the transfer function to
contain a complex conjugate factor. Since the effect of the other real

90r----r------~--~----,------,----~

60
~

30~--1-----_b~---~--1-----~--~

0.2 0.5 1.0 2 5 10

Fig. 4-22 Normalized phase contribution for the real distinct factor 1 + ja as 8.
function of a.
The analysis of linear systems 183

30

f.--- f,.--­ 1\
o

~-30

\
-60 ~ "­"
- ~
-90 i'-..
lIT, 11r2
W, rod /sec
Fig. 4-23 Total phase-shift characteristics for the transfer function Xk/Q. =
K(j7',,,, + 1)/(j7'2'" + l)(jr .... + 1) with 7'3 > 7'2 ? 7',.

distinct factor can easily be added to the complex conjugate factor, let
us restrict our considerations to the 1l1mple transfer function
XA: 1
(4-73)
Q, = (jw)2/WI2 + 2rl(jW)/WI + 1
containing a single complex conjugate factor. The magnitude of the
factor, in decibels, is given by

20 loglo I ~ 1 = -1010g lo {[ 1 (4-74)

after taking the square root of the sum of the squares of the real and
imaginary parts and putting the 1- exponent as a coefficient of the loga­
rithm. The phase angle of the transfer function (4-73) is obtained from

(4-75)

Notice that in both the magnitude and angle functions the driving
frequency W is always divided by the undamped natural frequency WI
of the complex factor. Therefore, by drawing the Bode plot as a function
of the normalized frequency W/WI, we shall have characteristics useful
for any value of WI.
Considering W/WI to be very small, the low-frequency asymptote
is obtained. From (4-74), we have

20 loglo I~: j "'" -10 loglo 1 o for w« WI (4-76)


184 Principles of electromechanical-energy conversion

When W/WI becomes large, Eq. (4-74) can be approximated by

for W » WI

(4-77)

thus giving the high-frequency asymptote. These two asymptotes can


readily be plotted on semilog graph paper, as shown in Fig. 4-24. The
low-frequency asymptote, from Eq. (4-76), has a value of 0 db. Since
W/WI is plotted on a log scale, the high-frequency asymptote of Eq. (4-77)
becomes a straight line with a slope of -40 db/decade in frequency.
This means that as W/WI goes from 1.0 to 10, the IXk/Qil decreases from
o to - 40 db. The actual magnitude of the transfer function therefore
reduces to 1/100 of its value over the decade range of frequencies.
The correction quantities for the asymptotic approximation to the
magnitude portion of the Bode response for a complex conjugate factor
depend not only on the normalized frequency W/Wl but also on the value
of the damping ratio .\1. Figure 4-25 shows these correction quantities
with a curve for each selected value of the damping ratio. For a par­
ticular damping ratio the proper decibel corrections can be added or
subtracted according to Fig. 4-25. For example, if the damping ratio
were equal to 0.6, at W/WI = 1.0, 1.8 db should be subtracted from the
asymptotic approximation of Fig. 4-24; at W/Wl 0.4, 0.3 db l!ihould be
added to Fig. 4-24; and at W/WI = 2.0,0.4 db should be added. Notice·

20

(,Lo..LreqJency LympJe
o

"" '"
r<.High-frequenc y asymptote
Slope -40 db/decade
:g -20
.
a
"'­
.;(-40

-60
'""" "­
-80
0.01 0.1 2
Normalized frequency
5
W/Wl
10
'"~ 100

Fig. 4-24 Asymptotic ma.gnitude plots for the tra.lIBfer function Xk/Qi =
1/[1 - (W/W,)2 + j(2(,w/w,)].
The analysis of linear systems 18S

201~.----~---,--~,-._~~------,_--r_,_-~I~~
I ~,=0.05
18r-~~~+-~-r--T--r-+-r~~~-----+--~+--+~+-+-~-+1

16 \__----+---+-_+_-\__+-+--1-11+1-- ~--\__-_+_- -'-+--+-+--l--l-l

5 6 7 8 10
Normalized frequency w/w.
Fig. 4-25 Decibel correction quantities to be added to the asymptotic magnitude
plot of Xk/Qi = 1/[1 - (W/WI)' + i(2 sl w/Wl)\ for various values of the damping
ratio SI ~ 1.0.

that /ilmall values of the damping ratio require large corrections, particu­
larly near W/WI 1.0. With no damping, a steady-state response of
infinite magnitude occurs when the driving frequency W equals the
undamped natural frequency WI. Only values of r 1 equal to unity or
less are shown, because for damping ratios greater than unity the complex
conjugate term can be factored into two real distinct factors.
Figure 4-26 shows the phase characteristics for Eq. (4-75), again
plotted as a function of the normalized frequency W/Wl. For each value
of rIa different phase characteristic is shown. Notice, however, that
all these characteristics start at 0 0 for W/Wl « 1.0 and approach -180 0
for W/Wl» 1.0. The phase characteristic for small damping ratios
changes very abruptly in the vicinity of W/WI 1.0, becoming more
gradual as r 1 increases.
The third type of factor which we might have in a transfer function
is a repeated factor either real or complex conjugate in form. The con­
tribution of such a term is the same as that of the single non!,"epeated
factor, except that it is included the same number of times as the factor
is repeated. Suppose the transfer function contains a real factor repeated
186 Principles of electromechanical-energy conversion

three times such as

(4-78)

Taking the magnitude of the transfer function of (4-78), expressed


in decibels, we have

20 loglo I~: I -3! 10 IOglO [(TW)2 +m (4-79)

Equation (4-79) is identical with that for a real distinct factor except
for the factor of 3 preceding the braces. The low-frequency asymptote
is still at 0 db up to the corner frequency w = l/T. Beyond the corner
frequency the slope of the high-frequency asymptote is -60 db/frequency
decade, or three times the slope for the distinct factor. The corrections
also are multiplied by 3. Thus at W = l/T we should subtract 9 db
(or 3 times 3 db) from the asymptotes. Also at w = 0.5/T and at w = 2/T,
3 db (or 3 times 1 db) should be subtracted.
The phase angle of Eq. (4-78) is given by

-3 tan- 1 TW (4-80)

or simply 3 times the phase shift of the real distinct factor. Thus, at
each corresponding frequency, the angular contribution is 3 times that
of the same nonrepeated factor. Figure 4-27 shows the complete Bode
plot of the transfer function (4-78). For a repeated complex conjugate
factor the same ideas apply. Simply find the magnitude and angle
contributions of the factor if it is not repeated and multiply these decibel

O~~~~~~~~UL~~~~~~IITTn
-20
-401---­
i-601---­
:t: -801-,-----1:-­
.c:: i

.
.. i
... -1001-'- - - + - ­
o
if. -120 t----t---t--tf--+-+--t-+-+-w~~-.;;;:-+---t--
-1401-----t--+--tf-t-+-+_
-160 t----t---t--tf--t­
-180'-­ _ _--'_ _...1.---''--.1....­
0.1 0.2 0.3 0.4 0.6 0.8 1.0 :3 4 10
w/Wj
Normalized frequency
Fig. 4-26 Phase-ahift characteristics for the transfer function Xk/Q. =
1/[1 - ("'/"'1)' + j(2ilw/"'I)) for various values of the damping ratio il :::; 1.0.
\ The analysis of linear systems 187

-20
o
r-­
-- ~

"
-I'--­
l""
~
,
o
-30
-60
90 ~

...
~ -40
'c "" ",,­
~

120£
:;::
150 :

"
V>

o ._­ -180,f
::0

-60
~ ~ -210
~ I~
-80
O.51r 1/r
W, rod /sec
2/r
--- .~
-240
270

Fig. 4-27 Magnitude and phase characteristics ofthe transferfunction 1/ (irw + 1)3.

and degree contributions by the exponential power of the factor. The


decibel corrections shown in Fig. 4-25 are also multiplied by the number
of times the factor is repeated, meaning the exponent of the factor. A
few numerical examples will help to clarify these procedures.

Example 4-7
A particular electromechanical system has only two ports, one being electrical and the
other mechanical. For a certain operating range, the following linear differential
equation in operator form describes the motion of the system:
(pI + 6.5p2 + I03p + 50)x = (1.5p + 3)v
where x is taken as the position response of the mechanical member, in meters, and v
is the voltage applied to the electrical port, in volts. Determine the frequency response
of the system in the form of a Bode plot.
Solution: The frequency response requires that the driving function v(t) be
sinusoidal in form. The magnitude and phase angle of the sinusoidal steady-state
response xCt) are to be determined as the frequency w of the driving function is varied.
By using operator notation, the transfer function between the response and the forcing
function can be put into the form

:: = 1.5p + 3
v pi + 6.5p2 + 103p + 50
which can be factored and manipulated to become

x O.06(O.5p + 1)

I) = (2p + 1)[pl/lOO + 2(O.3)p/10 + 1]

Factoring the third-order polynomial into the form shown can often be a difficult job.
The second-order factor in the denominator factors into two complex conjugate
factors, so we shall leave it as shown.
188 Principles of electromechanical-energy conversion

Substituting jw for p gives us the complex transfer function relating the sinor
forcing function V to the sinor response X. Thus, upon making this substitution, we
have
x (4-81)

The composite of the asymptotes for the magnitude of this transfer function is
shown in Fig. 4-28. For frequencies below j-, or 0.5, rad/sec the asymptote has a
magnitude equal to 20 loglo 0.06, or - 24.4 db. At w = 0.5 the real denominator factor
contributes a -20 db/decade slope. Then at w = 2.0 (which is 1/0.5) the real
numerator factor contributes a +20 db/decade slope, thus resulting in a combined
slope of zero. At w = 10, the complex conjugate factor contributes a -40 db/decade
slope, as shown in Fig. 4-28.
The corrections to the asymptotes are summarized in Table 4-2. For the real
factors, the corrections are 3 db at the corner frequencies and 1 db at half and twice the
corner frequency. Corrections for the complex factor are taken from Fig. 4-25, using a
damping ratio r = 0.3. By using the normalized phase characteristics shown in
Fig. 4-22 for real factors and in Fig. 4-26 for complex conjugate factors, we can con­
struct Table 4-3. The total phase characteristic is also shown in Fig. 4-28 referred to
the right-hand scale.
By using these two plots, the magnitude and phase shift of the response for any
sinusoidal forcing function can be taken by inspection. For example, suppose the
forcing voltage, applied to the electromechanical device, is given by
11(0 = 15 sin (4t + 60°) volts (4-82)
At a frequency of 4 rad/sec the magnitude and phase angle of the transfer func­
tion are seen from Fig. 4-28 to be -35 db and -36°, respectively. The magnitude

-20
-­ ~
~ ChL I
0

40

1­ y
,
.,'"
.0 ~ -Corrected "
"" ~ magnitude ;;
~ -40 -80 x
-

x
o
...
'0
Campo~ite of
the asymptotes
\

\
~
120
----
'0
~
'"
<=
i! -60
'"
'c

""
'"

::lE
\ '­ "~ ~
'"
.s::
a.
-80 160
.",
r--­
-100 200
0.1 0.2 0.5 1.0 2 5 10 20 50 100 200
w. rad/sec
Fig. 4-28 Magnitude and phase characteristics for the transfer function
X 0.06(jO.Sw + 1)
V .. (j2w + 1)[1 - (w/lO)2 + .i2(0.3)w/l01
The analysis of linear systems 189

TABLE 4-2 Decibel corrections for the factors


in Eq. (4-81)
(j",)2 2(0.3)j",
j2", + 1 jO.5", +],.." 100 + -10- + 1 Total
'"
0.25 -1 -1
0.5 -3 -3
1.0 -1 +1 0
2.0 +3 +3.0
4.0 +1 +1.2 +2.2
5.0 +1.7 +1.7
10 +4.5 +4.5
20 -1-1. 7 +1.7

TABLE 4-3 Phase-angle contributions, in degrees,


for the factors in Eq. (4-81)

(j",) 2 2 (0.3)j",
'" j2", + 1 jO.5w + 1
100 + -----w- + 1
Total

0.1 -11 0 0 11
0.25 -27 +7 0 -20
0.5 -45 +14 0 -31
1.0 -63 +27 -4 -40
2.0 -76 +45 -7 -38
4.0 -83 +63 -16 -36
5.0 -84 +68 -22 -38
10 -90 +79 -90 -101
20 -90 +84 -157 -163
450 -90 +90 -173 -173
1000 -90 +90 -175 -175

of - 35 db corresponds to
1O-3~. = 0.018 m/volt
Therefore, the response of the system to Eq. (4-82) is given by
x(t) = 0.27 sin (4t + 24°) meters

Example 4-8
As a second example, determine the Bode plot for the transfer function
T(P) = 5(0.011' + 1)
1'(0.002p + 1)1

Solution: Substituting j", for l' in the transfer function gives us

T( ' ) = 5(jO.01", + 1)
(4-83)
3'" j",ljO.OO2w 1)1+
190 PFinciples of electrolr1echanica/-energy conversio"

o
'< Slope =1- 20 db/decode JJ I
0

-20
~ Corrected magnitude
40
~k
com~osite of
0>

'"
-0

~
--..........
3-40
~ r-­ i'-..
the asymptotes
80 ~

'"'~"
I­ I­
....
o
....
0

'"
-0
.-Ec: -60
Phose)
~ '"c:
120 0>

'"
0
0>
o ~
::;;; 0
.J:;

"'r-­"
Cl.
-80
'" "-.......
160

-100 -200
10 100 200 500 10 3 10 4

w, radlsec

Fig. 4-29 Magnitude and phase characteristics for the transfer function

T(jw)
5(j0.Olw +1)
jw(j0.002w +1)2

The factor of j", in the denominator requires a slight extension of the techniques we have
developed for obtaining the frequency-response plots. The phase contribution of the
jw factor is -90 0 even at very low frequencies. Expressing the magnitude of the
transfer function' T(jw) in decibels, the jw factor contributes ~20 log,o w db to the
total magnitudj). Therefore, for each decade in frequency the factor subtracts 20 db,
or it has a slope of -20 db/decade even at very low frequencies. Remember, how­
ever, that the decibel contribution of this term is zero only when w = 1 rad/sec. If
the w axis starts at 0.1 rad/sec, then this jw factor contributes +20 db to the total­
magnitude plot. With w starting at 10 rad/sec, the contribution is -20 db.
Figure 4-29 shows the magnitude and phase plots of Eq. (4-83). At w = 10 rad/
sec the magnitude of T{,iw) is given by
IT(jl0)1 = 20 loglo 5 - 20 loglo 10 = -6 db
Observe that the magnitude plot starts at -6 db with a slope of -20 db/decade.
The repeated real factor has its corner frequency at w = 500 rad/sec (which is 1/0.002).
Past this frequency the factor contributes a slope of -40 db/decade, since it is raised
to the second power. Also, all its decibel corrections are double those of a nonrepeated
factor. The phase characteristic at w = 10 rad/sec starts at -84 0 because of a +6 0
contribution from the numerator factor. Also, all phase-angle contributions from
the repeated factor are double those of a nonrepeated real factor.

4-9 LAPLACE TRANSFORM TECHNIQUES

By using the linear differential operator, we have formulated sets of


equations describing lumped linear systems. After studying the prop­
erties of the constant-coefficient operator, we have shown that a single
linear differential eauation involvinl!" only one of the unknown coordinates
The analysis of linear systems 191

could always be obtained from the complete set of equilibrium equations.


In the preceding sections the solution of such an equation has been
studied under the conditions that all system forcing functions were either
constant in their value or sinusoidally time-....arying. The solutions were
further restricted to be valid only after steady-state conditions have
resulted. Therefore, after some period of time following application
of a d-c or sinusoidal forcing function, our solutions become valid.
Although these d-c and sinusoidal steady-state solutions are of great
practical importance in system analysis, we still require a more general
technique which will provide complete response information for a wider
class of forcing functions. These solutions are to be valid not only for
steady-state conditions but also for the complete time interval after
applying the forcing function.
For our purposes, one of the most useful techniques for solving a
linear constant-coefficient differential equation involves the Laplace
transformation. Given a time function f(O, the Laplace transform is
constructed by first weighting or multiplying the function by e- st and
then integrating the product with respect to the independent variable t
over the limits zero to infinity. Thus, symbolically the Laplace trans­
form of a time function is given by

.e[fCt)] = F(s) = fo~ f(t)e-· t dt (4-84)

The symbol .e[f(O] indicates the forming of the Laplace transform.


The definite integral is evaluated over the limits t = 0+ to t = 00. The
notation 0+ means that we approach time t = 0 along the positive time
axis, so t = 0+ indicates the time just after t = O. After the integral
is evaluated and the limits are substituted the resulting function involves
only the Laplacian variable s; thus, we use the notation F(s) for the
.e[f(O] in Eq. (4-84). Only certain time functions are actually Laplace­
transformable. All functions commonly encountered in engineering
work fall into this category. 1
Taking a few specific functions of time, we can find the correspond­
ing Laplace transform by simply substituting these f(O into Eq. (4-84).
A listing of time functions and their corresponding transforms can be
made as shown in Table 4-4. Transform pair 1 in Table 4-4 is for a
unit impulse oCt), which is defined as a pulse of width a and height l/a
in the limit as a goes to zero. Figure 4-30 shows such a pulse. Notice
that even though the base is of zero time duration and the height becomes
infinite the area of the pulse remains at unity. The impulse occurs at

1 For an excellent discussion of these ideas from an engineering viewpoint see


Wilbur R. LePage, "Complex Variables and the Laplace Transform for Engineers,"
McGraw-Hill Book Company, New York, 1961.
192 Principles of electromechanical-energy conversion

f(t)
au) =Iim f(tl
a-O
11--_---.
a

Fig. 4-30 The unit impulse fune­


o a t tion 6(t).

TABLE 4-4 Laplace transform pairs

No. I(t) i F(s)


1 6(t) 1
2 u(t) -s1
3 e-~tu(t) -1­
s+a
1
4 te-~'u(O
(s + a)1
(n - 1)1
5 tn -Ie-G'U( t)
(8 + a)'
+ '",)2 + ",2
6 e-~t sin cd u(t)
(s

7 e-G' cos cd u(t)


s+a
(8 + a)2 + ",'

8 dl(t) = pl(t) sF(s) - 1(0+)


dt
2
d /(t) = p2/(t) dl
9 s'F(s) - s/(O+) - - (0+)
dt' dt
dnl(O
10 - = pn/(t) s"F(s) - s'-I/(o+) - 8'-'1'1/(0+) ... 1".-1/(0+)
dtn

11 kl(t) kF(s)
12 klNt) + k.j.(t) k 1F I (s) + k,F.(8)
13 I(t - bluet - b) e""F(s)
-
14 f I(t) dt = I~)
F(s)
-+-­
s
p-1/(0+)
s
15 ff I(t) dt = I~t;
F(s)
-
S2
+ - 2- + - ­
1'-1/(0+)
8
p- 2/(0+)
8

16 II··· /../(t) dt = :.
F(s)
8"
p-1f(O+)
- + - - + ... + - ­
8"
p-nf(O+)
8
The analysis of linear systems 193

Fig. 4-31 The unit step function ..t


u(t).

the time when the argument is zero; thus oCt - b) would be a unit
impulse at time t b. On substituting the unit impulse oCt) into
Eq. (4-84), we have

.13[0(0] = 10: o(t)e~'1 dt 10: oCt) dt


= = 1 (4-85)

The only time when the integrand 0(t)e-B1 is not zero is at the time t 0+,
as seen from Fig. 4-30. At this ~ime instant the integrand, reduces to
simply oCt), which has unit area. In fact for any finite f(t) we can say

10: oCt - b )f(t) dt 'f(b) (4--86)

The second transform pair in Table 4--4 is for the unit step function
u(t) pictured in Fig. 4-31. The function u(t) is zero for all values of its
argument which are less than zero and positive unity for all positive
values of its argument. Therefore, u(t - b) is a unit step function, with
the step occurring at time t = b.
Transform pairs 3 and 4 are actually just special cases of transform
pair 5, with n = 1 and n 2, respectively. Notice that a u(t) is included
with' the time function to ensure our remembering that the transform
is defined only in the time interval t = 0+ to t = 00 and gives no infor­
mation for t < O. Thus, if we are given the function 1/(8 + a) and are
asked for the corresponding time function, the correct reply is e-"'u(t),
and not just e-"'. The two answers are equivalent only for t > O. By
letting the constant a in transform pair 5 be equal to zero, w.e have the
unit step for n = 1, the unit ramp for n = 2, the unit parabola t2u(t)
for n = 3, and so forth for higher powers of n.
Transform pairs 6 and 7 are exponentially damped sinusoids.
Notice that the transforms of these time functions lead to functions of 8
which have complex conjugate denominator factors of the form 8 + a + jw
and s + a - jw.
In transform pairs 8 to 10 we have the transform of the derivative
operation. With regard to the general nth derivative given by pair 10,
notice that the Laplace transform is obtained by substituting F(s) for
f(t) and 8" for the differential operator p". In addition, the transform
contains a series of initial conditions multiplied by s raised to various
194 Principles of electromechanical-energy conversion

ttt) u(t) tIt-b) u(t -b)

t
Fig. 4-32 A shifted
(al (bl time function.

powers. The initial-condition notation is defined as follows:


f(O+) = value of the time function f(t) at time t = 0+
pf(O+) = ~ (0+)
. dt
= value of the first time derivative off(t) at t = 0+

pkf(O+) = :;! (0+) = value of the kth time derivative off(t) at t = e+

Transform pairs 11 and 12 show the linear properties of the Laplace


transform. Multiplying the time function by a constant k simply multi­
plies the transform by k. The Laplace transform of the sum of two
time functions is the sum of their individual transforms. However,
remember that
(4-87)
The transform of the product of two time functions is, in general, not
equal to the product of their transforms. Transform pair 13 in Table 4-4
is for a shifted time function. Figure 4-32a shows a particular f(t)
which is zero for all timeless than zero, and thus it can be weighted by
u(t) without changing its value. In Fig. 4-32b the same function is
shifted b units in time and is therefore given by f(t - b)u(t - b). The
shifted step function u(t - b) must be included to ensure that the
function remains zero for all time t < b.
One additional transform pair will be found to be of great conven­
ience in our later work. If a function g(t) is the first time derivative
of an arbitrary function f(t), then in operator notation we can write
g(t) pf(t) (4-88)
On solving this equation for f(t), we have

f(t) = g(t) (4-89)


p

The original definition of the differential operator P, given in Sec. 4-2,


did not include the operation lip which appears in Eq. (4-89). Since
g(t) is the first time derivative of f(t), then f(t) must be the first time
integral of g(t). The operation IIp must therefore mean integration
The analysis of linear systems 195

with respect to time. However, the indefinite integral contains an


arbitrary constant, which means that ie-get) is a known function, then
Eq. (4-89) specifies J(t) only to within an arbitrary constant. On the
other hand, if we are given the function J(t), Eq. (4-88) specifies get)
precisely. Because of this arbitrary feature of the lip or integral
operation, we have chosen not to include its use form,ally in our previous
work. Wherever lip has appeared in an equation, we have always
been able to operate on the whole equation by p (multiplying the equa­
tion by p), thereby eliminating the integral operation.
For convenience in the use of the Laplace transform technique,
we desire the transform of g(t)lp. From Eq. (4-89) we have

,c [g~)] = ,c[f(t)] = F(s) (4-90)

Taking the transform of Eq. (4-88) gives us


G(s) = sF(s) j(O+) (4-91)
by using transform pair 8 in Table 4-4. Solving Eq. (4-91) for F(s) and
substituting into (4-90) gives us, finally,

.c [get)] = G(s) + p-lg(O+) (4-92)


p s s

where p-lg(O+) = J(O+) is the value of the integral of get) evaluated


at time t 0+.
By defining a function k(t) as the first time derivative of y(t), we
can write
.c [h(t)] = R(s) + p-1k(0+) (4-93)
p s s
and since h(t) g(t)/p, we have

(4-94)

Therefore, substituting (4-94) into Eq, (4-93) gives us, finally,

(4-95)

with p-1h(O+) p-2g(O+) as the second integral of g(t) evaluated at


the time t 0+, Transform pairs 14 to 16 summarize these integral
transforms with the notation p-kj(O+) standing for the kth integral of
jet) evaluated at time t = 0+.
196 Principles of electromechanical-energy conversion

_M_V_(_O_+_I~'~l____~_M_ _~r-___V_(S_)~ Fig. 4-33 A block diagram for the


. S+WM
transformed equilibrium equation.

Example 4-9
At time t = 0 a mass of M kilograms is moving with a velocity Va m/sec in the positive
x direction. A viscous force having a coefficient of D newton-sec/m opposes this
motion. Determine the velocity of the mass valid for all time t ~ O.
Solution: In operator notation, the equilibrium equation for the system valid for
all time t ~ 0 is given by
(Mp + D)v = 0

where we have used v = x as the velocity of the mass in meters per second. By
Laplace-transforming the equilibrium equation, we have
M[sV(s) - v(O+)j + DV(s) = 0

From Table 4-4 we have used pair 12, which equates the transform of the sum of two
time functions to the sum of the individual transforms. The capital V(s) stands for
the .e[v(t)j, and v(O+) is the value of v(t) at time t = 0+.
On solving for V(s), we have
Va
V(s) = 13 + D/M (4-96)

where we have substituted Va for v(O+) as the initial velocity. The time function
corresponding to Eq. (4-96) can be easily obtained by scanning the right-hand column
of Table 4-4 for an identical function of s. Pair 3 with a = D / M matches this
requirement, and therefore we have

(4-97)

as the x velocity for all time t ~ O. Notice that Eq. (4-97) says that v(t) is zero for all
time t < O. This mayor may not be true. However, the Laplace transform is not
valid in the time interval t < 0, so we do not expect a valid solution for values of
t < O. The inclusion of the unit step function u(t) helps us remember this fact.
Figure 4-33 shows a block-diagram representation for the transformed equation
(4-96). The input to the block is the initial momentum Mv(O+), and the output is the
transformed velocity V(s). The transfer function contained within the block is
obtained by substituting the Laplacian variable 8 for the differential operator p.
Remember, however, that the appropriate initial-condition inputs must be included in
the transformed block diagram as shown in Fig. 4-33.

Example 4-10
Figure 4-34a shows two rotating inertias 11 and 12 kg-m 2 coupled by means of a clutch
and a shaft of compliance K rad/newton-m. Each inertia has a windage or rotating
viscous friction coefficient of DI and D2 newton-m-sec/rad. At time t = 0 the inertia
I s is stationary and inertia I I is rotating at an angular velocity of ("'.}o rad/sec. The
clutch is then closed. Find the angular velocity of 12 for all time t ~ 0 after the
clutch is closed.
The analysis of linear systems 197
...

Solution: The equilibrium equations valid for all time t 2: 0 can be written by
using Lagrange's equation or, for a case that is this simple, directly by iUflpection.
Another technique also is found to be of some use. Consider the electrical network
shown in Fig. 4-34b with capacitors of J 1 and J 2 farads, admittances of D 1 and D. mhos,
a.nd a.n inducta.nce of K henrys. The node volta.ges are ta.ken to be WI and Wz with
respect to the datum node. Writing the nodal equilibrium equations for this network
would yield equa.tions identica.l with those we would write for the original system
shown in Fig. 4-34a. Such a network is often'called the electrical analog of the mechani­
cal system. Mechanica.l velocities become node voltages; forces and torques are
analogs of electrical currents. Mass and inertial elements are capacitors drawn from
their respective nodes to the datum node. Inertial velocities are always with respect
to a stationary reference, which in the a.nalog is the da.tum node. Viscous coefficients

wl(tl K

0,

Datum

(bl
p-l(W,-W2)(O+)

K$

~ Dotum
(e)
Fig. 4-34 (a) A lumped-linear rotational system. (b) The analog electrical net­
work. (c) The La.placia.n equivalent circuit.
198 Principles of electromechanical-energy conversion

become conductances, and compliances are drawn as inductances. Notice how the
same connections are used in the analog as in the original system.
Another way to show the equivalence of the analog circuit is to write the state
functions for the two diagrams. These state functions are seen to be absolutely
identical; thus the same equilibrium equations must result once Lagrange's equation is
employed. Writing the equilibrium equations for the system, from Fig. 4-34a or b, we
have

(JIP + DJ + ip)Wl - ipW2 0


-i p WI + (J.P + D. + i p ) w. 0

Laplace-transforming the equilibrium equations gives us

(Jill + D, + ill)01(8) - is fh(s) J 1Wl(0+) - ill [/ (WI - W.) dt]._o+

- ill 1 0 (8) + (J21l + D. + is) fh(s) J.w.(O+) - ill [/ (w. - wtl dt ]'_0+
(4-98)
The transformed equilibrium equations show the substitution of s for p, with the
initial-condition terms appearing as sources on the right-hand side of the equations.
The term [ / ("" - "'.) dt 1-0+ is simply the angular twist of the coupling shaft at
time t 0+. This quantity is analogous to the initial flux linkages in the inductor of
Fig. 4-34b. With the clutch open the angular twist can be taken to zero. The twist
cannot change instantaneously, because the angular positions of the two inertias can­
not change instantaneously without an infinite angular velocity. Therefore, these
two integral initial conditions are zero, as is the term ",,(0+). The initial condition
"" (0+) is given as (WI)..
From Eqs. (4-98) we can draw an additional analog circuit known as the Laplacian
equ.ivalent circu.it, as shown in Fig. 4-34c. Here the node variables are the transform of
the angular velocities. The initial conditions become Laplacian current sources.
All elements involving a p operator in the original network must have an initial­
condition source in the Laplacian circuit representation. By writing the nodal equa­
tions of Fig. 4-34c, we obtain Eqs. (4-98), thus verifying this circuital representation of
the equations.
The transform of the angular velocity of J. can be obtained from Eqs. (4-98) by
using determinants as follows,
1
J,Il + DI + Ks
1
- Ks o
where
1 1
Jls + D, + Ks
1 1
- Ks J 2B + D. + KG
After expanding these two determinants, the result can be manipulated into the form
(w,)o
KJ.
(4-99)
The Qnolysis of linear systems 199

The only technique we are going to use for finding inverse transforms (i.e., the
time functions corresponding to functions of 8) is that of identifying the particular F(s)
in a table similar to Table 4-4. If the particular F(s) is found in the table, then the
corresponding f(t) is obtained by inspection. Very often, however, the transform
functions of interest become quite complicated, and an enormous table would be
necessary to include all possibilities. Equation (4-99) giving [l.(s) has a third-order
polynomial in II as its denominator. An examination of Table 4-4 shows that it con­
tains no such function. Generally, if nume~ical values for aU the parameters are
known, then a mathematical manipulation known as a partial-fraction expansion can be
used to separate the function into parts which are always individually found in a table
as simple as Table 4-4. Even if values of the parameters are not known, the general
character of the response can be determined from the transformed function. In the
next section we shall study techniques which provide pieces of information concerning
the time function from the s function even though the inverse transform cannot
formally be determined.
Suppose, for a particular set of parameter values, Eq. (4-99) reduces to

which can be put into the factored form of

(4-100)
(8 + 2)[(8 + 2)2 + I)
This expression for [l2(8) is now rearranged into the sum of three fractions,

02(8) = ~ + B + --;~_-:::; (4-101)


("'1)0 s + 2 8 + 2 +;1 8
where A, B, and Care as-yet-unknown constants. Since the denominator of Eq.
(4-100) has three factors, three fractions in the form of Eq. (4-101) can always be
formed.
The constants A, B, and C can be found in the following manner. Multiplying
Eq. (4-'101) by the factor s + 2 and letting s = -2 would leave only r1 on the right­
hand side. Therefore, we can write

A = [(8+2)°2(8)]
("'1)0 0 __ 2
= (8
1
+ 2)2 +1 I
a=-2
=1

using Eq. (4-100) for 02(8)/(W1)o. Similarly, B is determined as

B [ (8+2+;1)02(8)]
(W1)0 .--2-i1
= 1
(8+2)(8+2 ;1)
I0=-2-;1
1 1
-2
and C is given by

C [(8+2_;1)°2(8)] =B* 1
(W1)o '--2+i1 - '2
where B* means the complex conjugate of B. To conjugate a complex number,
replace every j by -j.
On substituting A, B, and C into Eq. (4-101) and combining the last two fractions
we have
200 Principles of electromechanical-energy conversion

Transform pairs 3 and 7 in Table 4-4 are recognized to correspond to these two partial
fractions. Pair 12 tells us that the inverse transform of the sum of two 8 functions is
equal to the sum of their individual inverse transforms. Thus the angular velocity of
the second inertia is given by
W2(t) ,.. (wl).[e-"(l cos t)]u(t)
The term 1 - cos t oscillates between a value of zero and +2, with the coefficient e-2<
causing these oscillations to exponentially decay to zero as t goes to infinity. When I
reaches 2 sec, r2< has a value of rt "" 0.02; thus the angular velocity w. is very close
to being zero after four time constants, or 2 sec.

4-10 THE PARTIAL-FRACTION EXPANSION: POLES AND ZEROS

Linear systems with coefficients whose values are independent of time can
always be described by a relationship in the form
Q(t) (4-102)
where p = dldt is the differential operator and all a .. , an-I, . . • , au
coefficients are constants. The forcing function Q(t) is usually a specified
function of time, and the system response x(t) is desired. If the actual
system has many forcing functions, and correspondingly many response
coordinates, then we must first do some manipUlating to obtain a single
equation involving just one response, as given by Eq. (4-102). Section
4-3 gives a complete discussion of these manipulations.
Laplace-transforming Eq. (4-102) term by term gives us
-r als -r ao»)[(s)

Q(s) -r initial-condition terms (4-103)

with all initial conditions moved to the right side of the equation. The
response of the system is therefore given, by
)[(s) = Q(s) -r initial-condition terms (4-104)
ans" -r an_IS" -r . . . -r al8 -r ao
I

The time response x(t) corresponding to the transformed response )[(s)


can be found if we have a table of transform pairs containing the particular
function with the general form of (4-104). In Example 4-10 a technique
known formally as a partial-fraction expansion was used to break a
complicated function of s into a sum of simple functions which could
always be found in even so short a table as Table 4-4. However, in order
to use the partial-fraction technique, nth-order polynomials in s have to be
factored into their n factors. From a practical standpoint, the factoring
operation is often extremely difficult to .perform, and consequently we
become interested in developing techniques for obtaining response informa­
-
The oooly,;, of linear ,y,tems 201 '

tion from the Xes) function directly without having to actually obtain the
corresponding x(t).
The most general form that Eq. (4-104) will take, for our purposes,
can always be expressed by

(4-105)

Equation (4-105) indicates that any response can always be given as the
ratio of two polynomials in s. Further, we can always have the order of
the numeratQr polynomial, in this case m, at least one integer unit less
than the order of the denominator n. If this were initially not true
(that is, n were not greater than m), then the denominator could be divided
into the numerator to get Xes) as a polynomial in s plus a remainder in
the form of Eq. (4-105). The inverse transform. of each of the polynomial
terms can easily be found, and only the remainder of the long-division
process, given by (4-105), is of interest.
On factoring b", and an from the numerator and denominator,
respectively, we have
Xes) = b", s'" + (b",_I/b.,.)S_l + ... + (bI/b",,)s + bo/b"" (4-106)
a" s" + (an-I/a,,)s" 1 + ... + (at/an)s + ao/an
In theory Eq. (4-106) can be factored into the form
Xes) = b.,. (s + {Jl)(S + (J2) ••. (s + (Jm) (4-107)
a" (s + al)(s + at) . . . (8 + an)
where the numerator has m factors and the denominator has n factors.
As discussed in Sec. 4-7 in connection with the Bode plot, these factors are
real or complex conjugate in form and are either distinct or repeated.
If Eq. (4-106) can be factored into the form of (4-107), then a partial­
fraction expansion can be performed and Eq. (4-107) is split into a sUm of
terms given generally by

Xes) ~ + A2 + ... + ~ (4-108)


s + al S a2 8 + an
where each A. can be evaluated from
for i = 1, 2, . . . ,n (4-109)
If the kth factor is repeated r times, then the expansion given by
(4-108) is slightly modified to the form

Xes) = ... +~+ A,\:2


8 + ak (s + ak)2
A,.,.
+ ...+ (8 + ar)r + (4-110)
202 Principles of electromechonic;'l-energy conversion

Here the Aki coefficients can be found from

Aki = ( r· 2')1 {dd: [(s + ak)rX(s)]}


t. S i ._-a~

for i = 1, 2, . . . ,r (4-111)
If r = 1, meaning the factor is not actually repeated, then Eq. (4-111)
reduces to Eq. (4-109), since 1/0! equals 1.
The important thing to notice is that all of the terms in the expansions
of Eqs. (4-108) and (4-110) are given in Table 4-4. We have started with
a general response, Eq. (4-105), from which the complete time response
can be obtained if we can factor the polynomials in s. Names are com­
monly given to these all-important numerator and denominator factors
in Eq. (4-107). A value of s which makes Xes) equal to zero is liter­
ally called a zero of the function. This Xes) has m zeros located at
s -1'1, -1'2, ... ,-I'm' Similarly, a value of s which makes Xes)
have an infinite value is termed a pole of the function. The function
Xes) has n poles which are located at s -aI, -a2, . . . , -an. At
values of s near infinity we have
X( co) = lim Xes)

(4-112)

Therefore, we have (n - m) zeros at s co. Counting the zeros at


infinity, the total number of zeros always equals the total number of poles .

.4-11 TIME-RESPONSE CHARACTERISTICS FROM POLE-ZERO LOCATIONS


The Laplacian variable s can be considered to be ,a complex number with
the general form
s = IT + jw (4-113)
By use of a cartesian-coordinate system with IT as the abscissa and jw as
the ordinate, all possible values of s can be plotted, as shown in Fig. 4-35.
Such a graph is known as an s-plane plot. The poles, indicated by crosses,
and zeros, indicated by circles, of the function Xes) are plotted in Fig.
4-35. The first pole is located at s = -al and is drawn on the assump­
tion that a1 is a positive real number; thus s = -a1 is a point on the
negative real axis. Taking 1'1 to be a positive real number, the first zero,
located at s = -1'1, is also a point on the negative real axis. The second
and third poles and zeros are assumed to be complex conjugate terms with
negative real parts. This means that Re (a~) = Re (a3) is a positive
number and thus the poles are located in the left half of the s plane (LHP),
The analysis of linear systems 203

Lett-half plane Right-holt pia ne


(LHP) (RHP)

Fig. 4-35 Pole-zero locations plotted in the 8 plane.

as shown in Fig. 4-35. The fourth pole is taken with a4 equal to a nega­
tive real number, thus placing this pole at a point on the positive real axis.
In a similar fashion all the remaining poles and zeros of X(s) can be
plotted in the s plane.
Referring to Table 4-4 and Eq. (4-108), we see that the poles of the
F(s) functions determine the character of the response. For real distinct
or repeated poles transform pairs 3 to 5 summarize the relationships
between the terms in F(s) and their corresponding time functions f(t).
Thus;if the partial-fraction expansion of Xes), as given by (4-108) and
(4-110), contains poles locate~ on the real axis in the s plane, meaning
that a particular a, is a real number, then the transform pairs 3 to 5 give
the corresponding time function. On examining these transform rela­
tionships we see that poles located on the negative real axis (that is, ai is a
positive real number) contain a damped exponential in their correspond­
ing time function. Using L'Hospital's rule, we can show that!

for any k (4-114)


t-TOO

Therefore, poles located on the left-half reltl axis correspond to responses


that tend to decay with increasing time. Poles located farther from
the origin On the negative real axis give a faster decay because a in Eq.
(4-114) has a larger value. However, if the pole is located on the posi­

1 For example, if k = 2, we have


lim tSe- G' = lim ~ = lim ~ =0
1-". 1-+0. aeG ' I....... a'e'"

204 Principles of electromechanical-energy conversion

tive real axis, then the time response will contain a term of the form e'"
which grows exponentially to infinity as t goes to infinity.
For complex conjugate poles, transform pairs 6 and 7 give us a
similar result. For poles located in the left half of the s plane the time
response contains a decaying exponentiaL The total response oscillates
sinusoidally with an exponentially decreasing amplitude. On the other
hand, complex conjugate poles in the right half of the s plane (RHP) oscil­
.late sinusoidally with an exponentially increasing amplitude as t increases.
We can summarize these conclusions by two simple statements:
1. Poles in the LHP correspond to exponentially decaying time
responses.
2. Poles in the RHP correspond to exponentially growing time
responses.
Poles located on the boundary, or jw axis in Fig. 4-35, between the
RHP and LHP, give time responses which neither grow nor decay. If
a in transform pairs 6 L \d 7 in Table 4-4 were zero, then we would have
complex conjugate poles located on the jw axis. The corresponding time
responses are sinusoidal with amplitudes neither growing nor decaying.
For a distinct pole located at the origin, we have a step response u(t)
as the corresponding time function. This response neither grows nor
decays, but simply remains constant in time. Thus with an s-plane
plot the form of the corresponding time function can be completely
predicted simply from the pole locations. From Eqs. (4-109) and (4-111)
we see that the zeros enter into the calculation of the partial-fraction
expansion coefficients. Thus the zeros of the function serve only to
influence the magnitude of the responses.
In general, the transform of any time response can be expressed
as the ratio of two polynomials in s in the form of (4-105), A pole-zero
plot for Eq. (4-105) cannot be obtained unless these polynomials can
be factored. Any information that we can obtain concerning the factors
or, equivalently, the location of the poles and zeros without actually
factoring will provide information about the form of the time response.
For example, if we can show that all poles are located in the LHP, then
we know that the time response must decay with increasing time. For
a linear system this decay property will serve as a definition of stability.
A stable linear system has a transfer function of s containing only LHP
poles.

4·12 THE ROUTH CRITERION


A technique known as the Routh criterion gives a method for determining
how many roots of a polynomial have positive real parts. Choosing as
The analysis of lineor systems 205

our polynomial the denominator of Eq. (4-105), we have


First row
,--------------.--------------------~

pes) = a"s" +
Second row
(4-115)
First we must formulate an array of numbers known as the Routh array.
The complete array is as follows:
a" a..~2 a_4 a,,~6

a"~l a..~a a..~1i a,,-7


b1 bi b. a
(4-116)
Cl C2 C;
d1 di da

The first two rows of the array are composed of the coefficients in the
polynomial P(s). The first row starts with the coefficient of the highest
power of s, in this case an. The second term in the first row is the (n - 2)
coefficient, and so forth taking every other term in P(s). In the second
row of the array are the remaining coefficients in the polynomial (4-115).
The third row of the Routh array is formed in the following manner.
For the first term in the third row bi we have

The determinant-like structure is formed by taking the first elements


in the two rows above the term being formed and the two elements -in
the two rows above but one column to the right of the term being formed.
Thus, an and an-l are the first elements of the two rows above b1, while
a n -2 and a..-3 are the two elements in the two rows above but located
one column to the right of b1 • The determinant-like structure is evalu­
ated like an ordinary 2 X 2 determinant with a negative coefficient.
Thus the product of the two terms connected by the dashed arrow are
subtracted from the product of the two solid-arrow terms. The difference
of these two products is divided by the lower left-hand corner term of
the determinant, in this case an-I. The second term of the third row
is given by

_, I
Ia"'x
a
hi = a..-I a-. = a..-Ia..-, - a"a..-6

a..
_t a.._t

206 Princip/~s of electromechanical-energy conversion

Only the second column of the determinant changes to the terms in the
two rows above b2 located one column to the right of b2 • Similarly,
for ba we have

a,,_la,,_6 - a"a,,_l
an-I

The same rules are used for all remaining rows of the array. Thus,
for the fourth row we have

and

Always use the first two terms in the two rows above for the first column
of the determinant· and the two terms in the two rows above and one
column to the right for the second column of the determinant. Divide
the product of the solid-arrow terms minus the dashed-arrow terms by
the second-row first-column term in the determinant. Zero elements
do not have to be entered into the array.
The number of sign changes in the first column of the Routh array equals
the number of roots of P(s) with positive real parts. Thus; given any
denominator polynomial, we can formulate the Routh array. By count­
ing sign changes in the first column we know the number of RHP poles.
A stable system has no RHP poles 1

Example 4- J J
Use the Routh criterion to show that

x _ lO(s + 1)
(s) - (s + 2)(s - 1 + j3)(s - 1 - j3)(s + 3)
has two poles in the RHP.
Solution: The denominator of X(s) is given by

P(8) = 8' + 38 8 + 68 2 + 38s + 60


If the denominator of X(8) were originally given in the form of P(s), factoring would
be quite difficult and we would have no idea that two poles of X(s) are in the RHP.
The ana'r,i, of linear ,ystelnS 207
1
From P(,) the Routh array is given by
I Fir,k:olumn
signs RoutJr. array

+ 1 6 60
+ 3 38
-20/3 60
+ 65
+ 60
The terms are obtained a8 follows:
I. First row consists of a" aI, and ao in P(a).
2. Second row consists of a, and al in P(a).
3. Third row:

First term = 1 31~61


3
38 = 18 - 38 =
3-a­
20

Seoond term = \3
1 ><,:60 I
3 '" 0 -­ 60

.
Third term, etc. = I!><:~I
3 = 0

4. Fourth row:

First t
erm
= I 3 .... ><':'381
-20/3 '"60
'-20/3
= 38 + 27
.

= 65

. t -2gl;'<~1
Second term, etc. = -20/3 = 0

5. Fifth row:

-20/;.x 60 I

First term = 1 65 .... 0


- 60
65 ­
-20/3 0 I
1 65 0
Second term, etc. '"' 65 = 0

6. Sixth row, etc.:

65 01

First term, etc. =


1
~o 0
"" 0

7. And we get zero for all remaining terms in tliearray.


There are two sign changes in the first column, thus verifying that there are two
poles of X(a) in the RHP. Sign ohanges are eounted a8 follows:
208 Principles of electromechanical-energy conversion

First row plus, second row plus: no change

Second row plus, third row minus: change

Third row minus, fourth row plus: change

Fourth row plus, fifth row plus: no change

etc.

giving a total of two sign changes.

Example 4-12
As another example of the Routh criterion consider the angular-velocity response
function for the inertia J 2 in Example 4-10. This angular-velocity response is given
by Eq. (4-99), which for convenience is repeated as
("'1)0
KJ.

Use the Routh criterion to see if there are any real values of the parameters which
would yield an exponentia.lly growing response for ",.(t).
Solu.tion: The Routh array for the denominator is given by

D1 +Ds
J, J.
01

Dl +D.

KJ 1J.

where
01 = 1.. (.!. + 1..) + D D.
1 _ (D, + D.)/KJIJ.
K J1 J. J,J. D,/J, +D./J.

On simplifying 0" we have

01 = D,D. + (l/K)(D,/J,' + D./J.·)

J 1J. D,/J, + D./J.

Notice that for positive values of all coefficients every first-column term in the array
must be positive. Thus we have no sign changes and no RHP .poles in O.(s). For
non real negative coefficients we could have sign changes and an unstable response.
For example, if the shaft compliance K were negative, 0, could be negative and the
first column would have one sign change and hence one RHP pole. A negative K is
not a physically realistic parameter.

4-13 SUMMARY
In this chapter we have reviewed certain properties and techniques
applicable to a linear system. The homogeneity and superposition
The ollOlysis of linear systems 209

properties were found to be ·characteristic of all linear systems; thus,


responses to multiple inputs are handled individually by the system to
give a combined response. The linear differential operator provides a
useful shorthand notation for analytically describing the linear system.
For constant-coefficient operators all of the properties of ordinary
algebraic coefficients were found to be valid, thus allowing the differential
equilibrium equations for a system to be manipulated as a set of algebraic
equations. We therefore can always obtain a set of equilibrium equa­
tions, each containing only one unknown dependent variable.
The block-diagram notation has been included for convenience in
representing and manipulating a set of equilibrium equations. In our
later work these diagrams will significantly aid in actually formulating
an analytic description of electromechanical systems.
The d-c and sinusoidal steady-state responses of linear systems can
easily be extracted from the equilibrium equations. With all excitations
held constant, a linear system will always have a constant response,
provided some energy-loss elements (resistors, viscous dampers, etc.)
are present in the system. Thus we found that letting the differential
operator p = d/dt become zero in the equilibrium equations yielded a
set of algebraic equations whose solutions are the d-c steady-state
responses. Similarly, for sinusoidal excitations, we substitute jw for p
and sinor quantities for all excitations and responses. Now we have to
solve a set of complex algebraic equations whose sinor solutions can then
be related to corresponding sinusoidal time functions.
Since many physical systems are actually driven by periodic forcing
functions, plots giving the sinor responses as a function of the sinusoidal
driving frequency are very often of great interest. From a practical
standpoint one of the easiest plots to construct is the Bode plot. Here
the magnitude and phase angle of the sinor response~are individually
plotted as a function of the driving frequency. By the use of semi log
graph paper and decibel units for the magnitudes, a very simple. sys­
tematic scheme has been presented for obtaining these Bode plots.
For determining the complete response of a linear system to a large
class of practical forcing functions, the Laplace transform provides an
extremely powerful tool. By using this technique, we, can convert a
set of differential equilibrium equations into a transformed set of algebraic
equations containing sufficient initial-c~mdition information to obtain
complete transformed responses. Taking the inverse transform of these
Laplacian responses gives us the desired time-dependent sjtstem opera­
tion. By introducing the concept of a-plane poles and zeros, we are able
to discuss corresponding time functions directly in terms of the Laplace
functions without really having to find inverse transforms. In particu­
lar, the response corresponding to a pole located in the right half of the
210 Principle$ of electromechanical-energy conver$ion

8 plane (RHP) gives rise to an unstable or exponentially growing response


for any practical input. The Routh criterion provides us with a tech­
nique for determining whether or not poles or zeros are located in the
RHP without actually solving for their exact location.
Our later work with electromechanical systems will generally involve
nonlinear differential equilibrium equations. By extending the ideas
developed for linear systems, we can obtain information concerning the
operation of these nonlinear systems. Actually, the linear system should
always be looked upon as a mathematical idealization of our physical
world with which we are able to cope analytically. Therefore, we now
wish to see how this mathematical model, known as the linear system,
can be applied to what nature has created for the purpose of engineering
and scientific understanding.

PROBLEMS

4-1 The equilibrium equations for a group of physical systems, where x(t) is the
system response, are given by the following equations. Determine whether these
systems are linear or nonlinear by applying both the homogeneity and superposition
tests. Take Q(t) to be a known forcing function.
a. (5p2 + 6p + 2)x(t) = Q(t)
b. (16t2ps + 4t~~ sin 6t p + 3) xO) = Q(t)
c. pep + l)x(t) = Q(t)
d. [5p' + (px)(6p) + 2]x(t) = Q(t)

4-2 a. Show that (xp)' ;: X·p2.


b. Show that (xp)(x'p') ;: (x·p·)(xp).

4-3 Given the differential operators L,(p) = 6tp, L.(p) = 3p' + 2p, and
L3(P) 4pl.
a. Check the commutative and associative laws of addition.
b. Check the distributive and associative laws of multiplication.
c. Find the two operators which obey the commutative laws of multiplication.
Prove your answer.
d. Which of these operators are linear differential operators and which are classed
as constant-coefficient operators?
4-4 a. Explain why the differential operator p can be manipulated as an algebraic
coefficient in the solution of simultaneous constant-coefficient linear differential equa­
tions by means of Cramer's rule.
b. If n ... 4 in Eqs. (4-24), find the 412 and 422 cofactors required for the solution
of x•.

4-5 Reduce the block diagram shown in Fig. P4-5 to the form shown in Fig.
4-14a. Find expressions for G(p) and H(p).
The analysis of linear systems 211

Fig. P4-5

4-6 a. For the linear electrical network shown in Fig. P4-6 write the loop equa­
tions in the form of Eqs. (4-24).
b. From these loop equations determine the linear differential operator relating
i.(t) to vet). If i.(t) = L(p)v(t), find L(p).
C. Draw a block diagram representing the two loop equations.
d. Reduce the diagram of part c to a single block with vet) as the input and i.(t)
as the response output. Compare the operator in this single block with L(p) from
part b.

Rj

vI!)

Fig. P4-6

.....7 a. Suppose vet) in Prob. 4-6 is a d-c voltage source of magnitude v. volts.
Reduce the general loop equations to their doc steady-state form by replacing p by
zero. Do the same thing with the general block diagram for Fig. P4-6.
b. If vet) = V2 V sin wt, rewrite the equilibrium equations and the block dia­
gram for a steady-state description of the network. What is p replaced by, arid in
what form are the variables expressed?
C. Briefly justify the two different substitutions for p to obtain the d-c and
sinusoidal steady-state responses.
4-8 If the denominator polynomial in Eq. (4-65) is of seventh order, thus n = 7,
make a list of all possible combinations of factors. How many real factors and how
many complex conjugate factors could the polYll,omial contain?­
4-9 For the following linear differential equations in operator form determine
the frequency response in the form of a Bode plot. Assume vet) to be a sinusoidal
input function and x(l) to be the response function. For the magnitude plots, draw
the asymptotes and then add the correction quantities.
a. (p8 + 12p' + 20p)x = (1000p)v
b. (p' + 16p s + 100p)x = (lOp + 50)v
C. (p' + 16p· + 64 p 2)X (15p + 30)v
212 Prim:iples of electromechanical-energy conversion

4-10 For the asymptotic magnitude plot shown in Fig. P4-l0 determine the
transfer fUJlction GU",). Be Bure to evaluate the magnitude of the constant multiplier.

-40 db/decode

5 20 40
liJ. rod/sec
Fig. P4-10

4-11 The input to a certain linear system is given by


i(t) = 10 sin (wt + 30°) amperes
Express the transfer function X(jw)/I(jw) in terms of a magnitude, in decibels, and a
phase angle for the following responses:
a. x(t) ... 25 sin (",t 20°)
b. x(l) = 2.5 sin (wt - 20°)
c. x(t) = -2.5 sin (",I - 20°)
d. x(l) == -2.5 cos (wt - 20°)

4-12 By using the Laplace transform method, solve the following system
equations subject to the given initial conditions:
a. (pi + 4p)x = 6u(t) X(O+) = OJ x(O+) - 0
b. (p' + 4p)x = 66-"'14(0 x(O+) == OJ x(O+) ... 0
c. (p' + 4)x = 3 sin 2t '14(1) x(O+) ... OJ x(O+) = 0
d. (3p' + 15p + 18)x = 0 x(O+) = 2; x(O+) = 5
4-13 Expand the following fUJlctions in partial fractions and find the inverse
transforms:
5(s + 1)(s + 2)
a. G(s) = s(s + 3)(8 + 4)(s + 5)
10(s + 0.5)
b. G(8) == (48 + 1)(23 + l)(s2 + 48 + 5)

0.2(s + 3)"

t;. G(s) = 3'(8 + 2).

3+1

d. G(3) == (81 + 28 + 5)'


4-14 The poles and zeros for the transfer fUJlction representing a particular
linear system are loca.ted at the following points in the 11 plane:
The analysis of linear systems 213

Zeros at (-2 + jO), (-4 + jO)


Poles at (-3 +jO), (-1 +j2), (-1 -j2)
a. Draw a pole-zero plot on a sheet of graph paper.
b. What is the character of the impulse response from the pole-zero locations?
c. Is this linear system stable or unstable? Explain your answer.
4-1S The following polynomials repreS(\nt the denominators of various transfer
functions. Determine the number of RHP poles for each function and indicate
which of the systems is stable.
a. P(s) = 4s'+ 68 + 12s~ + 148 + 4
3

b. P(8) = 2s& + 8 + 198' + 68 + 8Isl + 25s + 25


6 3

c. P(S) = 8 6 + 8' 88' + 868' + 2508 + 168


4-16 The block diagram in Fig. P4-16 represents a certain feedback control
system.
a. Obtain the closed-loop transfer function VO(s)/VR(s).
b. Use the Routh criterion to determine values for K whereby the system is
stable. Assume K can range from - 00 to + 00 •

0.2 1'0(5)
(lOs +1)(15s + 1)

Fig. P4-16

4-17 For the following transfer function list the conditions on the coefficients
such that the system is stable.
G(s) _ K(b.8 2 + b,s + bo)

. - a,8' +
a38' + a,8' + alS + ao

Assume that all a and b coefficients are positive real numbers.


RESPONSE CHARACTERISTICS

5
OF ELECTROMECHANICAL SYSTEMS

In the preceding chapters energy techniques have


been developed to aid in the formulation of equilib­
rium equations for lumped-element electrome­
chanical systems. The lumped-element restriction
means that energy-state functions can always be
formulated in terms of terminal or port coordinates.
Lagrange's equation prescribes the proper com­
bination and manipulation of the state functions
to obtain system equilibrium equations. The out­
standing feature of the Lagrange technique is the
systematic inclusion of the electrical-mechanical
interaction terms. Thus, the mechanical forces of
electrical origin and the electrical sources due to
mechanical motion are automatically included in
the equilibrium equations.
If the equilibrium equations for a system can
be written by a simple inspection of the system,
then the Lagrange technique serves only as a check.
Therefore, for a simple electrical network the loop
or node equations can usually be written from a
direct inspection of the network. The Lagrange
technique is certainly not required for such a prob­
lem. However, in electromechanical systems, the
interaction terms are usually not obvious. Here
is where Lagrange's equations are a useful formu­
lation tool.
Very often the equilibrium equations are in "1
the form of a simultaneous set of nonlinear differ­
ential equations. Analytic techniques for solving
such a set of equations are,generally not available.
Our purpose in this chapter is to take particular
forcing functions, which are of practical interest,
and obtain approximate solutions which shed some
214 light on the operation of the electromechanical
Re'ponse 0' electromechanicol,y,tems 21 S

device. Since linear constant-coefficient differential equations can easily


be solved, we shall attempt to model the actual nonlinear system by a
linear system. Such a model can be reasonably valid over only a restricted
region of operation. The conclusions obtained from such a modeling
procedure must, therefore, be applied only if the system is restricted
to the specified operating region.
The discussions in this chapter will initially center upon a particular
magnetic-field electromechanical system. The device being selected has
equilibrium equations with a sufficiently general form to give a good
illustration of the linearizing procedures and the subsequent analysis
of the linear-system model. Once these ideas are developed for this
particular system, we shall be in a position to form certain general pro­
cedures and conclusions which ean then be applied to a wide variety of
electromechanical systems. The study of nonlinear systems allows for
much more ingenuity than the routine procedures involved with linear
systems. For analytic results we shall be matching our mathematical
wits against what nature has designed.

5-1 STATIC OPERATING POINTS


The electromechanical system we shall use in developing certain analytic
techniques is the magnetic-field device discussed in Examples 3-1 and
3-5. For convenience the device is redrawn in Fig. 5-1. Assume that
the device is cylindrical and that Fig. 5-1 is a cross-sectional view. A

Iron Slotor

Moss of iron
~Gro"itY plunger" M

Fig. 5-1 A croBll-8ectional view of 8.


o
ma.gnetic-field electromech8.nical system.
216 Principles of e/ectromechonico/.energy conversion

voltage source vet) is connected to the N-turn coil. The source has a
resistance R. ohms, and the coil resistance is R. ohms. The motion of
the iron plunger, of mass M kilograms, is restricted by a spring of com­
pliance K m/newton and a viscous damper with coefficient D newton­
sec/m. A brass pole face, with an area of A square meters and a thick­
ness of d meters, prevents the plunger from forming a complete iron
path. The force of gravity is assumed to be acting in the downward
direction as shown.
In Example 3-5 we determined a Lagrangian for this system to be

,(0 O)_!M 2+ aq2 _(X-b)2


d+x
o

J..: x,x,q - 2 x 2K (5-1)

where a ;= p.oN 2A/2 and b is the position of the plunger when the spring
is at its free length. When x = b, the spring does not exert a force
either up or down. The Rayleigh dissipa.tion function and the external
force terms are respectively given by .

!J(:t,q) =~Di2 + HR. + R.W (5-2)


Q. = Mg (5-3)
Q9 = vet) (5-4)

where g is the acceleration of gravity, in meters per second squared.


By using Lagrange's equation, the Lagrangian and dissipation func­
tion, together with the external-force terms, can be manipulated to give
the following two equilibrium equations:

x - bai'
M:f + Di + + (d + X)I Mg
(5-5)
2a , . 2aiX

d + x ~ + (R. + R.h - (d + X)2 = vet) (5-6)

The details in deriving these equations are given in Example 3-5. The
current symbol i has replaced Ii simply for convenience in the two equilib­
rium equations (5-5) and (5-6). For a specified voltage function vet) we
are interested in the position of the plunger given by x(t) and the current
in the coil i(t). However, the equilibrium equations are nonlinear, since
the unknown dependent variables x and i appear raised to powers other
than unity. Also appearing are products of the variables with their
derivatives.
In order to obtain some information concerning the response of the
system, let us assume that the applied voltage is held constant at some
value Vo. Since the system contains both electrical- and mechanical­
energy-loss elements, namely, R., R., and D, the steady-state response
Respona oleledromechonicol systems 217

must also be constant. Thus, if we have


v(t) = Vo a constant (5-7)
then
x(t) = Xo (5-8)
and
i(t) = io (5-9)
where Xo and io are also constants. I Substituting Eqs. (5-7), (5-8), and
(5-9) into the equilibrium equations (5-5) and (5-6) gives us
Xo - b
---x- + (d aio•
+ xo)1 = Mg (5-10)
(R. + R.)io == Vo (5-11)
with all time derivatives of the constants Xo and io equal to zero. Equa­
tion (5-11) expreeses the simple fact that the d-c steady-state current is
proportional to the applied d-c voltage according to Ohm's law. There­
fore, if Vo is specified, the corresponding steady-state current io can
easily be determined.
On solving Eq. (5-11) for io and substituting the resultinto Eq. (5-10)
we have
Xo - b avo·
---x-- + [(d + xo)(R. + R.)l' "" Mg (5-12)

Equation (5-12) represents the steady-state operation of the system for a


constant excitation. If a d-c voltage Vo is applied to the winding shown
in Fig. 5-1, the plunger will settle at a steady-state position Xo. Values
of the plunger position Xo corresponding to certain excitations Vo are
known as operating points. A steady-state value of the coil current io
alsO has an operating-point value very simply related to Vo by Eq. (5-11).
The complete ope~ting-point characteristic for this electromechanical
system can be obtained by plotting Eq. (5-12) on an Xo versus Vo plane.
Plotting Eq. (5-12) on an Xo-Vo plane is a difficult job. An inter­
mediate plot will aid in determining the desired operating-point charac­
teristic. Let us define the following two functions:

II = Mg -
Xo - b
---x- (5-13)
and
avo! 1
II = (R. + R.)1 (d + xo)1 (5-14)
Now if we plot It and It as functions of Xo for various fixed values of Vo,
the intersection of the two curves represents a solution of Eq. (5-12).
I In all of our work the aubeeript zero will be used to designate conat&nt values of the
variables.
218 Principles of electromechanical· energy conversion

Increasing Ivol

-d 0

Fig. 5-2 Graphical !!Qlution of the d-c steady-state operating-point equation.

In Fig. 5-2 the two functions are plotted. The first function Ii> defined
by Eq. (5-13), is a straight line with slope -1/K and intercepts of
Mg + b/K and KMg + b for the vertical and horizontal axes, as shown.
The second function h, defined by Eq. (5-14), is a set of equilateral
hyperbolas about the Xo = -d line. Positive and negative values of Va
give the same two hyperbolaa, since Vo appears raised to the second power
in I:. Four typical plots of h, shown in Fig. 5-2, are drawn with four
specific values for Vo. The direction of increasing the magnitude of Vo
is also noted.
For the curves labeled with l's, we have three intersections of the h
function with the II function. These three intersections are denoted by
dots. Increasing Ivol still further to the curves labeled with 4's, we have
only the one intersection indicated by a diamond. For each value of
Ivol the corresponding values of Xo, which m~ke It = h and therefore
satisfy Eq. (5-12), can be taken directly from Fig. 5-2. Plotting these
values of Xo as a function of the magnitude of Vo gives the curve shown
in Fig. 5-3. The intersection markings shown in Fig. 5-2 are also included
in Fig. 5-3. Notice that there is a dot at each of three values of Xo:
slightly less than -d, slightly greater than -d, and just less than
K M g + I). The triangle intersections corresponding to curve 3 show
only two values of Xo, as seen in Fig. 5-3. The value of Ivol at this limiting
point is designated as v::'u.
The value of v::,ax can be found by setting dvo/dxo equal to zero.
Equivalently, we can set dvo 2 /dxo equal to zero and thus get

dvo' = (d + xoHR. + Rc)2 (2KMg _ d + 2b - 3 x o) (5-15)


dxo aK
Relponle of eledromfICftan;ca/ly~teml 219

Setting this derivative equal to zero gives us as possible solutions Xo = -d


and

Xo = -32 (K M g + b) d
- -
3
(5-16)

From an examination of Fig. 5·3 the solution at Xo = -d is not the


desired peak; therefore, Eq. (5·16) must be the value of Xo at Vo == vo·".
Substituting this value of Xo into Eq. (5-12) gives us

v...." = 2(R. + R.) (KMg + b + d)% (5-17)


o 3v'3Ka

The point on the curve designated as VOl" can be found by substitut:­


mg Xo= 0 into Eq. (5·12). After some simplification, we have
(5-18)

as the value of the applied d·c voltage making Xo = 0 in the steady·state


characteristic given by Eq. (5·12).
Let us now examine the quasi-static operation of the electromechani­
cal system as the magnitude of the applied voltage is slowly changed.
With Ivol equal to zero, the plunger is at a position Xo = KMg + b, as
indicated in Fig. 5·3. This is simply the point where the spring force
balances the gravitational force. Remember that b is the position
where the spring force equals zero, so Xo - b is the distance the spring

(KMr;+bl r--'----__

Fig. 5-3 The static-<>perating.point characteristic.


220 Prim:;p/es 0' electromecha"ica/-energy co"versio"

."..,...", lIotn

---- -­
"" Fig. 5-4 Summary of the
stable and unstable static operat­
-------------- ing points showing the jump
phenomenon.

is stretched or compressed from its free length. As the magnitude of Vo


is slowly increased, always keeping the system in the d-c steady state,
the value of Xo decreases. Referring to Fig. 5-1, a decreasing value for
.:to means that the plunger is moving to close the air gap. As Vo continues
to increase to a value of vo", the plunger continues to close the air gap
with Xo decreasing to the value i;(KMg - b) - d13, as shown in Fig. 5-3.
Increasing Vo beyond the value v:ax causes the value of .:to to jump
from the top triangle to the bottom triangle located on the bottom curve
in Fig. 5-3. Physically, however, Fig. 5-1 shows that Xo can never be
negative. Therefore, as the plunger attempts to jump to the bottom
curve, it is stopped by the brass pole face at Xo O. If Vo is increased
beyond v:"', the plunger must remain at Xo = O. Now if Vo is decreased
v:
below u , the plunger will still remain at Xo = 0 all the way down to
the value V::,iD. If Vo is decreased below VOl., the plunger will jump back
to the top portion of the characteristic. Figure 5-4 summarizes the
possible operating points for the system, including the physical constraint
of the brass pole face. The arrows on the two vertical lines indicate
the jump direction from one portion of the curve to the other. The
system never operates on any portion of the dashed characteristic.
An electromechanical relay could be constructed similarly to this
device. Contact points must be placed on the stator and plunger in
such a fashion that an electrical connection is established when the
plunger snaps closed or when Xo equals approximately zero. The snap
action that occurs when Vo exceeds the value v;:a,. is a desirable feature,
giving a clean, fast switch closure. Notice that the relay will not open
until Vo is reduced below v~in, and again a desirable snap action opening
is present. In the case of a relay, v;,a" and v::,in are respectively known
as the pickup and dropout voltages.
R..,,~m. of ./eefromechankal .y.#_. 221

Suppose we desired to use the device shown in Fig. &.1 as an electro­


mechanical actuator to position some mechanical load attached to the
plunger. If the device is working properly, the position of the load
should correspond directly to the voltage lI{t) being applied to the winding.
Assuming that 1I{t) has a very slow time variation, the characteristic in
Fig. 5-4 shows that a linear response can be approximated over only a
limited operating region on the upper portion of the response curve.
Thus, if lI(t) is varying slowly enough that the entire electromechanical
. system is very nearly in the steady state, then x(t) can be proportional
to lI(t) only if the plunger has not snapped closed.
In the next section we shall discuss the concept of stable and unstable
operating points and prove analytically that the device could never
position itself along the dashed curve from A to B in Fig. 5-4. Then
in the following section we shall discuss the operation of the system in
some limited region about a stable operating point. Along most of the
upper characteristic the operation appears to be nearly linear, or straight
line, for a slowly changing lI(t). If lI(t) were changing more rapidly
about some d-c level, would the operation for small variations still be
linear? If linear operation could be attained in this manner, then, at
least for small motions, the device would give mechanical movements
corresponding to an electrical signal. This general problem is therefore
of great interest in many physical areas. Conversely, if a voltage
linearly related to the mechanical motion of the entire system or linearly
related to the velocity or acceleration of the system could be generated,
another class of useful devices would be obtained. Position, velocity,
and acceleration pickUps are commonly called mechanical transducers.
Our analysis is to be sufficiently general to include these devices.

5-2 LINEAR INCREMENTAL DIFFERENTIAL EQUILIBRIUM EQUATIONS

VALID ABOUT AN OPERATING POINT

Figure &.3 is a plot of all the possible d-c or static steady-state operating
conditions for the electromechanical system pictured in Fig. &.1. If a
d-c voltage 110 is impre8l!led on the N-tum winding, then the plunger will
finally come to rest at a position Xo as specified by Fig. 5-3 or, equivalently.
by Eq. (5-12), from which the curve was drawn. Values of Xo less than
zero are not physically realizable, since the plunger is stopped by the
brass pole face as shown in Fig. &.1.
The time-dependent operation of the nonlinear electromechanical
system can be studied by using the linear-system techniques discull8ed in
Chap. 4 if the motion of the system is limited to the neighborhood of a
static operating point. The equilibrium equations for the electro­
222 Principles of electromechanical-energy conversion

mechaniral devire shown in Fig. 5-1 are given by (1i-5) and (5-6), which
for cOllvenience are repeated as

b 2
lift + D:i: + x ai
+ -----
(d + X)2
= Mg (5-19)

2a, + (R + R ) . 2ai:i; (t)


d +xt , t - t (d + X)2 = V (5-20)

The electrical excitation v(t) is now assumed to be composed of a constant


value Va plus a small incremental time-varying value VI(t). Thus, we
have

vet) = Va + VI(t) (,5-21)


I
where the subscript 0 indicates a d-c or operating value and the subscript 1
indicates a small, time-dependent variation in the value of the variable.
In addition, let us assume that the d-c value Va is much greater than the
time-varying incremental drive VI(t).
The response of the system to such a forcing function is assumed
to have the same general form. Therefore, we set

x(1) (5-22)
and
(5-23)

where the subscripts 0 and 1 have the significance described with Eq.
(1i-21). Here we arc just guessing that the response of the nonlinear
system to a constant forcing furwtion plus an inercmental time-varying
forcing function will be in the form of a (~onstant plus an incremental
time variation. As VI(t) is made smaller and smaller, the assumptions
of Eqs. (5-22) and (5-23) become increasingly more accurate.
Not only will earh of the response coordinates x(t) and i(t) be com­
posed of a d-c component plus a small variation but each of the terms
in the equilibrium equations will have the same general form. For
example, the mechanical force of electrical origin given by

f. (5-24)

will be romposed of a d-c component plus small variations about this


constant level. Expanding the function in a Taylor series about the
operating value mathematically formulates such a composition. A
Response of electromechanical systems 223

Taylor-series expansion for a function of two variables proceeds as


follows:

f( X,l.) = f( Xo,lo
.) + af(x,i)
!> I (-
X Xo) + af(x,i)
!> . I (.l - to
.)
uX ro,i'o ut. zo,io
2 2
+ ! ' a f(x,i)
!> 2
I . (-
X Xo
)2 +! a f(x,i) I (- )(. - .)
2'!>!> . . X Xo t to
2. uX XO,IO • uX u1. ZO,IO

+ ~ a2f(~/) I (i - i o)2 + . .. (5-25)


2. al zo.',
where the notation aj(x,i)/axlzo.io is used to mean the evaluation of the
partial derivative at the operating point (xo,i o).
By the use of Eqs. (5-22) and (.5-23), the quantities (x - xo) and
(i - io) reduce to simply XI and iI, respectively. If the time-varying
quantities XI and i l are really incremental, then the terms in the expansion
of Eq. (5-2.5) involving x - Xo = XI and i - io = i l raised to powers
greater than unity or multiplied together can be neglected without
significant error. In other words, if the values for X and i always remain
in the neighborhood of the d-c levels Xo and io, the first three terms of
the Taylor series are sufficient to describe the function f(x,i). System
operation in such a fashion is termed incremental motion about an operat­
ing point.
Let us take each of the terms in the equilibrium equations (5-19)
and (5-20) and formulate a Taylor-series approximation retaining only
the linear terms in the expansion. Starting with the mechanical force
of electrical origin, we have
ai 2 ai o2 2aio2 2aio .
f. = (d + x)2 "'" (d + XO)2 - (d +
XO)3 XI + (d + XO)2l1 (5-26)

The first term is j'(Xo,io), or the force of electrical origin evaluated at the
operating point. The coefficient in the second term is the aj./ax evalu­
ated at X = Xo and i = io. All terms involving XI and i l as products
or raised to the second and higher powers are neglected.
The first term in Eq. (5-20) can similarly be approximated by
realizing that i = io + i l ; therefore, i = 0 + i l and thus the constant
operating level for i = di/dt is io = O. The term therefore is expanded as
2ai 2aio 2aio 2a:

d +X "'" d + Xo - (d + xo)2 XI + d + Xo II

which reduces to
2ai 2a,
(5-27)
d + x"'" d + Xo II

since io = o.
224 Principles of ./ectromechanica/.energy conversion

All of the terms in Eqs. (5-19) and (5-20) can similarly be expanded
to yield the following set of approximate equilibrium equations:
- b Xl aio' 2aiot

MXI + DXI + + K + (d + Xo)! - + Xo),XI

(d
2aio .
+ (d + Xo)! ~l Mg (5-28)

d !a XO II + (R. + Re)io + (R. + Re)i l - (d :;i;op XI = Vo + VI(t)


(5-29)
By Eqs. (5-10) and (5-11) we know that these equilibrium equations are
valid at the operating points (i.e., with both Xl and i l equal to zero).
Therefore, subtracting the two d-c steady-state operating equations (5-10)
and (5-11~ from (5-28) and (5-29), respectively, we are left with

2aio
2

[ Mp
2
+ Dp + K1 - (d + xo)' .]
Xl + [2aio] .
(d + xo)! ~l = 0 (5-30)

- [ (d
2aio
+ xo)! P ] Xl + [2a (
d + Xo P + R. + Re)] ~l. = VI(t
) (5-31)

as a set of linear incremental differ~ntial equations describing the opera.­


tion of the system in the neighborhood of a constant operating level Xo
and io. Notice that the constant coefficients in these two incremental
equilibrium equations involve the values defining the operating level.

S-3 STABLE AND UNSTABLE OPERATING POINTS

A stable operating point can be defined by considering the following


experiment. Apply a d-c voltage Vo and position the plunger at a value
of Xo corresponding to some point on the curve shown in Fig. 5-3. Theo­
retically, the plunger should remain right where you placed it. Now
displace the plunger very slightly while holding Vo constant. If the
plunger returns to the original point Xo, then this point is a stable operat­
ing point. If it does not return to this point, but rather moves to a new
operating position, then the original operating point is termed unstable.
A classic example involving a right-circular cone illustrates very
clearly the difference between a stable and an unstable operating position.
If it is placed on its base, the cone will always return to this position
when tipped up slightly. Therefore, the base-down position represents
a stable operating configuration. Theoretically, the cone could also be
balanced on its pointed peak. However, a slight disturbance from this
position would cause the cone to topple over onto its side. The peak­
down position is thus an unstable operating configuration. With the
cone lying on its rounded side, we have the case of a neutral operating
configuration, neither stable nor unstable. Upon displacing the cone
slightly (by rolling it over) we find that it neither returns to the original
point nor seeks a new point. The cone just remains at the displaced
position.
The stability or instability of a particular operating point for ~he
electromechanical system of Fig. 5-1 can now be studied by using the
incremental linear equations (5-30) and (5-31). Suppose that the system
is operating with an applied voltage v = Vo. The plot shown in Fig. 5-3
gives possible values of Xo for all values of Vo. Similarly, io can be found
by dividing Va by R. + Re according to Eq. (5-11). Therefore, specifying
Va fixes the possible operating levels of x and i. How does the system
respond if V is now changed by an amount VI(t)? On solving Eqs. (5-30)
and (5-31) for Xl, we have
2ai o
o + xoP
{d
(5-32)
2a
VI d + Xo P + (R. + Re)
where
2 1 2aio' 2aio
Mp + Dp + K - (d + xoP (d + xo)'
Ao(p) = 2a
2aio
p + (R. + Re)
- (d + xo)'P Xo
(5-33)
In a similar fashion a single differential equation involving only the
incremental current response i l can be formulated.
Expanding Eq. (5-32) and taking the Laplace transform gives as a
result •
(5-34)
with
(5-35)
and
2aM
aa = d + Xo

(5-36)
2aD

at + + (R. + Re) M
= d Xo (5-37)
2a
al = K(d + xo) + (R. + R.)D
(5-38)
2ai o2
aa = (R. + Re) [ K1 - (d + xo)' ]
(5-39)
226 Principles of electromechanical-energy conversion

The response XI(s) has it~ poles determined by the denominators


of V I(S) and the initial-condition terms, and also by the factors of the
third-order polynomial 4.(s). If 1/4.(s} has poles in the RHP, then we
know that XI(t) for any VI(t) and any set of initial conditions will exponen­
tially grow with increasing time toward infinity. In reality, the response
does not actually become infinite. The system simply moves to a new
operating condition generally outside the neighborhood of the original
operating point. Therefore, the stability or instability of a particular
operating level depends entirely on the location of the poles of 1/4.(s).
The Routh array, described in Sec. 4-12, for the third-order polynomial
(.')-35) is given by
aa al
a~ ao
al -
aaao
~~-
(5-40)
az

ao

The coefficients a; and a2 are always positive numbers, since Xo can never
be negative. Therefore, we must have

al -
aaao
--
a2
>0 (5-41)
and
ao> 0 (5-42)
for all of the poles of 1/4.(11) to be located in the LHP. When we sub­
stitute Eqs. (5-36) to (.'>-39) into the stability condition given by (5-41),
the following inequality, after some simplification, results:
4a D
2
+ 2aD2(R. + R.) + DM(R. + ReP

K (Xo + d)2 d + Xo

Inequality (5-43) is satisfied for all positive values of Xo and for positive
values of the parameters.
Stability of the operating points now depends only on satisfying
condition (5-42). From Eq. (5-39) this condition is equivalent to
2aKvo2 ]~' (5-44)
Xo > [ (R. + ReP - d
upon substituting io = vo/(R. + Rc) according to (5-11). The steady­
state characteristic shown in Fig. ,1)..,3 was originalty plotted from Eq.
(.1)-12) for vo 2, and 8ubst.ituting the result ill the stability condition (5-44)

RiVes.. > [2K(d + ••)' (Mg _" ~)]" _ d


le",on.. of electromechanical .y.tem. 2'17

which, for positive values of Xo, reduces to


2 d
xo> :3 (KMg + b) - 3 (5-45)

On examining Fig. 5-3, we see that the Routh criterion verifies the fact
that only the upper portion of the characteristic represents stable operat­
ing points as indicated in Fig. 1)-4. If the system were somehow posi­
tioned at an unstable operating point along the dashed A-B curve in
Fig. 5-4, then any incremental change in Vo would cause the system to
move from that position. Every practical voltage source has some
slight magnitude variation; therefore, the system could never remain
at an operating point on the dashed A-B curve.
With a constant or d-c voltage excit,ation applied to the electrical
port of the system shown in Fig. 5-1, the mechanical member moves to a
steady-state position. In Sec. 5-1 these steady-state positions or static
operating points were determined from the equilibrium equations as
functions of the applied d-c voltage. Figure 5-3 represents a plot of
the d-c steady-state plunger position as a function of the d-c applied
voltage. However, certain operating points are found to be physically
unstable. Moving the system very slightly away from such an unstable
operating point results in a gross motion to an entirely new operating
point. In this section we mathematically formulate a test for stability
or instability of the operating points. Assuming that the system remains
in the neighborhood of the point being tested, a set of incremental linear
differential equations is derived from t.he original set of nonlinear equilib­
rium equations. The incremental equations are a reasonably good
description of the physical system so long as the system remains very
near the operating point. If the linear incremental equations predict a
bounded responscr for an incremental change in the system excitation,
then we can predict that the operating point is stable. On the other
hand, if the incremental equations predict an exponentially growing
response, then the operating point is unstable. The incremental equa­
tions serve as a tool for studying stability or instability of operating
points.

5-4 INCREMENTAL OPERATION AS A


MECHANICAL-TO-ELECTRICAL TRANSDUCER
Very often in engineering work a device which has a response linearly
related to some exeitation is required. The measurement of mechanical
position, velocity, or acceleration is often accomplished by a transducer
which produces an electrical signal linearly related to the mechanical
variable. Such a device is converting a mechanical input signal into an
228 Principles of elec:tromec:honicol·energy conversion

electrical output signal. Negligible power must be absorbed from the


mechanical system being measured, since the instrumentation device
should not in any way influence the motion of the system being measured.
Mechanical actuators and audio loudspeakers are examples of devices
producing a mechanical output from an electrical input. Quite often,
the mechanical motion must linearly follow the electrical excitation over
some specified operating region.
In this section we shall see how the linear incremental equations
can be used to design a nearly linear system from a nonlinear system.
From the nonlinear device shown in Fig. 5-1, a wide variety of linear
mechanisms can be designed. By examining the characteristics of a few
of these devices, certain general procedures are to be formulated.
The ,lectromechanica! device shown in Fig. 5-1 can actually be used
as a linear position, velocity, or acceleration transducer under certain
restricted conditions if the device is properly designed. In Fig. 5-5 the
same basic mechanism is illustrated with the motion of the plunger with
respect to the stator restricted by a spring and viscous damper. Suppose
that the device were so mounted that its position with respect to a sta­
tionary reference is given by the coordinate y, as shown in Fig. 5-5. For
simplicity, only vertical motion is to be considered. The entire device
could be fastened on some vibrating structure whose vertical position,
velocity, or acceleration is to be instrumented. Figure 5-6 shows the
transducer mounted on a machine whose vertical vibration is to be
studied.

f'm=a--Brass pole
face

Nonma\jnetic Iran
support plunger

tJ

y
Stationory
reference Fig. 5-5 A mechanical-ro-electrical
transducer.
Response of e/ectromecltonicol systems 229

Vertical motion

Transducer

Machine

Fig. 5-6 The tra.nsducer mounted


on a. vibra.ting ma.chine.

The equilibrium equations for the system are identical with those
obtained when the stator of the device is stationary with the addition
of an external force on the mechanical equation. With the stator struc­
ture stationary the only external force acting in the vertical direction
is due to the acceleration of gravity given by g. With the stat.or position
given by y, the total external acceleration acting on the plunger becomes
g + 'Ii and the total external mechanical force is M (g + ii) acting in the
downward direction. The equilibrium equations, given by (5-5) and
(5-6), are therefore modified to

.... x-b ai 2 _
Mx + Dx + ---g- ~ (d + X)2 = M(g + y) (5-46)
2a, • . 2aix
d + x ~ + (R. + R.)~ - (d + x)2 = v(t) (5-47)

where a = ".ON 2Aj2, in which A is the pole face area, N is the number of
coil turns, and "'0 = 4r X 10-7 hjm is the permeability of air. The con­
stant b is the value of x when the spring is at its free length. All the
remaining parameters are shown in Fig. 5-5.
A linear transducer can be obtained by having the nonlinear electro­
mechanical system restrained in its motion to the neighborhood of an
operating point. All time-dependent forcing functions are made to
have the form of a constant plus an incremental time-dependent varia­
tion. Therefore, in Eqs. (5-46) and (5-47) we specify that

y(t) = yo + Yl(t) (5-48)


v(t) = Vo + Vl(t) (5-49)
230 Principle. of e/edromechanical-energy conversion

with the subscript zero representing the constant steady-state operating


level and the subscript 1 representing the incremental time-dependent
variation about this operating level. The term "incremental variation"
means that YI « yo and VI «!}o. The response of the nonlinear system
to forcing functions in the form of Eqs. (.5-48) and (1)-49) is also assumed
to have the same general form; therefore,
x(t) = Xo + XI(t) (5-50)
and
= io + ilU)
i(t) (5-51)
again taking Xl « Xo and i l «i o. All time derivatives of these responses
are equal to the corresponding derivatives of the incremental portions,
since Xo aIJd io are eonstants. Therefore, we have x XIt x Xl, I = II,
and so fort,h with all higher derivatives.
As discussed in Sec . .')-2, each term in the equilibrium equations also
is assumed to have the form of a constant operating value plus an incre­
mental variation. Therefore, for a term whieh is a function of one or more
coordinates a Taylor-series expansion can be formulated. For a function
of three variables this series takes the form

!(x,y,z) = !(xo,yo,zo) + C}[('J!.,Y,z) I (Xl)


AX %0.110.1'&

+ iJ!(x,y,z) I. (y) + iJ!(x,y,z) I (Zl)


iJy :.,•. 11.... iJz I1•.•• x ••

+ higher-order terms (5-52)


with Xo, Yo, Zo as the operating point and x - Xo Xl, y - yo = yt, and
z - Zo = Zl. If Xl, YI, and Zl are very small, then the higher-order terms
involving these incremental quantities raised to a power greater than
unity can be neglected.
Using Eqs. (548) to (5-52), the equilibrium equations given by
(.5-46) and (5-47) are approximated by
2
2ai o o
f··
[ 11 Xl + D'Xl + K
Xl
Xl + XO)2 tl. ]
+ (d 2ai
xo - b aio2] _
+[ + (d + xo)2 = [MYI] + [Mg] (5-53)

[ d-~
+ II + (R. + Rc)i
Xo
l - ~a~_o
(d + xo)2
Xl] + [(R. + Rc)io]
= [Vl(l)] + [vol
The equilibrium equations must be valid wit'h the system resting at
the operating point, which means that all variables with SUbscript 1 are
equal to zero. Therefore, at the operating point the first bracketed
terms on both sides of Eqs. (5-53) and (.5-54) are equal to zero and the
ResponM of electromechanical systems 231

d-c steady-state operating equations are given by


Xo - b ai o2
+ (d + xo)2 Mg (5-55)
(R. + Rc)i o = Vo (5-56)
By subtracting Eqs. (fi-5.'» and (fi-56) from (5-53) and (5-54), respec­
tivcly, wc have
.. + D' + 1 2aio2 + 2a1o M ..
M Xl XI K Xl - (d + xo)3 Xl (d + xo)2 tl. = Yl (5-57)

'+ (R •
2a Xo 11 + R c) h. - +-Xo 0)2 XI• =
(d 2ai VI
( )
t (5-58)

For a sct of nonlincar cquilihrium cquations dcscribing a system,


two sets of equations can always be obtained. One set describes the d-c
steady-state or the static response of the system to a set of constant forc­
ing funetiollS. In See. !i-a the stability or instability of these possible
operating points was investigated. For a stable operating point the
second set of incremental equations des(~ribes the dynamic motion of the
system ill the neighhorhood of the opera\.ing point. The nature of the
system's il\(~rcmental responsc depends on thc particular operating point,
since Xo and io appcar as coefficients in the linear incremental differential
equations.
The transducer is to be operated with a constant excitation voltage;
therefore let us set Vt(t) equal to zero. For simplicity in this analysis, let
us assume that the increment.al force of electrical origin, given by the last
term on the left side of Eq. (5-57), is negligible compared with the other
forces ill (!i-!i7). The incremental current response is therefore given by
1 2aio2
Mp.2 + Dp +-K- (d +XO)3
A(p)i l = 2ai o
(5-59)
- (d + XO)2 P o
where
1 2ai o2
Mp2 + Dp + K - (d + XO)3
o
A(p)
2ai o 2a
- (d + xo)2 P Xo
p + (R. + Rc)
(5-60)
By cxpanding Eqs. (5-59) and (5-60), we have
2aioM 3
i. _ ~______________~(d__+~X~o)~2.P~______________- ,

YI [ M p2 + Dp + ~ - (/~i~o)3] [d ~ XoP + (R. + R.)]


(5-61)
232 Principles of electromechanical-energy conversion

Equation (5-61) can be put into a more convenient form as follows:


kp'
(5-62)

amp-sec'/m (5-63)

(rad/sec)% (5-64)

(5-65)
2a Lo
+ xo)(R. + Re) (5-66)
'T ". (d (d + xo)(R. + R.) R. + R.
The pteady-state operating-point equation (5-55) has been used to
eliminate io in Eqs. (5-63) and (5-64). The quality X;:'h' is defined as

xff'in = ~ (KM g + b) - ~ (5-67)

and is the value of Xo corresponding to the bulge in the steady-state


characteristic as shown in Fig. 5·3. All of the stable operating points
have Xo greater than x:,ln, as indicated by Fig. 5-4. The time constant 'T,
given by Eq. (5-66), is the ratio of L o, the inductance of the coil with the
plunger held at the operating position, to the total resistance of the
electrical circuit. Equations (5-63) to (5-66) are so arranged that these
defined parameters can easily be calculated once an operating point Xo
has been chosen. With this value of Xo the corresponding operating cur­
rent io can be found from (5-55). By using Eq. (5-56), the required Va to
obtain the operating current can then be determined.
If the machine pictured in Fig. 5-6 is undergoing a small sinusoidal
oscillation, then for YI we have
YI = v'2 Yl sin wi (5-68)
where w is the frequency of these oscillations. The responding steady­
state current in sinor form using (5-62) is given by
11 k(jw)'
(5-69)
Yl = [(jw)'lw"% + 2t(jw)/w.. + l](jTw + 1)
This complex transfer function can readily be examined by using the
magnitude-ratio portion of the Bode plots discussed in Chap. 4. Figure
5-7a shows the asymptote for this function assuming that w.. is much less
than 1/'T. Notice that the asymptote starts at a slope of +60 db/decade
and then breaks over to 20 db/decade at w". Finally, at w = 1/'T, the
asymptote has zero slope. In the frequency range between w.. and 1/'T
Ituponse of electromechanical 'yM.",. 233

I - - Velocity Position
I transducer i - - t r a n s d u c e r -
I
I
I
I

1fT

(al

Acceleration 1.
.._ _ _, Pasition
tronsducer I transducer

I
I
I
I
40 db/deCOde
I
I I
I I
1 I
I I
60 db/decode I
I I
1 I
1 I
I/T &J

(bl
Fig. 5-7 '.Asymptotic magnitude plot for the transfer function
I, k(j... jI
Y, - [(jw)'/.....• + 2r(j"')/..... + l](fr", + 1)
(a) lI.IlIIuming ..... «!,
T
(b) lI.IlIIuming ! « "''''
T

the transfer function of (5-69) can be approximated by

ill'" A (jw) (5-70)

where A is an appropriate constant. Such a complex transfer function


results from the operator form

i2 ... Ap (5-71)
YI
234 Principles of e/edromechanical·energy conversion

whieh says t.hat. i l is proportional to Pl/l = ill, which is the incremental


veloeity of the oscillating machine. Thus, a slope of· +20 db/decade
corresponds to a veloeity transducer.
Similarly, for frequencies of oseillation greater than l/T the device
becomes a position transducer. For such high frequeneies the plunger
essentially remains statiolJary and the stator is being moved sinusoidally.
The sinusoidal incremental current is simply proportional t,o the position
of the stator given hy YI.
A second possihle Bode response is shown in Fig. fi-7b. Here l/T is
designed to he mU('h less than W'" therehy giving a frequency region with
a 40-dh/deeade slope. This region corresponds to an acceleration trans­
dueer. By eareful design, the incremental induced current it has a
steady-stttc form linearly related to the sillllsoidal position velocity, or
acceleration of t.he machine oscillating over a spec~ified frequen(~y range.
This form of transducer is useful only whell the stator oscillates about
a fixed position. Therefore, the increment 1/1 is usually a periodic time
fundion eorresponding to some oscillation about the position Yo. For all
physi('ally realizable functions, a Fourier sine-cosine series for YI can be
formulated. If all the harmonies required to reasonably approximate Yl
fall in the frequeney rpgion corresponding to one of the three types of
transdU('prs (i.e., position, velocity, or acceleration), then the device will
opNatc properly for this nonsillusoidal forcing fUlwtion. For example,
suppose .111 were a 2-cps triangular wave. The Fourier series for such a
waveform eontains only odd harmonic!s whose magnitudes are propor­
tional to {l/n)2, where n is the numher of the harmonic. Since the
pipventh and all higher harmonies are leg,., than 1 per eent of the funda­
melltai, they nail he IIPgle(·ted. Th(!r!·fore, the tranHdueer Khould have a
frcqueney range from 2 to 18 ellS, (,orresponding to the fundamental
through the ninth harmonic,.
Hemember that the frequeney ranges in Fig. tl-7a and b are not so
clear-cut as they are shown. The actual magnitude characteristics
change slope in a more gradual fashion once the eorrection quantities are
added to the asymptotes. By a very careful selection of t,he operating
point and the system parameters, a design whkh will produce a useful
transdurer can be synthesized. If a voltage output is desired, an addi­
tional resistance can be added to R. in Fig. ;>-IJ, and the voltage across
thesc resistors is proportional to i I .

5-5 MAGNETIC FIELDS FROM PERMANENT MAGNETS


The t.ransdu('cr shown in Fig. fi-;; requires a constant exeitation voltage
lia to establish a static air-gap field. If it is used as a mechanical-to­
Response 01 electromechanical sYItems 23.5

Contour

Fig. !i-H A cros!I-R()ctional view of a


"ylindrit"ld mflgnetie Htru("ture ("olltnininp:
n permnnent magnet.

electri('al transdueer, the out.put of the device iii! the ineremental change
ill the winding eurrcnt. The static magneti(~ field (~an be established in a
more convenient fashion by using a permanent magnet in the magnetic
cireuit, thus eliminating the need for an external d-c voltage source.
The system shown in Fig. »-8 is a ('ross-sertional view of a cylindrical
iron frame. The cross-hatched portion of the center post contains a
eylindri('al permanent. magnet, of le'ngth 1 and cross-seetional area A.
Magneti<, flux emanates from t.he nort.h pole of the magnet and circulates
around the iron path as shown in Fig. ii-8. A small air gap of length d is
cut in the iron path. The 1('II,,d,h of the air gap is nHwh smaller than the
radius of the gap a.
With the (~ontour shown in Fig. »-8 Ampere's cireuital law can be
used to determine the magneti(~-Rux densit.y in the air gap. Applying
Ampere's circuital law to the contour shown in Fig . .5-8 gives
Had + H",l o (5-72)

where 11 car is the magnetic-intensity veetor in the air gap and H ",a. is the
magnetic-intensity veetor in the permanent magnet. The permeability
of the remaining iron ('ircuit is assumed to be large enough that H can be
taken as zero along this part of t.he' (·olltour. Neglecting fringing fields,
the total magnetic flux in the air gap will equal the total magnetic flux in
the magnet. Therefore, we have
(5-73)

wheT(' En i~ the map:llcti(·-flux dpl)sity ill the air ~aJl and E", is the mag­
netic-flux density in the magnet. The quantity 211'ab is the nominal sur­
236 Principles of electromechanical-energy conver,ion

Shearing line

Slope = _ 21rp o abl 8:"

dA

Fig. 5-9 The hyster­


esis loop for the per­
manent-magnet material.

face area of the air gap through which B. flows, and A is the cross-sec­
tional area of the magnet through which B ... flows. I
The flux density in the air gap Bo. is related to the magnetic intensity
in the air gap Ho. by the relationship
(5-74)
where 11-0 = 4r X 10- 7 him is the permeability of air. On substituting
Eqs. (5-73) and (5-74) into Eq. (5-72) we have
21f'II-()Clbl
Bm - -""-<1X" H... (5-75)

after some slight rearrangement of terms. Permanent-magnet materials


have hysteresis loops relating B... to H ..., as shown in Fig, 5-9. If the
magnetic intensity in the permanent-magnet material H ... is made very
large, the material is driven into saturation at point S in Fig. 5-9. When
H ... is now reduced to zero, a large residual magnetic-flux density remains,
as shown at point R. In fact, the residual flux density for good perma­
nent-magnetic materials is very nearly equal to the saturation flux
density.
Since an analytic expression for the complete hysteresis loop is diffi­
eult to obtain, let us determine B... and H m graphically. Equation (.5-7.')
can easily be plotted, as shown in Fig . .'),.9. The straight-line plot of Eq.
(.')-75) is known as the shearing line. The intersection of the shearing
line with the hysteresis loop at point P determines the flux density Bm
I Since the air-gap length d is much leas than the radius Il, the radial dependence of B.
is being neglected.
Respome of electromechanical .y.tem. 237

and magnetic intensity H. in the permanent magnet. The intersection


in the second quadrant, rather than that in the fourth quadrant, is being
used because the flux density Bm in Fig. 5-8 is shown as being positively
directed.
Figure 5-10 is an expanded plot of the second quadrant of Fig. 5-9.
The exact characteristic intersects the shearing line at point P. Very
often a straight line can be used to approximate the exact B...-H...
plot in the second quadrant, as shown in Fig. 5-10. The equation for
the straight-line approximation is
BB
B ... = He H ... + Bs (5-76)

where BN is approximately the residual flux density and He is approx­


imately the coercive magnetic intensity. By using Eqs. (5-75) and (5-76),
the values of B... and H ... corresponding to point P' in Fig..5-10 can be
determined. By solving Eqs. (5-75) and (5-76) for B... and substituting
the solution into Eq. (5-73), we have for the magnetic-flux density in the
air gap
B _ Bs (5-77)
.. - 211"ab/A + Bsd/fJ.Jlcl
Equation (5-77) is an approximation for the air-gap flux density in terms
of parameters of the magnetic material and the physical dimensions
shown in Fig. 5-8.

8"

Shearin\! line
Slope" _ 21Tjl.pabl
4A

Fig, 5-10 A straight­


line approximation lor the
second quadrant of the H.. ,ampere - turns 1m
hysteresis loop.
238 Principles of electromechanical-energy conversion

For many electromechanical devices employing permanent magnets,


the magnetic fields are often concentrated in a small air gap by a suitable
iron structure. By using Ampere's circuital law and the magnetization
characteristics of the permanent-magnet material, the flux densities can
be determined for phYRieal eonfiguratioTlfI differing from Fig. 5-8. The
procedure used ill this section for a spceilic (~ollfiguration is intended to
demonstrate the general techniques for obtaining the magnetie fields due
to permanent magnets ill a multitude of practical situations.

5-6 A PERMANENT-MAGNET MECHANICAL-TO­


ELECTRI CAL TRANSDUCER

By usink the basic magnetic structure discussed in Sec. 5-5 and shown in
Fig. ,1-8, a wide variety of useful electromechanical devices can be con­
struded. One such device is shown in Fig. 5-11. A concentrated circu­
lar N-turn eoil of radius a is wound 011 a eylindrical coil form. The eoil
form and the N-turn ('oil are centered in the air gap by two springs each
having a ('omplialwe 2K. The ('ombined mass of the coil form and the
N -turn ('oil is M. The position of the coil with resped to the iron frame
is designated by the coordinate x, and the position of the frame with
respeet to a stationary reference is given by the coordinate y, as shown in
Fig. ii-II. If y is a specified time-dependent fUllction, the entire frame
wil! move with respc<,t to the stationary reference. In turn, the coil
form and I'oil could have some motion with respect to the iron frame.

Fig. 5-11 A perma­


nent-ma~net mechanical
z vibration transducer.
Response of electromechanical system. 239

Thus, the coil would be moving in the air-gap magnetic field established
by the permanent magnet, and a voltage would be induced in the coil.
The induced voltage bears some relationship to the specified coordinate
yet), and therefore the device can serve as a mechanical motion transducer.
The equilibrium equations for the device shown in Fig. 5-11 can be
determined by using a Lagrange formulation. Since there is magnetic
coupling between the permanent magnet and the coil, a loop formulation
is more convenient than a nodal formulation. Thus the magnetic­
coenergy-state function is required. If a current q in the coil is defined to
be positive as shown by the dot and cross in Fig. 5-11, the total coil flux
linkages are as follows:! '

A(q,X) = Lq + (2rabB o ) &N (5-78)

where L is the self-inductance of the coil. Magnetic flux passing through


the plane of the coil from left to right contributes to positive flux linkages.
The total flux emanating from the permanent magnet goes through the
air gap, and it is equallo the air-gap flux density Be. times the cylindrical
surface area (2rab) through whieh the flux passes. The fraction of this
flux linking the N-turn coil is equal to x/b times the total flux.
The magnetic-coenergy-state function is given by

W:(q,x) = foil A/(q',x) dq' (5-79)

By substituting Eq. (5-78), we have for the magnetic coenergy


W:(q,x) = iLq2 + 2raBe.Nxtj (5-80)
with Eq. (5-77) defining the air-gap flux density Bill'
The LagrangialJ for the entire electromechanical system is

£, = ~ M(iI + x)% + ~ Lq2 + 2raB..Nxq _ X2 (5-81)

Since y is to be a known time function, only the mass M of the coil form
and N-turn coil is included in the Lagrangian. The velocity of the
mass M with respect to the stationary reference is iI +:t. The two
springs each of compliance 2K are assumed to be at their free lengths
when x = O. The x = 0 point could be taken with the coil centered
axially in the air gap.
Assume a mechanical viscous damping to exist between the coil form
and the iron frame, as shown in Fig. 5-11. The source of the viscous
damping will be discussed later in this section. Taking the resistance of

I The dot in the top circle indicates current coming out or the paper, while the cr088

in the bottom circle stands for current into the paper.


240 Principle. of electromechanical-energy conversion

the coil to be R ohms, the system Rayleigh function becomes


~ = lD:i;2 + iRq2 (ti-82)
With no external mechanical forces, Lagrange's equation for the x
coordinate is

ft [~f ] - iJ2 + iJ~ = 0 (5-83)

On substituting Eqs. (5-81) and (5-82), we have, after some slight


rearrangement,

Mf + Di; + R- 211'aB"Nq = -My (5-84)

Notice that the mechanical force of electrical origin f. is simply the total
length of the coil (21I'aN) times the coil current q times the air-gap field
B". Since f. has a negative sign on the left side of (5-84), the force is in
the positive x direction. With B" directed radially outward in the air
gap, the crOHS produet q X B is a veetor to the left in Fig. !i-1l, which is
the positive x direction.
Whcn it is used as a mechanical position transducer, the coil is
normally open-circuited and the induced voltage is measured with a high­
impedanee device. The indueed coil voltage, which is also the voltage
of meehanical origin, is given by

(5-85)

where e is taken positive when opposing the assumed direction for q in


Fig. 5-11. The expression (5-8i» is obtained by selecting the particular
term in the electrical-equilibrium equation representing the motionally
induced voltage. On substituting the Lagrangian from (5-81) we have

(5-86)

as the open-eircuit induced voltage in the coil. Observe that this voltage
is the total length of the eoil (211'aN) times the relative velocity :i; multi­
plied by the air-gap field B".
By taking q = 0, Eqs. (!i-84) and (5-86) can be used to solve for the
indueed voltage e as a function of the forcing position y(t). On eliminat­
ing x between these two expressions we have

e =- M p2 + Dp + 1/K y
(5-87)
Response of electromechanical systems 241

Fig. 5-12 Magnitude :g


plots for thc transfer func- .>­
tion -W
E 40 db I decod e
y=

showing the asymptotes


and three corrected curves. w.

which can be put into the more convenient form


(l kp2
(5-88)
li - pq~:T+2fpi "';-+ 1
with
k =2raBQNMK volt-sec 3 /m (5-89)
1
w,,~ = M K (rad/secp (5-90)

r=~~ (5-91)

The magnitude frequency-response characteristic for the transfer


relationship (5-88) is shown in Fig. 5-12. For periodic oscillations of the
transducer at frequencies higher than w" the induced voltage is linearly
related to the velocity of the transducer. By making w.. as small as pos­
sible, the usable band of frequencies, known as the bandwidth, is increased.
At very high frequencies, second-order effects not included in this analysis
will place an upper limit on the bandwidth.'
In Fig. .')-12 three corrected magnitude characteristics are shown in
addition to the asymptotes. The deeibcl corrections are taken from Fig.
4-2.') and are a function of the damping ratio r. From Fig. 4-25 the
optimum value for r is approximately 0.6, thus giving the largest possible
bandwidth. The value for r as given by Eq. (5·91) can be adjusted by
altering the viscous damping D.

1 For 0. discussion of these second-order effects see H. K. P. Neubert, "Instrument

Transducers," cho.p. 4, Oxford University Press, London, 1963.


242 Principles 01 electromechanidal-energy conversion

Fig. 5-13 Detailed view of the coil form in


the radial air-gap field.

Visci->us damping can be obtained by making the cylindrical coil form


shown in Fig. 5-11 from a conducting material such as copper or silver.
Moving the cylindrical coil form in the radial air-gap field generates an
electric field in the conducting material. A circulating current is there­
fore established in such a direction as to oppose the motion of the coil
form. Figure 1}-13 shows the coil form with radius a, thickness h, and
width b. The radial air-gap field is designated as Bo. The coil form is
similar to the N-turn coil except that it has only one turn. The voltage
induced owing to motion of the coil form is therefore identical with Eq.
(5-86) with N = 1. Thus the induced voltage in the coil form is

feoil = 27raBo x
The ohmic resistance of the coil form is simply 27rap/hb, where p is
the resistivity of the coil-form material. The force opposing the motion
of the coil form is similar to f. in Eq. (5-84); thus we have

(.5-93)

where f Il is the damping force. The term in parentheses in (5-93) is the


circulating coil-form current obtained by dividing the induced voltage by
the path resistance. Notice that the damping force is proportional to
the translational velocity X, and thus the damping coefficient D is given by

(5-94)

By altering the physical parameters of the coil form, the damping coeffi­
cient D can be adjusted to obtain the desired optimum da.mping ra.tio
r = 0.6.
Response 01 electromechanical systems 243

5-7 THE MOVING-COIL AUDIO LOUDSPEAKER

A second electromechanical device can be constructed by using the basic


ma~netie structure discussed in See. 5-!) and shown in Fig. 5-8. An
electroacoustical transducer, commonly known as an audio loudspeaker,
is shown in Fig. 5-14. A nonconducting eoil form is attached to a paper
cone. The periphery of the cone has a corrugated or rippled edge
attaehed to a rigid outer support which is stationary with respect to the
magnetic structure. The cone and the attached coil form are arranged
to support a concentrated N-turn coil in the air gap of the magnetic cir­
cuit. The permanent magnet establishes an outward radial magnetic
field Bo in the air gap. As discussed in Sec. 5-5, the magnitude of Bo
depends on the physical characteristics of the permanent-magnet mate­
rial and on the dimensions of the magnetic structure. Equation (5-77) is
an approximate expression for B a •
H a positive current is inserted into the N-turn coil, then a force of
electrical origin is exerted on the coil (and consequently on the attached coil
form) in the positive x direction. The paper cone is therefore forced to
move in accordance with the coil current. Sound is generated by the action
of the cone Oil the surrounding air. The rippled edges of the cone serve as
a compliance K opposing the motion and tending to restore the coil to the
center of the air gap. The total translational mass of the moving system
composed of the coil, the coil form, and some portion of the paper cone is

Fig. 5-14 A permanent-magnet


moving-eoil audio loudspeaker.
244 Princip/e$ of electromechanical-energy convef$;on

designated as M. As the paper cone moves in the air, a viscous friction


force opposes the motion. For simplicity let us assume the viscous
damping force to be linearly related to the translational velocity, with the
damping coefficient given by D.
The equilibrium equations for the speaker can be written by using
the Lagrange formulation. The magnetic-coenergy-state function is
identical with Eq. (5-80) for the velocity transducer discussed in Sec. 5-6.
For convenience the coenergy-state function is repeated as
W~ = lL(F + 21raB Nxq
a (5-95)
where a is the radius of the air gap, L is the self-inductance of the coil,
and Ii is the eoil currellt taken positive as shown in Fig. ;)-14. The com­
plete La¥rangian for the loudspeaker system is given by

.c !2 M:i;2 + 2.~ Lq'2


. + 21raB a Nxq' - .!-=­
2K
where x 0 is so chosen that the compliance K of the paper cone is at its
free position. The Hayleigh dissipation function is simply
(5-96)
where R is the coil resistance. The external forces on the mechanical
coordinate are zero, and let us assume a voltage source vet) is applied
as an external source on the electrical coordinate. Thus we have
Qr o (5-97)
After Lagrange's equation is used for each of the two coordinates,
the followillll; equilibriulIl equatiolls, in operator form, are obtained:

( M p2 + Dp + k) x 21raB Ni a 0 (5-98)
(21raB a Np)x + (Lp + R)i = vet)
The current symbol i has been substituted for q. By solving for i as a
function of the forcing voltage vet), we have

A(p)i(t) = (UP2 + Dp + k) vet) (5-99)


where

A(p) = (1\.f p2 + Dp + k) (Lp + R) + (21raB..N)2p (5... lO0)

Since the loudspeaker is used for generating audio tones, the fre­
quency-response characteristics are of great interest. Therefore, assume
that I!(t) is a sinusoidal time funetion represented as sinor V. The ratio
of sinor V to the responding sinor I is the complex dynamic input imped­
Re$ponse 01 electromechanical $ydems 245

ance of the speaker. By substituting jw for p in Eqs. (5-99) and (5-100)


and sinor quantities for time-dependent functions, the dynamic input
impedance can be put into the form •

~ =Z = (R + R ...) + jw(L + L ...) (5-101)

where the motional resistance is defined as


R _ (lBa)2D
.. - D2 + (l/wK - wM)2 (5-102)

and the motional inductance is defined as


L = i!:!!a)2(I/wJS - w"!t (5-103)
... D2 + (l/wK - wM)2

The length of the coill has been substituted for the quantity 2raN in Eqs.
(5-102) and (5-103).
The reactance portion of the dynamic input impedance Z given by
(5-101) is usually much smaller than the resistance portion. Thus, as a
practical approximation the current input for a given applied voltage can
be represented as

(5-104)

The current i(t) is therefore in phase with the sinusoidal voltage v(t).
For a constant V the magnitude of I as a function of frequency has the
chara(~teristi(~ shown in Fig. 5-15. Th(~ millimum current occurs at the
mechanical resonant frequency

(5-105)

which makes the motional resistance R... have its maximum value. The
current at low and high frequencies has a magnitude roughly 3 times the

Fig. 5-15 The relative magnitude of


the coil current l1.li a function of frequency
o~--~---------------.
with the rms coil voltage held constant.
246 Principles of electromechanical-energy conversion

magnitude of the current at w = wn, as shown by the relative current


scale in Fig. 5-15.
From J<~q. (5-101), the average power supplied at the electrical port
is given by
(R + R",)J2 (5-106)

where I is the rms magnitude of the coil current. The quantity RJ2
contributes to the ohmic heating of the coil. The remammg average
power input is approximately equal to the radiated acoustical power.
Thus we have

pac"u. = R.../2 (5-107)


A porti~n of R",/2 is actually not radiated as acoustical power; instead it
contributes to such other factors as heating of the surrounding air and
hysteresis losses in thc compliance of the cone. These additional loss
effects arc neglected in this analysis. I
By using the approximation for / given by Eq. (.'>-104) and the
expression for Rm given by (5-102), the radiated acoustical power given
by (13-107) can be [lut into the form

P V2(lBa)2D[D2 + (l/wK - wM)2]


= ~....-.-~~.~~--... ~- ...~~---.~-.- (5-108)
RCQ"' {R[D2 + (l/wK - WM)2] + (lB.)2DI2
For most practical loudspeakers the last term in the denominator of
Eq. (:)-108) can be neglected. Therefore the output acoustical power
becomes

p.;:OOUI!I (5-109)

which can be manipulated into the more convenient form


kw 2
(5-110)
p aoou
• = [1 - (w/w .. )2j2 + (2tw/w..)2
where we have defined

w..' = J K (rad/sec)2 (5-111)

t=~~ (5-112)

(Vl~.K)
2
k = D watt-sec2 (5-113)

I For 0. more complete discussion of the o.udio loudspeo.ker see N. W. McLa.chlan,

"Loud Speo.kers," Dover Publications, Inc., New York, 1960.


Response of eleclromechanical .ryslems 247

i I i

+20 db/decadel~ /".c::±:::<: -20 ~b/decade


I
./' ~ 1\ ~ ........
! I I i

/'"
./'
: ['-Corrected
curve ~
I
1 I ""'-..
1 r---...
I , I
I , I ~
!
I ! I ............
i I i
Wit
Frequency, rod / sec
Fig. 5-16 The output acoustical power as a. function of frequency with the rms coil
voltage held constant.

The output acoustical power as a function of frequency can be


examined by plotting the power, in decibels, as a function of the driving
frequency on a logarithmic scale. Power can be expressed in decibels
by simply taking 10 times the logarithm to the base 10 of the power
quantity. I Thus the acousti(lal power given by (5-110) expressed in
decibels becomes
10 loglo k + 20 loglo W

10 loglo {[ 1 - (:)2r + e~~y} (5-114)

The asymptotes of Eq. (5-114) can be plotted by assuming w to


be either much less than or much greater than w". For w « w" the low­
frequency asymptote is given by
Paeou.(db)
, = 10 loglo k + 20 loglo w for w« w" (5-115)
and for w » w" the high-frequency asymptote is
w
P aco".(db) = 10 loglo k + 20 loglo W - 40 loglo­
w"
for w» w" (5-116)
The low-frequency asymptote given by (5-115) has a slope of 20 db/
frequency decade, as shown in Fig. 5-16. Thus, if w increases by a factor
of 10, the power increases by 20 db. The high-frequency asympto~
given by (5-116) has a slope of - 20 db/frequency decade. The last
I The definition of the decibel unit for power is different tha.n that of the decibel

unit for a tra.nsfer relationship as defined in Cha.p. 4. Quite often the power, in
decibels, is ta.ken with respect to some reference power level P e. ThUll we would take
10 log" (P fPc) - 10 loglo P - 10 loglo Po and say that P is a certain number of
decibels above the decibels of the reference power level.
248 Principles of electromechanical-energy conversion

term contributes -40 db/decade, whieh combines with the first two
terms to give a net slope of -20 db/decade. Notice that Eqs. (.5-115)
and (.'i-llfi) give jdentical results at W Wn•

At W w" the two asymptotes have their maximum error. The


eorre(~tioll quantitics, given by Fig. 4-2.5, depend on the damping ratio r.
By properly designing the damping ratio, a relatively fiat power vs.
frequency region can be achieved as shown by the corrected curve in
Fig..5-IB. Usually a damping ratio ncar unity or critical damping is
required to obtain the fiat power region.
The frequency capabilities of the speaker depend on the magnitude
of w.. For larger values of w" a speaker can radiate acoustical power
at higher frequencies. From Eq. (5-111) we can see that Wn is increased
by decr<Jasing the mass M or the compliance of the c·one K. The low­
frequeney eapaeity of the loudspeakN largely depends on the magnitude
of the gain constant J.. defined by Eq. (ii-Ila). The gain k can be
increased by increasing the length of the coill. However, such a change
also inereases the resistanee of the coil R, whieh tends to cancel thc ehange
in l. Irwreasing the air-gap flux density Ea makes the gain k larger.
A larger permanent magnet inereascs Ba and the lower-frequeney capa­
bilities of the speaker. The remaining two parameters, the viscous
damping coefficient D and the compliance K, can each he increased by a
larger-diameter paper ('one. However, increasing the cone size increases
M and euts the high-frequeney reSI)onse but boosts the low-frequency
response. Also, increasing K cuts the high-frequeney response, while
again boosting the low-frequency response.
In summary, we have found that a speaker with good low-frequency
capabilities must have a large magnet and a large-diameter, high-com­
pliance cone. For good high-frequency response the speaker needs a
small cone with low mass and small eompliance. In order to have a
speaker system covering the entire audio range, a number of speakers
are often used. The large low-frequency speakers are called woofers,
and the small high-frequency speakers are called tweeters. Good high­
fidelity speaker systems also use medium-sized speakers known as
midrange speakers.

5-8 AN ELECTROMECHANICAL-ACCELEROMETER TRANSDUCER I


The deviee shown in Fig ..'}-17 represents an electromechanical accel­
erometer. The magnetic circuit consists of t,~o pieces, each similar to
I This accelerometer has been studied in detail by C. J. Amato, Fourth Order

Oscillations in Pulse-torqued Pendulous Accelerometers, M.S. thesis, Case Institute of


Technology, Cleveland, Ohio, June, 1964.
Response of electromechanical systems 249

Fig. 5-17 A cross-sectional view of an accelerometer transducer.

the structure shown in Fig. 5-8, placed end to end about the center line
of the device shown in Fig. 5-17. The two permanent magnets establish
a radial magnetic field B" in the air gap. Since fringing fields were
neglected in the calculation of the air-gap field given in Sec. 5-5 and
approximated by Eq. (5-77), we have the same radial field in the con­
figuration of Fig. 5-17. Each magnet is establishing B" over half of the
air-gap length, which is designated as b. A cylindrical conducting coil
form with an attached N-turn coil is mounted in the air gap, as shown in
Fig. 5-17. The co~l form is supported by four identical springs each
having a compliance equal to 4K m/newton. The coil form is eon­
strained to move only in the axial direction. The position of the coil
form with respect to the center of the air gap is designated by the coordi­
nate x in .Fig. 5-17. When x = 0 the coil form is exactly centered an4
each of the four springs is stretched a distance c meters from its free
length. A positive value for x means that the coil form has moved to
the left with respect to the magnetic structure.
The mass of the coil form and the attached N-turn coil is equal to M
kilograms. The position of the magnetic structure with respect to a
stationary reference is designated by the coordinate y in Fig. 5-17. If
the magnetic structure is accelerating to the left (that is, Ii is positive),
then the coil-form mass M will be displaced to the right with respect to
the magnetic structure, thus making x negative. The displacement of the
coil form is related to the acceleration of the transducer. This displace­
250 Principles of electromechanical-energy conversion

ment is sensed by two capacitance rings mounted axially with the coil
form. The current in the N-turn coil is then automatically controlled
to return the coil form to the x = 0 position. The current required to
restore the coil form to the x = 0 position is related to the imposed
acceleration; thus, by measuring the coil current, the acceleration ii can
be measured. Let us now study the static and dynamic operation of the
accelerometer in detail.
The conducting coil form and the two axially mounted conducting
rings are illustrated in Fig. 5-18. The coil form is supported by the four
identical springs. The two conducting rings mounted on each end of the
coil form are rigidly attached to the magnetic structure by insulated
supports, as shown in Fig. 5-17. With x = 0 the coil form is exactly
centered between the two rings and the capacitances between the coil
form ahd each of the two rings are exactly equal. If x increases in the
positive direction, the coil form moves toward the left-hand ring. The
capacitance between the left-hand ring and the coil form is therefore
greater than the capacitance between the right-hand ring and the coil
form. For small displacements of the coil form, neglecting fringing,
these two capacitances can be written in the form
(5-117)

Fig. 5-18 A detailed view of the spring-supported coil form and the two
capa.cita.nce rings.
Response 01 electromechanical system, 251

Fig. 5-19 The electrical connections for


the coil form and the two rings in a capaci­
tance bridge circuit.

where Co is the value of the two capacitances with the coil form centered
at x = 0 and a is a constant giving the farads per meter displacement of
the coil form.
The two capacitances formed by the coil form and the a~ial rings can
be put into the bridge circuit shown in Fig. 5-19. Two fixed capacitors
with a value of Co farads make up the remaining two legs of the bridge.
The frequency of the sinusoidal voltage source used to excite the bridge
is selected to be much greater than the highest significant frequency
component in x(t). The sinor output voltage Eo for a given sinor driving
voltage Ei for the circuit shown in Fig. 5-19 can be put into the form
Eo COWl - C2)
(5-118)
Ei = (C 1 + Co)(C 2 + Co)
Notice that Eo equals zero if C l C2 , or the bridge is balanced. On
substituting Eqs. (5-117) for C l and C2 into (5-118), we have

Eo a
-=-x (5-119)
Ei 2C o
assuming that (ax)2 is negligible compared with C o2.
Equation (5-119) shows that the output of the bridge is a sinusoidal
time function with the same frequency as the input and with a magnitude
determined by the position x of the coil form. For positive values of x
the output is in phase with the input, and for negative values of x the
output is 1800 out of phase with the input. The output of the bridge is an
amplitude-modulated, or a-m, time function. The signal information can
be recovered by supplying the time function to a phase-sensitive demodu­
lation circuit, which supplies a d-c output proportional to magnitude of
Eo; it is positive when Eo is in phase with Ei and negative when Eo is
1800 out of phase with E i . t The combination of the capacitance bridge,

t See, for example, C. J. Savant, Jr., "Control System Design," 2d ed., pp. 345--347.
McGraw-Hill Book Company, New York, 1964.
252 Principles of electromechanical-energy conversion

the demodulating circuit, and possibly some amplification produces an


output voltage e(t) proportional to the position of the coil form. Thus,
we have
e(t) = f3x(t) (5-120)
where f3 is a constant having units of volts per meter displacement.
The ll('cclerometer shown in Fig. 5-17 is very similar to the velocity
transducer discussed in Sec. 5-6 and shown in Fig. 5-11. Thus, the
Lagrangian for the accelerometer can be developed along lines identical
with those for the veloeity transdu(:er given by Eq. (5-81). For the
accelerometer the Lagrangian is given by

1 M( y-
£ 1= 2 + x')2 + 21 L'2q + 21raB N'
axq - ----,rr -
(x - c)2 +
(x 4Kc)2

(5-121)
where a is the radius of the coil, L is the self-inductance of the coil, and c
is the distance each spring is stretched when x = O.
The Rayleigh funetion for the system is given by
(5-122)

where D is the viscous damping coefficient opposing translational motion


of the coil form and R is the coil resistance. Viscous damping is largely
obtained by eddy currents induced in the conducting coil form moving in
the radial air-gap field. Equation (5-94) relates the damping coefficient
D to the eoil-form parameters of Fig. ,'1-11. For the eon figuration shown
in Fig. 5-17 the damping coefficient becomes
D = 41rahbB..~ (5-123)
p

where h is the thickness of the coil form and p is the resistivity, in ohm­
meters, of the coil-form material. Notice in Fig. 5-17 that b is defined as
half the length of the coil form.
The externally applied forces are given by

(5-124)

where e(t) is as given by Eq. (5-120). The external force in the electrical
equation Qq must have a negative sign so that the coil current tends to
restore the coil form to the x = 0 position. Equation (5-120) has e(t)
positive if x goes positive. By examining Fig. 5-17, we can see that a
negative current (opposite to the dots and crosses) causes a force of
electrical origin in the negative x direction, and thus e(t) must be negative.
Response of electromechanical systems 253

By using Eqs. (5-121), (5-122), and (5-124) in Lagrange's equation,


the equilibrium equations for the system can be put into the following
form:

( M p2+ Dp + ~) x - lBoi = - M p2y


(5-125)
(lBop)x + (Lp + R)i = -e(t)
where the length of the coill is equal to 27raN and i has been substituted
for the coil current q. Figure 5-20a shows a block diagram for Eqs.
(5-12.''». Equation (.'i-120), expressing e(t) as a function of x(t), is also
included in Fig. 5-20a. The block diagram shown in Fig. 5-20a can be
reduced to the form shown in Fig. 5-20b, where the following quantities
have been defined:

M
kl = ­ amp-sec 2 /m (5-126)
lBo
k2 = lB.KfJ (5-127)
R
lBo
TI = If seconds (5-128)
L
T2 = Ii seconds (5-129)
2 _ 1
w" - MK (rad/sec) 2 (5-130)

r=Q~
2 M
(5-131)

The static operation of the accelerometer can be studied by assuming


a eonstant input acceleration. The d-c steady-state eoil current is
obtained by setting all p's in the block diagram of Fig. 5-20b equal to zero.
The input quantity p2y is not zero, since the acceleration has been taken
to be constant. l The d-c steady-state response current is

(5-132)

upon reducing the d-c steady-state block diagram formed by taking p = 0


in Fig. 5-20b. Since both kl and k2, defined by Eqs. (5-126) a·nd (5-127),
are constants, Eq. (5-132) shows us that the static output current is
proportional to the constant acceleration (p2 y )o.

I Simply supporting the accelerometer vertically in a gravitational field applies a

constant acceleration input equal to g - 9.8 m/scc·.

~- ------- - - - - - - - -
254 Principles of electromechanical-energy conversion

la)

(0)
Fig. .')...20 (a) Overall block diagram for the closed-loop accelerometer transducer.
(0) Simplified block diagram.

By substituting Eqs. (5-126) and (5-127) into (5-132), the d-c steady­
state transfer function can be put into the form
in MfJ
(p2y)O - R/K + lB.fJ (5-133)

If the quantity lB,.{J can be made much greater than R/K, then Eq.
(5-133) reduces to
in M (5-134)
(p2y)o "'" lB..

which is independent of the coil resistance R, the spring compliance K, and


the capacitance bridge coefficient /3. t The total mass of the coil form M,
the length of the coil l, and the air-gap field Bo should remain constant.
The use of feedback allows the operation of the transducer to be inde­
pendent of all the system parameters which are subject to change owing
. to environmental and aging factors, such as the spring compliance,
amplifier and demodulator gains, and the coil resistance.
The overall transient response of the device ·determines the frequency
limitations or bandwidth of the transducer. From Fig. 5-20b the closed-

t Remember that fJ is the net coefficient for the bridge, phase-sensitive demodulating
circuit, and an output amplifier.
Response of electromechanical systems 255

loop transfer function for the transducer is given by


l(s) k lk 2("ls + 1)
(5-135)
f(s) = (s2/",,,2 + 2fsl",,,+ 1)(r2s+ 1) + k 2(TIS + 1)
where f(s) is the Laplace transform of the system input p2y(t). Taking
all initial conditions to be equal to zero, the Laplace transform of Fig.
S-20b is taken by simply substituting the Laplacian variable s for the
differential operator p, as was discussed in Sec. 4-9. The denominator of
Eq. (5-135) can be expanded into the form
Closed-loop denominator = aas 3 + a2s2 + al8 + ao (5-136)
where

(5-137)

(5-138)

(5-139)
(5-140)
All the quantities in Eqs. (fi-137) to (fi-140) are defined by Eqs. (5-127)
to (5-131).
The Routh array for a third-order polynomial is given by (5-40),
Since all the polynomial coefficients are positive, the only criterion for
stability is

• for a stable system (5-141)

If the inequality given by (5-141) is not satisfied, then the closed-loop


transfer function (5-135) has two RHP poles and the system is unstable.
If the transducer is unstable, then the d-c steady-state response given by
(5-132) is not correct and satisfactory operation is not possible.
On substituting Eqs. (5-137) to (5-140) into (5-141), after some
manipulation, we have

2f"'n"2
- k2 - ( "2 + -w,..
2f + - -12 + k2'1'1 ) >
'TsWn
("2 - "1
)

for a stable system (5-142)

Notice that for n :> T2 the right side of (5·142) is negative and the
inequality is always satisfied, since all parameters are positive. From
Eqs. (5-128) and (5-129), this sufficient, but not necessary, requirement
256 Principles of electromechanical-energy conversion

means that
lBa L
~>­ sufficient for stability (5-143)
fJ R
Increasing the magnitude of the air-gap field Ba tends to stabilize the
system. Increasing the coil length 1 by adding more turns will not help
stability, since the self-inductance L (which is proportional to the square
of the number of turns) divided by the resistance R (which is proportional
to the number of turns) increases at the same rate as l. A higher­
capacitance bridge coefficient fJ causes instability. Using a smaller­
diameter coil wire increases the coil resistance R and thereby aids stability.
By examining (.')-142), we can see that increasing the damping ratio
r makes tfe left side larger and helps satisfy the inequality. Perhaps the
best. way to stabilize the system is to increase the eddy-current damping
in the coil form. From Eq. (5-123) we have the factors influencing the
damping coefficient D appearing in the expression (5-131) for the damping
ratio r. Remember that the static operation becomes independent of all
system parameters which are likely to change value if the quantity
lBafJ» R/K, as discussed with Eq. (.')-133). Increasing the air-gap
field Ba not only satisfies this condition but also increases the damping
coefficient D and thus aids stability. The design of a stable system, the
parameters of which adequately satisfy (5-142), is an involved process
requiring a large amount of judgment and experience.
Assuming that the system is stable, let us examine the closed-loop
frequeney response. A substitution of jw for 8 in the transfer function of
Eq. (5-135) yields the phasor relationship between sinor I(jw) and sinor
"i(jw) as given by
I(jw)
Y(jw) = [1 - (w/w n}2 + j(nw/wn)](jT2w + 1) + k2(jTIW + 1)
(5-144)

A Bode plot for Eq. (5-144) cannot easily be obtained without factoring
the third-order denominator. In general, the factoring of a third-order
polynomial is a difficult job. The subject of feedback systems develops
numerous techniques for adjusting the closed-loop system response by
dealing primarily with the open-loop transfer functions. The open-loop
transfer functions usually appear in factored form, so the conventional
Bode plots and polar plots can easily be drawn. Also, the locations of the
open-loop poles and zeros are known.
An adequate discussion of these feedback-system techniques is
beyond the scope of this chapter. The accelerometer transducer serves
to illustrate the very difficult and intriguing problems faced by the system
designer. The static system equations show us that, by making one
Response of electromechanical system, 251

group of parameters much larger than another group, the system becomes
insensitive to parameter variations. On the other hand, the dynamic
equations show us that the system can become unstable if only the static
requirements are met. In addition to these two factors, a good overall
closed-loop response must be obtained by the designer, who can work
conveniently only with the open-loop transfer "functions. The subject of
feedback theory or control theory develops detailed techniques for solving
these problems.

5-9 SUMMARY

By using an energy-state-function formulation, the equilibrium equations


for electromechanical systems can be systematically derived. The d-c
steady-state, or static, equations can be readily obtained from the general
set of nonlinear equations by setting all time variations of the coordinates
equal to zero. For many physical systems certain ranges of static
equilibrium points or d-c operating points are found to be unstable. With
a system positioned at an unstable operating point, any slight external
disturbance causes the system to move to an entirely different operating
point. In the case of a stable operating point the slight external dis­
turbance results in a motion of the system in the neighborhood of the
'operating point, with the system finally settling back to the original point
when the disturbance is removed.
By using a Taylor-series expansion for each term in the nonlinear
equilibrium equations, a set of linear incremental equations valid in a
small region about the operating point can be determined. The linear
incremental equ/lotions provide a sufficient test for the stability or insta­
bility of the operating point about which the expansion was made. In
addition, the linear incremental equations define a linear operating mode
over a specified operating region. Thus, a nonlinear electromechanical
device can be used as a linear transducer over a limited operating region
about a stable operating point.
A large variety of practical electromechanical transducers employ
permanent magnets to obtain a static air-gap field. In Sec. 5-5 a basic
eylindrical structure containing a permanent magnet is studied in order
to establish analysis teehniques for such a material. By using this one
basic structure, three different transducers are studied in detail; the first is
a velocity transducer which provides an electrical voltage linearly related
to translational velocity. The frequency response for the device illus­
trates how the magnitude of the parameters serves to influence the band­
width of the transducer. The second transducer, studied in Sec. 5-7, is
an audio loudspeaker converting electrical signals into sound. Our
258 Princ:iple. of electromechanical·energy conversion

interest in the loudspeaker centers on the acoustical power-frequency


spectrum. The factors influencing the limited bandwidth of the speaker
are discussed, and the need for multiple speakers with different parameters
in order to cover the entire audio range becomes evident.
The last device considered in the chapter is the accelerometer employ­
ing feedback in its operation. By increasing the magnitude of specific
parameters with respect to other parameters, the feedback allows the
system to become completely insensitive to most parameter variations.
Thus, the linear relationship between the acceleration being measured
and the output-current variable can be made largely independent of the
spring parameters, amplifier gains, etc., by using a feedback signal. How­
ever, we also find that system stability becomes a critical factor in a feed­
back system. Thus, by grossly improving the static or d-c response, we
run the risk of having a completely unacceptable transient response, which
in an extreme case becomes unstable. Feedback-control-system theory is
concerned with numerous techniques for obtaining an optimum design or
compromise between these opposing situations.
The main purpose of this chapter has been to illustrate how an
analytic approach to an electromechanical system can greatly aid in
establishing design guidelines. A careful paper-and-pencil (and even
computer) study of a proposed system preceding, during, and after actual
construction is of tremendous value in determining theoretical design
limitations and in optimizing system operations. The ultimate objective
of an engineering design is a practical piece of hardware. The construc­
tion of the device without a careful analytic study is usually found to
consume more time and money than if an analysis is at least attempted.
A practical design problem is best solved by a careful blending of the
"let's build it" people with the "let's analyze it" people.

PR08LEMS

5-1 A system is described, for 11.\1 t <?: 0, by the following two nonlinear dif­
ferential eqUl!.tions:
5x + 8x + x1y' - A(t) l2g + 20y + 4yx - B(t)

where
A(t) - 128 B(t) - 32
G. Find the steadY-f!tate values of x and y for the constant forcing functions A
and B.
b. The term x'y' ca.n be linearized into the following form:
Response 01 electromechanical systems 259

Find the value of a about the operating point:to = 3, yo = 2.


c. Find b at this same operating point.
d. For the equilibrium equations given above write a set of linearized incremental
equations about a general operating point X., yo.
5-2 By using a Taylor-series expansion, find a linear approximation to each of
the following functions valid in the neighborhood of a static operating point. Use the
subscript 0 for the operating point and the subscript 1 for the increments.
a. f(x,y) - 6x'Y' + 5x'y + 7y
b. f(x,8) - 4e-" sin 38
c. f(x,y,i) - 2e-~·'·
5-3 A particulnr singly excited magnetic-field electromechanical force trans­
ducer has the characteristic equation A ~ 60xi, where A is the flux linkage, X is the
position of the movahle mechanical memher, and i is the coil current.
a. Find an expression for the magnitude of the force of electrical origin f. as a
function of i.
b. Find a linearized approximation for f. valid in the neighborhood of i. = 60 rna.
c. Evaluate f. for i - 63 rna (thus i. = 60 rna and i. - 3 mil.) from the exact
expression found in part a and then from the approximation in part b. What is the
percentage error?
d. Find a value for the operating current i. such that the linear incremental force
coefficient equals 1.2 newtons/amp. (Thus if f. - f •• + f.1 and f .. - Ail, find i.
such that A - 1.2 newtons/amp.)
Notice that the nonlinear transducer can operate as a linear force transducer
for small increments in the exciting current.
5-4 The circuit shown in Fig. P5-4 contains a linear inductance L and a non­
linear resistance R. The resistance of R is related to its terminal current by the
relationship R - Roe"', where R. and a are constants. If the excitation voltage
vet) - v. + VI(t), where IVII «vo, find:
a. The d-c steady-state equilibrium equation for the system.
b. A time-dependent incremental linear equilibrium equation valid about the
operating condition of part a.
c. From the equation in part b determine the stability of all possible operating
points. •

Fig. P5-4

5-5 The simple pendulum shown in Fig. P5-5 consists of a stiff massless rod
of length 1connected to a point mass M. The pivot point is frictionless, but Ii viscous
friction torque proportional to the angular velocity 0 of the pendulum opposes the
motion.
a. By using Lagrange's equation show that the equilibrium equation for the
system is MI.(J + DO + MgI sin 8 - 0, where D is the viscous friction coefficient
and g is the acceleration of gravity.
260 Principle. of electromechanical· energy conyer.ion

b. By letting 6(1) -
(10, find the system's two static equilibrium positions.
c. Find a linear incremental equation involving (I,(t) which is valid about any
operating point (10_
d. By using the incremental equation found in part c, prove that one of the static
operating points found in part b is stable and the other is unstable.

I
I
I
I
I
/
/
/

Fig. P5-5

5-6 The three-port torque transducer shown in Fig. P5-6 has the following
parameter values:
Rotor self-inductance Li" - 25 mh
Rotor resistance R;r - 5 ohms
Rotor-to-stator mutual inductance M" - 0 cos'" henrys
Rotor moment of inertia J - 0.01 kg-m*
Rotor viscous friction coefficient D, - 0.5 newton-m-sec/rad
Spring compliance K, - 0.02 rad/newton-m

gr J
~-1
\ '

Fig. P5-6
Response of electromechanical systems 261

The spring is at its free length when fI' .. O. The stator winding is excited by a 5-amp
constant-current source. The rotor winding is excited by a voltage source such that
v'O) .. vo + v~(t)
and the mechanical port has an applied torque
Tr(t) = T~ + T~(t)

a. Write a Lagrangian for the system using a loop formulation on the rotor
winding.
b. By using Lagrange's equation, find the equilibrium equations for the rotor
winding and the mechanical port.
c. Separate the two equations in part b into a set of d-c steady-state equations
and a set of linear incremental different.ial equl\tions.
d. If the d-c rotor voltage v; .. 50 volts and the constant applied rotor torque
T~ "" 89.4 newton-m, find the operating-point values for the rotor current i; and the
rotor position 1/;.
e. By substituting the operating-point values from part d into the linear incre­
mental equations found in part c, draw a linear incremental block diagram for the
system. Show vW) and Tr(!) as inputs and 11';(1) as the response of interest. Reduce
the diagram to its simplest form.
f. Draw a Bode diagram for Tj(r) "" A sin wt and v;(t) .. O. The IT;(t)1 « T;.
g. Repeat part f for v;(t) = B sin wt and T~(t) = O. The Iv;(t)1 «v;.

h.. If v;(t) .. 5u(t) and T~(t) .. 0, find 1I'(t) for all t ;::: O.

i. If T;( t) - 5u(t) and v;W ... 0, find fir( t) for all t ;::: O.
5-7 The permanent-magnet vibration transducer shown in Fig. 5-11 has been
designed with the following specifications:

Coil form (see Fig. 5-13)


Radius a - 3 cm
Height b - 3 cm
Thickness h. - 1 mm
Cylindrical peffl,lanent magnet
Length I - 1.125 em
Radius r - 2 em
Residual flux density /J R .. O.!! wcber/m'
Coercive magnetic intensity He .. 0.8 X 10' amp-turns/m

Sta.f.or iron structure


Air-gap length d - 5 mm
Air-gap width b - 3 crn
Compliance of each spring 2K .. 0.024 m/ncwton
The coil form is constructcd of copper. A 10(}'turn winding of 23-gauge copper
wire is wound uniformly on the coil form.
Note: The density and conductivity of copper are given in the table of physical
constants shown in the Appendix. The diameter of 23-gauge wire is 0.573 mm.

a. Calculate the magnetic-flux density in the air gap B. established by the


permanent magnet.
262 Principle, of e/ectromechanica/·energy conver,ion

b. Calculate the total m&S8 of the moving system consisting of the coil form and
the 100-turn winding. AB8ume the springs to be m&S8IeB8.
c. Find the viscous damping coefficient D resulting from motion of the copper
coil form in the air-gap magnetic field.
d. Determine the complete transfer function between the open-circuit coil
voltage and the velocity of the stator frame with respect to a stationary reference.
See Eq. (5-88).
e. Plot the frequency response of the device and indicate the lowest frequency
over which it can be used &8 a velocity transducer.
,. What would be the effect of using an aluminum rather than a copper coil form?
5-8 The &ccelerometer transducer shown in Fig. 5-17 and discuB8ed in Sec. 5-8
haa the following design parameters:
Coillorm
Materi41 is aluminum

Radius - 1 em

Length - 1 em

ThickneB8 - 1 mm

Spring.

Four identical ma.ssleB8springs each having a compliance - 1.38 X lO-' m/newton

Coil
Material is 4O-gauge copper wire with a diameter of 0.08 mm

Number of turns - 100

Self-inductance - 20 jl.h

Pomw1'/. /ramduur
Gain constant of cap&citance bridge, demodulating circuit, and amplifier
(J - 6 X 10' volts/m

The two permanent magnets establish a radial air-gap !lux density of 10,000
gausa.
a. Calculate the resistance of the tOO-tum coil and the total m&S8 of the moving
system consisting of the coil form and the tOO-turn coil.
b. Draw the complete block diagram for the system in the form of Fig. 5-20a.
Substitute parameter values. Reduce this diagram to the form of Fig. 5-20b.
c. For a stea.dy-state &cceleration of 109, where g is the acceleration of gravity,
find the steady-state coil current.
d. Plot the open-loop frequency response between 1(j",) and k.V(j",) in the form of
a Bode plot.
e. On the sa.me graph, plot the same frequency response, taking TJ and Tl to be
negligible. DiscU88 the proposal that this second plot re8.llOnably approximates the
first.
I. Taking T, and Tl to be equal to zero, plot the closed-loop frequency response
between l(j",) and k,Y(j",) on the same graph. What is the bandwidth of the closed­
loop &ccelerometer? (Note: The bandwidth is the range of frequencies over which the
device approximates a constant-gain &ccelerometer by not attenuating its low-frequency
gain by 1eB8 than 0.7.)
g. Use the approximations in part 1 to find the closed-loop response i(t) to a 109
step change in the input &cceleration.
11.. Prove that the complete system (with Tl and TI not zero) is stable.
Response 01 electromechanical qstemll 263

5-9 A magnetic relay is designed as shown in Fig. P5-9. The stack height of the
iron laminations is 0.25 in. The other physical dimensions are as shown. Each of
the two series-connected coils has 12,020 turns of wire having a resistance of 2250 ohms.
The linear massle88 spring has a compliance K - 2.26 X lO-1 in./lb, and it is at
its free length when % - 0.011 in. Assume that the iron has infinite permeability
and neglect all leakage effects. The closed air-gap length is 0.001 in. owing to
imperfections in the pole faces and armature surfaces.
a. Derive the d-e steady-state equilibrium equations for the relay.
b. Sketch the steady-state position of the armature %0 as a function of the d-e
applied voltage Vo. Be sure to label the pickup and dropout voltages. What is the
minimum air-gap length with the relay contacts still open?
c. For Vo - 20 volts calculate the two steady-state air-gap lengths.
d. Develop a set of incremental linear differential equilibrium equations valid
about each of the operating points in part c and thereby prove one of the points to be
stable and the other unstable.

/l' Contocts

Stock height = V.."

Fig. P5-9 vI t I

5-10 The purI10se of this problem is to demonstrate the effect of considering the
iron structure in Prob. 5-9 to have a noninfinite permeability. A steady-atate volt­
age of 32 volts is applied to the two series windings of the relay in Fig. P5-9.
a. Calculate the stea.dy-state current in the windings.
b. With the armature held at its free position (i.e., the air-gap length is 0.0l1 in.),
find the force of electrical origin, assuming the iron to have infinite permeability.
c. Repeat part b with the armature in its closed position and the air-gap length
equal to 0.001 in.
d. Repeat the calculations of parts band c with the assumption that the iron hall
the following magnetization curve:
H., amp-tums/m o 80 249 478 800 2490 4780
B, webers/m. o 0.76 1.14 1.27 1.33 1.46 1.56
Hint: ExpreSB the force of electrical origin in terms of the flux density B. The
flux density can be determined by using Ampere's circuital law, which for a llon­
infinite permeability in the iron becomes
!!. 2%0 + ZH. - 2Nio
fIoO
264 Principle. of electromechanical-energy conve,..;on

where l is the mean path length through the iron circuit, Hi is the magnetic intensity
in the iron, and N is the number of turns on each coil. By Bolving this equation for B,
an appropriate load line can be plotted on the B versus Hi magnetization curve, with
the intersection giving the flux density B. The corresponding force of electrical origin
can now be found.
e. Compare the answers in part d with those obtained in parts band c. When is
the aasumption that the iron has infinite permeability valid?
5-11 Figure P5-11a shows the essential features of a circular electrostatic audio
loudspeaker. The speaker consists of two conducting plates separated by an elastic
dielectric material. The left plate in Fig. P5-l1a is stationary. When a voltage is
applied across the two plates, the right plate moves toward the stationary plate. If
the motion of the right plate follows the applied signal voltage, the device will act as a
loudspeaker converting the electrical signal to a corresponding acoustical output.
In order for the device to be a linear transducer, the signal voltage Vl(t) must be much
lC88 than a.p. applied d-c bias voltage Vo. The speaker has the following design
parameters:
Radius of the plates and dielectric - 10 cm
Frequency range: 3000 to 12,000 cps
Mechanical viscous damping coefficient D - 40 newton-sec/m
Permittivity of the dielectric material .... 2 X 10-1 farad/m
Young's modulus for the dielectric material Y - 5 X 10' newtons/m'
Effective maas per unit area of the right plate and the dielectric - 18.1 g/m'
Thickness of the dielectric with zero applied voltage d - 5 X 10-' m
Voltage-eource resistance R - 6000 ohms
a. Indicate the physieal origin of each of the lumped elements in the l\1mped­
linear model shown in Fig. P5-11b.
b. Find the effective mass M of the moving system. Show that the compliance
K - d/ A Y, where A is the area of the elastic dielectric. Find the value of K.

Fi~ed conducting plote


Spring torce '" 0 when x = d =5 X 10- 4 m

j)
R R
+ +
v( t) vi tl
(al Ibl
Fig. P5-11
Response 01 electromechanical sys'ems 265

c. Use Lagrange's equation to find the two equilibrium equations for the system
using a loop formulation for the electrical portion of the system. Separate these
equations into a set of static equations and a set of incremental linear differential
equations.
d. From the static equilibrium equations sketch the position of the system Xo as a
function of the d-c applied voltage v.. This curve is similar to Fig. 5-3. Calculate
v::'"' and the value of Xo at v:;'''. Indicate stable and unstable operating regions on the
X.,..Vo curve.
6. From the linear incremental equations, about an operating bias voltage
Vo - 3750 volts, find the sinusoidal steady-state input impedance Z - V1(j... )/II (j...),
where il(t) - q,(t). Arrange Z into the form
1 1
Z - R + R .. +.,-C
J"'.
+.,--C
J"' ..
where C• .. fA/xo. Find numerical values for R.., COl and C.. and thereby ehow that
Z "" R over the speaker's frequency range. .
f. From the results of part e show that the radiated acoustical power is given
approximately by P '" V,'R../R'. Find the radiated acoustical power for an rms
signal voltage V I - 250 volts.
g. With VI - 250 volts find the total input audio power, and UIIe the results of
part f to find the efficiency of the speaker.
chaptar

CONSTRUCTION OF THE

6 PRIMITIVE MACHINE

Let us now turn our attention to a different class of


electromechanical-energy-conversion device known
as electrical machines. Acting as generators, these
machines supply upwards of 99.99 per cent of all
our electric power. As motors they are one of the
most widely used sources of mechanical power.
The subject of electrical machinery encompasses a
large number of seemingly different electromechani­
cal devices. In the past an analysis technique has
been developed for each device. By using the
proper technique, the response of each machine
could be predicted in terms of certain parameters
and the applied excitations. A serious objection to
such a procedure is the requirement of learning a
new trick each time a machine with a different
name is encountered. Such a procedure corre­
sponds to learning the specific equations for a vast
number of electric circuits and then trying to apply
this bag of equations to a given circuit of interest.
Certainly, such a task is ridiculous and thus, by
analogy, the memorizing of a bag of techniques each
applicable to a specific electrical machine is also
foolish.
In this chapter a single model machine, often
called the primitive machine, is to be analytically
constructed. The construction of the primitive
machine is carefully arranged to allow for a complete
mathematical description. For our model machine
we shall be able to write equilibrium equations
completely. Unfortunately, these differential equa­
tions are nonlinear. Therefore, in the general case,
exact solutions cannot be obtained. However,
by using the incremental linea.riza.tion techniques
from the preceding chapter, a great deal of useful
266 response information can be obtained.
Co"lfrvdion of the primitiv. machi". 267

As formulated, our primitive machine does not exactly correspond to


any actual electrical machine. The primitive machine contains a switch­
ing mechanism known as a commutator, and therefore it would be similar
to a practical d-c machine in appearance. However, there are certain
construction details which would cause an error in our analysis of d-c
machines if a one-to-one correspondence between the primitive and actual
machine were used. Because of the construction similarity the primitive
machine can act as a powerful tool in the analysis of d-c machines, but
certain slight modifications have to be included in the analysis to give
practical useful results.
A-c induction and synchronous machines can also be analyzed by
using the primitive machine as a modeL These a-c machines are not con­
structed with the commutator mechanism. Our primitive machine must
be transformed to an entirely equivalent form more closely matching the
construction of the a-c machines before an analysis procedure can be out­
lined. Even the transformed model of the primitive machine does not
exactly match the actual physical a-c machine. As in the case of d-c
machines small adjustments are required to make the reSUlting analysis
more closely correspond to the actual machine of interest.

6-1 MAGNETIC STRUCTURE OF THE PRIMITIVE MACHINE


The stationary portion of the machine is called the stator, and the movable,
or rotating, portion is called the rotor. TLe air space between the stator
and rotor is termed the air gap. In many practical machines the air gap
is nonuniform owing to pole projections on either the stator or the rotor
of the machine. "for example, Fig. 6-1 illustrates a cross-sectional view
of a machine having a cylindrical rotor of radius a and a stator having
two pole pieces projecting from its inner periphery. The air-gap distance
g is not a constant. The term "salient" is used to describe ,a machine
with a nonuniform air gap. Saliency can also result from a smooth sta­
tor surface and a noncylindrical rotor surface. A salient machine is one
having a nonuniform air gap, regardless of whether the stator or rotor
has the pole projecti'>ns.
In our primitive machine only the stator will have a nonuniform
inner surface. The rotor is to be a perfect cylinder. If the actual machine
of interest should have the reversed condition, then our modeling process
will use the stator of the primitive machine to represent the actual machine
rotor and the primitive-machine rotor to represent the stator of the actual
machine. This inversion process will be clarified when such a case is
considered in Chap. 11 with the synchronous machine.
A cartesian-coordinate system is drawn through the machine shown
268 Principle, of electromechanical·energy converaion

. Quadrature oxis

Direct axis

Slotor

Fig. 6-1 CrolJ8-8eCtional view of a machine having a two-pole salient stator


structure.

in Fig. 6-1. The horizontal axis is denoted as the direct axis, and the
vertical axis is termed the quadrature axis. The symbols d and q are
respectively to be used for designating these two axes. Owing to the
machine geometry, a cylindrical-coordinate system is found to be more
convenient. For this purpose the polar angle (J is defined in Fig. 6-1 as
being zero along the positive direct axis and (J is taken positive in the
counterclockwise direction. The radial cylindrical coordinate r is
measured from the center of the machine, and the axial coordinate z is
positive in the direction out of the paper in Fig. 6-1. Using a right-hand
cylindrical-coordinate system, we have

ar X a, = a. (6-1)

where a represents the respective unit vectors.


Since our attention will be focused mainly on the air gap and the
rotor and stator surfaces adjacent to the air gap, a drawing of the machine,
more convenient than Fig. 6-1, is to be used. By cutting the machine at
(J = 0 and flattening the surfaces on either side of the air gap, a so-called
developed view of the machine is formed. Figure 6-2 is such a representa­
tion. The unit vector a, is directed along the horizontal axis, and the unit
radial vector a r , originally oriented from the rotor to the stator, is now
directed dpwnward in Fig. 6-2. The unit axial vector a. must therefore
be directed out of the paper, as seen by using Eq. (6-1) for a right-hand
coordinate system.
Con.'rvetion of Ih. primitive machine 269

Rotor

4,

~O~~--~~~==~==~~~~===f--oo
Stotor

Fig. 6-2 A developed view of the air gap of the salient two-pole machine.

Designating the air-gap length in Fig. 6-2 as g, we have the following


relationships

g = gtl. for - ! < 8 <! and


4 4
3,.. < 8 < 5r
4 4
(6-2)
for! < 8 < 3,.. and 5r < 8 < 7...
4 4 4 4
The air-gap length is a square-wave function of the angle 8 completing
two cycles as 8 goes from zero to 2,... When we expand the air-gap length
g in a Fourier series, we have as a result!

g(8) = f!o+ g~ - gq~-.f!!! ~ [cos (28) - Jr cos 3(28)


2 2 'If' ~
+ k cos 5(28} + ...J (6-3)

Only cosine terms appear because of the even symmetry. The argu­
ments of the cosine terms contain 28 owing to the fact that g(8) completes
two cycles as 8 proceeds from zero to 2'1f'. For air gaps which are not
square wave in form we can still develop a Fourier-series expansion in the
general form of (6-3). This conclusion is based on the physical fact that
the air gap must at least be periodic and progress through an even number
of cycles as 8 goes from zero to 2'1f'.
Retaining only the fundamental term in the Fourier-series expansion,
we have
g(8) "" go - gl cos 28 (6-4)

where for the square-wave variation expanded in Eq. (6-3) we have defined

gq + gd
go = --2- the average gap length (6-5)

I For an excellent review of this Fourier-series expansion see W. R. LePage,

"Analysis of Alternating-current Circuits," pp. 376-377, McGraw-Hill Book Company,


New York, 1952.
270 Principles of electromechanical-energy conversion

and
(6-6)

Since gl is usually much less than go, retaining only the first term in the
series of (6-3) is a reasonably good approximation. The higher harmonics
contain a coefficient lin, where n is the order of the harmonic; therefore,
for larger values of n these terms are less significant.
For a periodic air gap differing from a square wave, go and gl will not
be given by (6-5) and (6-6). However, an expression with the form of
Eq. (6-4) can always be obtained. The actual equations for go and gl, in a
particular case, depend on the exact air-gap geometry. For the structure
of our primitive machine Eq. (6-4) will be assumed to be exact. No
practical'machine is generally constructed with such an air gap. How­
ever, when it is used as a model, the primitive machine poses as a reasona­
ble first-order approximation for the salient machine. Notice that gl
becomes zero and Eq. (6-4) is exact for the case when no saliency exists,
or when gq = gd.

6-2 CONSTRUCTION OF THE STATOR CIRCUITS


On the stator of the primitive machine two windings are to be placed.
Let us consider each winding separately. Figure 6-3 shows a cross-sec­
tional view of the primitive machine. Notice that the rotor is cylindrical,
but the inner periphery of the stator is not cylindrical. The air-gap
length is given by Eq. (6-4), as indicated in Fig. 6-3. A current sheet is
now to be established on the inner periphery of the stator structure adja­
cent to this air gap. Before mathematically formulating an expression
for this current sheet, let us see how such a current distribution could
physically be obtained.
Axial slots can be cut in the inner periphery of the stator. A
continuous coil is then wound, running a conductor down one of these
slots from back to front, then across the front of the rotor to a slot "II" rad
from the original slot. The wire is placed in this second slot running from
front to back, then across the back of the rotor to a slot adjacent to the
starting slot. By continuing in this manner, a continuous winding
approximating a continuous current sheet on the entire inner periphery of
the stator is constructed. The sides of the winding running axially in the
slots are known as coil sides, and the parts of the winding running over the
front and back of the rotor are called end connect1·ons. By placing more
coil sides in certain slots, a desired current distribution can be obtained.
One of the stator circuits in our primitive machine is constructed in
this fashion. By grouping more connectors in the slots near 8 = 0 and
Construe/ion of the primitive machine 271

~--~~~------~----~--~--~----.d

Fig. 6-3 Cross-sectional view of the magnetie structure of the primitive machine
containing a quadrature-stator winding.

o = 11' in Fig. 6-3 and vcry few coil sides near 0 = 11'/2 and 0 = 311'/2, a
sinusoidal distribution of current can be approximated. For the winding
shown in Fig. 6-3 a magnetic field in the positive quadrature axis direc­
tion will be found in the rotor structure for the current direetions shown.
Therefore, this windillJ!: is known as the quadratu.l'e-slalor winding. The
current in this winding is denoted by i:.
where the subscript q denotes the
quadrature axis an!il the superscript s indicates the stator.
The eurrent sheet set up by this quadrature-stator winding is given by

amp/m
The factor K; cos 0 is a distribution term giving the number of conductors
per meter (in the 0 direction) each carrying a current of i; amperes.
Assuming that the air-gap length is negligible compared with the radius of
the rotor a, the quantity of eUlTcnt 011 the stator surface running axially
along the machine in a peripheral distance a dO is given by

dI = J;a dO amperes (6-8)

In order to further clarify the current-sheet description as set up by


the quadrature-stator windillg, consider the developed view of the primi­
tive machinc as shown in Fig. 6-4. For simplicity, the air gap is shown
to be uniform; however, in all calculations the functional dependence for
272 Principle, of electromechanical·energy convenion

r
G
r
~.
0
r-------- -,
I r(Z
- +-_-_--_--
-®--..;.--r--
-...
11"!
... (!.....!)...
. f-.. .
Rotor LOirected contour

3",/2 T Z11"
-_-:r"itr----'... -=t-l--~-tL->----"':8
L
9(8)=90-9, c0528
Stotor
Fig. 6-4 A developed view of the air gap of the primitive machine showing the
currents in a quadrature-stator winding.

g(tJ) gi\lCn by (6-4) will be used. Equation (6-7) gives the number of
amperes per unit length along the tJ axis established on the st.ator surface
by the current i:.
In the two quadrants 0 < tJ < 11"/2 and &r/2 < 8 < 211",
the current-density vector J:
is into the paper, or in the negative z-axis
j
direction, as verified by Eq. (6-7). For the other two quadrant.s J; is in
the positive a. direction.

6-3 STATOR MAGNETIC FIELD


The value of the vector-magnetic field in the air gap can be determined by
using Ampere's circuital law, which states that the line integral of the
magnetic-field-intensity vector around a closed contour is equal to the
current enclosed by that contour. Symbolically this law may be expressed
as
(6-9)

where H is the magnetic-field-intensity vector, dl is the incremental vector


path length, J is the current-density vector, in amperes per square meter,
and dS is the incremental vector surface capping the contour.
Throughout our work, the assumption is made that the permeability
of the material (generally iron) of which the stator and rotor are con­
structed is many orders of magnitude greater than that of air. Taking
the magnetic-intensity vector H in the iron to be zero, we may apply
Ampere's circuital law along the directed closed contour shown in Fig. 6-4.
Notice that the contour crosses the air gap at some angle 8 and then again
at 8 + 11". The length of the contour has been selected to be 11" rad to take
advantage of the odd symmetry of the current sheet. l Because of this

I A function f(x) possesses odd symmetry if f(x) - -f( -x). A sine function is an
example of odd symmetry. By the term "odd" we do not mean "peculiar."
Construction of the primitive machine 273

symmetry we must have

H(IJ) -H(IJ +.,.) (6-10)

Assuming the air gap to be sufficient.ly small that the magnetic-vector


intensity has only a radial component, Amp~re's circuital law for the
contour in Fig. 6-4 reduces to

H,(IJ)g(IJ) - ll,(IJ + .,.)g(IJ +.,.) = Is J . dS (6-11)

where we have used ll. as the symbol for the radial component of H. The
minus sign preceding the second term on the left side of (6-11) results
from the fact that at IJ + .,. the contour direction is opposite to the positive
radial direction specified by aT' The surface integral on the right side of
(6-11) represents the total current passing through allY surface capping
the contour.
For the contour direction shown in Fig. 6-4, current coming out of
the paper must be taken as positive according to the right-hand rule, and
therefore dS = d8 a.. By \ISC of F:qs. (fi-7) and (6-8), the total current
is given by
(B+ ..
f J . dS
S = J8 - K; cos 8 i;a d8
which reduces to

Is J . dS = 2K~ai: sin 8 (6-12)

Equation (6-12) <;an easily be cheeked by referring to Fig. 6-4. If the


left side of the contour is positioned at 8 = 0 the contour encloses zero
net current in the interval from zero to.,.. For the left side of the contour
at 8 = .,./2 the current enclosed is maximum and positive (j.e., out of the
paper), since the contour spans from .,./2 to 3.,./2.
By substituting Eq. (6-12) into (6-11), after some slight simplifica­
tion, we have for the magnetic-vector intensity in the air gap

' _ K;ai; sin 8 (6-13)


Hq - g(8) aT

with g(8) g(8 + .,.), as seen from Eq. (6-4). The constant K: is a
distribution factor with units of turns per meter, the rotor radius is a
meters, a.nd the air-p;ap length is g meters and is a function of (J. With the
coil current i:.
in amperes, t.he units of H; are ampere-turns per rrl:eter.
The magnetic-flux-density vector can be obtained simply by multiplying
274 Principles of electromechanical·energy conversion

Quadrature axis

Direct axis

Fig. 6-5 Uniform mag­


netic-flux density inside
the rotor cylinder due to a.
current in the quadrature.
stator winding.

H by the permeability of air, thus getting


B' = Io'oK;ai; sin e a,. (6-14)
9 g(e)

with 10'0 = 4n- X 10- 7 him and B: having units of webers per square meter.
In the air gap, Eq. (6-14) shows that the magnetic-flux density is also
radial and varies in magnitude as the sine of the angle e. Figure 6-5
shows only the rotor structure of the machine with the flux lines drawn
radially around its periphery. For a machine with no saliency (that is,
g( e) = go) and an isotropic, cylindrical rotor of a high-permeability material,
the flux density within the rotor member will be uniform and directed as
illustrated in Fig. 6-5. This uniform flux density, within the rotor
structure, can be expressed in cylindrical coordinates by
Io'oK'ai' .
--g-g
go
(sm e a, + cos e a,) in rotor (6-15)

Taking the direct axis as the x axis and the quadrature axis as the y
axis in a cartesian-coordinate system, Eq. (6-15) is equivalent to

in rotor (6-16)

with a". as the unit vector in the y direction.


Notice, in deducing Eq. (6-15) from Eq. (6-14), that the normal or
radial component of B at the air-gap rotor boundary is continuous. The
tangential component of B does not have to be continuous across the
Construction of the primitive machine '05

boundary. However, the tangential component of H should be con­


tinuous in the absence of boundary surface currents. In the air gap,
tangential H is zero, but Eq. (6-15) would suggest tangential H in the rotor
to have a magnitude of
H, = B, = K;ai; cos 8 ~ in rotor (6-17)
~ go ~

where ~ is the permeability of the rotor material. If ~ is many orders of


magnitude greater than the permeability ~o of air then, very nearly,
H. (in rotor) = H. (in air gap) = 0 (6-18)

For schematic purposes the single quadrature-stator coil is drawn on


the quadrature axis, as shown in Fig. 6-6. Current going into the dotted
side of the winding is assumed to set up a uniform field inside the rotor in
the positive direction for the axis on which the coil is placed. For the
quadrature axis the positive direction is taken upward, and for the direct
axis the positive direction is assumed to the right, as indicated by the
arrows on the two axes. Drawing the coil in Fig. 6-6 outside the circle
indicates it to be a stator coil. Rotor coils will be placed inside this circle.
In an entirely analogous fashion a stator coil is arranged to set up a
magnetic field in the horizontal direction. The additional stator coil is
also shown schematically in Fig. 6-6. The two fields set up by the com­
bination of these two stator windings are assumed to be independent of
one another, and the net field at any point in the interior of the rotor
member or in the air gap is the superposition of the two perpendicular
magnetic fields. All saturation effects in the iron are being neglected in
our primitive-machine model, thus allowing for this superposition.
For a current id in the direct-stator winding, current sheets identical
in form with tho~e for the quadrature-stator winding are set up on the

Quadrature axis

.,
4----1 q

Fig. 6-6 A schematic representation


of the stator windings in the primitive
machine.
176 Principles of e/ectromechanical·energy conversion

inner stator surface. Therefore, corresponding to Eq. (6-7) upon sub­


stituting 8 + 1f/2 for 8, we have l
Jd = Kd sin 8 z'da. amp/m (6-19)
as the surface current density due to the direct-stator winding. Follow­
ing a consistent notation, Kd sin 8 represents the effective number of turns
per meter each carrying a current i d. Using Ampere's circuital law, the
magnetic-field vector in the air gap would be
B' _ ~oK~i~ cos 8 (6-20)
d - g(8) a,
following a development identical with that used to obtain Eq. (6-14) for
the quadrature-stator winding. Entirely analogously to Fig. 6-5, the
direct-sthtor winding would create, in a nonsalient machine, a uniform
magnetic field inside the rotor structure directed to the right for positive i d.

6-4 COMMUTATOR-AND-BRUSH ROTOR WINDING


Constant l;urrents inserted into either of the two stator coils give rise to a
magnetic field that is stationary in space with respect to the stator. The
rotor circuits in the primitive machine are also going to set up a similar
magnetic field. This means the field of a rotor winding excited with a
constant current is to remain stationary in space with respect to the stator
Jor any orientation oj the rotor structure. Even if the rotor is continuously
rotating, the field due to the rotor winding will remain unaffected. A
rather interesting device known as a commutator is used to accomplish
this desired effect. The main point to keep in mind in the following dis­
cussion is the requirement of a stationary magnetic field dependent only
on the rotor winding current, and not on the position or velocity of the
rotor structure.
In order to build the idea of how a stationary field can result from
coils that are physically moving, consider Fig. 6-7. Here a cylindrical
tube has a single continuous winding wound in the manner shown. Two
brushes placed'll" rad apart on the front end of the cylinder make contact
with the end turns of this winding. If a current i is inserted into the
right-hand brush, a current of i/2 will wind around the top half of the
cylinder and a current i/2 will wind around the bottom half, with a total
current of i emerging at the left brush. The two paths are assumed to be
electrically identical. Notice that all conductors on the outside of the
I The direct-sto.tor winding is located 1r /2 rad hehind the quadrature-stator winding.

Simply substituting (; + ... /2 for (; in all quadrature-stator equations gives us the


corresponding equations for the direct-stator winding.
Construelion of the primitive machine '177

Fig. 6-7 A hollow cy­


lindrical tube holding a sin­
gle continuous winding. Cut

upper half of the cylinder carry currents out of the paper, as do all con­
ductors on the inside of the lower half of the cylinder. Similarly, all
conductors on the outside of the lower half and inside the upper half of
the cylinder have currents going into the paper.
While holding the brushes in a fixed position, consider rotating the
cylinder. The two brushes will each slide off one turn of the winding and
onto an adjacent turn. Notine, however, that the current distribution
would not be altered. Conductors located on the outside upper half and
inside lower half still have currents out of the paper, and all conductors
located on the outside lower half and inside upper half have currents into
the paper. If the brushes are held fixed, the current distribution is inde­
pendent of the rotary motion of the cylinder.
The winding shown in Fig. 6-7 is very inconvenient and impractical
for a number of reasons. One of the major objections would be the lack
of iron in much of the rotor structure, since Fig. 6-7 requires a hollow
cylinder. The magnitude of the magnetic field B available for a given
field intensity H would be seriously reduced. For this reason the solid
(or, in practice, laminated) rotor cylinder shown in Figs. 6-1 and 6-3 is
preferred. A solid cylinder can be used if all the conductors inside the
cylinder are moved to the outside periphery without changing any of the
relative connections. Figure 6-8 shows how one such conductor is moved.
Notice that the conductor I' is moved to a position 11" rad around to the
outside of the cylinder without any change in electrical connections. If
all conductors on the inside walls of the cylinder are similarly moved to the
outer periphery of the rotor, in a position 11" rad from their original location,
then the cylinder need no longer be hollow. Since no change has been
made in any electrical connections, the currents in each respective con­
ductor are the same as they originally were in Fig. 6-7. Therefore, all of
the currents on the outside of the lower half of the cylinder are into the
paper and all currents on the outside of the upper half are out of the paper.
278 Principle, of electromechanical-energy conve",ion

1'~~------------+-r-

Fig. 6-8 Repositioning the coil sides loca.ted inside the hollow rotor cylinder to the
outside or the cylinder.

The complete layout, illustrating all connections, may be illustrated


more conveniently by using a developed diagram of the rotor. This form
of circuit drawing is shown in Fig. 6-9. Imagine cutting the hollow
cylinder at the point illustrated by the arrow in Fig. 6-7 and then flattening
it out without distorting the position of each conductor on the walls of the
cylinder. Now when we look at the flat 8urfn.ce, the conductors appear as
in Fig. 6-9, where the dashed conductors are actually below the flattened
surface or inside the hollow cylinder. For clarity, the dashed conductors
are displaced slightly in Fig. 6-9, since they really should appear
directly under the solid conductors. The numbers on the conductors in

w'

\ I I 1 I 1 I 1 1 1
8'1 1'1 6'1 5'1 4'1 3'1 2'1 11'1 10'1
, , ,
9'1
11"

j ,I t
I I I
I I
I 1
I
1
I
I
I
I I

I1r
I

I
I
I
1
1
I ,
1
I
I
I
I
1
I
1
I
I
I
I
1
1
I
I
1 I I I
1
I I I I I I I 1 1
I
I 1 I I I 1 1 1 I
1
1 1 1 1 I I I 1 I
1
I I I I I I 1 I I
I
I 1 I I I I I I
I
9 I I + • 5 • 4
+ 12 I 11 I 10

Fig. 6-9 A developed view or the origina.l configuration or the rotor conductors.
Comtrvdion of the primit;". machine 279

Fig. 6-10 Rotor conductors rearra.nged to be loca.ted on the outer periphery of the
rotor cylinder.

Fig. 6-7 are duplicated in Fig. 6-9 to aid in verifying the similarity of the
two drawings. Also, notice the brush positions in the two figures are
identical. Current directions into the paper in Fig. 6-7 correspond to
currents up the page or away from the brushes in Fig. 6-9. Rotation of
the rotor counterclockwise with respect to the stationary brushes corre­
sponds to motion of the coil sides to the left in the developed view of the
same rotor.
The movement of all primed conductors 'It" rad around the rotor and to
the outside walls may easily be seen in this developed view. For example,
coil side I' will be moved above coil side 7, coil side 2' goes above 8, and so
forth without changing any electrical connections. This rearrangement
of the conductors is shown in Fig. 6-10. Again for clarity, where one coil
side is actually on top of another, it is displaced slightly and drawn adja­
cent to the original coil side. The developed view in Fig. 6-10 looks very
involved, but if you remember that no change in connections has been
made between Figs. 6-9 and 6-10, the tracing of the circuit is not difficult.
Observe in Fig. 6-10 that all conductors in the shaded area have currents
toward the brushes. The shaded area corresponds to the upper half of
280 Principles of electromechanical-energy conversion

the cylinder, and the nonshaded area corresponds to the lower half of the
cylinder.
We have now developed a mechanism for maintaining a stationary
current distribution on the surface of the rotor that is independent of the
position of the rotor structure. The sliding contact between the brushes
and the winding can no longer be made by having the brush slide over the
end connections, as shown in Fig. 6-7, since two coil sides are now located
in each slot on the rotor periphery in Fig. 6-10. This sliding connection is
usually made by attaching the conductor ends to copper segments
insulated from each other and mounted on the surface of a small cylinder
on the shaft of the rotor. The brushes then ride on these segments and
make a low-resistance sliding connection to the rotor coils. This struc­ "
ture composed of insulated copper segments is known as a commutator.
A devJloped view of the commutator is also included in Fig. 6-10.
The winding shown in Fig. 6-10 is only one configuration out of a vast
number of possible arrangements that will result in a stationary current
distribution that is independent of the rotor position. The purpose of
this discussion is only to show that the desired result is possible, and not
to give the vast details in the art of rotor winding.

6-5 ROTOR MAGNETIC fiELDS


The currents on the surface of the rotor have been arranged to be uni­
directional on the upper and lower halves of the rotor cylinder. A
developed view of the machine, looking in the axial direction, is shown in
Fig. 6-11. The air gap is drawn uniform simply for convenience. In all
calculations the sinusoidal variation of g indicated in the figure is to be
used.. On the air-gap periphery of the rotor surface, the current distribu­
tion established by the commutator-and-brush mechanism is shown.
Figure 6-11 is drawn with the assumption of many more coils on the rotor
than the 12 shown in Fig. 6-10. By placing these coil sides in slots cut

Rotor

Stator

Fig. 6-11 A developed view of the air gap showing the uniform surface-eurrent
distribution on the rotor periphery.
Con.truction of the primm.,. machine 281

axially and very close together on the rotor surface, the nearly uniform
current distribution shown in Fig. 6-11 can be approximated. Analyti­
('ally the rotor surface current can be expressed as

J:i = K:i'i:ia, amp/m for 0 < 0 < 'If'


and (6-21)
J:1 = - Kr,J t'da. amp/m for 'If' < 0 < 2'1f'
The distribution factor Kr,J expresses the number of turns, each carry­
ing a current of t'd amperes, per meter of rotor surface measured in the
direction of a.. The subscript d is being used because this current distribu­
tion will finally result in a field inside the rotor along the direct axis. The
superscript r indicates that the current distribution results from a rotor
winding. Remember that the current t'd is the terminal, or port, current
for the winding. For the winding in Fig. 6-10 the currents in the active
coil sides are half the value of the terminal current. Therefore, in
formulating Eqs. (6-21) the distribution factor K:i' must be properly
adjusted so that J'd is the correct surface-current density in terms of t'd,
the terminal current.
The square-wave surface-current distribution given by Eqs. (6-21)
can be expanded in a Fourier series as follows:

= 4Kr'i
r
J~ __01_01 (sin 0 + ! sin 30 + i sin flO + ...) a. (6-22)
'If'

This series is similar to Eq. (6-3) for the air-gap length, except that here
we have odd symmetry giving only sinc terms and the total period is 2'1f'
rad rather than 'If' rad as in the air-gap expansion. In the analysis of our
primitive machine only the first term in the series is to be retained.
When we define a slightly modified distribution factor
K~
4Kr'
= _ _
01 turns/m (6-23)
'If'

the direct-rotor winding surface current is approximated by


J:1 = K'd sin 0 t'da. amp/m (6-24)
on the primitive-machine model. Notice that the direct-rotor-current
distribution given by (6-24) has the same form as the current distribution
for the direct-stator winding as given by Eq. (6-19).
Ampere's circuital law, given by (6-9), can be used for determining
the magnetic field in the air gap. Using a closed contour 'If' rad long,
we have
('+'"
H'd(O)g(O) - H'4(O + 'If')g(O + 'If') =}, J'da d8 (6-25)
282 Principles of electromechanical-energy conversion

with H~ being radially directed and a as the rotor radius. According to


Eq. (6-4)
g(fJ) = g(fJ + '11") (6-26)
Also because of the symmetry of the current distribution
Wa,(fJ) = - H~({J + 71") (6-27)
By substituting Eq. (6-24) into the right side and integrating, we have,
finally,
• _ K;taiiJ cos fJ (6-28)
Hd - g(fJ) a.

The mp.gnetic-flux-density vector is obtained by multiplying Hd by the


permedbility of air, which gives us
B~ = l-'oKdaid cos (J a. (6-29)
g(fJ)

The direct-rotor winding on the primitive machine establishes a


magnetic field with exactly the same form as the direct-stator winding,
as seen by comparing Eqs. (6-20) and (6-29). Since both windings set
up identical surface-current distributions, we would certainly expect the
fields to be identical.
) One further point should, however, be made clear. The stator-wind­
ing sinusoidal current distribution can physically be obtained by properly
distributing the coil sides in the various stator slots. By thus stacking the
conductors, the sinusoidal surface-current density can actually be
achieved. In the case of the rotor winding, the commutator-and-brush
arrangement can only set up the square-wave surface-current distribution
illustrated in Fig. 6-11. Stacking more conductors in certain rotor slots
and fewer in other slots is of no help, since these circuits are continuously
being switched by the commutator and brushes as the rotor rotates.
Therefore, the sinusoidal rotor-surface current given by (6-24) is only an
approximation, since it is the first term in a Fourier-series expansion for the
actual distribution. Thus, physically the sinusbidal-current distribution
on the rotor is normally not a.ctiiaTIy achieved. Our primitive machine
can assume a sinusoidal rotor-surface-current distribution because it is
only a mathematical model related to physical machines in only an
approximate manner. In our later work with practical machines we shall
keep these approximations clearly in mind, and in this way the primitive
machine will serve as a useful tool.
So here, in Eq. (6-29), we have the desired field due to a rotor winding.
This field is stationary in space. It is dependent on the rotor winding
current only ~and is completely independent of the position of the rotor.
Comtl"Uction of the primitive machine 283

Quadrature axis

Direct
axis

Fig. 6-12 A schematic representation


of the rotor windings on the primitive
machine.

A schematic drawing of the rotor winding is shown on the direct axis in


Fig. 6-12. The winding is drawtl inside the circle to indicate that it is on
the rotor of the machine. Current into the dotted side of the winding is
taken to cause the field to be directed along the positive direct axis.
The schematic drawing is therefore representative of the winding we have
considered throughout this section, namely, one having a commutator
and brushes creating a stationary field with no regard for what the rotor
structure is doing. ----.--­
If the brushes in Fig. 6-7 were rotated by 1r/2 rad in the positive, or
counterclockwise, angular direction, then a magnetic-flux-density vector
similar to Eq. (6-29) in the air gap would be generated with the substitu­
tion of 0 - 1('/2 for 0, thus giving us
Br _ Vr 'r
1.It)1~,m,
.
sm 0
, - g(O) ar (6-30)

as the quadrature~rotor-winding magnetic field in the air gap. Thus a


rotor winding is placed on the quadrature axis of Fig. 6-12 corresponding
to a second pair of brushes located 1('/2 rad from the original pair of brushes.
If currents are simultaneously put into both pairs of brushes, the resulting
magnetic fields are taken to be the superposition of the fields experienced
from each of these_currents individually. The superposition principle
applies because sa.tu~ation effects in the stator and rotor material are
being neglected. Again, current into the dotted side of this quadrature
winding causes the field to be directed along the positive quadrature axis.

6-6 PRIMITIVE-MACHINE STATIONARY INDUCTANCE PARAMETERS


By using the schematic representation for the stator and rotor windings
introduced in the preceding sections of this chapter, the complete four­
284 Principle. of electromechallical·energy cOllyeNioll

Quadrature axis

Direct
axis

Fig. 6-13 A schematic representa­


tion for the complete four-winding
primitive machine.

winding primitive machine can be drawn as shown in Fig. 6-13. The


respective coil currents have their reference directions so chosen that for
positive currents the resulting magnetic fields are established in the posi­
tive direct- or quadrature-axis direction. For positive i~ and i~ magnetic
i:
fields are established to the right; for positive and i; magnetic fields are
directed upward inside the rotor strueture. In Figs. 6-6 and 6-12 dots are
placed on the windings to indicate this reference choice. However, since
we shall always adopt the same sign convention, these dots will not be
shown on further primitive-machine diagrams. All of our work in Secs.
6-2 to 6-.5 has used the same sign convention for the respective currents.
The two stator windings, shown in Fig. 6-13 outside the circle, are
the type described in Secs. 6-2 and 6-3. Each stator winding establishes
a sinusoidal-current distribution on the inner-' stator surface, which is
fixed in space with respect to the stator. Similarly, the two rotor wind­
ings, shown inside the circle in Fig. 6-13, are described in detail in Secs.
6-4 and 6-5. These rotor windings also establish a sinusoidal-current
distribution on the rotor surface. Since the rotor windings are supplied
through a commutator-and-brush mechanism, this current distribution,
like the stator distribution, remains stationary in space with respect to the
stator. The air gap between the stator and rotor, in general, is assumed
to be nonuniform and to have a sinusoidal variation as discussed in
Sec. 6-1.
In this section we shall develop expressions for the inductance
parametcrs of the primitive-machine windings with the rotor held sta­
tionary. Having these parameters defined allows our further machine
analysis to proceed from a circuital rather than a fields viewpoint.
Circuital variables (that is, voltages and currents) are more easily observed
and measured than field quantities. Suppose that all of the four winding
currents are increased from zero to some final set of values i~, t~, i;, and ~~.
Construction of tllft primitive machine 285

In general, the total winding flux linkages would then be given in matrix
form by
O

X9!]
X _ M~'
[LOqd
(6-31)
Xd - Md'a
X; M;'a
The L coefficients are the respective self-inductances of the four windings.
The remaining M coefficients are mutual inductances. The first set of
sub- and superscripts on the M coefficients designate the windings in
which the linkages exist, and the second set of sub- and superscripts
indicate the windings carrying the current causing these linkages. Thus
M~~i: are flux linkages in the direct-stator winding reSUlting from a current
in the quadrature-stator winding. From our previous work we know
that the inductance matrix in (6-31) must exhibit principal-diagona.l
symmetry in order that the energy- a.nd coenergy-state functions be
independent of the particular manner in which the final state of the sys­
tem is reached. This means that M~~ = M;, M'.tt = M'da, and so forth
for all other mutual coeffieients. Also, we shall see that many of these
mutual eoefficients are zero.
The magnetic coenergy of the primitive-machine system can be
obtained from

(6-32)

allowing the summing index k to represent each of the four windings


successively. On substituting the flux-linkage relations from Eqs. (6-31)
into (6-32), we have for the coenergy-state functions
*L~(id)2 + M:'idi~ +
!L;(i;)2
+ Mddi~id +Md~i:id +
!L:i(t"d)2
+ M;didt~ + M~~i:t~ +
M~dtdt~ + !L~(i;)2 (6-33)
Since the flux-linkage relationships given by Eqs. (6-31) are linear
expressions, the magnetic-coenergy-state function W~ equals the magnetic
energy W... stored in the system. The magnetic energy stored in the
primitive machine can also be obtained by integrating the total energy
density over all space. I This magnetic-energy density is given by

U... = B·H joules/m 3 (6-34)

1 See R. Ploneey and R. E. Collin, "Principles and Applications of Electromagnetic

Fields," sec. 8-5, McGraw-Hili Book Company, New York, 1961.


286 Principl•• of .I.dromechania;sl·energy convenion

where Band H are the total fields at each point in space. The iron por­
tions of the primitive machine have a permeability which has been assumed
infinite compared with the permeability of the air gap. For this reason
the I1tagnetic-vector intensity H in the stator and rotor iron must be zero.
The magnetic field B outside the stator structure is also zero because of
this assumed permeability. All of the magnetic energy stored is therefore
found only in the air gap of the machine, since everywhere else the energy
density U", is zero becauseB and/or H are zero. In the air gap H is related
to B by the permeability of air ~o = 4Ir X 10-7 him, and therefore Eq.
(6-34) reduces to
B2
U", = - joules/m a (6-35)
2~o

B
where is the magnitude of the total air-gap magnetic field.
With the (our windings of the primitive machine carrying the respec­
tive currents shown in Fig. 6-13, the total air-gap magnetic-field vector is
given by
B = Bd + B; + B~ + B; (6-36)
On SUbstituting Eqs. (6-14), (6-20), (6-29), and (6-30) into Eq. (6-36), we
have
~oa
B = g( e) (K'dtd'. cos 8 + K' '. sm
qti . 8 + Krd~dOr cos 8 + Kr Or sm 8) ~
iti °

(6-37)
with all components in the air gap radially directed, The magnitude of B
is not a function of the cylindrical coordinates rand Zo Thus the total
energy stored in the air gap is simply
( 2.,
10 U ...(e)g(e)la de (6-38)

The air-gap length g(e) times the length of the rotor cylinder l is the area
of a radially oriented rectangle in the air gap at an angle e. Assuming
that 9 is much smaller than the radius of the rotor, given by a, our
incremental volume is the g(e)l rectangle times the arc length (a de).
On mUltiplying by the energy density U"" we have the energy in this
incremental volume. Summing over e from zero to 21f gives Eq. (6-38),
which is the total energy stored in the air gap by the four windings.
Using the value of B from (6-37) in Eq. (6-35) and substituting into
Eq. (6-38), we have for the energy stored
W... (id,i;,id,i~)
= ~oa3l (2., (K~i~ COS 8 + K:i: sin 8 + K~i~ cos 0 + Kgi; sin 8)2 do
2 10 g(~
(6-39)
COIMtrvction of the primitive mochine 287

The air-gap length of the primitive machine g(fJ) is taken as being


g(fJ) = go - gl cos 2fJ (6-40)
Assuming that go» gl,t
_1_ "'" .!.. (1 + gl cos 2fJ) (6-41)
g(fJ) go go
Now, substituting Eq:(6-41) under the integral sign in Eq. (6-39)
and expanding the four-term squared parentheses gives us sixteen integrals
to evaluate. The following three definite-integral forms are used in the
evaluation of W... :
(2.. cos' fJ (1
10
+ gl cos
h
2fJ) dfJ = 1r (1 + ll.!.)
2~
(2.. sin' fJ (1
10
+ gl cos
go
2fJ) dfJ = 1r (1 _ ll.!.)
2go
(2.. cos fJ sin fJ
10 \
(1 + gl cos 2fJ) dfJ
go
0
After a great deal of manipulation the energy stored can be matched
term by term with Eq. (6-33), which gives the energy stored in terms of
the self- and mutual-inductance parameters. The results of integrating
and term matching for the two-pole (one-pole-pair) primitive machine
are as follows:
Ld = f.'oa*l1r(K d)'
go
(1 + ll.!.) 2go
(6-42)

L- == f.'oa*l1r(Igr (1 _1l!.) (6-43)


q go 2go
Ld = f.'oaal1ll(K~)2 (1 + ll.!.) (6-44)
go 2go
U = f.'oa*l1r(Kg)2 (1 _1l!.) (6-45)
q go 2go

M~d = M~ = Md' = f.'_


",,3l1r K'd K'd ( 1 + -2.
g ) (6-46)
go 2go

Mrr = M"qq = Mfrq = f.'oaal1rK'K'(


_ _:x::!l 1 - -gl) (6-47)
qq go 2go
~=~=~=~=~=~=~=~=O~~
t The function !(.) - 1/(1 - .) expa.nded in a. power series becomes
!(.) _ (1 + • + .' +.' + ...)
For values of • « 1 the first two terms of the series give a ree.aonable approximation;
thus
1
!(.) - 1 _ .... 1 +• for. « 1
288 Principles of e/ectromechanical.energy conYers ion

The eight mutual-inductance terms which are all zero in Eq. (6-48) are
mutuals between two windings in quadrature or at 1r'/2 rad from each
other. The mutual inductance between either winding on the direct axis
and either winding on the quadrature axis is zero. The mutual induct­
ance between two windings on the same axis is not zero, as indicated by
Eqs. (6-46) and (6-47).
On substituting these results into the linear inductance matrix given
in (6-31), we have
o
X~] [L~0 o ] ~:.r]
M';{
X' L~ 0 M~'
X~ = M,;{ o L~ o i:; (6-49)
X; I 0 M~' 0 L; $~
with Eqs. (6-42) to (6-47) as expressions for these parameters in terms of
the primitive-machine dimensions and the winding-distribution factors.
Notice from these parameter equations that
M~r = (L4L:;) >i (6-50)
and
(6-51)
as must be the case when all leakage magnetic flux is neglected. Also,
ob8erve that all of these inductance parameters are independent of the angular
p08ition or angular velocity of the rotor structure. This angular independ­
ence should be expected, since both the rotor- and stator-current dis­
tributions are independent of the rotor angle.

6-7 PRIMITIVE-MACHINE ROTATIONAL INDUCTANCES


The voltage induced in a closed circuit existing in a mag' .etic field B is
given, in general, by

eiDd = - fs aB.
at dS +,f..
'f' c
v X B • dl (6-52)

where v is the velocity vector for a vector increment of length dl and the
surface S is bounded by the closed contour C. The integrals on the right­
hand side of Eq. (6-52) are a general form of Faraday's law, which states
that a voltage is induced in a closed circuit if the circuit flux linkages are
changing with respect to time. The change in the surface integral occurs
owing to the flux-density vector B being nonstationary, or a function of
time. Thus the iJB/iJt would be nonzero. The closed-contour integral in
Eq. (6-52) has the flux linkages changing because of motion of the closed
circuit with respect to the magnetic field. Thus Eq. (6-52) expresses the
Construction of the primitive m<Jchine 289

general law that changing the flux linkages in a closed circuit causes an
induced voltage regardless of how the flux linkages are changed.
The air-gap magnetic fields in our primitive machine can vary with
time only if the winding currents setting up these fields are themselves
time-varying. The contribution to the voltage induced in anyone of
the windings owing to a time-changing field is taken into account by the
inductance parameters developed in Sec. 6-6. Thus the first integral on
the right side of (6-52) reduces to

r aB. dS -+ ax (6-53)
J8 at at
with the respeetive flux linkages X given by Eqs. (60049) in terms of the
self- and mutual-inductance paramet.ers. Since all of the winding currents
are functions only of time and are not functions of any position coordinates,
the ax/at reduces to a sum of terms with the form L di/dt and/or M di/dt,
using the total instead of the partial deriva.tive of the current. The next
chapter is concerned with equilibrium equations for the primitive machine,
and the sense or direction of these induced voltages. At this point let us
just realize that the sense of the L di/dt voltage terms must be such as to
oppose an increase in the winding current i.
In this section we wish to examine the effect of the second integral in
Eq. (6-52). The total radial air-gap field due to currents in all four
windings on the primitive machine is given by Eq. (6-37). For conven­
ience, let us repeat this equation as

B I/ofIJ (K'"
= g(8) dtd cos 8 + K ~l.,'. sm
. 8 + Kr'r
dId cos 8 + Kr'r . 8) a
.,t" sm r

(6-54)

where a is the radius of the rotor cylinder, g(8) is the air-gap length as a
function of the polar-coordinate angle 8, and the respective K's represent
the sinusoidal-distribution factors with units of turns per meter.
The quadrature-rotor winding conducting a current i; is assumed to
establish a current density on the surface of the rotor as given by

- K; cos 8 l~a. amp/m (6-55)

Figure 6-14 shows the rotor cylinder with the quadrature-rotor-current


distribution indicated along curved sides of the cylinder. At IJ = 0 the
current density (in amperes per meter of surface, measured in the a.
direction) is maximum and is directed in the negative z-axis direction,
Similarly, at B = r/2 the current density J; has zero magnitude. The
length of the current lines in Fig, 6-14 indicates the relative magnitude of
the current density at that place on the rotor surface. The arrowheads
290 Prinr:iples of eledromec:honica/·energy conversion

Fig. 6-14 The primitive­


machine rotor cylinder,
showing the suriace-current
density established by the
r quadrature-rotor winding.

on the current lines give the direction. Carefully verify that Eq. (6-55)
does indeed represent the current distribution pictured in "Fig. 6-14.
The quadrature-rotor winding is constructed from a certain number
of coil turns wound on the rotor cylinder. One such turn is shown in
Fig. 6-14. Notice that this single turn is composed of four sides in a
rectangular form. The two sides running axially on the curved surface
of the rotor are often called the active sides, because the currents in these
two sides of the coil establish the air-gap fields. The two sides along the
front and back of the rotor cylinder simply serve to conduct the winding
current from one active side to another. The front and back sides are
termed end connections. The effects of these end connections are being
neglected in the formulation of the primitive-machine parameters. The
magnetic fields in the vicinity of the end connections are usually quite
small, since the mt'diur..... urrounding the ends of the rotor is air.
If the rotor is revolving a.t a.n angular frequency w', as shown in
Fig. 6-14, then the velocity of the active sides is given by

v = aw'a, (6-56)

where a is the rotor radius. Using Eq. (6-54), we can form the vector
product v X B and get as a result

1J~2W'"
v xB = -
, .,
g(8) {Kd~d cos 8 + Ktt, *sm
t
.."
8 + KdZd cos 8
r .,.

+ K;i; sin 8)a. (6-57)


Remember that a, X a.. = - a•.
Construdion of the primitive machine 291

The voltage induced in the single coil shown in Fig. 6-14 can be
determined from the second term in (6-52), which is

(6-58)

Selecting the contour direction shown by the arrows on the coil in Fig.
6-14 gives us for vector length dl
dl = -dl a. for the active side at tJ
and (6-59)
dl = dl a. for the active side at tJ + ,..
However, since B(tJ) = -B(tJ + ,..), owing to the odd symmetry of each
current distribution, Eq. (6-58) reduces simply to
1 (I
Cind = -2 10 [v X B(tJ)] • dl a. (6-60)

where l is the length of the rotor cylinder or the length of each active side.
Since v X B, as given by (6-57), has only a z component, the scalar or dot
product under the integral in Eq. (6-60) is easily taken. By substituting
(6-57) into Eq. (6-60) and integrating, we have for the voltage induced in
this single coil

due to the currents in all four primitive-machine windings. The contour


direction leading to Eqs. (6-59) was chosen to correspond to the direction
of the quadrature-rotor currents on the rotor surface. Thus, if e:nd as
given by (6-61) is positive, then the indueed voltage is in a direction to
aid i;. These sign considerations are studied in greater detail when the
equilibrium equations are formulated in the next chapter.
The complete quadrature-rotor voltage induced in all of the coils of
the winding can be obtained by summing the induced voltages in the
individual coils. This summing operation gives the proper total voltage,
because we have pictured the winding as being a series of many coils
carrying the same current i;. Referring to Eq. (6-55), we see that the
number of active coil sides contained in an increment of arc dtJ is given by
K; cos tJ a dtJ, since the quantity a dtJ is the increment of arc length and
K; cos tJ is the turns per meter at an angle tJ. Equation (6-61) gives the
voltage induced in one coil at an angle Q. Therefore, the total voltage is
simply the volts per turn times the number of turns in an increment of arc
at an angle tJ. On forming this summation, we have as the total induced
292 Principle. 0' electromechanical-energy conver.ion
r'

voltage in the quadrature-rotor winding ,/

,,/2 1
e~ = / -0'/2 eiJ\dK~ cos fJ a dfJ (6-62)

The sense of the quadrature-rotor rotationally induced voltage is e;


taken to be positive when aiding the assumed direction of the quadrature
rotor current t~. The limits on the integral need only proceed over 71' rad,
since Eq. (6-61) was formulated to give the voltage induced in both
active coil sides with one side at an angle fJ and the other at fJ + 71'.
Therefore, summing over half the rotor eylinder from one quadrature­
rotor brush at -71'/2 to the other brush at 71'/2 includes all of the active
sides.
) Equa,ion (6-41) gives us a useful approximation for the factor l/g(O).
Assuming that the saliency coefficient g! is much less than the average
air-gap length go, we have

_1_ .,. ..!.go (1 + gl cosgo 2fJ)


g(O)
(6-63)

Using the approximate expression (6-63) for the reciprocal of the air-gap
length in Eq. (6-61) and then substituting into Eq, (6-62) will give us four
integrals to be evaluated. After some amount of manipulation these four
integral expressions can be combined and reduced, finally, to give us as the
total induced voltage!

(6-64)

where we have defined

(6-65)

(6-66)
.
The G coefficients have units of henrys. They are usually termed
rotational inductances with the induced voltage being a function of the

I The following two integrals aid in obtaining this result:

,,/2 cost
/ -,,/2
6(1 +!l.! cos 26) d6 "':':2 (1 +
(10
2(11)
(10

and

f ,,12 cos
-./2
6sin 6(I +!l.! 26) dB
(10
COil 0
Construction of the primitive machine 293

angular velocity of the rotor w'. The first sub- and superscripts denote
that the rotational voltage is induced in the quadrature-rotor winding.
The second sub- and superscript.s denote the winding establishing the
field which causes this indu('ed voltage. Thus G';a refers to the quadrature­
rotor rotationally induced voIt.age due to current in the direct-stator
winding. From Eq. (6-64) it can be seen that the quadrature':'rotor rota­
tionally induced voltage e; results from the two windings on the direct axis
located at 11'/2 rad to the quadrature-rotor winding. Remember that in
See. 6-6 we found that the so-called transformer voltages induced through
the mutual or AI coefficients, owing to a winding current changing with
time, occurred between windings on the same axis. Rotationally induced
voltages occur only in rotor windings owing to currents in rotor or stator
windings in quadrature, while mutual or transformer voltages are between
windings located on the same axis. For positive w' and positive i~ or i:l,
Eq. (6-64) shows that the induced voltage e; is also positive, indicating
that the rotational voltage is aiding the winding current t~ to be positive.
The rotationally induced voltage in the direct-rotor winding due to
currents in all four primitive-machine windings can be obtained by a
similar development. Equation (6-61) for the voltage induced in a
single turn is still valid, since this voltage is due only to the four com­
ponents of the total air-gap field. The direet-rotor current-density
vector from (6-24) is given by

J~ K~ sin e i~. (6-67)

In this case the number of active coil sides contained in an increment of


arc de is given by (K; sin e)(a de). The total voltage in all of the direct­
rotor turns is therefore given by

e~ = - J/e~nd Kd sin 0 adO (6-68)

The limits on the integral again span 11' rad. The direct-rotor surface­
current dhtribution given by Eq. (6-67) is in the positive z-axis direction
for 0 < 8.< 11'. Referring to Fig. 6-14, we can see that the contour
direction for the single coil is opposite to the direction of Jd as the angle e
varies from one brush at () = 0 to the other brush located at e 11'.
Therefore, in order to have ed positive, when directed to aid i:i, a negative
sign must be included in Eq. (6-68). In Eq. (6-62), for the quadrature­
rotor rotationally induced voltage, the contour direction matched the
direction of J; over the angles -11'/2 < 8 < 11'/2 between the two brushes,
so the integral has a positive coefficient.
When Eq. (6-61) is substituted into (6-68), again using the approxima­
tion (6-63) for the reciprocal of the air-gap length, four integrals must be
294 Principles of elecfromechanical-energy conversion

evaluated. The final result can be put into the form I


(6-69)

(6-70)
and
Grr = p.oa3l1rKqK'd
dq go
(1 - 2go
~) (6-71)

In the ease of the direet-rotor winding the rotationally induced


volt.age!> r<!8ult from ellrrenLS in the t,wo will()in~s on the quadrature axis.
Again notire that a current in a winding, on either the rotor or stator, at
11"/"2 rad to a rotor winding, causes an induced rotational voltage. Equa­
tion (6-69) indi(:ates that e~ is negative (opposing the flow of i~) for positive
wr and positive quadrature currellts i; and/or i~. As with the case of all
the self- and mutual roefficients developed in Ser. 6-f>, all the rotational G
roeflieienh; arc rompletely independent of the rotor position. Equa­
tions (6-6;j), ((i-fili), (fi-70), and (6-71) do not depend on O. Thus for a
eonstant rotor angular velocity and for a constant. quadrature-stator
current, a constant d-e voltage with a magnitude of G~~wri; is induced in
the direct-rotor winding. This commutator-and-brush mechanism is
thus the foundation of all d-e machines. We have no rotational voltage
in the stator windings, sinee they are not rotating.

6-8 MULTIPOLE PRIMITIVE-MACHINE INDUCTANCE PARAMETERS


III Sec. (i-I the general structure of the primitive machine was introduced.
Figure fl-l shows a cylindrical rotor mounted within a salient stator
strudurc. A dcveloped view of t.he air gap of the manhine is shown in
Fig. 0-2. The air-gap length g is, in general, a function of the angle 8,
and for the two-pole st ruc'ture shown in Figs. U-l and 6-2 this function has
a period of 7r rad. Tl]{'refore, a FOllrier-series expansion for g(O) has the
form
g(O) = go - gl ('os 20 + g.. cos 3(20) + ... (6-72)

I To obtain this result the following two integrals are required:

- 10" sin' IJ (1 + ~ cos 2IJ) dIJ = - i (1 - i~)


and
- 10" sin IJ cos IJ (1 + ~ cos 29) dIJ = 0
Con.trvetion of the primitive machine 29S

Rotor

.. 1

,
..12" ,
r/l! ,
3..kl! 2",In
I ..0
C-t-' Stator
(al

Rotor

"0
[i[:/~~
r
CL

I Stator
(ol
Fig. 6-15 (0) A developed view of the air gap of the salient n-pole-pair machine.
(b) The surface-current distribution for an n-po\e-pair winding.

Very often machines are constructed with a salient structure having


an angular periodicity less than 11" rad. In order to model these machines
by using the primitive machine, the more general air-gap construction
shown in Fig. 6-15a will now be considered. The number n is any selected
positive integer greater than zero. For n = 1 the air-gap length g(8) has
a periodicity of 11" rad, and we have the structure previously considered.
Similarly, for n ::: 2, the stator has twice the number of pole projections
and g(8) ~as a period of 11"/2 rad. In general, the Fourier-series expansion
for the air-gap length now becomes
g(8) = go - gl cos 2n8 + ga cos 3(2n8) + ... (6-73)
As an approximation we shall retain only the first two terms of the
expansion; thus
g(8) "" go - gt cos 2n8 (6-7,4)

For a nonsalient air gap g(8) becomes equal to go, since the saliency factor
gl is zero.
In Sees. 6-2 and 6-3 the surface currents established by the primitive­
machine stator windings were assumed to have a periodicity of 211" rad.
After expanding the surface currents in a Fourier series and retaining only
296 PrillCip/e, of .Iectromechanical.energy conve",ion

the fundamental terms, we have used

J~ = -K: cos 0 i;a. amp/m (6-75)


J4 = Kd sin () ida. amp/m (6-76)

as approximations for the quadrature- and direct-stator surface-current


distributions. Figures 6-3 and 6-4 illustrate the assumed distribution for
the quadrature-stator winding.
The vector-magnetic fields resulting from these two stator-current
distributions are given respectively by Eqs. (6-14) and (6-20), which for
convenience are repeated as

B' ~oK;ai~ sin 0


q =f --g(O-)- ar (6-77)
B' ~o K'·'
datd cos 0
d = g(O) a. (6-78)

For a nonsalient air gap, Eq. (6-16) and Fig. 6-!} show us that a magnetic
field in the air gap with the form of Eq. (6-77) results in a uniform field
inside the rotor cylinder in the positive quadrature-axis direction. The
cross-sectional view of the rotor, as shown in Fig. 6-5, resembles a bar
magnet with a single north and south pole. The quadrature-stator wind­
ing establishes a two-pole field inside the rotor. In a similar fashion the
direct-stator winding establishes a two-pole field along the direct axis.
Thus current distributions with a periodicity of 2'lr rad establish a so-called
two-pole field inside the rotor.
Machines having a salient stator structure with an air-gap periodicity
of 1I'/n radians have corresponding surface-current distributions with a
periodicity of 211'/n radians, as shown in Fig. 6-1.')b. Therefore, the funda­
mental terms in the quadrature- and direct-stator surface-current Fourier­
series expansions are given by

J~ =- K~ cos nO i;a. amp/m (6-79)


Jd = Kd sin nO ida. amp/m (6-80)

By use of Ampere's circuital law, with a contour 'lr/n radians in length,


the magnetic fields corresponding to the currents in Eqs. (6-79) and (6-80)
are given by

(6-81)

(6-82)
ConstlVct ion of the primitive machine '197

r Quadrature axis

Oirect axis

Fig. 6-16 The air-gap ma.gnetic field


estahlished hy a. six-pole quadrature­ N
sta.tor winding. I

The magnetic field inside the rotor cylinder with the quadrature­
stator air-gap field, given by (6-81), would resemble a magnet with a
total of 2n poles around its periphery. Figure 6-16 represents the case
when n = a, showing only the maximum (both positive and negativej
values of the air-gap field speeified by Eq. (6-81). By convention, nux
emanates from the north poles and enters the south poles, as shown by the
arrows in Fig. 6-16. The integer n thus specifies the number of pole pairs
established by the current dist,ribution. For n = 3 we have six poles, or
three pole pairs, as shown in Fig. 6-16.
For stator-surfaee-current distributions having the general form of
(6-79) and (6-80), the primitive machine has rotor-surface-current dis­
tributions with the same form. Because of the commutator-and-brush
switching, the rotor-surface currents are actually uniform with a periodicity
of 2r/n radians. As usual, only the fundamental term in the Fourier­
series expansion is retained, giving

J; = - K; cos nil i;a, amp/m (6-83)


J~ = K~ sin nil ida, amp/m (6-84)

as approximations for the quadrature and direct-axis current distributions.


On examining the two-pole armature winding shown in Fig. 6-7, .we
can see how a multi pole winding can be established. Suppose two addi­
tional brushes were placed on the front-end connections joining sides 4-4'
and 10-10' making a total of four brushes, two on the horizontal axis and
two on the vertical axis. Inserting a current i into each of the horizontal
brushes and extracting a current i from each of the vertical brushes will
establish the current distribution shown in the developed view of Fig. 6-17.
298 Principle, of e/ectromecbcll'licu/••nergy conver,;on

w'

\
GI/
'3'1
I
81
I
7~
I 6"1 5"
I
4'1
1
3"
I
2~
I
1~
1 11"I 10~
1
I I I t t t I I 1 t
-L~ I
I
1
,,
I I
I
, ,, ,
I
I
1 I
I
I
I
I
1
1
,
I
I
I
t
I
1
I I I I I I
,,
I I
I ,
I I
I I
I
I
I
I ,
I I
1 ,
I I
I
, I
,
I I I I 1 1 I
,I 1

•, ,
I I I I
l~
I I
I I

I '3 I 8 I
I I
I
I
I I
I
,
I
, I
, 5 , 4 I 3 I 2 I 1
I
I
I
I
7 I 6 12 I 11 I 10

ut ~2i

Fig. 6-17 A four-brush two-pole-pair primitive-machine rotor winding.

Conductors on the outside of the rotor have the following eurrent


distribution:

Conductors 2, 3, 4 = + 2i a.
~
Conductors 5, 6, 7 = - 2 a.

(6-85)

Conductors 8, 9, 10 = + 2i a.

~
Conductors 11, 12, 1 = - 2 a.
I
Now moving all of the conductors, located on the inside of the rotor,
r/2 rad around to the outside gives a stationary current distribution
having a periodicity of r rad. With four brushes on the same commutator,
we have established a four-pole (i.e., two-pole-pair or n = 2) direct-rotor
winding. By using additional pairs of brushes, multi pole windings of
higher orders can be established. The conductors located on the inside of
the rotor cylinder are moved r/n radians around to the outside of the rotor,
where n is the number of pole pairs established by the winding. Brushes
Construdion of the primitive machine 299

are '/fIn radians apart on the commutator, with alternate brushes con­
nected together to make n parallel rotor circuits. A multipole winding
constructed in this manner is known as a lap winding.
By rearranging the active coil sides o'i'ilj()th, the inside and outside of
the rotor cylinder shown in Fig. 6-7, multipole windings using only two
brushes can be established. This alternate winding scheme contains only
two parallel paths and is known as a wave winding. Since the active coil
sides for a wave winding are connected in series in only two parallel paths,
this winding can generate higher voltages than an equivalent lap winding.
However, the lap winding with its n parallel paths between the terminals
is capable of handling larger port currents. l
Notice in Fig. 6-17 that the currents in the active sides of the winding
arc i/2 amperes, whereas the terminal winding current is 2i amperes. In
Eqs. (6-83) and (6-84) the currents i; and i:i are the terminal or port
currents for the respective windings. -~ 'The winding distribution factors
K; and K:i must be calculated such that J~ and Jd represent the funda­
mental term in the Fourier expansion for the actual current distribution
on the rotor surface. l"or a specific rotor winding the surface-current
distribution should be expressed in terms of the terminal current, and not
the '!urrents in the actual active coil sides. A Fourier-series expansion for
the surface-eurrent distribution as a function of the terminal current will
yield the required expressions for the distribution factor. This formula­
tion of the distribution factors allows us to assume that the priluitive­
maehine mulLipolc winding is a single continuous winding actually carrying
the terminal current in the active coil sides. Thus the winding" establish
the respective n-pole-pair sinusoidal surfa{~e-current distributions. 2
The air-gap magnet.ic fields resulting from the rntor-surface distribu­
tions given by (6-~3) and (6-84) are found by Ampt:.e's circuital law to be

B' _ J.'oK;ai; sin nfJ (6-86)


II - ng(fJ) a,
' _ J.'oK:lai:i cos nfJ
Bd - ng(fJ) a, (6-87)

With currents in all four primitive-machine windings the total


magnetic field in the air gap is the sum of the component fields given by
Eqs. (6-81), (6-82), (6-86) , and (6-87). The total magnetic energy stored
in the air gap can still be found by using Eq. (6-38). Analogously to

I For a. complete di8cu88ion of the commuta.tor-brush multipole rotor winding see

A. S. Langsdorf, "Principles of Direct-currcnt Ma.chincs," 6th ed., McGraw-Hill Book


Company, New York, 1959.
• Problems 6-3 and 6-5 expand on these poin{.d for a two-pole-pair wave and lap
winding.
300 Principles of electromechanical-energy conversion

Eq. (6-39), the energy stored is now given by

t'i!:/J: I02"(!(~~~ _~()_s.~~_:±-~_;~;_s.in n:(t__~~~~_c()s_n~±~;~~!,in ~r d8


(6-88)

Assuming that gl go
« and by using Eq. (6-74), the reciprocal of the air­ ?
gap length can be approximated by

-1-
g(8)
~ ]
-,
go
(gl
1 + -- cos 2n8)
go
(6-89)

Afte~ integrating Eq. (6-88), by using the approximation of (6-89),


the resulting energy expression can be compared with Eq. (6-33). The
following self- and mutual-inductan{'e parameters are thereby defined:

Ld = ~o(],3[.If(Kd) 2 (1 + Jl!. ) henrys (6-90)


n2go 2go

n2yo
3
U = l"oa 11l'(K;)2 (1 gl) henrys (6-91)
q 2go
L~ = lJ.oa l1l'(Kd)2
3

n2go
(1 + Jl!.
2go

) henrys (6-92)

Lr =
q
l"oailr(Krp
n 2go
q
(
1 Yl)
2go
henrys (6-93)

M~r = lJ.oaalrK~Kd
2
(1 + Jl!. ) henrys (6-94)
n yo 2yo
Mor = l"oa311l'K~K; (1 _ _g~ ) henrys (6-95)
q n 2go 2go
The self- and mutual-inductance parameters for the multi pole wind­
ings are redu{'ed to the expressions found in Sec. 6-6 for the two-pole case
simply by setting n = 1. By using a derivation paralleling Sec. 6-7, the
rotational inductances for the multipole windings are found to be

(6-96)

(6-97)

(6-98)

(6-99)
ConS#rvction of the primitive machine 301

Again notice that these expressions all reduce to the one-pole-pair case
originally considered in Sec. 6-7 when n is set equal to L
In deriving the self- and mutual-inductance parameters given by Eqs.
(6-90) to (6-95) the following three definite-integral forms are used
repeatedly:

/02r cos2 n6 (1 + gl c~: 2n6) d6 ,.. (1 +i;o)


//r sin 2 n6 (1 + gl c~: 2n6) d6 11" (1 - i;o)

{2r
10 cos n6 sm n6 1
. (
+ gl cosgo 2n6) d6 = 0
The rotational inductances require the evaluation of similar integrals over
appropriate limits spanning only 11" instead of 211" rad. Dividing the integral
values by 2 provides the results for these reduced limits.

6-9 PRIMITIVE-MACHINE RESISTANCE PARAMETERS

The resistance of the stator winding can easily be determined by using the
relationship
R = pL (6-100)
A
where R is the winding resistance, in ohms; p is the resistivity of the con­
ductor material, in ohm-meters; L is the total length of the winding
conductor; and A is the cross-sectional area of the conductor. For a
sinusoidally distributed direct-stator winding the number of turns in an
increment of arc l~ngth a d6 is given by Kd sin n6 a d6; therefore, the total
number of direct-stator turns in an angle spanning 11" In radians is

None = {r/n K~ sin n6 a d6 = turns (6-101)


..,ne 10 n
The distribution factor Kd: is formulated on the basis that the port current
id actually proceeds through all of the winding turns taken in series.
Therefore for an n-pole-pair winding, we would have n such zones on the
stator inner periphery, and the total number of turns is simply n times
the number of turns in one zone. Thus we have for the total number of
turns
N = 2aK~ turns (6-102)

The length of each turn is simply (2l +


4a) meters, since each turn
has two active sides of length 1and two end connections each with a length
302 Principles of electromechanical-energy conversion

equaling the rotor diameter. The assumption that the rotor radius is
much greater than the air gap is again being used. Using Eq. (6-100),
the resistance of the winding is given by

' _ 4aK~(l
Rd - A
+ 2a) ohms (6-103)

Similarly, the quadrature-stator resistance is given by

R' = 4aK:zp(l + 2a) ohms (6-104)


9 A I
I
since the quadrature-stator winding distribution is similar to the direct­
stator diStribution. In fact, all of the windings on the primitive machine,
I
both on the rotor and on the stator, are assumed to have a sinusoidal ;J
turns distribution. Therefore, by symmetry, the rotor-winding resist­
ances are given by :/
!
il
' _ 4aKdP(l + 2a) "f!
Rd- A ohms (6-105)
R' = 4aK;p(l + 2a) ohms (6-106)
9 A

The resistivity p and cross-sectional area A in each of Eqs. (6-103) to


(6-106) are shown as being identical. For correct results the actual p and
A for each of the windings must, of course, be used. Quite often the end
connections are not placed diametrically across the ends of the rotor
cylinder as we have assumed. The winding length may therefore have to
be adjusted in these equations. Also in the case of the rotor windings,
the brush and commutator resistances, which in a practical case are not
negligible, must be added to R~ and R~. For windings which are not
sinusoidally distributed Eq. (6-100) must be applied directly, and Eqs.
(6-103) to (6-106) will give approximate results only.

6-10 IDENTICAL ROTOR DISTRIBUTION FACTORS


A very interesting set of relationships exist between the rotational induct­
ances and the self- and mutual-inductance coefficients if the two rotor
windings are established by simply placing two pairs of brushes, at 7r/2
rad, on the same commutator. With such an arrangement the distribu­
tion factors KiJ and K; will be equal. Also, by constructing two identical
separate windings, with two separate commutators, KiJ can be made equal
to K;. In either event, if K'd = K;, then from Eqs. (6-92) to (6-99) we
Construction of the primitive machine 303

Fig. 6-18 Rotation d


of the quadrature-axis
brushcs illustrating the
relationship betwecn the
rotational inductances
and the self- and mutual
inductances.

have the following equalities:


G~~ = nM'; (6-107)
G;~ = nM',l (6-108)
Gd.~ nL; (6-109)
G;~ = nL'd. (6-110)
The significance of these relationships can be seen by examining Fig.
6-18. The quadrature-rotor winding can have rotationally induced
voltage from currents in the direct-stator and/or the direct-rotor wind­
ings, as indicated by the coefficients G;d and G;d, respectively. If the two
quadrature-rotor brushes were rotated on the commutator 1f /2n radians
in the negative angular, or clockwise, direction, then effectively the quad­
rature-rotor winding would be on the direct axis. Since we have assumed
K'd. = K;, the quadrature-rotor winding, when relocated to the direct
axis, appears identical with the direct-rotor winding. Now between the
direct-stator winding and the relocated quadrature-rotor winding we
would have a mutual inductance. This mutual inductance is identical
with M',l, because the relocated quadrature-rotor winding is identical with
the direct-rotor winding. Similarly, the mutual inductance between the
relocated quadrature-rotor winding and the direct-rotor winding becomes
L'd, since leakage fluxes are being neglected. I
Thus for K'd K;, Eq. (6-108) shows that the rotational coefficient
between the quadrature-rotor and direct-stator winding equals n times
the mutual inductanee between the direct-stator winding and the quad­

1 The mutual inductance between two identical windings with unity coupling equal8
the self-inductance of either winding. These two self-inductances are also equal to
each other.
304 Principles of electromechanical-energy conversion

rature-rotor winding with its brushes relocated to the direct axis. Equa­
tion (6-110) shows that the rotational coefficient G';d is equal to n times
the mutual induetanr:e hetween the direr:t-rotor winding and the relocated
quadrature-rotor wiuding, with this mutual inductance becoming a self­
inductance because of unity coupling. Equations (6-107) and (6-109) pose
a similar relatioIlship upon rotation of the direct-rotor brushes 1r/2n
radians in the counterclockwise direction.

6-11 SUMMARY
A primitive-machine model is to be used as an analysis tool for investigat­
ing a wide class of electric machinery. The stator and rotor windings of
this machine are constructed in a fashion which is sufficiently general to
allow for an analysis of all machines commonly encountered. The rotor
structure is a smooth cylinder of high relative permeability with windings
placed in slots around its periphery. Connections to points between the
various loops of the rotor winding are made by means of a commutator
"lrudurc and stationary pairs of brushes. The purpose of this arrange­
ment is to create a magnetic-flux-density veetor which is stationary in
space and completely independent of the rotor position. The stator
windings, being placed on the stator structure, create' a similar vee tor
field completely independent of the rotor position.
The analytic form of all magnetic-flu x-density vectors is taken to be
radially directed in the air gap between the stator and rotor. The form
of the space dependency is assumed to be sinusoidal with a maximum
value equal to the actual maximum of the fundamental component in a
Fourier-series expansion of the periodic magnetic fields. In the designing
of many actual machines an attempt is made, as nearly as possible, to
achieve this result. If the air-gap fields are radial in direction with a
magnitude varying as a sinusoidal function of the polar angle, then the
magnetic-flux-density vector inside the rotor cylinder of high-relative­
permeability material has a uniform magnitUde directed along either the
direct or quadrature axis.
Neglecting leakage fluxes and saturation effects, expressions for the
various self-, mutual, and rotational inductances are developed. Certain
interesting relationships between these coefficients, given by Eqs. (6-107)
to (6-110), exist under the assumption of sinusoidally distributed, radially
directed air-gap fields and identical rotor-winding-distribution factors.
In the next chapter these coefficients are used in formulating the equilib­
rium equations for the complete primitive machine. The approximations
used in the construction of the n-pole-pair primitive machine and the
derivation of its parameter are as follows:
Construction of the primitive machine 305

1. The permeability of thc rotor and stator iron structures is suffi­


ciently large compared with 1'0, the permcability of air, that the magnetic­
intcnsity vcdor H in the iron mm be tnkcn as bcing zcro. The iron is
assumcd to have negligible saturation cffects for the range of fields com­
monly cncountered; thus supcrposition of scparatc fields is valid. All
leakage or fringing fields are also neglected.
2. The air-gap length between the stator and rotor varies periodically
and can be expanded in a Fourier series. If only the first two terms of the
series expansion are retained, the air gap ean be approximated by
g(O) "" go 01 (~os 2nO (6-111)
The rotor is a smooth cylinder, and the stator has the salient-pole projec­
tions. The saliency factor gl is much less than the average air-gap length
go, so that
1
g(O) ~ go
1 (gl
1 + fio cos 2nO
) (6-112)

Also, the radius of the rotor a is mueh larger than the air-gap length; thus
a + g(8) "" a (6-113)
a. The direel- and quadraturc-stator windings are assumed to have
their active eoil sidcs sinusoidally distributed around the inner periphery of
the stator. The cffects of the slots in the stator structure holding the coil
sides arc ncglccted. The sinusoidally distributed stator winding estab­
lishes a surface eurrcnt which is axially diredcd with a sinusoidal angUlar
variation. The dircet- and t[uadrature-stator windings carrying currents
ij and i;, respc(·tivcly, ('stahlish eurrent shcets as follows:

J~ = Kj sin no i~a, amp/m (6-114)


J: = - K~ C08 nO i:a, amp/ill (6-115)

where the distribution factors K~ and K; have units of turns per meter.
4. The direct- and quadrature-rotor windings are assumed to estab­
lish similar current sheets on the rotor surface. These rotor windings are
supplied through a commutator-and-hrush mc('hallism which maintains
the rotor-eurrent shects stationary in spaee completely independently of
the rotor angular position or velocity. Thus we have taken

Jd = Kd sin nO ida. amp/m (6-116)


J; = - K; cos nO i;a. amp/m (6-117)
The rotor-current sheets given hy (6-116) and (6-117) are actually the
first terms in the Fourier-series eXJlansioll of the adual periodic rotor­
current sheets. The active coil sides of the rotor windings cannot be
306 Principles of e/ectromechonical-energy conversion

sinusoidally distributed because of the commutator switching mechanism


required to maintain the current sheets stationary in space.

PROBLEMS

p-l A four-pole machine hlLll an air gap as shown in Fig. P6-l.


a. Expand thc air-gap length (1(8) in a Fourier aerics.
b. Compare the magnitude of the third term in the series with the magnitude of
the second term.
c. Find g(l1) in the form of a Fourier serieR for a machine having 2n salient poles
on the stat<1r. The integer n is the numher of pole pairs.
I

Rotor

d, j "4 2 9, t ~ 3"{2 1- •e

C-i fl ~2

Slolor

Fig. P6-1
1
6-2 For the stator-eurrent distrihutions shown in Fig. P6-2 sketch the surfaee­
current distribution and the air-gap magnetic-flux density determined graphically by
using Arnp/,zoe'l! eircuital law. Since the atator has eight slots, a!!8ume the current
density to be uniform over each 21</8 rad or ± (2,../16)a meters (where a is the radius
of the stator) around the center of each slot. Assume that each of the active coil
sides shown hy a dot or cross contains N conductors. Each conductor carries the
terminal current of i amperes in the directions indicated. List all of the approxima­
tions you use in solving for the air-gap flux density vector.

Fig. P6-2
Construction of the primitive machine 307

6-3 Figure P6-3a shows a. developed view of a stator with a four-pole double­
layer full-pilch wave winding. The stator strueture has 12 equally spaced slots with
slot 1 at 6 = 0, slot 2 at 6 "./6, etc., for the remaining slots. The coils are so wound
that cach active side contains 15 turns connected in series. The coils are con­
'nected into two separate circuits labeled A-A' and R-B'. The term double layer
means that each slot contains two active coil sides. 'Full pilch indicates that the
active sides for each coil are 'lrln radians apart, where n is the number of pole pairs
developed by the winding. In this case n = 2 and thus the coils with sides in slot
1 have their other sides in slot 4, which is 11'/2 rad away. The term "wave" refers to
the way the end connections are formed. See Fig. 1>6-00 for an illustration of a lap
winding.
a. For the connections shown in Fig. P6-3b show that the stator-surface current
can be approximated by
I!lOi ,.. 3".
J = - -".a a. for II < 6 < 2 and ... < 6 < '2
J = 18~ a. f or ".
2 < 6
a.r <
< 11' an d '2 6 < 211'
'Ira
where a is the mean radius of the stator.

11' 211'
, • IJ

A
8 (a)

-,-...
I
-,-.
I

li/z

..i­
0
- i

(b) (e) (d)


Fig. P6-3
308 Principles of electromechanical-energy conversion

b. By expanding the stator-surface current in a Fourier series, show that the


fundamental term is given by

J "" ?20i sin 211 a,


,..'a
and thereby show that the distribution factor is
720
K turns/m
Tn
c. Use the steps outlined in parts a and b to find the distribution factors for the
connections in Fig. P6-3c and d.
d. Find an expression for the air-gap magnetic-flux-density vector for each of the
three connections.
Note: The winding in Fig. PG-3a is a modified wave winding. In practice a
thirt('entll slot would he required.
6-4 The machine in Proh. 6-3 hal! a. rotor diameter of 0.3 m, a rotor length of
0.7 m, and a smooth air-gap length of 0.002 m. For a terminal current i = -2 amp
in each of the connections in Fig. P6-3b to d:
a. Caleulate the fundamental component of the air-gap magnetic-field vector.
List all approximations used in your calculations.
b. Find the self-inductance at the two terminals.
c. Model this stator winding in terms of the primitive-machine stator.
6-5 The stator winding shown in Fig. P6-5« is a four-pole full-pitch double­
layer lap winding. The term full pitch means that the two coil sides for each coil
are separated hy ,../n radians, where n is the number of pole pairs. For this two-pole­
pair (or four-pole) winding, n "" 2 and the coil sides are .../2 rad apart. Thus the
coil with one side in slot 1 has its other side in either slot 4 or slot 10,1<'/2 rad away.
The winding is double layer, since two coil sides are in eaeh slot. The term "lap"
refers to the way the end connections for this type of winding overlap each other.
The stator has a total of 12 slots. Each of the slots contains two coil sides, as shown
in the figure. The individual coils are constructed with 10 turns in scrips and arc
joined as shown in the figure into four separate circuits A-A', B-R', C-C', and D-D'.
For each of the connections indicated in Fig. P6-5b to e:
a. Draw an axial developed view similar to Fig. 64.
b. Write a set of zonal equations, with the form of Eqs. (6-21), expressing the
stator-surface current in terms of the terminal current i.
c. Expand the equations of part b in a Fourier series and from the coefficient
of the fundamental term find the distrihution factor K.
6-6 Repeat Prob. 6-4 for the four connections of the lap winding in Prob. 6-5.
6-7 Figure P6-7 indicates how the four-pole full-pitch double-layer lap winding
in Prob. fJ.. 5 can be connected to a 12-segment commutator and used as a rotor winding
to establish a stationary current distribution. The small coils in Fig. P6-7 represent
each of the 24 active coil sides. The numhers on the coils represent the rotor slots
containing the coil sides. The lowercase u and I indicate upper and lower coil sides
in each of the 12 rotor slots. The small numbered squares in Fig. P6-7 stand for
the 12 segments on the commutator. The four separate windings in Fig. P6-5«
are connected as follows: A to D, B to C, A'to R' , C' to D', as shown in Fig. P6-7.
Four stationary brushes ride on the commuta.tor, as indicated by the shaded squares.
The two positive brushes' are connected together, as arc the two negative brushes.
Condruclion of the primitive machine 309

TT/Z TT 2."
I I I .0

c
o (a) 0'

-i

(d)

(b) (e)

(e)
Fig. P6-5

The series of arrows around the windings indicate motion of the coil sides with respect
to the stationary brushes for positive angular rotation of the rotor.
a. Carefully examine Fig. P6-7 and verify its oorrespondence to Fig. P6-OO.
Make 0. drawing similar to Fig. 6-10 for this rotor winding.
b. Indicate the brush positions for a r/6-rad rotation of the rotor, and show that
the rotor-surface-eurrent distribution is stationary with respect to the stator.

310 Principle, of e/ectromechanical.energy conversion

Ir= A lu 41 2u 51 3.. 61 A'

-Brush II
\
8'

Fig. P6-7

c. Determine the distribution factor for the winding; assume that each coil
contains 10 turns in series. (Hint; See Prob. 6-5, particularly Fig. P6-5e, for an
outline on finding K.)
d. Repeat part c with the rotor turned through 1</2 rad.

6-8 Repeat Prob. 6-4 for the rotor winding shown in Fig. P6-7.
6-9 a. By expanding the integral given by Eq. (6-39), develop the self- and
mutual-inductance expressions given by Eqs. (6-42) through (6-48).
b. Why can Eq. (6-39) for the total magnetic energy stored in the air gap be
equated to Eq. (6-33), which represents the coenergY-Btate function?
c. Give a physical explanation showing that all the mutual inductances in Eq.
(6-48) are zero.
6-10 Derive Eqs. (6-96) to (6-99) for the rotational inductance coefficients for a
multi pole winding. In this derivation explain why the integrals are evaluated over
limits spanning 1< rad. (Nok; These limits are equivalent to using a span of ... /n.
radians and then multiplying the result by n.)
6-11 The rotor structure of a nonsalient primitive machine has 56 slots. A two­
pole double-layer full-pitch lap winding consisting of 56 coils is placed in the rotor slots
and connected to a 56-segment commutator. (Note: This rotor winding is identical
with the winding in Fig. 6-10.except that 56 coils are used instead of 12.) Each rotor
coil contains five turns. The stator contains two identical sinusoidally distributed
two-pole windings in quadrature. The stator distribution factor is K' - 2900
turns/m. The physical dimensions of the machine are as follows: rotor core aD ­
4.585 in., stator ID (bore) - 4.631 in., rotor core length - 3.0 in., stator core length ­
3.0 in. Both the rotor and stator are wound with double-stranded No. 20 copper
wire having a diameter of 0.032 in. per strand. One pair of brushes ill placed on the
commutator ... rad apart, and they are so aligned that the rotor circuit has an axis in
Construction of the primitive mochine 311

line with one of the stator windings. A second pair of brushes is placed in quadrature
with the first pair, thus establishing a second rotor circuit in line with the other stator
winding. The total primitive machine can thus be represented by Fig. 6--13.
Owing to the slots used to hold the windings on the rotor and stator structures,
the air-gap length must be increased. By using a flux plot, the reluctance between
the actual rotor and stator surfaces can be determined for the actual slot configura­
tion. An effective air-gap length is defined as the length between two smooth rotor
and stator structures giving the same reluctance. For the primitive machine in this
problem the effective air-gap length is 2 times the distance between the rotor and
stator teeth.
a. Calculate the resistallce R' of each of the rotor windings. The resistivity of
copper is PCu - 1.72 X 1O-8 ohm-m. Hint: Between either pair of brushes there are
two parallel paths of 28 coils each.
b. Calculate the resistance R' of each of the stator windings by using (6--103) or
(6--104).
c. Determine the distribution factor K- for a sinusoidal approximation for the
rotor winding. Hint: Follow the steps outlined in parts a and b of Prob. 6--3.
d. Determine the self-inductances of all four primitive-machine windings.
e. Determine the mutual inductances between each pair of primitive-machine
windings.
f. Determine all four rotational inductances for the machine.
g. With a 2-amp current in each of the stator windings, find the magnitude of
the open-circuit voltage across each pair of brushes when the rotor is stationary and
when the rotor revolves at 126 rad/sec. Note: The parameters used in this problem
have been taken from the Westinghouse generalized machine. If the actual machine
is available, your answers could be checked experimentally.
chapter

EQUILIBRIUM EQUATIONS
7 FOR THE PRIMITIVE MACHINE

In Chap. 6 the form of the individual stator and


rotor windings is introduced. The self~ and mutual
inductances for the four windings of the primitive
machine, in addition to the rotational inductances,
have been formulated in terms of the physical con~
stants of the model machine by using the approxima­
tions listed in Sec. 6-11. In this chapter the cir­
cuital equilibrium equations for the primitive
machine are to be studied. These equations are
nonlinear in their general form.
The linearization technique discussed in Chap.
./') will be used to obtain incremental transient opera­
tion about an operating point. Using this procedure
on the complete pr,mitive maehine allows us to
formulate, in a general way, the steady-state4per­
ating equations and the linear incremental differ­
ential equations to be used in our later work. By
utilizing a special form of the ideal transformer, a
complete equivalent circuit representing the prim­ "
itive machine can be formulated. This circuit is
quite useful in the analysis of actual physical
machines.

7-1 REFERENCE DIRECTIONS


The primitive machine shown in Fig. 7-1 has four
windings, two on the stator and two on the rotor.
The stator windings are of the type described in
Sec. 6-2, creating a magnetic field in the rotor
cylinder along the axis upon which the winding is
placed. The direct axis has a positive reference to
the right, and the quadrature axis has a positive
sense directed upward, as indicated by the arrows
312 on each axis in Fig. 7-1. It is assumed by conven­
Equilibrium equations for the primitive machine 313

Quadrature axis q

Oirect
axis d

Fig. 7-1 The complete four-winding


d-q primitive machine.

tion that the magnetic field in each stator winding is in the positive axial
direction for positive winding currents. For example, id, the direct-axis­
stator current, if actually flowing as indicated ill Fig. 7-1, creates a field
directed to the right in the rotor of the machine. The dots originally used
on the windings in Chap. 6 are being omitted with the understanding that
reference currents are always to be chosen into the aots.
The two rotor windings are of the commutator-and-brush variety
discussed in Sec. 6-4. These windings create magnetic fields which are
stationary in space and eompletely independent of the rotor position.
The rotor fields are assumed to depend only on the currents in· the respec­
tive rotor windings. As before for the stator windings, the direction of
the rotor currents is chosen to create a uniform magnetic field in the rotor
cylinder oriented in the positive direction for the respective axes. Thus
i;, the quadrature-rotor current, if flowing in the assumed direction, creates
a uniform field in the upward direction in Fig. 7-1.
The terminals of the four electrical windings and the mechanical
shaft of the machine represent places where energy can flow in or out of
the primitive machine. The term port is commonly used to denote these
energy entrance and exit points. The assumed positive sense for all
winding terminal voltages is 80 chosen that energy is flowing into the
winding or port if the respective port power is positive. Thus if the
winding current is actually flowing in the assumed direction, then a posi­
tive terminal voltage indicates an instantaneous energy flow into the
winding. For an electric motor, which converts electrical energy to
mechanical energy, we would expect that average port power would
actually be positive at the electrical ports of the device. However, with a.
generator, converting mechanical energy to electrical energy, average
power will
be negative at certain electrical ports.
314 Principles of elec:tromechonicol-energy conversion

For the single mechanical port, physically the shaft of the machine,
the assumed positive direction for angular velocity wr is taken to be
counterclockwise, as shown in Fig. 7-1. This direction corresponds with
the usual direction for positive angle_ In order to again have a positive
mechanical-port power, corresponding to energy flow into the mechanical
port, the assumed direction for the externally applied torque Tr is also
taken to be counterclockwise. If energy, in reality, is flowing out of the
mechanical port, then either wr or Tr will turn out to be negative, or
opposite to the assumed positive direction, and the mechanical port
power will be a negative quantity.
If all of the terminal quantities, namely, Tr and w' for the mechanical
port, and all the v's and i's for the electrical ports are positive, then all
port powers are positive and energy is flowing into all ports of the prim­
itive mJchine. This would mean that the device must be either storing
or dissipating (in the form of heat) all this accumulating energy. For an
energy-conversion device such operation does not seem very satisfying;
consequently, we shall be expecting energy actually to flow out of at least
one of the five ports.

7-2 ELECTRICAL-EQUILIBRIUM EQUATIONS


With the terminal reference conditions shown in Fig. 7-1, we shall now
formulate the equilibrium equations for all the electrical ports of the
machine. Since magnetic mutual coupling is present, the loop formula.­
tion is more convenient than the nodal formulation. Our job is to set
down the appropriate loop equation for each of the four windings in the
primitive machine.

Direct-stator Winding
The equation for this winding is given by
• - R'"
Vd - dId
+V d
di~
dt + M" de
d
did (7-1)

where R~ = resistance of the direct-stator winding


L~ = self-inductance of the direct-stator winding
M~' = mutual inductance between the direct-stator and the direct­
rotor windings!
Notice that the mutually induced voltage is positive in Eq. (7-1). This
is a result of having the positive directions of t~ and t~ both creating a
I Equations for all machine parameters in terms of the construction details of the

primitive machine are summarized in Sees. 6-8 and 6-9.


Equilibrium equations for the primitive machine 315

field in the same direction. If the conventional dot notation were included
on the two windings, both currents would be into the dots, resulting in a
positive sign on the mutual term. The positive sign preceding the mutual
term indicates a voltage opposing the flow of the direct-stator current i~.
No further terms are needed in Eq. (7-1), since all other windings are in
quadrature with the direct-stator winding.

Quadrature-stator winding

Following the reasoning for the direct-stator winding, the equation for
this winding is given by

(7-2)

where R~ = resistance of the quadrature-stator winding


L~ = self-inductance of the quadrature-stator winding
M~' = mutual inductance between the quadrature-stator and quad­
rature-rotor windings

Quadrature-rotor winding

The equation for the rotor windings becomes slightly more complicated,
since rotational voltages must be included. The loop equation for the
quadrature-rotor winding is given by
d" d"
v'q = R'i'
q q + U::::!.J.
q dt + M"~
dt -
q Grow'i'
qd d - G"w'i'
qd d (7-3)

where R~ = resIstance of the quadrature-rotor winding


L~ = self-inductance of the quadrature-rotor winding
M~' = mutual inductance between the quadrature-rotor winding
and the quadrature-stator winding
a-;d rotational inductance between the quadrature-rotor winding
=
and the direct-stator winding
G';d = rotational inductance between the quadrature-rotor winding
and the direct-rotor winding
Note that the first sub- and superscripts always represent the winding in
which the voltage is induced and the second sub- and superscripts stand
for the whiding conducting the current which induces this voltage. The
four-letter notation is present on all G's. The sign on the mutual voltage
term M;' di;/dt is positive, since the currents i~ and i;, as shown in Fig.
7-1, cause magnetic fields in the same direction. Let us first determine
the sign on the rotational terms by an intuitive discussion and then check
these results with the development in Sec. 6-7. The sign on the rotational
terms can be determined by examining whether the rotationally induced
316 Principles of electromechanical-energy conversion

q Direction of induced

currents due to 0 field

in the +d direction

Assumed

di rection of i;

~~~---4------ ~------~d

Fig. 7-2 The rotationally induced


currents in a two-pole quadrature-rotor
winding due to rotation in a direct-axis
field.

voltage iends to aid or hinder the flow of l~. If the induced voitage
hinders the flow of t~, then the sign of the term should correspond to that
for the resistance and self-inductance terms in Eq. (7-3), which certainly
oppose the flow of i~. Figure 7-2 shows the distribution of currents on the
periphery of the rotor set up by a two-pole quadrature-rotor winding and
i;. t As shown, the left half of the rotor has currents out of the paper and
-the right half has c:urrents into the paper. With both the direct-rotor and
direct-stator windings creating a field directed to the right, the induced
currents in the rotor conductors, moving in the assumed direction for the
rotor velocity, would flow as shown by the inner circle of dots and crosses.
This direction is determined by forming the vector cross product of the
velocity of the conductors with the direction of the direct-axis field.
SiIlce the induced currents are in the same direction. as the assumed
quadrature-rotor current, the induced-voltage term is aiding the flow of
~~. Consequently, the rotational terms are negative on the right-hand
side of Eq. (7-3). Equation (6-64) indicates that the voltage induced in
the quadrature-rotor winding from the direct-stator and direct-rotor wind­
ings has a positive sign. The reference sense on this induced voltage
was originally chosen in the development of Sec. 6-7 to be positive in a
direction aiding the conventional direction of t~. Thus the intuitive
argument presented here is consistent with the original derivation of the
terms in Sec. 6-7. ...

Direct-rotor winding

In a similar fashion the eq'lttion for this winding is given by


r + U did + M" did
lJrd = Rri
d d d dt dt + G"wri' + Grrwrir
d dq q do 0
(7-4)

t For the purposes of this qualitative discussion neglect the iact that the surface­
current density has a. sinusoidal variation.
Equilibtium equations for the primitive machine 317

where R~ = resistance of the direct-rotor winding


L~ = self-inductance of the direct-rotor winding
M"J = mutual inductalwe between the direct-rotor winding and the
direct-stator winding
G~~ = rotational inductance between the direct-rotor winding and
the quadrature-stator winding
G';q = rotational inductance between the direct-rotor winding and
the quadrature-rotor winding
The rotational terms for the direct-rotor coil are positive, as may be seen
from Fig. 7-3. Notice that the rotationally induced currents are opposite
in direction to the assumed positive direction of i:i. Therefore, the induced
voltage is hindering the flow of i~, and thus it must have a positive sign on
the right side of Eq. (7-4). The original development of these two rota­
tional terms in Sec. 6-7 gives a corresponding result, as can be seen by
checking Eq. (6-69).
A simple rule that is helpful in remembering the sign associated with
these rotational-voltage terms is as follows: If rotation of the rotor coil
axis by r/2 rad into the axis of the eoil inducing the voltage is made in the
assumed reference direction for rotor angular velocity, then the rotational
term is positive; if opposite to this diredion, the term is negative. Thus
referring to Fig. 7-1, the d-axis rotor winding must be rotated in the posi­
tive w' direction to line up with the q axis, so all rotational voltages in the
direct-rotor winding contain positive signs. However, the quadrature­
axis-rotor winding must be rotated oppositely to the direction of positive
w' to become aligned with the d axis, so negative signs appear with rota­
tional voltages in the quadrature-rotor equation. Obviously this rule is
still valid if the assumed direction for w' is reversed.

Directions ot induced
currents due to a field
in the +q direction

Assumed
direction of i/

~~+---+------e~-----.d

Fig. 7-3 The rotationally induced


currents in a two-pole direct-rotor wind­
ing due to rotation in a qua.drature-a.xis
field.
318 Principles of e/ectromechanica/.energy conversion

The four electrical-equilibrium equations (7-1) to (7-4) are now col­


lected, and for convenience they are written in matrix form as follows:

where the symbol p is used for the differential operator d/dt. t By using
an abbreviated matrix notation, Eq. (7-5) can be written more con­
veniently as

~; = (61 + ~p + Wr~)id~ (7-6)


I
where

~r = v~
dq ":J
v~

vq
r

(7-7)

td
.,
·.r _ 'to'

i;'J
Idq - 'r (7-8)
td

6l =
[R: 0
0
R'q
0
0
(7-9)
0 0 R~
. 0 0 0 iJ
~~]
0 M';
[ L:
~ = ~';
L'q 0
(7-10)
0 L~
M"q 0 Lrq

~]
0 0

~= [ 0
0 0 0

(7-11)
Gd~
-~;d
0
0 -
Grrqd

Equation (7-6) is a very general form representing the loop equations for
the electrical ports. If more than four ports were present, the form of
this equation would not change. The number of terms in each of the
individual matrices would increase, but no additional matrix quantities
would be required.

t This symbol p is not to be interpreted as the Heaviside or Laplacian variable. It


stands for d/dt and nothing else, as discussed in Chap. 4.
Equilibrium equations for the primitive mom;ne 319

7·3 MECHANICAL-EQUILIBRIUM EQUATIONS


The fifth and remaining port in the primitive machine, shown in Fig. 7-1,
is the mechanical PQrt, or the shaft of the machine. The equilibrium
equation for this port is formulated in precisely the same fashion as the
electrical equations were. The torque externally applied to the shaft
must be opposed by certain generated torques. These generated torques
include an inertial effect, a viscous or windage torque, a compliance torque
due to the twisting of the shaft, and lastly a torque exerted on the rotor
from an electrical origin. This torque of electrical origin is expected,
since the rotor periphery contains currents which exist in a magnetic field.
If the mechanical torque of electrical origin were not present, then there
would exist no interaction between the electrical and mechanical ports and
there would be no possibility of any energy-conversion process.
The mechanical.port equation is, therefore, given by

T' = J dw' + Dw' + -K1 JI- .. w' dt + T. (7-12)

where T' = externally applied torque


J total polar moment of inertia of the rotor
D = angular viscous friction coefficient
K = compliance of the rotor shaft
T. = torque of electrical origin
In any practical machine, the twist in the shaft is small enough to be con­
sidered negligible. Therefore, the compliance torque is to be neglected
in all subsequent torque equations. Figure 7·4 shows the reference direc­
tion for all of the torques in Eq. (7-12). Positive T' means an external
torque is applied in the direction of positive angular velocity w'. All the
remaining torques oppose this applied torque and are taken as being posi­
tive when opposite to the direction of positive angular velocity.
The torque of electrical origin T. can be determined by applying the
first law of thermodynamics. This basic physical law states that the net
total energy supplied to a system can only increase the system's stored
energy. In formulating our primitive machine, only electrical and

Fig. 7-4 Reference directions for all


shaft torque quantities.
320 Principles of electromechanical-energy ~onversion

Electricol Mechonicol

ports port

r-------,
p;
pi Loss
pi' If

{p;
dq Storoge Fig. 7-5 The first law of thermo­
p'It dynamics applied to the d,..q primitive
machine.

mechanical ports have been included. A heat port has been neglected,
because with electric machines energy flow, in the form of heat, is only
outwardly directed.' The quantity of energy leaving the primitive­
machine system, in the form of heat, must be supplied from either the
mechanical or electrical ports or from the stored-energy supply. There­
fore, for lour system, the first law can be expressed as follows:

The algebraic sum of the energy flow into the electrical and mechanical ports
of the primitive machine is accountable in either energy storage or energy
loss in the form of heat.

This statement of the first law is valid at each instant In time;


therefore
P';[o dt + P;,. dt = (power into storage + power into loss) dt (7-13)
where Pdq and P;" are the total instantaneous powers at the electrical and
mechanical ports, respectively, and dt is the differential time increment.
Figure 7-'/) illustrates this interpretation of the first law. When dt is
canceled from both sides of Eq. (7-13), the first law results in an instan­
taneous energy flow or power balance.
The total instantaneous power into all the electrical ports is given by
~=~+«+~+« ~~
where we have used the sign convention established in Fig. 7-1. Sim­
ilarly, the total instantaneous power into the mechanical port is
P' (7-15)
'"
Equation (7-14) can be written in a simpler form by using the matrix forms
established by Eqs. (7-7) and (7-8); thus,
(7-16)
where (idq)! means the transpose of the column i4 matrix, in other words,
the row iik matrix.
The total instantaneous power supplied at all five ports is the sum of
Eqs. (7-15) and (7-16); therefore,
(7-17)
Equilibrium equations for the primitive machine 321

in which p~t"l is the total instantaneous power at all ports of the machine,
Pd~ is the portion at the electrical ports, and P;" is the portion at the
mechanical port. When Eq. (7-6) for v~~ and Eq. (7-12) for TT are sub­
stituted into Eq. (7-17), the total power is expressed by .
p~~tAl = (ido)tlR(id~) + (id~)t£p(i~~) + (id;)twT(J(id~)
+ wTJpwT + wTDwT + wTT. (7-18)
The various terms in Eq. (7-18) can readily be identified as follows:
Power loss = P~":' + P:;:·· (7-19)
where (ido)llR(id"q) is ohmic or electrical power loss and wTDw T is viscous
mechanical or windage loss.
Power into storage = P'd~()red + p~or.d = (idq)'£p(i do ) + w'Jpw'
(7-20)
where (ido)t£p(i::q) is power going into storage in the magnetic fields and
w'Jpw' is storage in the form of kinetic energy in the rotating inertia. A
compliance torque would result in mechanical storage in the form of
potential energy in a twisted shaft. I
The remaining two terms ill Eq. (7-18), since they are not contained
as contributors to either loss or storage, must equate to zero. Therefore
(i::.)tw'(J(id~) + w 7'. = 0
r (7-21)
and thus the torque of electrical origin is given by

T. = - \
w
(id~)twr(J(i:rq) = - (i~~)I(J(i~~) (7-22)

When Eq. (7-2~) is expanded in terms of the matrix expressions for i~~ and
(J, given by expressions (7-8) and (7-11), the torque of electrical origin for
the primitive machine becomes

T. =
o
o
0]
0 "i
G~~:~
d
]

-G';a 0 i;
(7-23)
For the special case when the two rotor windings have the same turns
distribution factors (that is, K~ = K~), Eqs. (6-107) to (6-110) give
equivalent relationships for the rotational inductances in terms of the

I Capacita.nce effects in the windings would a.ccount for electric-field energy storage.
On the basis that this quantity of energy is negligible, these electric-field effects have
also been neglected.
322 Principles of electromechanical· energy conversion

rotor self-inductances and the stator-to-rotor mutual inductances. With


the condition that Kd = K; the torque of electrical origin given by (7-23)
becomes
(7-24)
For unequal distribution factors Eq. (7-23) must be used.
A somewhat simpler procedure for formulating T. is available by
defining the quantity p. as
p. (i~~)IW"~(i~) (7-25)
in which case p. is the portion of the total power, supplied at the electrical
ports, not going into either loss or storage. The quantity p. may be
formed Py expanding the triple matrix product of Eq. (7-25). However,
a more convenient method is simply to take all the terms in the four
electrical-equilibrium equations involving rotational voltages and multiply
by the port current for which the equation was written. The electrical
torque is then readily determined as

T = - p. (7-26)
• W,.

If the quantity p. is instantaneously positive, then T., according to


Eq. (7-26), must be negative. A negative T. means that the torque of
electrical origin is in the direction of positive angular velocity. This
means the electrical currents are generating a torque which is trying to
turn the rotor in the direction of positive speed. If the rotor actually
turns in the direction of T., we' have motoring operation, or the conversion
of electrical power to mechanical power. The quantity p. is the portion
of the electrical power supplied to the electrical ports that is converted to
mechanical power.

7-4 MULTIWINDING IDEAL TRANSFORMER


Before considering certain practical machines, a rather unusual, idealized
circuit element is to be introduced. Discussions involving equivalent
networks for our primitive machine will make use of this device. The
usual ideal transformer is shown in Fig. 7-6. The port currents are defined
positive when taken into the dotted sides of each winding. In addition,
the port voltages are taken positive in such a fashion that energy flow is
into the port for positive current and voltage. This convention is con­
sistent with that selected for the primitive machine in Sec. 7-1. The ratio
of turns is marked in Fig. 7-6 as 1: N 2, indicating that winding 2 has N 2
times the number of turns as on winding 1.

_
.. - - _.. ......- .---_._..._ - •....... - - - _ .

Equilibrium equations for the primitive machine 323

Fig. 7-6 The two-winding ideal trans­


former.

By definition the ideal transformer imposes the port constraints


V2 = N,.vl (7-27)

i2 = - ~ i
N2 l
(7-28)

following the reference directions of Fig. 7-6. Multiplying Eq. (7-27) by


i, gives
(7-29)
upon substituting Eq. (7-28). These three equations indicate that the
device transforms the magnitudes of the applied voltage and current var­
iables while keeping the power constant or invariant. Receiving a power
PI vdl at port 1, the device transfers this energy, without any loss, and
delivers an equal power P 2 = -v,i 2 at winding 2. The minus sign
indicates energy flow out of port 2, provided this flow is into port 1. The
ideal transformer can be very nearly approximated by a properly designed
physical transformer for voltage and current functions having a limited
magnitude and frequency content.
An extension of the two-winding ideal transformer is shown in Fig.
7-7. One winding is drawn on the left side of the vertical core lines, with
(k 1) windings ,on the right side. All the port variables are defined by
using the standard convention described for the two-winding ideal trans­
former: .. Also notice in Fig. 7-7 that the turns ratios have been defined in
a fashion similar to that for the two-winding device of Fig. 7-6.
The defining port constraint equations for this multiwinding trans­
former are as follows:

(7-30)
v. = Ni;VI

and
(7-31)
If energy flow is into port 1, then Eqs. (7-30) and (7-31) give
" (7-32)
324 Principles of electromechanical-energy conversion

Nz

[~}

+{:Ji.."
"'1
,
t} :
I

F.}'-,
Ideol Fig. 7-7 A multiwinding ideal transformer.

showing that the power is again invariant. All of the energy supplied to
port 1 is found emerging from the remaining ports. The device is ideal
in the sense that it loses no energy and simply transforms the magnitudes
of the voltages, aceording to Eqs. (7-30), and imposes the current con­
straint of Eq. (7-31). Notice that only one port voltage can be inde­
pendcnt.ly specified. However, (k - 1) currents could be arbitrarily
selected.
One further point should be mentioned with regard to the constraint
definitions given by Eqs. (7-30) and (7-31). These relationships are
taken to apply without regard to the form of the voltage and current time­
dependent functions. Thus, for example, if VI(t) were a constant d-c
value, then Vk would still be given by Nk times this value and would also
be a constant provided Nk was a constant. As we shall use this multi­
winding transformer in following sections, the number of turns on certain
windings will be found to be functions of certain port currents.

7-5 PRIMITIVE-MACHINE EQUIVALENT CIRCUIT

The complete set of equilibrium equations fQr the primitive machine


shown schematically in Fig. 7-1 is obtained from Eqs. (7-5) and (7-12).
After substituting the torque of ele()trical origin from Eq. (7-23) into Eq.
(7-12), the five equilibrium equations, one for each of the five ports, are
repeated for convenience as follows:
Equilibrium equations for the primitive machine 325

'l]
u'q
Ud
-
-
[R: +0 L;p
lW;p
R~
0
+
G~~u{
L;p
R~
M~rp

0
+ D'dp
on
M';p
G'daw'
i~
i'd
(7-3:!)
v~ -G;'aw r M;'p - G"
qd W r R; + L;p i;
T' (Jp + D)w' + (G;~ - G'da)i'di; - Gd~i;~~ + G;~id~~ (7-34)

By using the multiwillding ideal transformer, introduced in Sec.


7-4, the five equilibrium equations given by (7-33) and (7-34) can be
represented by the equivalent circuit of Fig. 7-8. Notice that the com­
plete circuit involves ordinary mutual couplings between the direct-axis
stator and direct-axis rotor windings and also between the quadrature­
axis stator and quadrature-axis rotor windings. Small square and tri­
angular markings are used to represent the sense of these windings. In

Direct
stator port

Direct
I':i rotor port

Rotor
port
w:{ J

o---..I--...i-----l
+

Quadrature
rotor port

Ideol . 1.1",/
f f

~
~
}:! ,
Quadrature
stator po,.t

R:
Fig. 7-8 The complete equivalent circuit for the five-port do{] primitive maehine.
326 Principles of electromechanical-energy conversion

addition, a five-winding ideal transformer is included in Fig. 7-8. Dots


are indicated on these windings to indicate induced voltage and current
polarities. The turns ratios between the windings on the right and the
single winding on the left are indicated just above each winding. Observe
that these ratios are functions of the electrical port currents with the
left-hand winding taken .as unity. Thus the number of turns on four
of the five windings is variable, and in general it would be time-dependent.
The loop equations for the four electrical ports are seen to be identi­
cal with Eqs. (7-33). The voltage across the rotor port winding is w',
the rotor angular velocity, thereby fixing the voltages on the other four
windings of the ideal transformer. For example, the top right-hand
winding in the direct-rotor loop has an induced voltage, taken positive
on the dotted side of the winding, equal to (G:i~i:)w'. This voltage term
has th~ same reference sense as the voltage across R:i or L:i, and thus it
takes the same sign in the direct-rotor port-loop equation. The third­
row second-column term in Eq. (7-33) verifies this conclusion. Carefully
check that the right side of the circuit in Fig. 7-8 does indeed yield the
four equations of (7-33).
For the rotor port, a nodal equation must be written. The sum
of "currents" through the capacitor J and conductance D gives the first
terms on the right side of Eq. (7-34). The applied torque T' is a source
appearing on the left side of Eq. (7-34). The remaining terms in the
equation represent the torque of electrical origin, given by T. in Fig. 7-8.
Equation (7-31) gives the ideal transformer equation whereby T. may
be calculated, thus verifying the remaining terms in (7-34). Since T.
is into the dotted side of the left-side winding, the remaining currents,
when out of the dotted side of the right-side windings, contribute positive
terms in the equation for T.; when into the dotted side, they contribute
negative terms. Thus, considering the center two windings, (G;~id)i;
is a positive contribution to T. and (G~2~)id is a negative contribution,
which corresponds to Eq. (7-34).

7-6 D-C STEADY-STATE AND INCREMENTAL OPERATION


The equilibrium equations for the primitive machine given by (7-33)
and (7-34) are nonlinear differential equations in their general form.
Terms involving products of the rotor angular velocity w' and various
port currents are contained in the direct- and quadrature-rotor equilib­
rium equations. Also, the rotor-port equation (7-34) contains products
of the electrical-port currents. Unknown dependent variables appear­
ing as products make the equations nonlinear.
In Chap. 5 we studied a linearization technique whereby all system
Equilibrium equations for the primitive machine 327

forcing functions were assumed to be composed of a d-c or static level


plus an incremental t.ime variation about this level. By using such an
assumptioll, all 10 port variables for the primitive machine can be
expressed by equations with the following form:

Vd = VdO + Vdl id = ido + i d1 (7-35)


v~ = v~o+ V;l i~ = i~o + i~l (7-36)
v~ = v~o+ V~l i~ = i~o + ~~l (7-37)
v; = v;o+ V;l i; = i~o + i~l (7-38)
Tr= To + T~ w r = w~ + w~ (7-39)

where the subscript 0 refers to the cOllstant quiescent or d-c portion and
the subscript 1 indicates the time-dependent incremental portion.
These five sets of equations can now be substituted into the port
equilibrium equations (7-33) and (7-34). Since the resulting equations
must be valid at the quiescent points, with all increments equal to zero,
two sets of port equations result. The first set of five equilibrium equa­
tions involve only the steady-state quiescent portions of the port vari­
ables. The second set of five equations involve the incremental portions
of the variables as unknowns with the quiescent values appearing in
certain coefficients. If all products of incremental quantities are neg­
lected, the resulting sets of equations are as follows:

Steady-state quiescent equations:

o
o (7-40)
R~
-G~dw~
(7-41)

Incremental linear equations:

o Mdrp
R~ + L~p o
G~;w~ R~ + L~p
Jlf~rp -G~dw~
-G~;i~o (G;~ - G~~}i~o - G~;i;o

o
M";p
G~~wo
(7-42)
R~ + L~p

(G;~ - Gd~)i~o + G;~idO

328 Principles of electromechanical-energy conversion

Equations (7-40) and (7-41) supply information concerning the doc


steady-state operation of the four-winding primitive machine. Equation
(7-42) can then be used to examine how the machine operates about
this steady-state or quiescent condition.
The doc steady-state operation of the complete primitive machine
can be studied by reducing the complete equivalent circuit, shown in
Fig. 7-8. In the d-c steady state all inductors are replaced by short
circuits and capacitors by open circuits. All port voltages and currents
for the electrical ports, and torques and speeds for the mechanical port,
are assumed to be constant at their quiescent value. Five port termina­
tion conditions, or constraints, must be specified to allow a network solu­
tion to be computed. As we shall see in the following chapters, the
selection of these constraints determines the mode of operation.
Ad this point a word of "confidence building" is probably necessary.
Equations (7-40) to (7-42) represent a total of 10 equations required to
specify the linearized operation of the complete primitive machine. At
first glance the solution of such an array of equations might appear to
be a hopeless job. However, when considering practical machines some
amount of simplification is generally possible. For example, the two­
winding doc machine to be considered in the next chapter is actually
just a special case of the more general formulation of the complete primi­
tive machine. By eliminating appropriate terms in Eqs. (7-40) to (7-42),
the doc steady-state and incremental time-dependent equilibrium equa­
tions to be used for the two-winding case can be obtained.
The purpose of formulating the equivalent network of Fig. 7-8 and
Eqs. (7-40) to (7-42) for the five-port machine is to demonstrate the
general technique whereby the operation of any machine ean be exam­
ined. The generalized philosophy of examining the operation of a non­
linear device by considering the device to be linear for small perturbations
about a steady-state quiescent point is a very useful and practical tooL
Where such an approximation is not valid, the nonlinear differential
equations given by (7-33) and (7-34) must be studied.

7-7 SUMMARY
A general analytic model machine has been formulated to aid in the
analysis of a wide class of electromechanical energy-conversion devices.
This model, named the primitive machine, consists of two magnetic
axes in space quadrature. On the stator of the model machine two
windings are placed, one on each axis. The rotor structure also has one
winding on each axis. The rotor windings are connected to their external
Equilibrium equation. for the primitive machine 329

terminals by means of brushes and a commutator. The purpose of the


commutator is to maintain rotor currents stationary in space and inde­
pendent of the rotor angular position or velocity.
Equilibrium equations for the four electrical ports and one mechani­
cal port are formulated in terms of a rigorous sign convention specifying
the assumed positive sense of all terminal quantities. The interaction
between mechanical and electrical quantities consists of certain rotational
volta.ges as functions of the rotor angular velocity in the electrical
equations and a torque of electrical origin in the mechanical equation.
Through these interaction terms the energy-conversion process occurs.
A multiwinding ideal tra.nsformer is introduced to assist in formulat­
ing an equivalent lumped-element network representation for the primi­
tive machine. The ideal transformer changes the magnitude of the
applied-voltage and -current variables in accordance with the respective
turns ratios. Thus if one winding voltage is specified, then all other
voltages are fixed according to Eqs. (7-30). However, all but one of
the winding currents can be independently specified corresponding to
Eq. (7-31). For a k-winding ideal transformer, (k - 1) currents may
be picked in addition to one independent voltage. These k choices, in
addition to the k constraints of Eqs. (7-30) and (7-31), make up the
required 2k conditions to specify all k currents and k voltages. Further­
more, the constraint equations are so chosen that the device is completely
lossless, and the power remains invariant. As used in· the primitive­
machine equivalent circuit, the turns ratios are functions of certain port
currents. Therefore, in general, these turns ratios are actually functions
of time, since these currents are time-dependent.
The general set of nonlinear differential equilibrium equations for
the five ports of the primitive machine are expanded into a set of d-c
equations and a Set of linear incremental differential equations. Such a
linearization technique provides a tool for obtaining pieces of response
information. Very often, practical-machine problems require only a
steady-state solution. For machines used in many control systems only
the incremental transient response is required for a stability analysis.
Consequently, the linearized set of equations are, from a practical stand­
point, very important.
The following chapters use the primitive machine as a starting point
in unfolding the operation of many outwardly different forms of machines.
Through certain coordinate transformations these other machines can
be correlated with the single primitive machine and suitable equilibrium
equations can be determined. By inverse transformations the quanti­
ties thereby determined can be converted back to the actual machine
being studied.
330 Principles of electromechonico/-energy conversion

PROBLEMS

7.1 For each of the following conditions on the primitive machine shown in
Fig. 7-1 describe the vector-magnetic field inside the rotor cylinder. Assume the
stator structure to be nonsalient. Assume also that all four windings have identical
distribution factors. All currents are in am peres, and the speed is in radians per
second.

.,
Condition i:' i~ t. i"
• ",'

4 1 1 1 1 0
b 1 1 I 1 100
c 1 1 -1 -1 100
Ii 0 0 C08 t sin t 100
e I cos t 0 0 sin t -100

7-2 The primitive machine shown in Fig. 7-1 has four windings with distribu­
tion factors K~ = K: and K~ - K; and the following parameter values:

Rd = R; = 10 ohms R~ = R; = 1 ohm

I'd = L: = 6h L'd = L; .. 1.5 h

M:r = M:' 3h G~~=G;~ .. 6h

J = 20 kg-m t G,:.=G;'d=3h

D = 0.5 newton-m-sec/rad

4. Is the stator of this primitive machine salient or nonsalient 1 Why? Is the


rotor salient or nonsalient?
b. How many pole pairs does this primitive machine have? (Find n.)
c. Write the three parameter matrices contained in Eq. (7-6) in terms of the
parameters of this primitive machine.
7.3 The primitive machine described in Prob. 7-2 has the following port
constraints :

ido = t~o .. 2 amp (d-c) t-;' "" 6 amp (d-c)

i do - 4 amp (d-c) "'0 4 rad/sec

4.Find the four steady-state port voltages.


b. Find the total steady-state power being supplied to the two stator ports.
Where is this power going?
c. Find the magnitude and direction (Le., clockwise or counterclockwise) of the
torque of electrical origin T,.
d. Determine the total steady-state power at the two electrical rotor porta.
e. Determine the appli"lfi torque To
and the power at the mechanical rotor port.
/. Make a complete summary of the power transferred through the machine,
including input or output at each port, all losses, and the converted power.
7-' A IO-volt d-c voltage source is suddenly connected to the direct-stator
winding of the primitive machine described in Prob. 7-2 at time t .. O. The rotor
speed is maintained constant at "'~ = -7.0 rad/sec. All other electrical ports are
open-circuited.
Equilibrium equations for the primitive machine 331

a. Find the direct-stator current id as a function of time.


b. Solve for the voltages appearing at the terminals of the remaining three elec­
trical ports as time functions.
c. Determine the external torque required to maintain the rotor speed constant
at -7.0 rad/sec.
7-5 Repeat Prob. 7-4 with the addition of a short circuit on the quadrature­
rotor port preceding the application of the 10-volt d-c source on the direct-stator port.
7-6 The direct-stator port of the primitive machine described in Prob. 7-2 is
excited by a voltage source such that

v:'(I) - ~ 10 sin 1.51

The rotor speed is maintained constant at "'~ = 4.0 rad/sec. With a short circuit on
the quadrature-rotor port and the other two electrical ports open-(lircuited:
a. Find all of the steady-state port currents.
b. Determine all of the steady-state port voltages.
c. Find th.e average torque of electrical origin and the average torque required
to maintain the rotor speed at 4.0 rad/sec.
d. Calculate the average steady-state power at each of the five ports of the
machine and prepare a complete tabulation of the power transferred through the
machine. Include in this tabulation the input or output at each port, all losses, and
the converted power.
7-7 a. By expanding the matrices in Eqs. (7-19) and (7-20), show that these
equations correctly represent the power into storage in the primitive machine.
b. (Optional) For sinusoidal excitation on the electrical ports of the primitive
machine show that the steady-state ohmic power loss is given by

P~":' = Re [(I:[.)':R(I d.) oJ

where I::.is the column matrix for the port currents in sinor form and the asterisk
indicates conjugation of the matrix.
7-8 The two totor windings on the primitive machine, with the rotor revolving
at constant speed, can serve as a transformer capable of handling direct voltages
and currents.
a. Write the d-(l steady-state equations for the two rotor windings.
b. Derive an expression for the resistance of a load RL connected to the quad­
rature-rotor port as seen from the direct-rotor port under d-(l steady-state conditions.
c. For the connections of part b derive an expression for the efficiency of the
d-(l transformer. Neglect rotational windage losses, and assume that R'd and R;
are very small.
7-9 The following parts are to be answered for' an ideal transformer having
five windings.
a. How many port variables (voltages and currents) describe the state of the
transformer?
b. How many constraint equations between port voltages does the transformer
impose?
c. How many constraint equations between currents does the transformer
impose?
332 Principles of e/ectromechonico/-energy conversion

d. Are these constraint equations holonomic? Why?


e. How many voltage and current variables can we choose arbitrarily?
J. If resistance loads are connected to four of the windings, how many voltage
and/or current variables can we choose arbitrarily?
7.10 The three-winding transformer shown in the circuit of Fig. P7-10 is an
ideal trantJJormer. Find the voltages across each of the windings and the two loop
currents.
III

2n

rdeal Fig. P7-10


7·11 a. Explain why all windings on an ideal transformer cannot, in general,
be connected in series carrying the same current.
o. Also explain why all windings cannot, in general, be connected in parallel
in a network.
c. Determine the conditions under which statements a and 0 are not va.lid.
7-12 A metadyne generator can be represented by the primitive-machine model
shown in Fig. P7-12. The machine rotor is being driven at a constant speed "'0
rad/sec.
a. Draw an appropriate equivalent circuit representing the metadyne generator
including all port constraints.
o. From your equivalent circuit write the equilibrium equations for the meta­
dyne generator in matrix form.
c. Indicate the torque of electrica.l origin on the circuit diagram of part a. Find
an expression for T. by using the ideal transformer-eurrent constraint equation (7-31).

Gqf

v;=-.f2 Vsin wt
RL
Fig. P7-12 Fig. P7-13
Equilibrium equations for the primitive machine 333

7-13 Figure P7-13 shows a primitive-machine model of a polyphase induction


machine. Assume. that the shaft ill coupled to a constant load torque; thus T' ..
- TL, where TL ill constant.
a. In terms of the conventional primitive-machine parameter notation, draw a
complete equivalent circuit showing all port constraints.
b. From the equivalent circuit, write the five equilibrium equations in matrix
form. Are these equations linear?
c. Would the specification of T' - T L be correct for a passive load torque?
Explain your answer.
Direct-current commutator machine, (II) 401

Fig. 9-16 Complete bloek-diagram representation for the equilibrium equations


for the amplidyne.

where we have defined

Tf =1'1
- seconds (9-36)
R,
(9-37)

seconds (9-38)
ohms (9-39)
By carefully designing the compensating winding, the feedback and
feed-forward paths in Fig. 9-16 can be eliminated. Following the form
of Eq. (6-94) the mutual inductances between the control field f and the
compensating and armature windings, for a two-pole machine (n = 1),
are given by

M,. = p.oa' l1l'K,Kc


go
(1 + 1l.!.)
2go
(9-40)

M,. = p.oa' l1l'K,Kd (1 + 1l.!.) (9;-41)


go 2g0
where K, and Kc and Kd are the respective winding distribution factors.
Similarly, the rotational inductances G"c and Gqd from Eqs.(6-97) and
(6-99) hav;e the form

G,,< = p.oa'l1l'K"K c
go
(1 + 1l.!.)
2go

(9-42)
3l

Gqd = p.oa 1l'K"Kd


go
(1 + .1!)
2go
(9-43)

where K, is the quadrature-winding distribution factor.


402 Principles of electromechanical·energy conversion

By designing the eOlllpensatillg winding such that Kr = K d , we have


M/, = MJd (9-44)
and
(9-45)
If K, = K d , the maehine is lermed fully eompellsated. Also, with
K, = K d , Eqs. (6-90), (6-92), and (6-94) ean be used 10 show that,
(9-46)
thus making Ta 0, as seen by examining Eq. (9-:~8).
By substituting the equalitirs (0-44) to (9-46) into the blo('k diagram
of Fig. !)-Hi, we have a fully compensated amplidyne without the feed­
forward a1nd two feedhaek paths. Sin('e the two feedback paths involving
the load ('urrent if. are now eliminated, a more convenient transfer func­
tion relating the amplidyne open-circuit armature voltage to ('ontrol field
voltage eall now be written as follows:
I'. Ko
Ilf = (~~p+-lrG~p~+-fj

where
K. = fiq!!h7~:'o)2 volts/volt (9-48)

and I'. is the open-eircuit or internally generated amplidyne armature


voltage. The open-('ireuit transfer fUIH,tion given hy Eq. (9-47) ('an
be IIsed with a ('onneeted armature load only if the! machine is fully
compensated.

Example 9·3
The Ward Leonard system discussed in Example 9-1 is to be controlled by adjusting
a. small potentiometer. Since the generator field winding requires more power than

•I Molor
v~
!
Molor rI field

~--.~-~ \ ~II~
230vol15
w.:;..:v
J(£ r,

Fig. 9-17 A Wa.rd Leonard speed-control system using an llmplidyne for an addi­
tional stage of amplification.
Dired-current commUfator machines (II) 403

('8n t'allily he ohtaim'd from a small potentiofll!'ter, a fully ('ompensnted amplidyne


is used 8S a rotational amplifi!'r as shown in Fig. \l-17. The amplidyne is bt'ing
driven at a ('onRtant "pcI'd "':" = 11'00 rpm. Tht' amplidyne pammctt'T!\ aTe as
follows:
Control field re~istan('e /{fo = 121111 ohllls t
Control field induf'! an('e l,f. = 211() h
Control fipld rot.ational indul'tan('t' G:!, = 1.1) h
Quadmture winding rcsistan('e R~ = 1.0 ohm
Quadrature winding indu('tanf'e L~ = O.G h
Quadrature wi'nding rotational indm'tnncc G~· = O.2fi h
Totl11 armaturf' rp>listanl'.t' /{~ + N~ = Z.O ohms
AI\I\unlC unity ('oupling o!'twe!'n th!' C 11Ild d winding"
Thl' gl'nl'rnl.or and motor )1arnm!'t!'rH IIr!' Rh'!'n in Exampll' n-I. For ('om'!'ni!'n"!'
t.h!' ll;l'npmtor fit'ld rt'Histan!'e and induptllnN' MI' rpp!'ntpd Ill<
Generator field rpRist.ance N£ = 71'> ohms
(,pnl'rator field induptance l,~ = !I.O h

n, /)etprmint' the potentiometer voltage "f. n!quired to have thl> amplidyn!'


Rupply rated voltage of 230 \'olts to the ll;!'nerator field winding.
h. /)esip;n a potent.iomet('r and hattpry circuit with sufficient adjustment to
allow for 120 ppr ,'('nt of th!' ml.!'d p;('nl'rntor fil'ld voltall;('. Call'ulate the power p;ain
of tilt' amplidynt'.
('. By IlHinp; thp rpslllt" of I';xnmpl!' !l-I, dl,tprlllinl' t.1l(> o\'prall frcqut'nf'Y reRponse
of the ('ompi!'te ..yst{'111 shown in Fig, !I-17.
Solulion: a. From E'1. (9-4S) the open-I'ircuit jl;ain of II fully eompensated
amplidyne, usinll; the notation of this ('xample, is given hy

By 8uhRtitutinjl; param('ter valucR, with the !!pced "'~. in radians per sC(,ond, we have

K. I.;; X.tJ:.t~_.X IS9' = 11 6 volts/volt


120n X 1.0 .,

Thc generator fi('111 winding is in a volt.ajl;e-divider ('ireuit "dth the amplidyne armature
!'ir!'uit. r!'1'i8tanc'('. TheTpfore. the st,('ndY-fltate jl;pnerator field voltage is

II:.....R, + R'
R: + R: K ".
~.----."----.- vI

For IJ~. - 230 volts we have

II~. -
2:~() X (7i)
-"
+ 2.0)

..' - - 20.4 voltfl

71) X 11.6

as the required potf'ntiomcter output.

t The suo.~pript a rcfcrs to the amplidyne, and tht' :mpt'Tseripts follow the notation
used in this !!('('tion. Thifl notation is therefore I'onsistent with that of Example 9-1.
where suhscripts g and", refer to the gf'nerator lind motor.
404 Prirn;iple' 01 eledromechanicol.energy conver,ion

I,. If the generator field voltage is to he 120 per cent of the rateel value, thp
battery V in Fig. 9-17 must have a value not Ipss than
V ~ 1.2 X 20.4 ... 24 ..'> volts
The potentiomt'ter R, for a linear voltagf' adjustmf'nt, shoulel be approximately
10 pt'r eent of the {'ontrol ficld resistan!'!'. Thprefore
R - 0.1 X 1200 == 120 ohms
With the potentiometer tap slightly down from its maximum setting, the current in
the pot!'ntiomeh>r above thp tap is approximately
v
I - -
R
+ -R!t" ~
24.5
-120 24.5
+ -1200 == 0.22 amp

The potentiometer wattage rating eorr""pondinp: to thi" WOTllt ron <lit ion ill thprf'forl'
Watt,Jge == (O.ZZ): X IZO "" 5Jl watt!!
Thl' powl'r gain of the amplidyne can be caleulated from
. output power (v!o)'IR~

Power gam - Trlp~t-~W& == (~~o)'/R~

By substituting for v~., we have


R~R~K.·
Power gain
(R~ + R: + R:)t

== 1200 X 75 X 01.6)' = 2040

(75 + 2)'
Notice that the voltage gain of the amplidyne is only 11.6, but the power gain for
this circuit is 2040.
c. The overall sinusoidal steady-state transfer function for the complete system
shown in Fig. 9·17 is Bimply a cascading of the functions given by Eq. (9-10) for the
generator and motor and Eq. (9-47) for the amplidyne. However, the generator
field resistance R;
must include the total resistance of the amplidyne armature­
generator field loop. Thp generator field inductance I!! not altered owing to thf' unity
coupling between the d and c windings of the amplidynp. ThuB for R: in Eq. (I)- 11) Wf'
should use 77 rather than 75 ohms.
The overall transfer function, using the parameter values in Example 901, is
therefore
Q;'(j",) 11.6 0.80
---- = -~-- X .. .-----,-­
V!(j",) (0.167j", + 1)(0.6j... + 1) (0.117}", + 1)(0.44j", + I)
For the amplidyne we have substituted
L~ 200
T'• - -R! = - - == 0.167 sec T' = L! = 0.6 '" 0.6 sec
1200 • R! 1.0

The generator-motor portion of the transfer function is similar to that obtained in


part b of Example 901 with only the value of R: being changed.
The frequency response of the system is similar to that of Fig. 9-8 with the
inclusion of two more break frequencies at 1.67 and 6.0 rad/sec. For higher fre­
quencies the asymptote has a negative slope of 80 db/decade and a total phase shift
Direct·current commutator machine, (IIJ 405

of -360·, The amplidyne has added a significant power gain to the system. How­
ever, the bandwidth has been reduced from 2.3 to 1.67 rad/see. Notice that the
amplidync's quadrature-winding time constant governs the bandwidth figure. In
order to have II. large quadrature field, Il! is made as small as possible by an external
short circuit. Thus the quadrature-winding time constant,.: = IJ:/R! becomes large.
The ('omhination of large gains and large time constants tends to make the system
unstahle ir reedback is included.

9-6 THE METADYNE CONSTANT-CURRENT AMPLIFIER


Machines employing two rotor fields are known in general as metadyne8.
The amplidyne ('ollsidered in See. 9-;' is a met,adyne with the addition
of a properly designed eompcnsating winding, The amplidyne is the
most commonly used of all the variations of the metadyne, In this
section we shall examine the steady-state eharacterist.ics of the basic
metadyne. Simply eliminating the compensating or c winding on the
amplidyne gives us the basic metadyne. When driven at constant
speed, the steady-state block diagram of the metadyne is given by Fig.
9-18, which is derived from Fig. 9-16 by setting p = 0 for the doc steady
state and eliminating all parameters involving the compensating winding.
Reducing Fig. 9-18 to a single block, we have as the steady-state
transfer function of the basic metadyne

(9-49)

Since the quadrature-winding resistance Rq is usually made as small as


possible, we generally have
Rq(Rd + RL) « GqdGtlq(<>JcY (9-50)
for a wide range of practical loads R L • Substituting the approximation
given by (9.;')0) into Eq. (9-49) gives us

i LO "'" Gql = K. amp/volt (9-51)


lIfO GqdRf
as the steady-state relationship between the load current and the control
field voltage. Notice that the metadyne without compensation acts as

ItO

Fig. 9-18 D-c steady-atate block diagram representing the bBllic metadyne.
406 Principle. of electromechanical-energy conver$ion

a corlstant-current generator, since Eq. (9-.'H) shows that i r.o is independent


of the load Rr. and is only a function of UfO.

Example 9-,4
A constant-('urrent metadyne jl;f'nE'rator is constructpd hy removing the compensating
winejin!!: from thE' amplidynl' in Example 9-a, with all other parameters rl'maininjl;
f'onstant.
a. Assumin!!: that th!' machine is supplying field current to the same jl;enerator
as in Example 9-:3, ('hpC'k the approximation of Eq. (9-50). Compute the gain K.
of thp metadynf'.
h. Calf'ulate thp powpr lI:ain of the metadyne and make a comparison with the
amplidynp.
Soiulion: Il. Thp ampli<iynp IIn!1 nJ!'tadyne are constructed with a nonaali!'nt
!;tator anll two pairs of hrushps on the Hame commutator. Therefore, for the machine
in EXRJUI>h' !I-a WI' havl'
R, = H. = I.n ohm G•• = G.. 0.26 h
Equation !!l-IlO) ill ...hpckpd as follows:
]("iU.( + /( tJ = 1.0 X 11.0 + 75) = 76 ohms
r;.,I;".lw~)· = m.2G X lK!J)2 24200hmo;

Rim'e 76 is mudl 11'88 than 2420, Eq. (\)-.')0) is certainly reasonahle for this particular
pxampll'.
The metndyne gain from (9-1>1) is

K G,/ 1.5 4 / I
i = "(:•• Iij = n.2!i X 120() = .R rna vo t
Hat!'d field ('urrent for the generator load is \~ = 3.1 amp. With a gain K. "" 4.8
rna/volt the metadyne would require If/o = 650 vo\t1'l to supply rated load current,
as opposed to approximately V/O 20 volts in the <'ase of the amplidyne. The con­
stant-current feature of the metadyne is achieved by large <,urrent feedback at the
expense of rorward gain.
b. The power gain is calculated rrom

= (4.8 X to-,), X 75 X 1200 = 2.1


The metadyne has a power gain of 2.1 compared with 2040 for the amplidyne.

9-7 NONLINEAR CHARACTERISTICS: SATURATION EFFECTS


The primitive-machine model fonnulated in Chap. 6 assumes an iron
structure with infinite relative permeability. The magnetic fields estab­
lished by the various windings are therefore linear functions of the cur­
rents in the windings. Because of this linearity the machine inductance
parameters are constants independent of the winding currents. The
Di,ed·ct/rrent COllllllutator lIIachines (II) 407

Fig. (j.1!I The separately e"rit('d d-c g('n­


('rator h('inll: driv('n at ('onst ant spe('d.

operation of many d-e machines can be predicted reasonably well with


the linear primitive-machine model. ~Particularly if time-dependent 1
operation is restricted to a relatively limited rango'!' the linear parameter
valul's can be carl'fnl\y rhosl'n to give very close analytic correspondence
to the physical machinl'. A very important rlass of d-c generators known
as Bell-excited machine8 depend on the nonlinearities for stable operation.
For the self-excited generator the linear model completely fails to provide
even steady-state operating characteristics. In this section we shall
examinl' the partirular I'fi'ect saturation has on d-c machine steady-state
opl'rating ('harac'tl'rifltiNI and how this form of nonlinearit.y can be incor­
porated into our analysis of the self-exeited machine.
Figure 9-19 illustrates t.he separatcly exrited d-c generator with the
rotor being drivcn al a eonstant speed of Wo rad/ser. The open-circuited
armature voltage is given by
with ia = 0 (9-52)
Equation (9-.52) specifies a linear relationship between the armature
voltage I'. nnd the field cnrrent if. Figure 9-20 illust.rates typical results

Fill:. 9-20 Typical curves showing


the opf>n-circuit armature voltage as
a function of field current for three
values of rotor speed.
408 Principle, of electromechanical· energy conversion

Jo'ig. 9-2l A typical plot showing the


rotational inductance a.'! a function of
the field current for a two-winding d~
machine.

obtained by actually driving a d-c machine at constant speed and meas­


uring the open-circuited armature voltage as a function of the field
current. For 1, = 0 the generated armature voltage is not zero because
of the rJsidual magnetism of the stator iron upon which the field circuit
is wound. Thus, reducing the field current. to zero does not reduce the
stator magnetic field to zero.
As the field current is increased to larger values, the generated
armature voltage tends toward a constant value. At these higher field
values the stator iron is becoming saturated. Saturation means that
an increase in the magnetic intensity H does not cause the same increase
in the magnetic field B. Thus, increasing i, does not change the strength
of the field and the generated voltage does not greatly increase. For
values of field current below the knee in the curves shown in Fig. 9-20,
the linear approximation of Eq. (9-52) is reasonably satisfactory. How­
ever, for operation beyond the knee, Eq. (9-52) cannot be used except
in an incremental sense.
The curves shown in Fig. 9-20 can be approximated by an expression
in the form
(9-53)
where k. is a constant whose value depends on the residual magnetism
of the field iron and Gf and Go, are functions of the field current. From
Fig. 9-20, G.., is the slope of the curve divided by w;. A typical plot of
Ga' as a function of the field current would look like Fig. 9-21. Most
d-c machines are usually operated with field currents above the knee.
Therefore, as i, is increased, G.., tends to decrease as saturation begins
to take place.

9-8 SELF-EXCITED SHUNT GENERATOR

The self-excited shunt genera.tor shown in Fig. 9-22 is one of the most
commonly used of all doC machines. Until the introduction of the alter­
Dired-eulT'lln' commutator machine, (II) .co9

Fig. 9-22 A self-excited shunt gen­


/
erator being driven at constant speed and
supplying power to a resistive load. ~o
nator, all automobile electrical systems employed this generator as a
source of d-c power. The machine shown in Fig. 9-22 is being rotated
at a constant speed of wli rad/sec. With the field current if equal to zero
the machine generates an armature voltage due to residual magnetism.
Now if the field is properly connected across the armature, the residual
voltage generates a current if which adds to the residual field. A
higher armature voltage which in turn further increases the field
current if is then generated. In this manner the machine builds up to
some final terminal voltage and thus supplies power to the load RL.
The open-circuit voltage generated by the armature for a specified
field current, as given by Eq. (9-53), is opposed by the load voltage and
the resistance and inductance of the armature. The armature circuit
equilibrium equation is therefore given by

(9-54)

The remaining equilibrium equations for the self-excited shunt machine


are combined with (9-54) in the block diagram shown in Fig. 9-23. By
reducing the block diagram, the following differential equation relating

Fig. 9-23 Block-diagram repreaentation of the equilibrium equations for the self­
excited shunt generator.
410 Principles of electromechanical-energy conver.ion

I
I
I
1
I
I RaRf
Slope = R, + Ra + -n
I L
I
I
I
I
I

Fig. 9-24 Graphical determination of the stl'ady-state performance of the self­


excited shunt generator for various load resistan('es.

the field current to the residual voltage is obtained:


1.,,/"1
{-RL, p2 + (I-I + L. +--~
I"IR" + -..
RL
1""Rf)
RL
~- P

+ [ RI + R(J. + l!k~~l - wr,G'(i/ ) - W'il}(J.f(i / )]} i l = krwo (U-S!»

Equation (9-55) is nonlinear because G' and Ga , are functions of the field
current if.
The static equilibrium points are obtained by assuming i, = i,o,
where i/o is a constant or d-c value. Equation (9-5.1) thus reduces to

( R, + Ra + !i~~f) i/o = I.:rw~ + woifoG'(ifo) + w[,ifoGaf(ifo) (9-56)

The right-hand side of Eq. (9-.')6) is simply the open-circuit armature


voltage for the particular speed w;. In Fig. 9-24 the open-circuit voltage
is plotted as a function of the field current i /o • The left side of Eq.
(9-S6) is a straight-line funet·ioll of the field current going through the
origin with a slope equal to RI + R. + RaRdRL' Figure 9-24 illustrates
four typical straight-line plots for various values of the load resistance
R L • Notice that, as R r, decrea.'lCs, the slope of these straight lines
increases.
Direct-current commutator machine. (II) 411

The operating points for the machine are obtained when the left-hand
side of Eq. (9-t>6) equals the right-hand side, or when the straight line
in Fig. 9-24 intersects the open-circuit characteristic. For example,
when RL = R LI , the intersection occurs at point A and the field current
has a value i/o = i A • The steady-state terminal voltage of the machine
is simply
(9-57)
so if i/o is known, the d-c load voltage VLo can be calculated. Now by
reducing RL to a value Ru in Fig. 9-24 we see two intersections at points
Band C with corresponding field currents i8 and i e . By multiplying
iB and ic by R" the corresponding load voltages VLO are obtained.
Figure 9-2.1 shows a plot of the load voltages as a function of the
load currents iLO = vLo/R L for the same values of RL as used in Fig. 9-24.
Thus for RL = RLI the load voltage is R,iA , as shown in Fig. 9-25. By
decreasing RL to RL2, we have two possible values of VLO, as shown. In
a similar fashion the remaining load values Ru and RL4 are transferred
from Fig. 9-24 to Fig. 9-25. Notice in Fig. 9-25 that for a load ~L
between Ru and RL2 three load voltages seem possible. We shall show
that the intermediate voltage value, between points C and D, is unstable
and physically never occurs.
In order for the self-excited machine to build up to an appreciable
output voltage, the load RL must be greater than R L2 . Figure 9-25
shows the output voltage to remain fairly constant for all loads greater
than R L2 • If the load is less than IlL2, then the voltage builds up only
to the low value at point C. Further decrease of RL to RLI and RL4

~~ --------­ ---­
~ia

i£Ol s.c.

Fig. 9-25 Output-port characteristics for the self-excited shunt generator.

412 Principles of electromechanical-energy conversion

causes the load voltage to reduce even further to points E and P, respe(~­
tively. For RI- = 0 the field current reduees to zero, the load voltage
is zero, and the load current is given by
.I
tLO =k.wo
_. (9-58)
•• r. R"
as shown in Fig. 9-25. The seif-exeited shunt machine is thus current­
limited and tends to protect itself from a short-circuited load.
The stability of the static operating characteristic show II in Fig.
9-2.5 can be checked by obtaining the linear incremental differential
equation for the machine. Subtracting the steady-state equation (9-56)
from the general equation (9-55) gives us

{ Lt.LI pi + (LI + L.. + LIR" + L..RI.) p


R I• RL RL

+ [ RI + R,. + RR~I - W'{,GGI(il 0) ]} ifl = 0 (9-59)

where ifl is the increment in the field current. Notice in Eq. (9-,19)
that Gal is to be evaluated at the operating point.1 Now R(I. (9-f>9)
gives rise to a stable incremental response provided all the coefficients
are positive. Thus stable operating points are obtained for

R, + Ra + RR~I > w~GQ,(i,o) (9-60)

The inequality of (9-60) is satisfied when the slopes of the straight lines
in Fig. 9-24 are greater than the slope of the open-circuit characteristic
where the two intersect. Thus, for intersections between points C and D,
(9-60) is not satisfied and the operating points are unstable. For all
other intersections, (9-60) is satisfied and we have stable operating points.

9-9 TRANSIENT RESPONSE OF THE SELF-EXCITED GENERATOR

Section 9-8 has developed the steady-state operating characteristics for


the doc self-excited shunt generator. The output-port characteristics
I From Eq. (9-54) the open-circuit generated voltage all a function of the field current ill
l1(i,) ." k..w~ + ",~iP'(i,) + w'Qip.,(i,)
Expanding v(i,) in a Taylor series about an operating field current i,o. we have

v(i,) .. v(i,o) + d~ I
d~, iI.
i,1 + ...
The term v(i,p) hall been used in the steady-state equation (9-56), The coefficient
dl1/di, evaluated at if .. if. is simply ",'Il., (i,p) , and therefore only the term
",'Il.Ai,.)i,1 appears in Eq. (9-59).
Direct-current commutator machine, (II) 413

are shown in Fig. 9-2;'). For a load resistalll'c greater than R1.2 the output
voltage reillains relativt'ly (·onstan!.. Bclow a eriti('ai load resist,allce
the machine call not huild up to its rated voltage and remains at a low
output voltage. In this section we shall examine a graphical technique
for solving the nonlillcar equilibrium equation for the sclf-exdted machine.
For small excursions about an operating condition the linear incremental
equation (9-59) can be used to predict transient. changes from one operat­
ing point to another. However, for the initial buildup of the generator
voltage, from zero to its final operating value, an incremental technique is
very tediou!<.
With most self-excited shunt generators the armature inductance
is very small and can he neglected. Therefore, with L.. "" 0 t.he equilib­
rium equation (9-5.''» for the machine becomes

{(LI + Lk~!!) 11 + [ RI + R" + !f;lt! - w~G'(i,) - w~G.,1(i/)]} i,


= krw~ (9-61)

After sollie slight rearrangement Eq. (U-B1) ('an he put. int,o the form

(9-62)

From Fig. 9-24 we call see that the right,-hand side of Eq. (9-62) is simply
t.he vertical distance from the straight line at, a slope RI + Ra + R.RdR,.
to the open-drcuit. l'haracteristic curve. Referring to Fig. 9-26a, the
right side of (9-62) is the voltage increment t.v.
By separation'of variables, Eq. (9-62) can be rewritten in the form

1~! + 12.Ral ~f:. dil = dt (9-63)


t./·(t,l)

Assuming that. il is zero at. t = 0, Eq. (9-63) can be integrated to give

(
1.1 + I:..IRa) (il(/) di'. = t (9-64)
R ,. 10 t.1I('/)

The left side of (9-64) can be integrated graphically as seen by considering


Fig. 9-26b. The time required to reach the field current at point A is
simply the shaded area under the curve from i l = 0 t.o i l at point A.
Thus by using Fig. 9-26b, Eq. (9-64) can be integrated graphically to
give the field current as a. function of time, as shown in Fig. 9-2&. In
forming Eq. (9-64) we have assumed the field inductance to be constant.
Actually, L f is also a function of the field current if. By knowing the
414 Principles of electromechanical-energy conversion

if
(al

Fi~. 9-26 Graphical determination


of huildup transient for a self-excited
shunt generator heing driven at ('on­
Icl stant speed.

functional relat ionship, we can appropriately alter Fig. !J-2nlJ. The


graphical integration ean still be conduded in the sallle manner.

9·10 AIR-GAP FIELDS IN ACTUAL o-C MACHINES


All the d-c machines we have examined in Chaps. 8 and 9 have been
modeled after the primitive machine. Analytic~ results obtained by
using the primitive-machine model show a dose correspondence to those
obtained with the a<'tual d-c ma(~hine. However, the parameters of the
model, as calculated from machine dimensions hy using the equations
in ReI's. 6-8 and 6-!J, must he altered slightly for most d-e machines.
The parameter corre<'tions arc Ilcc'cseary be{'ll.llse the Btator and rotor
Burfac'e ('urrents in d-c maehines are significantly different from those
distrihutions assumed for the primitive mR{'hine.
Two additional problems arise ill {'onstrul'ting prac·tical d-(' ma(~hines.
Both these problems are con('erned with the ('onIll1utator-hrush switching
Direct-wrrent commutator machines ell} 415

mechanism. Actual Illodern~dny d-(· Illlwhines contain two additional


windings designed to alleviate the comlllutation problems. In this
section we shall study the nature of these two problems. Once the two
addit.ional windings are incorporated into the actual d-c machine, the
commutation difficulties no longer exist and need not be taken into
account in the llnalysis problem. Therefore, the primitive machine
with its ideal commutator-brush mC('hanislll is a fine approximation to a
physieal d-c machine with the two additional windings and a nonideal
commutator·brush mechanism.
Figure 9~27a illustrates, in a developed view, the construction of
most d~c maehines. For simplicity only a two-pole machine is shown.
The stator strul'ture is construetcd with salient pole pieces, including pole
tips, spanning more than r /2 rad. Windings arc placed around these
stator poles as shown in Fig. 9-27a. By use of Ampere's circuital law and
t he assumption that the stator and rotor iron have infinite permeability,
the Illajl;netic~vector intensity H in the ai~ gap wilf be in the form of a
square wave with a ('Onstant. magnitude over the poles and zero magnitude
between the salient stator poles. However, because of the pole tips t,he
vector-magnetic field B in the air gap due to the stator current has t,he form
shown in Fig. 9~27b. The outward radial direction a. is downward in
Fig. 9-27; thus the stator winding as shown appears to be a direct-axis
stator winding on t he primitive machint'o At (J 0 the field is maximum
and in the positive a. direction, and at (J = r the field is in the negative
a. dire(·tion. When the waveform in Fig. 9-27b is expanded in a Fourier
scries, t he fundament al terlll is a reasonable first, approxi mat,ion.
The ('olllmlltl\tor~hrU8h Illcehanism maintains a specified rotor-cur­
rent disl ribul ion I hat is ('olllpletely independent of the angular position or
velocity of the rotor. The uniform current distribution shown on the
rotor ill Fiji;. 9~27~ would establish 1\ triangular magnetie~vector intensity
H in the air gap, as shown by the d-ashed lines in Fig. 9~27c, if the stator
were nonsalienl. Because of the stator saliency, the magnetic field in the
aIr gap due to the rotor currents takes the form shown by the solid curve
in Fig. 9~2ic. N'otice the decrease in field strength in the interpolar
region, at r/2 and :~r/2, where the air~gap length is great. Also observe
that the armature field is not symmetrical in the interpolar regions.
Where the armature field in Fig. 9~27c is in the same direction as the stator
field shown in Fig. U~27b, the pole tips become saturated, thus tending to
redu('e the field strength R for a given field int,ensity H. The pole tip!!
10('uted he tween zero and r /2 and het ween rand 3r/2 are saturated as
indicated by the redu('ed-field curve in Fig. 9~27c. The rotor~current
distribution shown in Fig. 9-27a ('orresponds to a quadrature-rotor wind­
ing on the primitive lIla('hinc. The rotor field tends to peak in the pORi­
five a. dirertion at (J r/2 and in the lIegative a. direction at (J = 3r/2.
"16 Principle. of electromechanical-energy conversion

Rotor

a, I,
0
+ l
1712 1r
I
¥
31r/2
I
+
I

I
~1r
I •

k r 1 [ ]
9

I Stotor
lal

Slolor 0 ." 2."


field
9

, ..
(b)
,I
Rolor 2."
field
9

(c)

Totol
field
9

Fig. 9-27 (a) Developed view of the air gap for a practical two-pole d.c machine.
(b) The radial air-gap magnetic field established by currents in the stator field windinp;.
(c) The radial air-gap magnetic field established by the armature currents. (d) The
total radial air-gap magnetic field due to both field and a.rmature currents.
Direct-evmmt commutator machine. (II) 417

For currents in both the stator and rotor windings, the total air-gap
field is obtained by adding Fig. 9-27b and c, as shown in Fig. 9-27d. The
effect of the armat.ure or rotor currents in distorting the total air-gap
field is known as armature reaction. If the armature current is zero, then
the total field reduces to Fig. 9-27b with zero field at 8 = .../2 and 8 = 3.../2.
With currents in the armature, the armature reaction shifts the magnetic
neutral, or zero field point, to 8 = 8" and 8 = 8" + ... in Fig. 9-27d. The
value of 8. depends on the magnitude of the armature current.

9·11 COMMUTATING OR INTERPOLE WINDINGS


Now that we know the general form of the d-c-machine fields, let us tt,lrn
our attention to the commutation problem. Referring to Fig. 6--10 and
the discussion in Sec. 6--4, the commutator-brush mechanism switches
the currents in the rotor coil sides in such a fashion that the rotor-surface­
current distribution remains stationary in space. Figure 9-28 shows the
switching process in detail, taking one rotor coil undergoing commutation
from Fig. 6-10. In Fig. 9-28 the brush is stationary and supplied with a
constant. current i. The commutator segments and rotor coil sides are
moving with an angular velocity w' rad/sec in the direction shown.
In Fig. 9-28a the brush is on the left-hand commutator segment and
the brush current i distributes in the manner shown. The coil current i.
is equal to i/2. Figure 9-28b shows the middle of the commutation period
where the commutator and rotor have so moved that the brush is centered
between the two segments. Sin(~e the brush provides an electrical path
between the two segments, the coil current i. can circulate around the

i/Z

ti ti
(al (bl tel
Fig. 9-28 A detailed view of the commutation procel!8 (0) just previous to the
commutation interval, (/I) at the center of the interval with the brush equally spaced
between adjacent commutator segments, and (c) at the concluaion of the commutation
interval.
418 Principle. of eleclromechanical-energy conver,;on

tl/2

v
. -
ie

Lc Re
8

fi
1
i2 t (l-{JRb
te

9 1i Fig. 9-29 A circuit model of a. single


coil undergoing commutation.

coil. Figure 9-28c illustrates the conclusion of the commutation period


with the brush on the right-hand segment and the coil current i, -i/2.
The closed loop illustrated by the heavy lines in Fig. 9-28b has a certain
amount of inductance. If ie is not equal to -i/2 at the instant the left­
hand segment moves from under the brush, then an arc wiII occur when i.
is forced to change rapidly to -i/2, as shown in Fig. 9-28c. This com­
mutation sparking wiII cause deterioration of both the eOllllllutator seg­
ments and the brushes and greatly shorten the operating life of the maehine.
The commutation cycle shown in three stages in Fig. 9-28 can be
modeled by the circuit shown in Fig. 9-29. The constant current i sup­
plied to the brush is shown as a constant-current source of strengt h i.
The two paths, each taking i/2 amperes to other windings on the rotor,
are modeled by two current sources. The brush resistance is designated
88 R~. The time t. is the total time the hrush is in eontad. with two
adjacent segments. The inductance I.e is the self-inductance of the closed
circuit formed when the brush contacts two segments as shown in Fig.
9-28b. Similarly, Re is the closed-circuit resistance. When t = 0, the
resistance (t/t.)R b is zero and the current i 1 = i and i2 = O. Thus i, at
time t = 0 is equal to i/2, since

(9-65)

As t proceeds to a final value t. at the end of the commutation period, the


right-hand resistor (1 - t/t,}R~ becomes zero, corresponding to the brush
being located completely on the right segment, as shown in Fig. 9-28c.
The voltage source v shown in Fig. 9-29 in series with the coil is
required because a rotational voltage is induced in the closed cireuit
formed when the brush contacts two segments. For a quadrature-rotor
armature winding the brushes are usually so placed that the right-hand
active ('oil side in Fig. 9-28 is at (J = 11"/2 and the left-hand adive coil side
is at (} 311"/2 during the center of the commutation period. Figure
Direc/·current commutator machines (U) 419

9~28a shows the right-hand coil side just previous to the commutation
period carrying i/2 amperes in the - a, direction. In Fig. 9-27a the a.
direction is out of the paper alld the rotor moves in the a, direction.
Thus the right-hand coil side is located at an angle (J just less than 1r/2.
Similarly, t he left-hand eoil side is located at all angle (J just lel'll'! than 31r/2.
Figure 9-27d shows, the total air-gap field at (J = 1r/2 to be in the a.
direction. Similarly, at (J = 31r/2 the total field is in the -a. direction.
With the rotor revolving at an angular velocity of w' rad/sec a voltaga is
induced ill the dosed loop as given by

v = ¢r. (aw'a,) X B· dl (9-66)

where a is the rotor radius and B is tlw total air-gap field. By evaluating
Eq. (9~6(), we have

" = 2alw'B. (9-67)


with L' taken positive when aiding the assumed direction of ie, as shown in
Fig. 9-29. The points (J 1r/2 and (J = 31r/2, where the rotor-surface
currents arc eommutated, are known as t.he mechanical neutrals; thus B ..
is the air-gap field at the neutral positions. Figure 9-27 shows that the
magnitude of B. is a function of the armature current for a constant field
current.
From Fig. 9-29 the following equilibrium equation is obtained:

(Ltp + R, + Rh)i, = " - iRbG - t) (9-68)

Equation (9-68) is derived by using Kirchhoff's potential law around the


loop and eliminating i l and i2 with the current law at nodes A and B.
By taking i and I} as constants, Eq. (9-68) can be solved for ic in the form

whereT = L./Rbseconds, and the assumption has been made that Re« Rb,
which for carbon brushes is usually a valid approximation.
Figure 9-30 shows characteristic plots of Eq. (9-69) for various values
of v and T. At t = 0 t,he coil current it starts at i/2 amperes. When
t = ie, t.he current is forced to be. -i/2 amperes as the brush leaves the
left-hand commutator segment in Fig. 9-28. The instantaneous change
in ie at the end of the commutation period t ie is equal to

(9-70)
420 Principles 01 electromechonicol·energy conversion

Fig. 9-30 I The coil current as a function of time during the commutation interval.

If Ai. is not equal to zero, then a spark occurs between the brush and the
left-hand segment as they separate. The intensity of the spark is a func­
tion of tli, and the magnitude of the coil inductance L •.
One method of eliminating the current discontinuity tli. is to shift the
brushes on the commutator. For the armature current assumed in con­
structing Fig. 9-27d, the total air-gap field is zero at 8 = 89 and 8 = 8.. + 'If'.
Shifting the brushes so that the coil undergoing commutation has its
active coil sides located at some angle slightly greater than 89 and 8a + 'If'
would cause BN in Eq. (9-67) to be negative. Thus the rotationally
induced voltage v is also negative. By properly adjusting the new brush
position, tli., given by (9-70), can be made equal to zero and sparkless
commutation can be achieved. In the early days of d-c machinery,
brush-shifting mechanisms were incorporated on the machines. The
major difficulty with brush shifting as a means of achieving good commu­
tation is the dependency on the magnitude of the armature current. AI'!
the armature current increase!', the shape of the air-gap field in Fig. 9-27d
changes shape. The magnitude of v therefore ehanges alld a new brush
pOl'!ition mURt be obtained to make tli. again equal to zero.
On modern d-c machines nearly perfect commut.ation is achieved by
adding small stator poles between the main stator poles, as shown in Fig.
9-31. These additional stator poles, usually ealled interpoles because of
their location between t he main field poles, have windings eonnected ill
series with the armature winding. For the armature-<:urrent direction
used in Fig. 9-27 the interpole currents would be as shown in Fig. 9-31.
The interpoles establish a localized air-gap field at 8 = 'If'/2 and 8 3'1f'/2.
The brushes are so plaeed 011 the machine that the coil undergoing com­
mutation is sweeping past the interpole. An additional rot,ational voltage
due to the interpole field is thus induced in the coil being commutated.
Examining FigR. 9-27 and 9-31, the additional voltage /Jr due to the inter­
Direct-current commutator machines (II) 421

Rolor

o "/2 11' 211'


I I I I • 8

[jJ •



+
+
+
+
Slolor
[jJ +
+
+
+



Fig. 9-31 The inclusion of interpole windings on the d-e machine stator.

poles opposes the original rotational voltage v given by Eq. (9-67). The
coil-current di!'!eontinuity at the end of t.he (~Olllll\ut.at,ion period t = t. is
now given by

!J.i, = - (..!::..
llb
+ !t, i - ..!!)
Rb
(1 - e-I"r) (9-71)

upon adding the interpole rotational voltage to Eq. (9-70). By properly


designing the interpoles and positioning the brushes, VI can be adjusted to
make !J.i, approximately zero for a wide range of armature currents and
thus provide sparkless eommut.atioll with t.he brushes in a fixed position.

9-12 COMPENSATING WINDINGS


For t.he practical doe machines diseussed ill Chaps. 8 and 9 we often
assumed the armature indU(~tall('e to be negligible. In t.his section we
Rhall !;lee that. fhiR slllall armature induct.a.nce, particularly on large d-c
machine!'!, can cause a I»erious ar('ing condit.ion between comnlutator seg­
ments. Large modern d-c mal'hinel» are usually constructed with an
additional winding on the stator connected in !'!eries with the armature to
minimize the pORsibility of commutator arcing or flashover.
The armature of a d-e maehine is constructed with a number of coils
conne('ted in series. These series conne('t.ions are made at, the commutator
segments as shown in Fig. 9-:12. The total self-induetance of the armature
winding is the combination of t.he indu('tallees of t.he individual coils.
Normally, at least. two parallel paths exist between each pair of brushes,
so a series-parallel ('ombinatioJl of t.he individual coil induetances must
be made to obtain the total armature induct.ance La. A rapidly changing
load on either a motor or generator causes the armature current of the
ma('hine also to change rapidly. The voltage of self-induction, given by
La di./dt, can therefore have a relatively large value beca.use di,,/dt is
large. The total self-induced voltage is divided between the individual
coils and therefore appears between adjacent commutator segments. If
422 Principles of electromechanical-energy conversion

the voltage hetween segments exeeeds the breakdown voltage, then a low­
resistanee are will form between segments. The total voltage between the
hrushes now appears acro!';s the remaining segments, thus continuing the
arc formal iOIl unt ill he hruflhes arc l<hort-cireuit cd hy a cOlllplete flashover
or "ring of fire" at the COlHlllutator. The large di,,/dt can thus trigger a
('haill reo.<'tioll of arcing hetween adjacent eOlllmutator seglhents until the
entire commutator is short-ein'uited hy olle ('olltinuoul'l arc.
The possibility of commutator flashover can be greatly minimized by
the addition of the compensating winding shown in Fig. 9-33. The
compensat iug winding is a dist ributed winding wound on the stator t.o
establish a eurrenl. distribution on the st.ator surface that is as dose as
possible to the ('urrent distribution established by the armature. Now
the armature and eOlllpensating winding arc ('onllceted ill series in such
a fashion that the two ('urrent distrihutions on either side of the air gap
are cqual but opposite in direction. Therefore, the voltage appearing
aeross the brushes owing to self-induetion is now given by

I'. C~ I, ~i:! _ M or di" (9-72)


., " rlt dt

By proper design of the ('ompensating winding t he mutual induetanee


it! "r ('an he made nearly cqual to '.0'
thus making the brush voltage t'b
duc to self-illdudion approximately zero.
The total armuture-('ircuit indudunee from Fig. 9-33 is given by

(9-73)

Wit h the compensating winding very similar to the armat.ure winding, L.


is approximately equal to La. and thus with La = At oc the total inductance
given hy (9-73) is very nearly equal to zero. The interpoles discussed in
Sce. 9-11 are also conneeted in series with the armature. The inductance
of the interpoles should be included with t he total armature inductance.

}<'ig.9-32 The series connection of the individual armature coils at the commutator
segments.
Direct-cum'"t commutator machine. (II) "23

COmpensating
winding

Fig. 9-33 A primitive-machine


model showing the compensating wind­
ing.

On large doc machines the compensating winding is usually placed in


slots stamped into surfaces of the stator pole faces. The armature is
wound with many coil sides in each armature slot. The compensating
winding must be so arranged that the same total current as on the arma­
ture flows in each slot. For ea.se in construction, heavy copper strips are
placed in the stator slots, and they are connected in such a fashion that
they conduct the proper total current per slot and thus match the arma­
ture-current distribution.

'·13 SUMMARY
The Ward Leonard system is used where precise control of large amounts
of mechanical power is required. The source of this power can be an a-c
electrical supply, which is commonly available. The combination of
these two factors (i.e., precise control and readily available power) make
the Ward Leonard system a widely used industrial drive. In formulating
the equilibrium equations for the Ward Leonard system the formalism of
the primitive-machine notation was modified for greater convenience.
By inspection of the system, the interconnected equivalent circuit, shown
in Fig. 9-1, was drawn. If the general energy-flow direction is known, the
reference sense of all current and voltage variables as well as the trans­
former dots can be assigned for a more compact formulation. From the
equivalent circuit, both the doc steady-state operation and the transient
operation are readily obtained. The response capabilities of the Ward
Leonard system are studied by using step- and frequency-response char­
acteristics. Even though the system equations appear to be linear, an
incremental analysis is used to emphasize the fact that parameter values
are constant only in the neighborhood of an operating point.
424 Principles of .Iedromechonicol·energy conversion

The compound d-c generator has two stator fields in quadrature with
the armature. One of the stator fields is constructed of many turns of a
light-gauge wire and is separately excited. The other winding is made up "

of a relatively few turns of a heavy-gauge wire and is called a series field


because it is connected in series with the armature. With the separately
I
excited field and the series field properly designed and cumulatively con­
nected, the machine generates a nearly constant voltage for a wide load
variation. The analysis of the compound generator requires an extension
of the basic primitive machine, since both the series and shunt fields are
l
on the same stator axis.
The amplidyne and metadyne are often referred to as cross-field
machines. Their operation depends on two pairs of brushes placed on
the mac,ine's commutator. The armature circuit in quadrature with the
main stator winding has its brushes short-circuited. A rotationally
induced voltage in the quadrature-rotor winding generates a large short­
circuit current, and therefore a large quadrature field is established. The
quadrature field induces a substantial rotational voltage in the direct­
rotor winding on the same axis as the stator field. By adding a properly
designed direct-stator compensating winding, we have a voltage amplifier
with a very large power gain, known as an amplidyne. When the com­
pensating winding is omitted, the machine becomes a controlled-current
generator called a metadyne.
The seIf-excited shunt generator relies on the nonlinear characteristics
of the d-c machine for its operation. The machine requires no external
source of electrical power, and therefore it is often used as an exciter or
d-c field supply for larger machines. The a-c alternator, which generates
all of our commercial a-c power, very often receives d-c excitation from a
self-excited shunt generator. Automobile electrical systems use a self­
excited machine to eliminate the battery drain required by the field
of a separately excited generator. The residual magnetism of the stator
field and the saturation of the iron are required for a stable useful output
voltage of a reasonable magnitude.
The commutator-brush switching mechanism on an actual d-c
machine has two very serious shortcomings. As the brush slides from
one commutator segment to the next, a short-circuited coil is formed on
the armature. At the instant the brush breaks connection with the
original commutator segment an arc is forn~ed if the coil current is forced
to have a discontinuity. By adding interpole windings in series with the
armature, a rotational voltage is induced in the coil being commutated.
This induced voltage tends to eliminate the coil-current discontinuity and
thereby reduces the commutation sparking over a wide range of armature
currents.
Rapid changes in armature current cause a large voltage of self­
Direct·current cOlllmutator lIIaelaine, (IIJ 425

induction in the armature coils. Since these coils are connected in series
at closely spaced commutator segments, a voltage breakdown or flashover
between segments can occur, particularly on large d-c machines. Com­
mutator flashover is eliminated by adding a distributed compensating
winding in series with the armature. The compensating winding induces
a voltage equal and opposite to the self-induced voltage and thus tends
to cancel the voltage between commutator segments caused by rapid
changes in armature currents. The compensating winding is closely
coupled to the armature winding, making the net inductance of the series
combination practically zero. .
In the analysis of the various d-c machines in Chaps. 8 and 9 an
ideal commutator was assumed. Also, we very often assumed the anua­
ture inductance to be negligible. Both the ideal commutator and zero
armature inductance are reasonable assumptions because interpole and
compensating windings are included in all machines in which brush
sparking and flashover are encountered.

PROBLEMS

'.1 A Ward Leonard speed-oontrol system is shown in Fig. P9-l. The ampli­
fier input voltage is the difference between a reference voltage and the tachometer
voltage. Assume the voltage amplifier to be linear with a very high input impedance
and negligihle output impedance. If the motor speed decreases, then the tachometer
output voltage decreases and the net input voltage into the amplifier increases.
The amplified volta.ge applied to the generator field increll.lles, and 80 the genera.tor

Ra
9
Amplifier

Fig. P9-1
,(26 Principles of electromechanical-energy conversion

output voltage increases, thus raising the motor armature voltage and tending to
increase the motor speed. The generator is heing driven at constant speed, and the
field current of the motor is constant. Neglect the armature inductances of the two
machines. Other system parameters arc as follows:

Amplifier gain K. = 90 volts/volt


Generator field resistance R! = 183 ohms
Generator field inductance L:= 92 h
Generator and motor armature resistances R: = R:, = 0.42 ohm
Generator voltage constant G:'w;o 177 volts/field amp
Motor torque constant G:[i!.. = 4.89 newton-m/armature amp
Total motor plus load inertia J ~ = 4.0 kg-m 2
Total windage coefficient D~ = 1.14 newton-m-sec/rad
Tachometer voltage constant K. = 0.04 volt-scc/rad

a. Drflw a complete block diagram for the closed-loop spccd-control system.


b. If the reference voltage VII is set at 10 volts and the stcady-state motor speed
is 100 rad/sec, find the generator field current and the applied constant load torque
on the motor.
c. Determine the incremental frequency response for the system in the form of a
Bode plot relating 0;" (jw) to VR(jW).
d. The system is operating under the steady-state conditions of part b when,
at t "" 0, the reference voltage is changed suddenly to II volts. Find the motor
speed as a function of time.
9-2 The speed-control system discussed in Prob. 9-1 can be converted into a
position-control system by removing the tachometer and coupling the rotor to a
pOtentiometer on the motor shaH as shown in Fig. P9-2. The potentiometer output
voltage subtracts from the reference voltage. The constant load torque on the motor
shaft is removed, with all other system parameters remaining the same as in Prob.
9-1. The feedback potentiometer has an output of 0.2 volt/rad rotation of the motor
shaft.
a. Draw a complete block diagram for the closed-loop position-control system.
Assume the load torque to be zero.
b. If the reference voltage VII is changed from zero to 10 volts, find the steady­
state change in the motor shaft angle. Assume the system to be stable.

G
1 >
-
Amplifier
----~
~
0",

\
\~

I
Fig. P9-2
Direct-current commutator machines (II) 477

c. From the complete block diagram of part a find the overall transfer function
8~(8)/VR(8). Use the Routh criteria to find the number of RHP poles. Is the
system stable?
d. Sketch a frequency response for the overall transfer function.
e. What would be the elTect of including the tachometer in the position-control
system, thereby producing a voltage subtracting from VB for positive motor speed?
9-3 A lO-kw 125-volt 1750-rpm compound generator has an armature resistance
of 0.06 ohm. With the shunt field separately excited at rated voltage, the generator
produces rated voltage when driven at rated speed with the series field short-circuited
and the armature open-circuited. The scries field has a resistance of 0.04 ohm and
induces 0.1 volt in the armature per ampere of series field current at rated speed.
The shunt field resistance is 75 ohms.
a. Find the two rotational inductances Ga , and G....
b. If the shunt field is excited at rated voltall:e and the machine is driven at rated
speed, find the output voltage when the machine delivers rated load current first
with the series field cumulatively compounded and then with the field differentially
compounded.
c. Find the overall efficiency for each case in part b.
d. Repeat parts band c for a generator speed of 2500 rpm.
9-4 A 220-volt 5O-kw 120ft-rpm compound d-c generator when running at rated
speed produces rated no-load voltall:e with a shunt field current of 5.0 amp. At full­
load current 7.2 amp is required in the shunt field to produce rated output voltage
at the same speed.
a. Calculate the rotational inductance G., and the armature resistance R. for
the machine.
b. Assuming that G•• = 0.02R., where G•• is the series field rotational inductance,
in henrys, and R. is the series field resistance, in ohms, find the values of Ga. and R. to
flat-compound the machine at rated speed.
c. If a 0.2-ohm resistance is placed in parallel with the series field, find the output
voltage at full-load current with 5.0 amp in the shunt field.
9-5 A fully compensated amplidyne being driven at constant speed has the
following parameters:

Control field: R, = 75 ohms, L, 7.5 h


Quadrature field: R. 1.04 ohms, I•• = 0.318 h
Armature: R. = 1.04 ohms, I•• = 0.318 h
Compensating winding: R. = 1.8 ohms, L. = 0.318 h

The machine has an open-circuit output voltage of 250 volts with 0.1 amp in the
control field.
a. Write the open-circuit transfer function between the armature voltage and
the control field voltage.
b. If the control field voltage is suddenly changed from 4 to 6 volts, find the
open-circuit voltage as a function of time.
c. For a total armature load of 12 ohms in series with 25 h, find the steady-atate
load current with 4.0 volts applied to the control field.
d. For the conditions in part c find the load current as a function of time following:
a change in the control field voltage from 4.0 to 6.0 volts.
e. Find the d-c power gain of the amplidyne.
....28 Principles of electromechanical· energy converlion

9-61 Industrial applications requiring very accurate speed control of large drive
motors often employ the extended Ward Leonard system shown in Fig. P9-6. The
voltage at the input terminals of the electronic amplifier is the difference between
the preset reference voltage VII and the tachometer output voltage. If the motor
speed decreases, owing to an increased load, the tachometer output voltage decreases
and the amplifier input voltage increases. Tire amplifier supplies increased current
to the control field of the fully compensated amplidyne. The amplidyne supplies a
greater current to the field of the doC generator being driven at a constant speed

Amplidyne

~er
I ...r--~~/
~
/
/'--w;o
/

ill Fig. P9-6


of "';.. The generator, in tum, feeds more current to the armature of the separately
excited doc motor whose field current is held constant at i!,.o amperes. The motor
speed therf'fore tends to remain constant. By varying the value of VR, an operator
can set the desired speed. The system parameters are as follows:
Amplifier
Gain K. = 10 volts/volt
Input impedance is infinite
Output impedance is zero
AmpUdyne
This is the amplidyne discuBBed in Prob. 9-5.
Generator
Field resistance R~ = 12 ohms
Field inductance L~ = 25 h
Armature resistance R:= 0.002 ohm
Armature inductance is negligible
Voltage constant G~·",;o = 20 volts/field amp
,l'otor
Armature resistance R:' = 0.010 ohm
Armature inductance is negligible
Torque constant G~·i!..o 40 newton-m/armature amp.
1 Answers to this problem are listed at the end of the book.
Direct-current commufofor mochi"., (II) 429

Tachomder
Gain K, .. 1 volt-see/rad
Motor ahaft load
Inertia J.. - 6000 kg-m'
Windage is negligible
Constant load torque TL - 40,000 newton-m
a. Draw a complete block diagram for the system. Reduce the diagram showing
VB and TL 8.8 inputs and the motor speed ...:;. as the response.
b. Find the d-c steady-state relationship between two inputs VR. and TLO and the
steady-state speed ...:;... What reference voltage is required for a motor speed of
15 rad/sec? With this reference voltage find the change in speed for a 10 per cent
increase in load torque.
c. The system is operating under the steady-state conditions described in part b.
Draw an incremental Laplacian block diagram showing operation about the steady­
state conditions. Determine the transfer function relating 0:;'.(8) to V R.(8). Is the
system stable?
9-1 By running a lIO-volt d-c shunt motor 8.8 a separately excited generator
at a speed of 1800 rpm, the following data are obtained:
Field current 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65 0.70
Open-circuit armature
voltage 70.0 78.0 84.5 91.0 96.0 101.5 106.6 111.0 115.0
The armature resistance is 0.64 ohm, and the field resistance is 192 ohms.
a. If the shunt motor has 115 volts impressed across its field and armature, find
the no-load speed of the machine.
b. Determine the rotational inductance G./ for an incremental linear model
valid around the operating conditions of part a.
9-8 A d-e generator being driven at its rated speed of 1200 rpm has the following
saturation curve:
Open-circuit armature voltage 4 21 40 58 72 89 105 115
Field current o 0.1 0.2 0.3 0.4 0.6 1.0 1.5
The field resistance is 75 ohms, and the armature resistance is 0.8 ohm. If the
generator is driven at rated speed find:
a. The open-circuit armature voltage with the field separately excited across a
nO-volt supply.
b. The open-circuit armature voltage with the field shunt-connected. Is the
sense of the field connections important?
c. Repeat parts a and b for half rated speed.
d. Find the short-circuit armature current for the connections of part.s a and b
for rated speed and half rated speed.
9-9 A 5O-kw 220-volt d-c generator, being driven at its rated speed, has the
following saturation curve:

Open-circuit armature voltage 11 61 110 145 170 190 206 220 228 233
Field curren t O O . 5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.4

The armature resistance is 0.04 ohm, and the field rheostat is 80 adjusted that the
machine, operating as a shunt generator, produces 233 volts at no load.
430 Principles of .tecfromechonicol.energy conversion

a. Find the total resistance of the field circuit.


b. Graphically determine the steady-state output voltage as a function of the
output current for thc shunt generator connection.
c. What is the terminal voltage when the output current is at its rated value?
What is the maximum current the machine can deliver?
9-10 The SO·kw g!'nerator in Prob. \l·1l is being driven at rated speed with no
load on the armature. The total field circuit resistance is set at 60 ohms, and the
field inductance is 25 h.
a. At time t 0 the field is suddenly shunted across the armature. Find the
armature terminal voltage as a function of time.
b. After the armature voltage has reached its steady-state value in part a, a
15.0-ohm load resistor is connected across the armature. Find the armature termi­
nal voltage as a function of time. Hint: Consider using an incremental linear analysis
for part b.
9-I1 I A two-pole doe generator has 40 closely spaced commutator segments.
Each of the two brushes is wide enough to completely cover one segment. The self­
inductance of each armature coil is 0.1 mh, with the resistance being 0.001 ohm. For
the approximate circuit model shown in Fig. 9-29 use a brush resistance of 0.2 ohm.
The total armature current being commutated at the mechanical neutral is 175 amp.
The machine is being driven at 1800 rpm. The radius and length of the rotor are 0.3
and O.S m, respectively.
a. Find the total time available for commutation.
b. The radial air-gap field at the mechanical neutral is 3 X to-I weber/mt.
H each armature coil contains JO turns, find the voltage induced in the armature coil.
c. Find the coil current i, as a function of time during the commutation interval.
What is the coil current discontinuity at the end of the commutation interval?
d. What should be the strength of the interpole field for perfect commutation?
chapter

THE A-B PRIMITIVE MACHINE

10

The primitive machine we have been using supplies


energy to t he rotor eil'cuits by means of a commu­
tator and bl'Ushel>. As the brushes slide from olle
1'0nllllutatOl'scgnlf'llt to the next, the al'fangement
of the rotor circuits maintains a current dist.ribution
on the surface of the rotor stationary in space with
respect to the stator. Thus, for a constant current
supplied tot he eOllllllU tator-brush mechanism, a
stationary Ilul.glleti(~ field is generated in the rotor
stl'Uel me nnd in the air gap without regard 10 rotor
position. The d-(' machines discussed in Chaps. 8
and n are l\<'Iually ('ollstrueted with COllllllutator­
brush mc{'hunisllls. In I he ('ase of the amplidyne
and Illetadyne I wo pairs of brushes are placed on
I he sallie {'ollllllutator, thus establishing both direct­
and quadrat nre-rotor windings.
A seeolld broad {'atl'gOl'y of rotating machines
using 1\-(' excitations are often eonstructed in a
slightly different manner, The 1'0101' eircuits, in
1he ('asc of a-(~ llIa('hines, often receive energy
through simple sliding contacts known as slip rings.
Since no circuit switching is performed, the gen­
erated magnetic field due to slip-ring windings
lIloves with the rotor structure. Therefore, a-c
maehines do not fit direQUy into the framework of
our primitive lIlll('hine beeause of the difference in
the rolor sliding-('ontaet. meehanism. The general
{'OllstrUC'j ion of the a-e machine is similar to that of
the primitive machine exeept for the use of slip
rings illstead of a commutator.
A second form of the primitive machine, often
known as the a-b machine, will be developed in this
chapter. The a-b machine has a rotor structure
with two windings receiving or delivering power
431 through slip rings, Actual a-c machines can thereby
432 Pril!Cip/es 01 electromechanical· energy conversion

be readily modeled in terms of the a-b machine. By means of a ('oordinate


transformation, the a-b machine can be transformed into our original d-q
primitive machine. We can already handle the d-q machine problem,
and thus by an inverse transformation we can ohi ain the solution to the
a-b problem corresponding to the aetual a-c machine of interest. Syn­
chronous machines and a-c induetion machines are analyzed in the next.
two ehapters hy using this transformation teehnique.
A careful review of the d-q primitive-machine formulation presented
in Chap. 6 will greatly add to your understanding of this chapter.

10·1 THE SLIP-RING WINDING


I
The direr!- and quadrature-axis rotor windings developed in Chap. 6 have
their individual ('oil sides joined together at eOllllllutator segments.
Suppose we construrted exactly the same winding except that the coil
sides were joined together without the {'omplication of I,he eommutator
segments. Taking the developed view of the winding, as shown in Fig.
6-10, we have simply removed the commutator without changing any of
the connections. In order to supply power to the winding, or extract.
electrical power in the ease of a generator, two leads are eonnected at the
junction points where the bmshes are shown in Fig. 6-lO. The main
problem with this arrangement is that these two leads will wind together
as the rotor revolves. To alluvial e I his diffkulty, t he sliding contaets
shown in Fig. 10-1 arc used. Two ('onducting eyliudrieal disks, insulated

Fig. 10-1 Slip-ring con­


nections to the rotor on the
a-b primitive machine.
The a-b primitive machine 433

II

Fig. 10-2 Hchematic representation


of two slip-rinK rotor windings in
quadrature.

from the rotor shaft, are so mounted that t.hey revolve with the rotor
structure. Leads are connected from the proper two points 011 the rotor
winding to these two disks, whieh are known as slip rings. Brushes
sliding on the slip rings provide the required ('onnection to the rotor
circuit.
Slip-ring windings on the maehine rotor ean be represented as shown
in Fig. 1()"2. The two axes are designated as the aT and bT axes, symbol­
izing the a and b rotor axes on the a-b ma(~hine. Since the two windings
are on the rotor, they are drawn inside the cil'cle, as shown in Fig. 10-2.
The a axis is drawn at an angle a to the d axis of our fixed d-q axis frame.
As the rotor turns, the aT and fir llxes llnd the slip-ring winding turn right.
along wit h it. Therefore, the a-fl rotor axes fraille is essentially "glued"
to the rotor structure. Rotor windings supplied through slip rings are
drawn on the a-b rotor frame (Le., on the aT Ilnd br axes) and revolve wit.h
the rotor. Rotor windings on the d-q frame have their coil sides physically
moving with t.h& rotor, but because of the commutator-brush switching
mechanism the magnetic fields est ablished by currents in these windings
remain stationary in space and are completely independent of the rotor
position. For positive rotor currents i:
and fb the respectiv"l magnetic
fields are established in the positive axis directed as indicated by the
arrows in Fig. 10-2.
In Chap. 6 the current on the rotor surface is assumed to be sinus­
oidally distributed. For the commutator-and-brush rotor windings of
our original d-q primitive machine the assumed sinusoidal distribution is
only the first term in a Fourier-series expansion of the actual square-wave
distribution of surface currents. The direct- and quadrature-rotor­
winding surface ('urrents for all n-pole-pair machine, as given by Eqs.
(6-83) and (6-84), are repeated for ('ollvenience as

J~ K~ sin n9 i~az amp/m (10-1)


J; = - K; cos n9 i;a, amp/ill (1()"2)
~4 Principle. of electromechanicol-energy conversion

with K~ sin n8 and K; cos n8 being the respective turns per meter at an
angle 8 for each of the two windings. Equations (10-1) and (10-2) are
independent of the angular position or velocity of the rotor. The current
distributions have a periodicity of 27(/n radians.
The slip-ring windings shown in Fig. 10-2 are also going to establish
a sinusoidal current distribution OIl the rotor surface. Since no switching
is involved with the slip-ring windings, the active coil sides could actually
be grouped to establish a sinusoidal surface current in the same manner as
for the stator circuits. If the a-b rotor windings are uniformly distributed,
then a Fourier-series expansion is made and only the fundamental term
for the current distribution is retained. For the a-b rotor windings the
rotor-surface currents for an n-pole-pair machine are given by
J; :oJ K; sin n(8 - a) i;a. amp/m (10-3)
J~ = - K~ cos n(8 - a) iba. amp/ill (10-4)
where a is the angle measured counterclockwise between the stationary
d axis and the movable a axis or, equivalently, from the q axis to the b
axis, as shown in Fig. 10-2. The (:onstants K~ and Kb are the respective
distribution factors. The a-b rotor windings establish (mrrent sheets
identical with those of the d-q rotor windings upon substituting 8 - a for
8 in Eqs. (10-1) and (10-2).
For a changing a, Eqs. (10-3) and (10-4) will indicate that the (mrrent­
sheet distributions move with respect to the stator. If the rotor is
revolving at a constant angular frequeney w~, then a w~ and

aCt) = w~t + a(O) (10-5)


where a(O) is the position of the a-b frame at time t O. For simplicity,
let us assume a(O) is zero. Upon substit.uting Eq. (1O-i» into (10-3) and
(10-4) we have
J; = K~ sin n(8 - w~t) i;a. (10-6)
Jb - Kb cos n(8 - "'~t) i'a. (10-7)

The a-axis rotor-surface-current density J; is a maximum when

n(8 - "'lit) = ; (10-8)

By solving for 8, we have

(10-9)

where 8max is the value of 8 where the maximum surface current occurs.
The value of 8max changes linearly with time at an angular rate of w~
rad/sec. Therefore, the two slip-ring surface-('urrent distributions are
The a-b primitive machine 435

rotating with respect to the stator at an angular frequency of w~ rad/sec,


assuming a constant or direct current in the windings.
By use of Ampere's circuital law the magnetic fields in the air gap
due to the a-b rotor-winding currents are found to be given byl

B~ = ~(8) K;i~ cos n(8 - a) a, (10-10)

B br /.loa Kr'r'
= ng(8) btb Sill n
(0 - a
) ar (10-11)

where a is the radius of the rotor and g(8) is t,he air-gap length having a
periodicity of "KIn radians as given by Eq. (6-74). If a = w~t, then these
air-gap fields will revolve in a counterclockwise direction at an angular
frequency of w~ rad/sec.

10-2 ROTATING D-Q ROTOR SURFACE-CURRENT DISTRIBUTIONS

Suppose the t.wo d-q rotor windings are excited by sinusoidal currents with
t,he following phase relationship:
i~ V2 I~ cos (wt + "') (10-12)
where I~ and I; are the rms values of the currents, w is the radian fre­
quency of the excitations, and", is a constant arbitrary phase angle. By
substitution of these two d-q rotor current.s into Eqs. (10-1) and (10-2) the
d-q surface-current distributions (~an be put. into the form

J~ = Ji [sin
Krlr
(n8 + wt + "') + sin (nO - wt - ",)]a. (10-13)

J; = 0' [sin (n8 +


KrI'
wt + "') - sin (nO - wt - ",)]a, (10-14)

by using the trigonometric identity for the product of the sine and cosine
of two angles.
When Eqs. (10-13) and (10-14) are combined, the total d-q rotor­
surface current is given by

J4, = [!C~I4~K;I; sin (n8 - wt - "') ] a.

+ [K~I4:;2K~I; sin (nO + wt + "') ] a. (10-15)

In all of our work with the a-b and d-q primitive machines we shall find it
convenient to have all winding distribution factors equal. Thus, when
I Refer to Sec. 6-5, where a similar development is performed for the d-q rotor wind­
ings having sinusoidal current distributions, with n - 1.
436 Principles of e/edromechanical-energy conversion

we set K~ = K; = Kr, Eq. (10-15) becomes


Kr
JdV = V2 [(l~ + I;) sin (nO - wt - Ifl)]a.

Kr

+ y2 [(ld - I;) sin (nO + wt + Ifl»)a. (10-16)

Notice that Eq. (10-16) is composed of two main terms. The first
term represents a current distribution rotating around the rotor in the
positive angular direction (countercloekwisc) at an angular frequeney of
win rad/sec. The second term in 00-16) is a similar surface-eurrent
distribution rotating in the clockwise or positive angular direction at the
same radian frequency win rad/sec. For the special ease when I~ = I;,
only a c~untercIockwise rotating current distribution exists. Similarly,
when -ld = I;.
only a elockwise rotating distribution is found on the
rotor surfaee. Notice again that the angular position or velocity of the
rotor does not afi'eet t he surface currents for the d-q windings.
A sinusoidally distributed current sheet rotating in the positive or
counterclockwise dire(~t,ion at an angular velodty of win rad/sec can also
be established by using the slip-ring windings. By supplying a direct
current to either t,he a- or b-axis rotor winding, a sinusoidal surface current
is established on the rotor. Revolving the rotor at an angular velocity
of win rad/sec will ('RUSe the current, whieh remains stationary with
respect to the rotor strueture, to revolve at the same angular frequency
in the same direction. The technique of establishing identical current­
sheet distributions with either fOl'llI of rotor winding is considered in
detail in the next se(~tion.

10-3 EQUIVALENT A-8 AND o-Q ROTOR


SURFACE-CURRENT DISTRIBUTIONS
In the preceding sections we have discussed two met,hods of establishing
current distributions on the surface of the rotor. With t.he d-q rotor
windings currents are supplied through a commutat.or and brushes and
the resulting distributions given by
Jd = K~ sin nO i~a. (10-17)
J; -K; cos nO i;a. (10-18)
are stationary ill space with respect to the stator. The a-b rotor windings
are supplied currents through slip rings with the resulting distributions
given by
J: = K: sin n(fJ - a) t:a. (10-19)
J;; = - K;; ('os n(fJ - a) i;;a. (10-20)
Th. a-b primi,iv. machine 437

Fig. l0-3 Rotor windings on the d-tl


and a-b axis frames.

Figure 10-3 schematically shows the d-q and a-b rotor windings and
their respective axis frames. Let us now examine what currents should
/
be supplied to the d-q rotor windings ill order to establish the same current
distribution on the rotor surface as that set up by a given set, of currents
in the a-b rotor windings. Thus, if i: and ''b are specified, we wish t.o find
values for i~ and t-; which result in idenHcal rotor-surface currents. The
total surface current due to i~ and i; is obtained by summing Eqs. (10-17) .
and (10-18), thereby gettingl
J~Q = (K~i~ sin n6 - K~i; cos n6)a. (10-21)
Similarly, the total rotor-surface current resulting from i: and ib is t,he
sum of (10-19) and (10-20), which becomes
J~ = [K:i: sin n(6 - a) - K/,ii: cos n(6 - a}Ja. (l0-22)
By using the trigonometric identities for t.he sine and cosine of the
difference of two angles, Eq. (10-22) can be expanded and rearranged to
the form
J~ = [(K:i: cos na - K;i; sin na) sin n6
- (K:i: sin na + Ki:i~ cos na) cos n61a. (10-23)
By comparing Eqs. (10-21) and (10-23), we can see that the two total cur­
rent distributions J'aIJ and J'oJ, will be identical provided
(10-24)
and
(10-25)
I The double SUbscript and one superscript on J:I. indicate that both the d- and q-axis
rotor windings are being considered.
438 Principles of electromechanical-energy conversion

Practical a-c machines are so constructed that all windings on the


rotor are identically wound. In addition, all stator windings are iden­
tically constructed, but the rotor and stator windings may be different.
Therefore, the two rotor windings on our a-b primitive-machine model can
be taken to be identical. The corresponding d-q machine has the same
rotor windings. The four stator windings on the two primitive machines
are also all identical. Thus from practical considerations we can write
for the rotor distribution factors
K~ (10-26)
When Eqs. (10-26) are substitu~into (10-24) and (10-25), the two condi­
tions reduce to
i:i .. cos na i; - sin na i;; i; = sin na i~ + cos na i; (10-27)

This relationship between the d-q and a-b currents can be expressed in
matrix form as follows:
(10-28)
where

(10-29)

(10-30)

and a~q.Gb expresses the transformation to the d-q rotor-current coordinates


from the a-b rotor-current coordinates and is given byl
a' _ [cos na -sin naJ (10-31)
dQ.Gb - sin na cosna
Similarly, if the currents are expressed in the d-q reference frame and
the equi valent a-b currents are desired, they can be found from the equation
(10-32)

where a~.do is the inverse of a~Q.Gb and is found to be

a~,dq
(a r )-1 = [ c~s na sin na] (10-33)
do,Gb -sm na cos na

The currents in the d-q coordinate frame, specified by Eqs. (10-12),


set up a sinusoidal surface-current distribution with two components, one
I The matrix transforms the Becond pair of subscripts into the jirBt pair of subscripts;

tws the a-b currents are being transformed into the equivalent d-q currents. The
single superscript indicates that only the rotor currents are being transformed.

"

The a-b primitiVe machine 439

revolving in the positive angular direction and the other revolving in the
negative angular direction. Let us now examine what would be the
equivalent currents in t.he a-b reference frame. For convenience, Eqs.
(10-12) specifying the direct- and quadrature-axis rotor currents are
repeated as follows:

ir = ~"d] = V~ I~ cos (wt + <p)] (10-34)


dQ i; V2 I; sin (wt + <p)
using matrix notation. Equation (1O-32) is now used to find the equiva­
Je~t a-b rotor currents:

'r
1""
= i:]i~ = V2 [ cos na sin na] I~ cos (wt
-sin ncr cos na I; sin (wt
+ <p)]
+ <p) 00-35)

which, upon expansion of the matrix product, gives us

i: = I'd+ I' cos (wt + <p -


Q ncr) + I'dv12
- I'
"'cos (wt + <p + na)
V2
(10-36)
ih =
I'+ I'' sin (wt + <p -
dv12 na) -
I' - I'
d fl sin (wt + <p + na)
V2
(10-37)
after we use the trigonometric identities for the product of the sine and
cosine of two angles.
For simplicity in examining Eqs. (10-36) and (1O-37) let us begin by
assuming that d-q excitation is balanced or that
I~ = I; l' (10-38)
This assumption, ,as shown in the preceding section, yields a rotor-surf ace­
current distribution rotating in the positive angular direction. Substitut­
ing Eq. (10-38) into Eqs. (10-36) and (10-37) reduces the a-b coil currents to
i~ = vI2 r cos (wt + <p - na)
(10-39)
i~ = vI2 l' sin (wt + <p - ncr)

The angle a, as illustrated in Fig. 10-3, defines the angular rotation between
the two coordinate systems. Suppose we allow the a-b coordinate frame
to rotate in the positive angular direction with a constant angular fre­
quency of win rad/sec. This condition is expressed by setting '"

aCt} = ~ t
n
+ f;3 (10-40)

where f;3 is a constant phase angle defining the angular displacement of the
two coordinate frames at time t O. Substitution of this expression for
440 Prindp!e, of electromechanical-energy conver,ion

a into Eqs. (10-39) gives the currents in this rotating a-b reference frame as
~: = V2 r cos ('" - fJ) i~ = V2 r sin ('" - fJ) (10-41)
with both", and fJ as constant phase angles. Equations (10-41) show
that the currents in the a and b coils on this rotating coordinate system
are not time-dependent but are actually constant or direct currents.
Direct currents in stationary coils set up a constant or static surface­
current distribution in space. By physically rotating the coils at a con­
stant angular frequency, this curren~ distribution is caused to rotate in
space. Thus, a rotating current distribution can be generated by station­
ary windings, such as the d-q windings, with time-varying current,s or,
equivalently, by rotating a-b windings with constant currents.
SUllpose that the original d-q windings set up a surface-current dis­
tribution in the negative-angular direction. As shown in the preceding
section, this can be accomplished by having balanced excitation with
-14 = 1; = r (10-42)
Substituting these currents into Eqs. (10-36) and (10-37) gives the a-b
axis coil currents as
i: = - v'2 Ir cos (wt + '" + na)
(10-43)
i~ = V2I r sin (wt + '" + na)
As before, consider the a-b axes to be rotating in a positive-angular direc­
tion with angular frequency win rad/sec. Therefore, substituting Eq.
(10-40) for a into (10-43) gives
i; = -V2Ir cos (2wt + '" + fJ) (10-44)
i~ = V2 Ir sin (2wt + '" + fJ)
Equations (10-44) indicate that the equivalent a-b axis currents are now
time-dependent at twice the radian frequency of the original d-q currents.
In fact, the currents, defined by Eqs. (10-44), are setting up a current
distribution rotating in the negative-angular direction at an angular
frequency of 2w/n rad/sec. Since the a-b axes are rotating at win rad/sec
in the positive-angular direction, the net angular frequency of the field is
win rad/sec in the negative direction, as it must be if this transformation
procedure is valid.

10-4 A-8 MACHINE EQUATIONS AND PARAMETERS


The complete four-winding d-q primitive machine is shown in Fig. 10-4.
The equilibrium equations for the original d-q primitive machine were
formulated in Chap. 7, with the complete results given by Eqs. (7-33) for
q

1
If'
i{

Fig. 10-4 The complete four-winding


d-q primitive ma.chine.

the four electrical ports and by Eq. (7-34) for the mechanical port. In
forming a d-q machine equivalent to the a-b primitive machine we shall
assume that t,he two d-q stator windings are identical. Also, the two
d-q rotor windings are identicaL Thus we have
K~ = K; = K' turns/m (10-45)
K~ = K; = Kr turns/m (10-46)
By use of the d-q parameters developed in Sees. 6-8 and 6-9, Eq.
(10-4.') or (10-46) can be substituted to give the following results:

Rd = R'
q
= R' = 4aK'p(lA + 2a) ohms (10-47)

R~ = R; R' = 4aKrp~ + 2a) ohms (10-48)

G~~ =
q ngo
3
nMor ;. Jl.oa l1fK'Kr (1 _~)
2go
henrys 00-49)

G'~ nM~r Jl.oa 11fK·K' (1 + ~)


3
= = henrys (10-50)
II nyo 2go
Gd~ nL; = n(L' - L.) henrys (10-51)
G;;' = nL~ = n(l1 + 1...) henrys (10-52)

In Eqs. (10-51) and (10-52) the inductance terms 11 and L. are defined by
8
L ' -_ Jl.oa 11r(Kr)2 henrys (10-53)
n2yo
Jl.oa 3l1f(Kr) 2g1
L. = 2 n 2go 2 henrys (10-34)

The inductance L. is termed the saliency inductance, since a nonsalient


machine has Yl = 0 and thus L. is also zero.
442 Principllu of fliectromechonical-energy conversion

Upon substituting Eqs. (10-47) to (10-52) into the elec!.rical-port


equilibrium equations for the d-q machine given by (7-33), we have as a
result in matrix form
(10-55)
where we have defined

(10-56)

(10-57)

and for the impedance matrix we have


R' + L~p 0 M~rp
Z" _ 0 R' + L:p 0 o
M';p ]
aq - [ M~rp nM;rw r Rr + (U + L.)p n(Lr - L.)wr
-nM:;w r Af';p -n(U + L.)w r Rr + (U - L.)p
(10-58)
The d-q impedance matrix, given by (10-58), is the result of having
identical windings on the stator and identical windings (which are in
general different from the stator windings) on the rotor.
In Fig. 10-5 the four-winding a-b primitive ma(~hille is drawn.
The a-b machine is identical with the d-q machine ill every detail with
one main exception. The rotor windings on the a-b machine are excited

a'

as

Fig. 10-5 The complete four-winding


a-b primitive machine.
The a·b primitive machine 443

through slip rings rather than through a commutator-and-brush mecha­


nism. Equation (10-32) tells us how to transform a set of d-q rotor
currents to an equivalent set of a-b rotor currents. The stator currents
in the two primitive machines are identicaL Therefore, the following
matrix equation transforms all four d-q currents into equivalent a-b
currents:\
(10-59)
where

(10-60)

I-CO-:~O 'in~
n -a-- ""]
-sin na cos na
(10-61)

and i~~ is defined by (l0-57). Expand Eq. (10-59) and convince yourself
that this equation is correct. The matrix given by Eq. (10-61) is referred
to as the complete d-q to a-b transformation matrix. Sometimes it is also
called the stationary to rotating transformation matrix. Notice that the
matrix a~.dq given by (10-33) is only a part of the complete transformation
matrix.
Premultiplying Eq. (10-55) by the d-q to a-b transformation matrix
yields
a·'
aI>.dqv·' - a·'
dt -
7..'j"
'f'.dq-drdv (10-62)
where a~.dV is defined by expression (10-61). The d-q axis currents
can be expressed in terms of equivalent a-b axis currents by using
(10-63)
Substituting these a-b currents into Eq. (10-62) gives us
(10-64)
For convenience the following two definitions are made:

v:; = a:h.c/qv:,; (10-65)


Z:l = a:h.dqZ~(a:h.dt)-\ (10-66)

I The order of the SUbscripts indicates that the d-q currents are being transformed
into the equivalent a-b currents. The double superscript means that botA the stator
and rotor currents are being transformed.
444 Principles 0' e/ecfromecMmicQI.energy conversion

Thus Eq. (10-64) can be written simply as

(10-67)

and has precisely the same form as Eq. (10-515) in terms of the d-q coordi­
nates. The transformed voltage matrix v''; defined by Eq. (10-65) can
be considered as a set of coil voltages in the a-b reference frame equivalent
to the original coil voltages in the d-q coordinate frame; therefore, we have

~'"" = :iv:]
Vb
(10-68)

I
If the d-q voltages are given, then, by using (10-65), all four of the a-b
voltages can be found.
The transformed impedance matrix Z~ gives the relationship between
the coil currents and voltages in this new a-b frame. After a great
amount of manipulation Eq. (10-66), usingZ~~ defined by Eq. (10-58)
and a~.dq defined by Eq. (10-61), can be expanded to the following form: 1

In obtaining Eq. (10-69) the assumption that 0: = w' has been made.
Thus the rotor of the a-b machine has an angular velocity equal to that
of the rotor of the d-q machine.
The various terms in the impedance matrix for the a-b primitive
machine can easily be identified. For the two a-b axis stator windings
(10-70)
as we would expect, since the stator windings on both machines' are
identical. Between adjacent stator and rotor windings of the a-b machine
we have the following mutual inductances:
M:;'(a) = M"; cos na M~(a) = - M"; sin na
(10-71)
Mb;;(a) M:r cos na M~(a) = M"; sin na
Notice in each case that the mutual inductances are functions of
the angle a. The a-b machine has its parameters as functions of the
rotor coordinate, while the d-q parameters are always independent of the
I Remember that or is a function of time and therefore, for example, the first..row­
third-column term M';p cos nor is equivalent to M'; cos nor p - M';na sin nor.
The a-b primitn,e machine 445

mechanical coordinate. If a is held equal to zero, then the ar and br


windings are identical with the d r and qr windings, respectively. With
a = 0, M: = M:; and M:;' = M~r as would be expected, while
M:'; = M::; = 0
since these windings are in quadrature.
The rotor-winding resistances are seen, from (10-69), to be
R: = R;' = Rr (10-72)
The self-inductances of the rotor windings are
L~ = L' + L. cos 2na [",I. = [1 - L. cos 2na (10-73)
with a double angle variation due to the machine saliency. With no
saliency, L. = 0 and L~ and Lr. would be constants equal to Lr. Also
because of saliency a mutual inductance between the two rotor windings
is given by
M;(a) = - L. sin 2na (10-74)
For a = k7r/2n, with k equal to any positive or negative integer including
zero, M~ "" O. For these values of a t.he rotor windings are symmetrical
with respect to the air gap and t.he mutual inductanee between t,he two
rotor windings is equal to zero. However, between these values of a
the air gap is not symmetrical with respect to the windings, and thus
the mutual induetance between the ar and br windings is not zero.
In using the a-b machine for the modeling of an a-c machine, the
various windings of the actual mnehine can be matched with windings
of the a-b model. Dimensions of the actual a-c machine can also be
included in the a-b model; thus all of the a-b parameters can actually be
calculated. Equations (10-70) to (10-74) give the relationships between
the a-b parameters and the d-q parameters which together with the a-b
to d-q transformation matrix completely define t.he equivalent d-q model
machine. The d-q equations are easier to work with, which simplifies
the analysis. These points will be clarified by examples in the following
two chapters.

10-5 LAGRANGIAN FORMULATION OF THE A-8 MACHINE


The electrical-equilibrium equations for the a-b primitive machine have
been formulated in Sec. 10-4 by transforming the equilibrium equations
of an equivalent d-q primitive machine. An alternative approach for
obtaining these equations is to start with the a-b machine and formulate
the equations directly by using Lagrange's equation. If the Lagrangian
4.c.6 Principles of electromechanical·energy conversion

for the a-b machine can be written, then the equilibrium equations can
be obtained very systematically. Lagrange's equation applies to the
a-b machine because the slip rings represent a holonomic coordinate
constraint. As discussed in Sec. 2-4, holonomic constraints between a
set of coordinates allow us to obtain a set of generalized coordinates
sufficient to describe the system completely. The number of generalized
coordinates is equal to the number of degrees of freedom possessed by
the system. The slip ring represents a holonomic constraint because
t.he actual currents ill each of the rotor coils can be defined in terms of
t,he brush or port (·urrents. The brush eurrent. can then serve as a
generalized eoordinatc for the slip-ring winding. Once the brush current
is specified, each of the <'oil currents is then also specified.
T\le ('olllll1utator-brush switehing me(~hanism on the rotor of the
d-q primitive machine represents a nonholonomic constraint. If the port
current for a cOlHlllutator-brush rotor winding is specified, then eaeh
of t he individual rotor eoil currents is not determinable until the position
of the rotor is also specified. Thus the constraint equations for a com­
mutator-brush winding give the individual coil eurrents as a nonanalytic i
function of the port or brush eurrent and t,he rotor position. The func- ~
tion is 1I0nanalytic because the coil currents jump from one value to
another as the coil is cOllllllutated. The rotor position is often an
unknown coordinate whieh we are interested in finding. Since the
constraint equations for the d-q ma('hine are nonholonomic, a set of
generalized coordinates cannot be obtained and Lagrange's equation
cannot be used for obtaining the d-q equilibrium equations. In Chaps.
6 and 7 the d-q equilibrium equations were obtained by resorting directly
to Faraday's law.
Heferring to Fig. 10-;,), a loop formulation must be used because we
have ll1utual-induetive coupling. The flux linkages, !IS functions of the
coil ('urrents, are as follows:
o M~:(a}
L: Mb:(a}
(10-75)
It;:(a) L:(a}
lIf;;(a} lIf:;;(a}

Reciprocity has been assumed in the inductance matrix of (10-7.5), and


the angular dependence of the mutual and self-inductances is indicated.
The magnetic-coenergy-state function is given symbolically by
II

It''('''}
.. tab,a =
\' Jo
L" (i."(,,
"k f ,a
}d"fk (10-76)
i-\

where i~ IS an abbreviation for aU four a-b machine currents. Upon


The a-b primitive machine 447

substituting the flux-linkage relationships, given by Eqs. (10-75), we have


Cor the magnetic-coenergy-state Cunction
W:(i:b,a) = j-L:(i:)%+ jL:(i:)2 + M:'(a)i:i;
+ Mb~(a)ibi: + iL;(a)(i;p + M:b(a)i:ii:
+ Mbi:(a)i:ii: + M:h(a)i;i; + iL;(a)(1--;}2 (10-77)
The rotor-surCace-current distributions Cor the a-b machine windings
are given in Sec. 10-1. The a-b stator distributions are identical with
the d-q stator-surCace currents. If these current distributions are known,
the c;orresponding air-gap fields c;an he d('termined by using Ampere's
circuital law. From the field expressions the magnetic energy stored
in the air gap can he determined. The a-b machine is linear, and there­
fore the magnetic: eoenergy is equal to the magnetic energy stored. If
the field-derived expression for the coenergy is matched to Eq. (10-77),
the various self- and mutual inductances will be specified in terms of the
machine geometry. Equations (10-70) to (10-74) give all of the a-b
inductance parameters in terms of the d-q parameters. Equations
(10-47) to (10-54) give a summary of the d-q parameters in terms of the
machine geometry; t.herefore, by substitution, we have the a-b param­
eters defined in terms of the machine geometry. The field approach is
identical with t.he method used in Chap. 6 for finding the d-q parameters,
so it will not be repeated here.
Substituting the d-q equivalent parameters from Eqs. (10-70) to
(10-74) for all of the a-b parameters in Eq. (10-77) gives us
W:(i:b,a) = ·l!.d(i:)2 + jI.;(iZ)2 + M·,t cos na i:i:
+ M"; sin net ii,i; + HJ.r + J,. cos 2na) (i:P
- ¥:; sin na i:ii: + M;r eos na ii,ib
- L. sin 2na i;i; + i(U - L. cos 2na) (1--;)2 (10-78)
Equation (10-78) is the magnetic-coenergy-state Cunction for the a-b
machine. The d-q parameters are summarized by (10-47) to (10-54)
in terms of the geometry of the primitive machine. The equivalent
a-b machine has the same geometry, and thus the results of (10-47) to
(10-54) are valid in Eq. (10-78).
The a-b machine-coenergy function is
(10-7.9)

where J is the polar moment of inertia of the rotor. Neglecting the sha.ft
complia.nce, the system energy function 'l) is zero. The system's La.gra.n­
gian is therefore

(10-80)
448 Principles of electromechanical-energy conversion

where W: is defined by Eq. (10-78). The a-b machine has a Rayleigh


dissipation function given by
5(i:';,a) = lR'[(i~)2 + (i~)2J + jRr[(i~)2 + (ib)2} + j-Da 2 (10-81)
The external forces
Q.. = Tr (10-82)
are voltage sources for the electrical coordinates and an applied torque for
the mechanical coordinate. The external torque applied to the shaft of
the a-b machine is taken to be the same torque Tr applied to the equiv­
alent d-q primitive machine. The reason for using the same Tr is dis­
cussed in the next section.
Thf equilibrium equations for the a-b machine can now be obtained
by using Lagrange's equation, which for convenience is repeated as
~ [O.c(i!t,a,a)] _ 0.c(i:};,a,a2 + 05(i!t,a) Qk

dt OEk O~k OEk

for k = 1, . . . ,5 (10-83)
where E! = i:, ~2 = i~, ~3 i:, E4 i b, and h a. Using k = 1, 2, 3,
and 4 in Eq. (10-83) will give Eq. (10-67) as a result. The mechanical­
equilibrium equation for the a-b machine is obtained by setting k = 5,
thus giving

~ [o.c(i!t,a,a)] _ O.c(i!b,a,a) + ~(i!b,a2 = Q.. (10-84)


dt oa oa oa

From Eqs. (10-80) and (10-81) we have

~ (:~) Ja (10-85)
o.c
oa
+ 111-; sin na i:ib + L. sin 2na (i:)2 + 2L. cos 2na i~i;;
- L. sin 2na (iiYl (10-86)
05 = Da (10-87)
oa

Substituting Eqs. (10-85) to (10-87) with Q.. = Tr into Eq. (10-84) gives
us for the mechanical-equilibrium equation of the a-b machine

, Ja + Da + T. = Tr (10-88)
'. where the torque of electrical origin T. is given by Eq. (10-86).
If equivalent d-q and a-b currents are respectively applied to the two
equivalent primitive machines, then the same torque of electrical origin
450 Principle. of electromechanical-energy conver.ion

By use of the d-q to a-b transformation given by Eq. (10-63), the equiv­

alent d-q currents are


(10-93)
Upon substituting Eq. (10-93) into (10-91), we have
P~":' = (i:i) '[ «id~ .•b) '!R(i:[q.abli:;; (10-94)
after using the matrix identity
«id~,ubi:i)' = (i:i)'«id~.ab)' (10-95)
for the transpose of the product of two matrices.
Equation (10-94) is precisely equal to P~-::·, the ohmic losses in the
a-b macthine, if we ('an show that
(10-96)
By simply substituting the inverse of Eq. (10-61) for (i~~."b and Eq. (10-92)
for !R, Eq. (10-96) can be verified. Thus,
(10-97)
whieh means that the powers in either the a-b or d-q primitive machine
lost in ohmic heating are equaL
Let us now examine the total power supplied to all five ports of the
two primitive ma('hines. For the d-q maehine the toial power is given
by Eq. (7-17) as
(10-98)
By inversion of Eq. (10-65), the d-q port voltages are
(10-99)
By substituting Eqs. (10-93) and (10-99) into (10-98), we have
(10-100)
after using t he identity given by (10-95). Equation (10-100) is equal to
the total power into all five ports of the a-b machine if we can show that
(10-101)
or equivalently
(10-102)

where 'U stands for the unit matrix, with l's on the principal diagonal and
O's in every other position. If a transformation matrix premultiplied by
the transpose of the transformation matrix equals the unit matrix, then the
The a-b primitive machine 449

should be generated in each case. Equation (7-24) expresses the d-q torque
of electrical origin for the case of identical quadrature- and direct-axis
rotor windings. For convenience (7-24) is repeated as
(10-89)
with the substitution L~ - 1..; = 21... from Eqs. (10-51) and (10-52).
The d-q currents equivalent to the a-b currents can be obtained from the
matrix (10-59) by premultiplying both sides by (1~~."" = «(1~.tl'l)-l. Upon
expanding this equation, we have
i~ = i; cos na - i;; sin na (10-90)
i; = i; sin na + i;; cos na
Substituting the equivalent currents of (10-90) into the d-q machine torque
of electrical origin gives us precisely Eq. (10-86), which is the a-b machine
torque of electrical origin. Thus, as we would expect, both primitive
machines generate the same torques for equivalent currents. In Sec. 7-3, we
obtained T. for the d-q machine by using the first law of thermodynamics.
Here we have obtained T. for the a-b machine by using Lagrange's equa­
tion. The equivalence of these different formulations is certainly evident.

10-6 POWER INVARIANCE


The a-b or slip-ring primitive machine is constructed to be identical with
the d-q commutator-and-brush primitive machine except for the manner
by which currents are supplied to the rotor circuits. Equation (10-61)
gives a 4 X 4 matrix which allows us to specify the four a-b currents
establishing stator' and rotor current distributions identical with those
established by a set of four d-q currents. In Sec. 10-5 the torques of
electrical origin for equivalent a-b and d-q currents were shown to be
identical. Since the angular velocities of the two machines are equal, the
converted powers, given by T.w', must also be equal.
For the d-q machine the total power lost in the form of ohmic heating
is given by Eq. (7-19), which for convenience is repeated as

where the resistance matrix is defined as

~. ~. 1]
R'
iR = 0 (10-92)
[ o
o
The a·b primitive machine 451

total power is invariant or constant under this transformation. Upon


inverting Eq. (10-61) we have

[1oo 0
-L.]
0 0 0
1
0
0
cos na
oo ]
sin na
X [ 00 1
0
0
cosna
o 0 -sin na cosna 0 0 sin na cosna

[1o 00 0]
1 0 0
(10-103)
o0 1 0
000 1

Since Eq. (10-101) is satisfied for the a-b to d-q transformation matrix
a 4 ."", the total electrical powers supplied to each machine are equal; thus
we can write

(10-104)

Using the first law of thermodynamics, these total powers can only con­
tribut.e to energy lost or stored. Therefore at each instant in time

(10-105)

Equation (10-97) shows the power-loss terms to be equal, and Eq. (10-104)
equates t.he total powers. Therefore, the portions of the total power
supplied to each machine adding to t.he stored energy are also equal.
Thus we have shown that the total power is invariant under the a-b to d-q
transformation. Also, each component of the power is invariant.
The commutator of the d-q machine has been replaced by slip rings
to obtain the a-b ~)achine. As a compensation for this constructional
change, transformed currents are supplied to the a-b rotor windings in such
a manner that all current distributions remain the same. Since the
windings are the same and the currents are the same between the two
machines, it is not too surprising that the powers should be invariant.

10-7 THREE-PHASE TO TWO-PHASE TRANSFORMATIONS


On both the a-b and d-q primitive machines t.he stator and rotor windings
have been placed on two axes 90° apart. Figure 10-6 shows two identical
stator windings on the a and b axes of the a-b machine. Since the stators
of the a-b and d-q machines are identical, Fig. 10-6 could also represent the
d-q primitive-machine stator. Inserting two currents in these windings
establishes sinusoidally distributed current sheets on the stator inner
452 Principles 01 e/ectrome<:hanical·energy r:onversion

periphery. These current sheets for an n-pole-pair winding are given by

J: K! sin nO i:a. amp/m


(10-106)
JZ = - K: cos nO iZa• amp/m

. For two identical stator windings we would have

K! = K~ = K:O (10-107)

a.-b primitive-machine rotor windings establish similar distributions, as


given by Eqs. (10-3) and (10-4), with the substitution of 0 - ex for 0, and
the appropriate rotor distribution factors for K:O. The angle ex takes into
account the movement of the rotor-surface-current distribution if the rotor
is turne4.
The a-b primitive machine has two circuits on the stator and two
circuits supplied through slip rings on the rotor. Each of these stator or
rotor circuits is known as a phase. The a-b machine has a two-phase
stator and a two-phase rotor. Most common polyphase a-c machines
have three-phase windings rather than two-phase windings. A three­
phase stator is shown in Fig. 10-7, with three identical windings on the
ex, ~, and 'Y axes. I Each of these three windings establishes an n-pole-pair
sinusoidal current distribution along its respective axis. However, each
distribution is 271/3 rad or 1200 from the other two distributions, rather
than 1r /2 rad as with the two-phase winding. Therefore, the current
sheets established by the currents in the three windings shown in Fig.

I The symbol a is al80 used to designate the angular position of the a-b machine rotor.
The three symhols a, {3, and "I used as phll.8e-winding designations will usually appear
8.8 subscripts, 80 no confusion should result.

K'b

Fig. 10-6 A two-phase set of stator


windings on the a and b axes.
The a-b primitive machine 453

Fig, 10-7 A three-phase set of


stator windings on the or, {J, and l'
axes,

10-7 are given respectively by


J~ = K~ sin n8 i~a. amp/ill
J;, = K~ sin (n8 + 2;) i~. amp/m (10-108)

J•')<- . (8
- K'')<sm 211') 1,''.)<a.
n -3 amp/m

For three identical windings the distribution factors are all equal; there­
fore we may write
(10-109)
The total surfa('c-current. distrihution duc to the three-phase stator
willdings ean be found by simply adding the individual distributions.
Thus we have as the total three-phase stator distribution
J~#')< = J~ + J~ + J~ (10-110)
By subst.ituting Eqs, (10-108), the t.otal distribution J~#')< can be put into
the form
,
J all')< = K'a(h [(.t~ + cos 3211' 1,.& + cos 211'
3 1,;.), sm n 8

- ( - sm 3 t~ + Sill
,211'. . 3211'.)
t; cos n 8] a. (10-111)

where we have used the trigonometric identities for the sine and cosine for
the sum of two angles.
In a similar fashion the total surface-current distribution established
by the a and b axis windings is given by

J~ = J~ + Jt (10-112)
454 Principle. 01 electromechanical-energy conversion

Substituting Eqs. (10-100) for the individual current distributions leads


us to
(10-113)
For conveniellce in subsequent expressions let. us assume that the two
windings on the a-b maehine have distribution factors which are equal to
the distribution factors for the three-phase windings. Thus we have
(10-114)
The total current distributioll established by the three-phase wind­
ings can also be established by the two-phase windings. However, for
eonvenie!l("e in later expressions t he two-phase eurrent distribution is to
be lI1aqe equal to two-thirds of the total three-phase current distribution.
Compahng Eq. (10-1ll) with Eq. (10-113), we can make J~ = iJ~fj1 pro­
vided the a-b currents are so chosen that
'J'
~t~ = '+ cos 211'.
t! 3' t~ + cos 211'.
3' t~
(10-115)
a' . 211'. + . 211',
'It: = -sm - t: sm - t'
3 " 3 ~

The relationship between the distribution factors given by (10-114) has


been used in establishing Eqs. (IO-l!.,}). For a given set of three-phase
stator currents i~, i~, i; a parti(~ular current distribution is established on
t he inlier periphery of the stator surface. Equations (10-115) specify the
set of a-b winding currents establishing t.wo-thirds of this current
distribution,
By substituting the values for sin (211'/3) and cos (211'/3), we can
rearrange Eqs. (10-11.5) into the matrix form l
i~ = a~.../lJ~'l' (10-116)
where
(10-117)

(10-118)

and

a'""'.,.{joy -- i
[
1
0 -']
~~ (10-119)

I This transformation will he used only for balanced three-phase excitations and ill

therefore sufficient for our needs. For Ii more complete discU88ion see A. E. Fitz­
gerald and C. Kingsley, "Electric Machinery," 2d cd., pp. 229-231, McGraw-Hill
Book Company, New York, 1961.
The o-b primilive mochine 455

The three-phase to two-phase transformation matrix given by (10-119)


relates to work done originally by Andre Blondel in France, and the tech­
nique, when applied to synchronous and induction machines, is often
referred to as the Blondel two-reaction method.
Equation (10-116) transforms t,he three-phase currents into an equiv­
alent set of two-phase currents. Following a similar development for the
transformation relating the a-b machine to the d-q machine in Sec. 10-3,
the same transformation (10-116) can be used for the respective port­
voltage variables. Thus if we have a set of three-phase stator voltages
v!, v~, v~ applied to the three windings in Fig. 10-7, an equivalent two-phase
stator voltage set is given by
(10-120)
where

.
V_L = V!] (10-121)
- Vb

v!jJy = v'vi ] (10-122)


v'..,
and Eq. (10-119) defines the transformation matrix.
Let us now examine the power supplied to the two-phase windings
and compare this power with that supplied to the equivalent three-phase
winding. In matrix form the two-phase power supplied is given by

P:'" = (i:"')'V:'" (10-123)


where the superscript t is used to indicate t,he transpose of the column­
current matrix, which gives the row-current matrix. Substituting Eqs.
(10-116) and (10-120) into (10-123) leads to
(10-124)
after we use the matrix identity for the transpose of the product of two
matrices. Upon expanding Eq. (10-124) by using Eqs. (10-118), (10-119),
and (10-122), we have

p:.,. = t[(v! - !v~ - !v~)i! + (-!v! + v~ - !v~}i~


+ (-jv! - jVg + v~}i~l (10-125)

as the total power supplied to the two-phase windings in terms of the


equivalent three-phase voltage and current quantities.
Equation (10-12.') has a very interesting interpretation if the three­
phase voltage quantities are obtained from a balanced three-phase sinus­
456 Principles of e/ectromechonicol.energy conversion

oidal supply. With a balaneed supply we have

t'~ = V2 V· cos wt v~ = V2 V· cos(wt _ 2;)


v~ = V2 V· cos (wt + ~;) (10-126)

The three quantities in parentheses in Eq. (10-12.'» can therefore be


reduced in the following manner: 1

v~ - !v~ - i(V2 V· cos wt) = ~IJ~


!v~ =

- ~vr + vii - ~IJ~ = ~ [ V2 V· eos (wt - ~:) J = ~lla (10-127)

-il'~ - !II~ + v; j [V2 V· cos(wt + 2;)] = .~v~


Substituting (10-127) into Eq. (10-125) gives us

(10-128)

where P!-y(j is the total power supplied to the three-phase windings.


With the balanced three-phase voltages given by (10-126) applied to
three identical windings we would expect that the same average power is
absorbed by each phase. All polyphase a-e Illaehines arc constructcd in
sueh a fashion that the three windings are identical except for their relative
angular displacement. The average power per phase is therefore simply
one-third of the total average power; thus we have

(10-129)

where a bar is used to indicate average value and the subscript q, to


indicate phase power.
Substituting the bululH'cd three-phase voltages into Eq. (.10-120), the
equivalent two-phase voltages are found to be

": = V2 V- cos wt v; V2 V· sin wt (10-130)

Notice that the two-phase voltages are also balanced, but they are only
90° out of phase. Since balanced two-phase voltages are applied to the
two identical a-b windings, the average power supplied to each phase is

I Equations (10-127) can easily be checked by first rorming the sinor equivalents of
the three sinusoidal voltages and then adding the appropriate parts or the complex
numbers either analytically or graphically.
The a·b primitive machine 457

simply olle-hulf of the totnl average power supplied to both windings.


Thus we have

(10-131)

Taking the average value of both sides of Eq. (10-128) and dividing
by 2 gives us

(10-132)

Upon substitut.ing Eqs. (10-129) and (1O-1:J1), we have, finally,

Equation (10-1:J3) tells us that the average power supplied to each phase
of the two-phase system is equal to the average power supplied to each
phase of the three-phase system.
By using the transformation given by Eq. (10-119) and the distribu­
tion factors given by (10-114), a set of two-phase windings can be con­
structed and excited in such a fashion that current sheets whieh are
equivalent to but not identical with those established by a balanced three­
phase system are established. Equations (10-130) show us that v~ = v~,
and therefore we ('an easily show that i~ i!. Thus the a phase of the
two-phase system has port variahles ident.ietll with those of the a phase
of the thl'ee-phase system. A trallsforllllttioll b:wk to three-phase quan­
t.ities from the t.wo-phllSe quantities is re!tdily aceolllplished, since the
/J- and 'V-phase quulltit.ies arc simply displaced by -2'11'/3 and 2'11'/3 rad,
respect.i vely.
All power quantit.ies eal(~uhtted in terms of the two-phase system can
be written on a per-phase basis. For the three-phase system a multiplica­
tion by 3 yields the corresponding total t.hree-phase power quantity. By
using the results of this section, all symmetrical three-phase a-e machines
can be reduced to equivalent. t.wo-phase machines and thus fit the a-b
primitive-machine model.

10-8 IMPEDANCE MATRIX TRANSFORMATIONS


Let us now examine the effeet of the aIJ-afJ'y transformation on the
impedanee nmtrices for Cl\(~h system. We wish to obtain t.he a-b imped­
ance mat rix equivalent to a given a-/J-'V impedance matrix. This discus­
sion is similar to that of Sec. 10-4, where the d-q impedance matrix was
transformed to an equivalent a-b impedanee matrix.
458 Principles of electromechanical-energy conversion

By comparing Eq. (10-128) with Eq. (10-124), we can see that for
the special ease of balanced three-phase excitations the matrix product

1 0 0]
(a:u,.a~.,)'a:b.a~., = i
[00 01 10 = i'U (10-134)

where 'U is the unit matrix as shown in (10-134). Premultiplying Eq.


(10-120) by the transpose of the three-phase to two-phase transformation
matrix leads to

(10-135)

For bah.~nced sinusoidal exdtations, the result. of (10-134) can be sub­


stituted and thus gives us

(10-136)

By expanding (10-136) with the use of (10-119), we can again see that the
a-phase quantities are identieal with the a-phase quantities. A similar
expression can be written in terms of the eurrents; thus

(10-137)

For the three-phase stator windings shown in Fig. 10-7 assume that
we can write a set of equilibrium equations in the form

(10-138)

By using Eq. (10-137) for i!~., and premultiplying by the ab-aI'J'Y matrix,
we have

(10-139)

The bracketed quantity is recognized as being the corresponding a-b


machine impedance matrix; thus

(10-140)

Equation (10-140) tells us how to transform the three-phase impedance


matrix into an equivalent a-b machine impedance matrix.
Assume that the three windings shown in Fig. 10-7 are all identical
and are wound on a nonsalient stator structure. The resistance and self­
induetance of each winding are R' and V, respectively. These windings
are to be excited by a balanced set of three-phase voltages. Neglecting
The a-b primitive machine 459

any leakage effects, the equilibrium equations for the three windings are
given by Eq. (10-138), where

L' L'
R' + Vp -2 P -2 P
L' p L'
Z!/I.., = R' + Vp -2 P (10-141)
V V
-2 P -2 P R' + Vp
The mutual inductance between any two of the windings is equal to
La cos (211'/3), or -V/2 wit.h no leakage fields.
By using Eq. (10-140), the equivalent a-b impedance matrix can be
found with the ab-a{fJ'Y transformation matrix defined by (10-119). After
a good deal of manipUlation, we obtain

· - [R' +0 lLap
Zoo - R' +0 iL'p J (10-142)

as the a-b machine equivalent impedance matrix. The interesting thing


about Eq. (10-142) is the factor of i preceding the inductance coefficients.
By simply examining the two a-b stator windings in Fig. 10-6, the imped­
ance matrix we would probably write would be identical with (10-142)
without the i factors. This factor of i appears beeause we arbitrarily
made the a-{fJ~'Y total surface-current density i times the a-b total surface­
current density in formulating the ab-a{fJ-y transformation.
If another set of a-{fJ--y windings were now placed on the rotor of the
machine, an equivalent set, of a-b rotor windings could be obt.ained. Each
of the a-b rotor windings is identical with the three identical a-{fJ--y rotor
windings. Figur~ 10-Sa shows the complete three-phase machine with
three slip-ring rotor windings each having a distribution factor K' and
three stator windings each having a distribution factor K·. In Fig. l0-8b
is shown the equivalent two-phase a-b machine having the same rotor
distribution factor K' and stator distribution factor K·. By writing the
six equilibrium equations for the three-phase machine shown in Fig. 10-8a,
we can obt,run the 6 X 6 impedance matrix for the machine. By gen­
eralizing the transformation equation (10-140), we can obtain the imped­
ance matrix for the equivalent a-b machine of Fig. lO-8b from

(10-143)
where

(10-144)
460 Principles of electromechanical-energy conversion

yS

)L----j--.pC¥::q--L-_ _ (x/

/3'
(a)

bS

Fig. 10-8 (a) Com­


plete symmetrical three­
phase machine. (b)
Equivalent symmetrical
two-phase a-b model ma­
(b) chine.

Remember that (10-144) is a 4 X 6 matrix with a 2 X 3 matrix in each


corner. I
The net result of expanding Eq. (10-143) would be identical with Eq.
(10-69) with a factor of! appearing before each self- and mutual-induct­
ance coefficient. Therefore, we can always avoid writing the equations
for the three-phase machine by simply writing the equivalent two-phase
I By using the power invarianee condition of (10-102), the d-q to a-b impedance trans­

formation given by (10-66) becomes


Z:b = (l:b.~~(<<:b.dq)1
This form is identical to the three-phase to two-phase transformation given by
(F -I. 10-143).
The (Job primitiYe m(Jd!;ne 461

equations and inserting a factor of f before each inductance coefficient


involving an equivalent two-phase variable. If the a-b machine is
transformed to an equivalent d-q machine, the factor of J appears before
each indu(·tance coefficient (including self-, mutual, and rotational induct­
ances) involving a d-q variable equivalent, through transformations, to a
three-phase variable.

10-9 SUMMARY
The study of a-c machines requires a model primitive machine having
rotor windings supplied currents through slip rings rather than a com­
mutator-and-brush meehanism. The a-b model machine provides the
desired new primitive ma(~hine. The iron struct,ure of the a-b machine is
identical with that of the original d-q primitive machine. The two stator
windings of each machine are also identical with those of the other.
Since most a-c machines are c01lstrll(~ted with identieal windings on each
stator axis, the winding distribution faetors of the two stator windings on
the a-b machille are taken to be identical with each other, and they are
also identical with the two stator windings on the equivalent d-q machine.
In addition, the two a-b rotor windings are identical with the two d-q
rotor windings. However, t.he rotor and stator windings are, in general,
different.
The only constructional differe1lce between the two machines is the
mechanism by which current is supplied to the identical rotor windirigs.
The a-b machine uses slip rings, whereas the d-q primitive machine
employs the commutator-brush mechanism. In Chap. 6 we found the
commutator-and-brush switching on the d-q machine maintained the
rotor-surface currents stationary in space irrespective of the angular posi­
tion or veloeity of the rotor. No switching takes place with the slip rings
on the a-b machine, and therefore the rotor-surface currents move with
the rotor.
By using a matrix transformation, a-b and d-q currents which main­
tain identical rotor and stator surface-current distributions on each of the
two primitive machines can be found. The form of the transformation
giving the a-b currents in terms of the d-q currents is

<l:.b.dQ = [~o !
0
11 La 'in no]
-sin na
g
eos na
(10-145)

The subseript notation indicates that the transformation gives the first
two subscripts from the second two subscripts, or that we obtain the a-b
462 Principles of e/ectromechanical-etlftrgy conversion

currents from the d-q currents. The two superscripts indicate both the
stator and rotor currents to be involved in the transformation. The
inverse of matrix (10-145) is written as ad~ ...I" giving the d-q currents in
terms of the a-b currents.
The electrical-equilibrium equa.tions for the a-b machine are obtained
by transforming the original equilibrium equations for the d-q machine.
The a-b parameters can be identified in terms of the d-q parameters. In
Chap. 6 all the d-q parameters were determined as functions of the
primitive-machine geometry. By this identification process, the a-b
parameters are thus related to machine geometry.
For equivalent currents the torques of electrical origin for the two
machines arc found to be identical. Also because of the parameter rela­
tionship~, the components of power (or the two machines are the same.
The (ollbwing list of powers can be equated:
Total power supplied p~~~al = p:~al
Total electrical power supplied Pd~ = P:;'
Total ohmic loss p~~.. = P~~"
Total electrical storage Pd~o'ed = p~ored
Total mechanical power supplied P:" = w'r'
Total viscous or windage loss P!:'" = w'Dw'
Total mechanical storage p:ored = wrJpw'
All rotor-port quantities of one primitive machine are identical with those
of the other. The power that each machine converts from electrical to
mechanical or mecha.nical to electrical is also identical with that converted
by the other.
::\lost practical polyphase a-c machines have three-phase stator or
rotor windings rather than the two-phase windings on the a-b primitive­
machine model. An equivalent set of two-phase windings with the same
number of poles can be constructed with the same distribution factors as
are used for the three-phase windings. By using the transformation

~l (10-146)

all three-phase variables can be converted to equivalent two-phase var­


iables. The energy-conversion properties of the machine remain
unchanged since the transformation maintains equivalent surface-current
distributions.
For balanced three-phase quantities the average power per phase of
the two-phase system is equal to the a.verage power per phase of the
three-phase system. The a-phase varia.bles of the two-phase system
exactly match the a-phase variables of the three-phase system, so a. solu­
tion of the two-phase problem solves the equivalent three-phase problem
The g·b primitive machine 463

at the same time. In writing the impedance matrix for either the a-b or
d-q machine a factor of j must be inserted before each inductance param­
eter involving a term containing a current related through transformations
to the three-phase currents.

PROBLEMS

to-t Starting with the rotor-surface-current distributions given by Eqs. (10-6)


and (10-7), show that F..qs. (10-10) and (10-11) correctly represent the magnetic
fields in the air gap of the a-b primitive machine.
10-2 In Eq. (10-16) verify that the first term represents a current distribution
revolving counterclockwise at win rad/sec and that the second term represents a
current distribution revolving clockwise at the same angular velocity.
10-3 The rotor-surface-current distributions shown in Fig. PIO-3 are uniform
over the respective angular segments. The effect of the individual slots on the rotor
periphery is being neglected. For es!'h of these distributions having equal areas over
:r /n radians derive the fundamental term in a Fourier-sine-series expansion and thereby
determine the distribution factor K. The winding terminal current equals 1 amp.

omp/m
10 ,-­

00 Trln 2Trln 9

-10
(a)

amp/m
20

00 9

-20
(0)

amp/m
15.0
7.5

00 1T/3n 2~3n 5~3n 21T1n 8


% 4;3n ,
-7.5
-15.0
Ie)
Fig. PIO-3
464 Principle. 0; eledromec:honicol-energy conversion

10-4 A four-pole commutator-brush primitive-maehine rotor has its two wind­


ings excited hy various current sources. An equivalent pair of slip-ring windings
are to he excited in such a way that identical surfacc-()urrent distrihutions are to be
developed on the rotor. For each of the following d-q rotor currents find the appro­
priate a-b rotor currents with the given angle between the two framcs.
a. id - l~ - 10 amp; a = 30°
b. i:; ,.., 10 amp; l~ = 20 amp; a 30·
c. ~"d = -7.07 sin 377t; i; "" 7.07 cos 377t; a = 377t/2
d. id -7.07 sin 377t; i; ,.., 7.07 cos 377t; a == -377t/2
e. t"d = -7.07 sin 377t; i; = 14.14 cos 377t; a 377t/2
f. Give a short physical interpretation to the answers for parts c and d.
g. What is meant by an "equivalent pair" of a-b and d-q rotor windings?
Describe similarities and differences.
10-5 Construct an a-b primitive machine which is equivalent to the d-q primitive
machine wltose parameters are listed in Prob. 7-2.
a. Draw a diagram for the a-b machine.
b. List al1 of the winding parameters, including resistance, self-inductance, and
mutual-inductance values. Also include the mechanical parametcrs of the equivalent
a-b machine.
.-:. Write the five equilibrium equations for the a-b machine, including an ex pres­
Ilion for the torquc of elcctrical origin in terms of the equivalent parameters.
JtQO-6 Assume that thc primitive machine shown in Fig. to-5 has the fol1owing
parameter values:
R- ohms
=.'} n = 3 pole pairs

L' 3h R' = 3 ohms

J = 10 kg-m' U = 3h

D = 0.1 newton-m-sec/rad M" ,.., 3 h (max)

a. Are the stator and rotor structures of this primitive maehine salient or non­
salient? Explain your answer.
b. Write the five equilibrium equations for the machine, including an expression
for the torque of electrical origin.
c. Formulate an equivalent d-q primitive machine. List al1 d-q parameter
values.
d- 10-7 A 2-amp direct-current source is connected to the a-axis stator winding
of the primitive machine in Prob. 10-6. The speed of the rotor is maintained con­
stant at 100 rad/sec in the positive angular direction. All other ports are
open-()ircuited.
a. Calculate the steady-state voltage across the current source.
b. Find the open-()ircuit voltage on each of the other three electrical ports.
c. Find the corresponding voltages on the four electrical ports of the equivalent
d-q primitive machine.
d. How much' external torque is required to maintain the speed at 100 rad/seef
10-8 The rotor or the primitive machine in Prob. 10-6 is held stationary at an
angle a 10°. The following constraints are plaeed on the four electrical portli:
Vb = V2 50 sin 377t volts II: "" lib = 0 volts
a. Formulate the three electrical-equilibrium equations in complex or sinor form
necessary to yield the steady-state port currents.
The a-b primitive machine 465

b. Solve the equations of part a for the three steady-state port currents. Express
your answers as functions of time.
c. Find the set of port constraints for the equivalent d-q primitive machine.
d. Formulate the equilibrium equations for the equivalent d-q machine and solve
for all port currents.
e. Transform the d-q currents, found in part b, to equivalent a-b currents. Com~
pare your results with the currents found in part b. .
,. Find the magnitude and direction of the external torque required to maintain
the rotor at a fixed angular position.
~,10-9 The rotor of the primitive machine in Prob. 10-6 ill now being driven in
the positive angular direction at a constant speed of 377/3 rad/sec. The two stator
windings are open~ircuited, and the two rotor windings have the following constraints:

v~ = V2 50 cos 377t volts V2 50 sin 3771 volts

Assume that ",(0) = 0, where ",(t) is the angle between the rotor and stator axes.
a. Derive an expression for <>(/).
b. Find the open-circuit port voltages for the two stator windings.
c. List the set of port constraintl! for the equivalent d-q primitive machine.
Explain these constraints qualitatively.
d. Solve for the open-circuit port voltages for the two stator windings on the
d-q primitive machine. Compare thCS(l rC8ultll with the results in part b.
e. Hepeat parts b to d for v~ = - V2 50 cos 377/, with all other conditions as
previously given.
10-10 A reluctance synchronous motor can he constructed as shown in Fig.
PlO-lO. The single rotor winding is supplied through slip rings. The stator is salient
with two pole projections as shown, and it has no electrical windings. The rotor
has an inertia J and a windage coefficient D. The resistance of the winding is Rr,
and the self-inductance ('an be approximatffi hy Lr L. cos 2a. No external torque
is applied to the rotor shaft, and a voltage source v(/) is ('onnected through the slip
rings to the rotor winding.
a. Draw a primitive-machine model of the machine in the form of Fig. 10-5.
b. Formulate a Lagrangian, a Rayldgh dissipation function, and the external
forces for the machine.
c. By using Lagrange's equation, formulate the two equilibrium equations for
the machine.

Fig. PIO-IO
.c66 Principles of electromechanical.energy conversion

d. If v(t) = v'2 V sin wt, determine the speed at which the rotor must revolve
so that the average torque of electrical origin is nonzero. (Hint: IC v is sinusoidal,
then in the steady state the rotor current must also be sinusoidal with the same
frequency.)
e. If the stator structure has 2n poles (where n is the number of pole pairs)
reformulate an approximation for the winding inductance and repeat parts b to d.
Note: Common electric clock motors are reluctance synchronous machines. Usually,
the single winding is placed on the stator and the rotor structure is made to have
salient poles. Since our primitive-machine model can only have a sa.lient stator with
a nonsalient rotor, the model of the actual machine must be inverted. A similar
procedure is followed in Chap. 11 in discussing synchronous machines with d-c fields.
10-11 a. Prove the matrix identity (10-95) by expanding both sides of the
equation.
b. Verify that Eq. (10-96) is valid and thus prove that the total ohmic losses in
the two ertfivalent primitive machines are identical.
10-12 o. For the following set of three-phase currents circulating in three
identical stator windings, arranged as shown in Fig. 10-7, find an equivalent set of
a..o stator currents which result in two-thirds the current sheet established by the
three-phase curren ts.
i~ = v'2 120 sin (wi + 30) i~ = v'2 120 sin (wi - 90)
i~ v'2 120 sin (wt + 150)
Hint: Sinor notation is very useful in performing this transformation.
b. What is the relationship between the magnetic fields established by the two­
phl\S(> windings and those established by the three-phase windings?
10-13 Show that the transformation given by (10-140), when applied to (10-141),
yields Eq. (10-142).
10-14 Each of the three identical stator phases shown in Fig. 10-7 has a resist­
ance of 3 ohms and a self-inductance of 2 h. Sinusoidal three-phase currents at a
frequency of 2 rad/sec and with an rms magnitude of 5 amp are established in these
stator windings.
a. By using sinor notation, determine the port voltages on each of the three
phases. Take i~ as reference; thus, i~ v'25 sin 2t.
b. Find the avcrage power supplied per phase.
c. Determine the equivalent set of two-phase stator currents and use (10-142)
to find the appropriate two-phase port voltages. Compare the a-phase and a-phase
quantities.
d. Find the average power per phase being supplied to the two-phase windings.
Compare this resulting per-phase power with that obtained in part b.
coMpte.

THE SYNCHRONOUS MACH~E


11
The transmissioll of electric power over long dis­
tall('es is accomplished most economically by min­
imizing the line currents. The power transmission
losses are due largely to ohmic heating; therefore,
by reducillg the currents for a given size cable, these
losses can also be reduced. In order to transfer
large quantities of power with reduced currents, as
high a voltage as possible and practical must be
used. Usually the consumer buying the electric
power cannot uSe the very high voltages required
for economical transmission. Thus the power
utility must be able to adjust the voltage of the sup­
plied power to meet the customer's requirements.
For this reason most electric power transmitted
over some distance is generated in the form of a-c
power. Conventional power transformers are osed
at the generating station to increase the voltage
before the power is supplied to the long-distance
transmission cables. At the receiving or customer
end of the system transformers are again used to
.step down or reduce the line voltage to a safer, more
desirable value. The power transformer can be
used only for a-c power, thus making voltage adjust­
ment very economical with the power in a-c form.
The synchronous generator, or alternator, is
present;ry~used to generate most of the a-c electric
power. The tern1 itsynchro-n-ous"-Tmpliesihat the
machine revolves aCa constant speed. The gen­
erated a-c power is usually sinusoidal in form, and
because of the synchronous operation of the alter­
nator the frequency is held fixed. Most large a-c
alternators generate power into a balanced three­
phase distribution system. Balanced three-phase
distribution systems transmit power in such a fash­
467 ion that the instantaneous power transmission
468 Principle. of electromechanical-energy conver,jon

equals the average power transmission. Single-phase systems have an


instantaneous power transmission which pulsates and is found to be less
economical than the three-phase system. The three-phase system offers
further advantages to the consumer. Three-phase synchronous and
induction machines have more desirable characteristics than single-phase
machines. If he uses a wye-connected system, the consumer has a
choice of voltages by making line-to-line connections or line-to-neutral
connections.
Both the synchronous alternator and synchronous motor are often
constructed with a polyphase nonsalient stator and a salient single-phase
d-c excited rotor. The polyphase stator is called the armature, and the
d-c rotor is known as the field. The a-b primitive machine formul;;'ted in
Chap. 10 fas saliency only on its stator structure; the rotor is smooth, or
nonsalient. Our model of the synchronous machine will therefore be
arranged with the field on the stator and the armature on the rotor.
Synchronous machines are rarely built in this inverted fashion because the
large power quantities associated with the armature circuits would have
to be transferred through slip rings. The field requires a small portion of
the rated power and is much more eaSily supplied through slip rings.
Synchronous machines are most often operated at a constant steady­
state speed and are connected to a fixed polyphase supply system. After
formulating a genera) primitive-machine model of the synchronous
machine, we shall study the steady-state operation in great detail.
Under normal circumstances a large majority of the synchronous alter­
nators and motors are operated in the steady state. As discussed in Secs.
10-7 and 10-8, an equivalent two-phase model is used to represent the
actual polyphase (usually three-phase) system. Per-phase results are
then obtained, and they can be converted to actual results simply by
multiplying by the actual number of phases. Certain transient operating
modes of practical interest are also included in this chapter.

11-1 AN ELEMENTARY VIEW OF THE SYNCHRONOUS MACHINE


Before proceeding with a formal analysis of the synchronous machine, let
us consider a rather elementary physical picture of the device. A
simplified diagram showing the basic'operation is given in Fig. 11-1-
The figure is drawn for the particular case of the machine being used as a
motor, but the general philosophy of the device can still be explained.
The stator consists of a number of windings or phases whose prime job is
to set up a rotating magnetic field in the air gap of the machine. The
precise manner in which this Interesting phenomenon takes place is dis­
cussed in the preceding chapter. Assume that the north and south poles
The 'Ynchronous mochine 469

FiR. 11-1 An elem('ntary


diagram of a three-phase two­
pole synchronous mar,hine.

of a sinusoidally distrihuted field are revolving in the air gap of the devi-ce.
The rotor of the machine consists of a winding to which a d-c source is
connected, thus establishing a constant field depicted in Fig. 11-1 by the
north (N) and south (8) on the rotor structure.
The rotor north pole is attracted to the stator south pole, and in a
similar fashion the I'Otor sout.h pole is aUraded t.o the stator north pole.
This magnetic bond between the stator and rotor fields causes the rotor
to revolve in synchronism with the stator field. The stator field is
revolving at a ("onstant speed and thus the rotor, being magnetically
locked to this field, revolves at the same constant speed. The position
of the rotor with respect to the stator field is given by the angle 0 in Fig.
11-1. If the machine is operating at a constant speed, the angle 0 remains
constant. If the mechanical load on the shaft of the machine is increased,
the rotor will tend to lag farther behind the stator field and thus increase
the angle o. A simple, but useful, model of the magnetic bond is obtained
by considering the bonding or locking force between the stator and rotor
fields to be provided by rubber bands between the two fields. Loading of
the rotor causes an increase in the angle 0 that stretches these rubber
bands, thus providing an increased torque output to drive the increased
load. Extreme loading of the rotor shaft will finally cause these rubber
bands to break and thus stall the machine. With this simple moGel, much
of the steady-state machine performance can be predicted.
If mechanical power is put into the mechanical port of the synchro­
nous machine in Fig. 11-1 and the direct current is maintained in the
rotor structure, an alternating voltage will be induced in each of the
470 Principles of electromechanical-energy conversion

st.ator coils. The relative phasing of these induced stator voltages will
be different in each coil owing to their physical displaeement on the stator
structure. With the stator windings mechanically separated hy 211"/3 rad, .
the respective induced stator voltages will have a phase relationship of
this same magnitude if the rotor has the single north and south poles
shown in Fig. 11 -1. If electrical loads are connected to the stator
windings, electric power may be extracted from the electrical ports of the
machine.

11-2 A PRIMITIVE-MACHINE MODEL


The synchronous machine shown in Fig. 11-1 can he represented sche­
maticall~ as shown in Fig. 11-2. The three-phase stator winding is
represented hy three identical windings on the a-{l--y stator axes, as shown
in Fig. 11-2. The d-c field winding is drawn as an a-axis rotor winding.
The a-b rotor axes must be used because the field winding is supplied
through slip rings and not hy means of a commutator-and-brush arrange­
ment. Section 10-7 shows us how to convert a three-phase winding into
an equivalent two-phase winding. Equation (10-114) specifies that the
distrihution factor of the two identical two-phase windings should be
equal to the distribution factor of the three identical three-phase windings.
Thus, the three a-{l--y stator windings shown in Fig. 11-2 can be replaced
by two identical windings placed on a set of a-b stator axes, as shown in
Fig. 11-3. Any voltage or current variables present. in the three-phase
windings can be converted to equivalent two-phase quantities hy using
the transformation matrix given by (10-119).

Fig. 11-2 Schematic representation


of a three-phase n-pole-pair syn­
chronous machine.
Th. synchronous momin. 471

Fig. 11-3 Equivalent two-phase n­


pole-pair synchronous machine with
two stator windings constructed
identically to the three-phase stator
windings.

The stator structure of the equivalent two-phase synchronous machine


shown in Fig. 11-3 is usually nonsalient, whereas the rotor structure can
often have salient poles. The a-b primitive-machine model, formulated
in Chap. 10, is assumed to have saliency only on the stator structure.
Our two-phase model shown in Fig. 11-3 can be made to fit the standard
a-b model machine by simply inverting the machine's construction. The
doc field can be placed on a salient stator structure, and the two-phase
armature can be constructed on a nonsalient rotor structure supplied
through slip rings. This inverted two-phase machine is shown in Fig. 11·4.
The synchronous machine is to be observed by effectively standing
on the rotor. In Fig. 11-3 the positive direction for rotor angular velocity
is chosen to be clockwise. However, if one were standing on the rotor,
the actual stator of the machine would appear to be rotating in the
counterclockwise ,direction with the same angular velocity",'. There­
fore, in Fig. 11-4 the two armature windings on the rotor are shown

Fig. 11-4 The inverted equivalent


two-phase n-pole-pair synchronous ma­
chine.
472 Principles of electromechanical-energy conversion

Fig. 11-5 A d-q primitive-machine


model of the inverted two-phase n-pole­
pair synchronous machine.

revolving eountewlockwise at. a speed w' with respect to the stator field
winding. A synchronous machine is not usually constructed as shown
in Fig. 11-4. However, t he machine could be built in this fashion with
the two a-b rotor windings supplied power through slip rings. The chief
drawback tot he inverted arrangement lies in the relative magnitudes of
the armature power and the d-c field winding power. The field winding
operates at a much lower power level and thus is more easily supplied
through the slip-ring mechanism.
The analysis of the a-b primitive-machine model shown in Fig. 11-4
can be readily formulated by first developing an equivalent d-q primitive
machine. Equations (10-70) to (10-74) summarize the relationships
oetween t he parameters of the two machines. The electrical-port vari­
ables are related by the d-q to a-b transformation given by Eq. (10-145).
,
The mechanical-port variables for the two primitive machines are
identicaL J.<'igure 11-5 shows the d-q model of the inverted two-phase
synchronous machine.
Let us briefly review how we have modeled the three-phase syn­
chronous lIlachine shown in Fig. 11-1 and schematically in Fig. 11-2.
The three-phase set of stator windings is first converted to an equivalent
set of two-phase stator windings. Because the field is often constructed
on a salient strudure, our view of the machine is inverted with the two­
phase armature placed on the rotor and the d-c field put on the stator.
We now have a ('oflventional a-b model which is transformed to an
equivalent d-q model machine. Equations for the d-q primitive-machine
model ('an be readily written and solutions can be obtained for whatever
constraints are placed upon the ports of the primitive machine. Con­
straints pla{'cd upon the ports of the actual synchronous machine must
first be transformed to correspond to equivalent constraints on the d-q
primitive machine. When the primitive-machine response has been
The synchronous machine 473

obtained, inverse transformat.ions call be used to examine the response


of the aetual synehronous ma(·him'.
From the primitive-machine model shown in Fig. 11-5 the following
electrical-port equations can be written in matrix form.

v~l
v'd
Hf~~p
R'd + iLdP (11-1)
v'Q -~G;dw'

The factor I is inserted before each inductance parameter involving the


currents i'd and i;, as discussed in Sec. 10-8. Both the direct- and quad­
rature-rotor windings have distribution factors K' identical with the origi­
nal three-phase windings. However, the factor of ! must be inserted
because of the form selected for the three- to two-phase transformation.
Thc mechanical-port equation ean also be written in the following
form:

T' = (.Jp + D)w' + T. (11-2)

The torque of electrical origin T. is obtained from the eleetrical-port


equations (11-1) and is given by

.-
T - - P.
w'
(11-3)

which, for convenience, is rearranged to be

T. = G;~idi; +. I(G;~ - Gd;)i;i:t (11-4)

Now, upon specification of the required number of additional operating


constraints, the four port equations (11-1) and (11-2) together with the
torque of cleetrieal origin, spedfied by (11-4), can he used to study the
operation of the synchronous machine.
The convenience resulting from the requirement that the power
be invariant under a transformation of ('oordinates can now be appre­
ciated. If this property were not valid, then the power quantities cal­
culated in the new coordinate system would have no relationship to
power quantities calculated in the old set of coordinates. The converted
power p. is identical for the d-q primitive machine and the equivalent
a-b primitive machine. Since the speeds w' are also the same, the torques
of electrical origin are identical. Similarly, the machine losses are also
identical. The power invariance of the d-q to a-b transformation allows
us to determine much operating information without the necessity of an
inverse transformation.
---- - _.._- __.._--_.,..,........_....

....

474 Principle, 01 electromechanical.energy conversion

11·3 CONSTRAINT EQUATIONS FOR STEADY-STATE OPERATION


ON A BALANCED INFINITE BUS
The synchronous machine is very commonly used as a steady-state
energy-conversion device. When it is used as a generator or alternator,
mechanical energy is supplied to the mechanical port and the machine,
revolving at a constant steady-state speed, pumps electric energy out
of its stator windings (Le., the original stator windings of Fig. 11-2).
Very often the individual alternator is one of a large number of such
generators supplying electric power to an extensive power system.
Under these circumstances the stator-port voltages are fixed by the
power I\Ystem into which the machine is operating. When the stator
termina\ voltages are specified in this fashion, we say the machine is
operating into an infinite bU8. The magnitude of the power each indi­
vidual machine supplies can be adjusted, but the voltage at which the
electric power is available is fixed as an operating constraint.
When the machine is used as a synchronous motor, the same stator
terminals are connected to an electric voltage source. In order to deter­
mine the motor's operation, the voltage source function must be given.
For operation as either a motor or generator the electrical-port voltages
are therefore to be specified as known constraints. The machine can be
considered to be operating from an infinite bus regardless of whether
energy is converted from mechanical to electrical as in a synchronous
generator or alternator or from electrical to mechanical as in a syn­
chronous motor.
Let us assume that the three-phase stator of the synchronous
machine pictured in Fig. 11-1 or 11-2 is connected to a balanced three­
phase infinite bus. Thus, the three stator voltages could be given by
Eqs. (10-126), which are repeated for convenience as

v~ = V2 V cos wt v6 = V2 V cos ("'t _2;)


v; = V2 V cos ("'t + 2;) (11-5)

where V is used as the rms value of the applied phase voltages. The
three-phase to equivalent two-phase transformation matrix is given by
(10-119) as being


(!ab.G<lh -
_ 2
1r
[
1
0
f]
By applying this transformation to the balanced three-phase bus voltages
(11-6)

given by Eqs. (11-5), we obtain a set of balanced two-phase voltages


.- » rn .. • .'S$O p 57
7

The .ynchronOil' machine 475

given by
v: = y'2 V cos wt v: = -y'2 V sin wt (11-7)
Since the stator and rotor of the machine are inverted, the two-phase
armature voltage constraints given by (11-7) are the rotor-port voltages
in Fig. 11-4. Therefore, the rotor-port variables are specified as
v: = y'2 VII cos wt vi; = -y'2 VI> sin wt (11-8)
where VII and Vb are the rms or effective values of the respective rotor
voltages. The corresponding constraints for the d-q primitive machine
can be found quite simply by using the linear transformation introduced
in Sec. 10-3. Thus, the equivalent d-q rotor voltages are given by the
matrix equation
(11-9)

where the transformation matrix is given by


a 41.ab
r _ [cos ncr
sin ncr
-sin ncr] (11-10)
- cos ncr
with cr representing the instantaneous angular displacement between
the a-b axis frame and the d-q axis frame as measured from the d axis to
the a a)j:is or, equivalently, from the q axis to the b axis, and n is the num­
ber of pole pairs established by the windings. Substituting Eqs. (11-8)
into Eq. (11-9) and expanding gives
Vd = y'2 (VII cos wt cos ncr + Vb sin wt sin ncr)
v~ -y'2 (- V.. cos wt sin ncr + Vo sin wt cos ncr)
(11-11)
By use of trigonometric identities for the product of the sine and cosine,
these equations can be rearranged to the following form:

v:; = y'2 [Vo -; Vb cos (wt _ ncr) + V.. ; Vo cos (wt + ncr)]
V~ .. [V.. +
= - v_'2 2 Vb sm
. (wt - ncr) - VII -2 Vb'
sm (wt + ncr)]
(11-12)
With balanced phase voltages we have
V.. = Vb = V (11-13)
and the d-q primitive-machine equivalent rotor-port voltages are given by
v~ = y'2 V cos (wt - ncr) v~ = -0 V sin (wt - ncr) (11-14)
The third constraint on the synchronous machine is the specification
of the field winding excitation. In the usual operation of this type of
J,76 Principles of electromechanical-energy conversion

machine, a constant or d-c voltage source is applied to the field winding.


For simplicity, let us assume a constant-current source rather than a
constant-voltage source. The field winding current is expressed by

(11-15)

If the field winding current remains constant, then whether a d-c current
or voltage source is chosen becomes equivalent. However, a time-depend­
ent voltage induced in the field circuit would give an additional field
current component if a voltage source were used on the field winding.
A fourth constraint is available by limiting our considerations to
the steady-state operation of the machine. Under steady-state condi­
tions, tre rotor angular velocity remains constant and is equal to the
angular velocity of the rotating magnetic field. Equations (11-7) define
the original stator voltages to be sinusoidal with an angular frequency
of w rad/sec. Therefore, in the sinusoidal steady state, the original
stator currents must also be of this same frequency. In Sec. 10-2 ba.l­
anced stationary currents of frequency ware shown to produce a magnetic
field rotating in space at an angular velocity equal to win, where n is
the number of pole pairs established by the winding. The angular speed
of the rotating magnetic field win is known as the synchronous speed.
Notice that the synchronous speed is dependent only on the excitation
frequency and the number of pole pairs established by the windings. In
the steady state the rotor must rotate in synchronism with this revolving
field; thus the rotor of the actual machine must rotate at an angular
velocity of win rad/sec. This constraint can be expressed by

w' = ~ (11-16)
n

where w is the constant electrical angular frequency of the port voltages.


The primitive-machine model for the synchronous machine consists
of three electrical ports plus one mechanical port. Therefore, a total
of eight equations are necessary to specify all eight port variables com­
pletely. Four equations are given by the port equilibrium equations
(11-1) and (11-2). Two constraints are imposed by specifying two of
the port voltages by Eqs. (11-14). The remaining two constraints are
obtained by specifying the field winding current and the rotor angular
velocity to be constants, as given by Eqs. (11-15) and (11-16), respectively.

Example J 1- J
A IO-hp 230-volt four-pole 6O-cps three-phase wye-connected synchronou8 machine has
the line-to-line voltages shown in Fig. 11-00 applied to the armature windings. In
The .ynchrol'lOll' machine 477

V'Z
. .
V23

V31
.
3

(al

V31

Fig. 11-6 (a) A three­


phase wye-connected syn­
chronous-machine stator.
(b) Line-to-line and line­
to-neutral voltage sinors. (bl

sinor form these voltages are given hy


v .. "" 2:W/O V. , ~ 230/2,../a
Find:
a. The machine synchronous speed
b. The equivalent set of phase voltages
c. An equivalent set of two-phase a-b voltages
d. An equivalent set of d-q voltages for the inverted model machine
Solution: a. The synchronous speed of the machine is given by Eq. (11-16) as
being
w

where w = 2,..1 is the excitation frequency and n = t is the number of pole pairs
established hy the armature windings. For this example we have
?,77
(jJ"=- 188.5 rad/sec = lROO rpm
2

b. The three-phase armature-port voltages corresponding to the given line-to­


line sinor voltages can he determined hy multiplying the magnitudes by 1/ V3
QS Principle. of electromechanical-energy converlion

and shifting the sinora by -'If/6 rad. Thus we have

V' - VI< ,. ---= 1 -'If/6) ... -230~


Vu (/ - ... 6
<II V3 - - y'3
V~ _ Vu " Vu (1/-... /6) = 230 / -5..-/6
~ -- y'3--

V'., - Va. ". -VII (1/ - .../ 6) ... -230 / ... / 2


y'3 -- y'3-

Figure 11-60 shows a sinor diagram of the three-phase line-to-line and line-to-neutral
voltages. The diagram can be checked by noting, for example, that

VI< - Vu - Vu + V., - Va
The corrrponding time forms of the three-phase stator voltages are

v~ ... v'2 X 133 sin ( 377t - ~)


V; .. v'2 X 133 sin ( 377t - ~)
vy - v'2 X 133 sin ( 3m + n
where 230/ y'3 ,. 133.
c. The equivalent set of two-phase a-b voltages can be obtained by using the
transformation matrix given by (11-6). Thus we have

The expansion of the matrix transformation can be performed most easily if all time
quantities are expressed in sinor form. Thus we have

V' ,.

i X 133 (II-----
- ~ -!I-----
6
- ~ - ~/~) ­
2 6 2 2
---
V'~T
- III X 133 ( v'3 I s... + -v'3 I"')
- -2- - -6 2 2- I'" = 133 ­3
--- ---
The corresponding time forms for the equivalent two-phase voltages are

V~ .. v'2 X 133 sin ( 3m - n


v: == v'2 X 133 sin ( 377t + i)
Notice that v~ and v: are .../2 rad out of phase and that If. is equal to v;,.
d. The inverted model machine places the two-phase armature windings on the
rotor and the field on the stator. Therefore, the voltages v~ and Vb now become v:
and v~, respectively. The equivalent d-q rotor voltages can be obtained by using the
The qnchronour machine 479

matrix (11-9). Expanding (11-9) leads to

Vd] [cos 2a -sin 2a] II~]


v; = sin 2a cos 2a II;
with n .. 2 as the number of pole pairs.
Substituting the expreesion for v~ and Vb, we have

Vd .. v'2 X 133 [sin (377t - n cos 2a - sin ( 377t + D


sin 2a ]

II; .. v'2 X 133 [sin ( 377t - ~) sin 2a + sin ( 377t + i) C08 2a ]


By recognizing that

sin ( 377t + D.. cos ( 377t - n


these expreesions can be reduced to become

Vd - v'2 X 133 sin (377t - ~ - 2a )

II; .. v'2 X 133 cos ( 377t - ; - 2a)

If the speed of the machine is equal to synchronous speed, then a - .lI.¥ and thus
a(t) .. .lp t + a(O)
The expressions for v:i and v; are then d-c quantities, since they are independent of
the time t.

11-4 SYNCHRONOUS TORQUE OF ELECTRICAL ORIGIN


Since the direct-stator winding or field current is taken as a known
quantity, the equilibrium equation for this port serves only to determine
the port voltage, or the voltage across the constant-current source. The
remaining two equilibrium equations of (11-1) are expanded and repeated,
for convenience, as follows:

v~ = (R~ L;P)i4+ iGd;w'l~ + (11-17)


v; = -G;dwridO - 1G;:iw'td (R~ + + iL~ph~
with the derivative of the constant field current equal to zero in the
direct-rotor equation. The direct- and quadrature-rotor voltages, for
balanced phase excitations, are given by Eqs. (11-14). The angle Q in
these equations defines the instantaneous angular position of the a-b
reference frame with respect to the fixed d-q reference frame. Referring
to Fig. 11-4, the a-b frame is seen to be rotating at a constant speed in
480 Principle, 0' e/edrome<:honico/-energy con"e"ion
the positive-angular direction. Therefore, a is increasing at w' radlsec
and ('an be given by

a Wt_...!.+~ (II-IS)
n 2n n

since the synchronous speed w' = win according to Eq. (11-16). The
constant angle (l/n)(o - 71'/2) defines the value of a at time t O.
The physieal significance of the phase angle a is considered in Sec. l1-S,
Substitution of Eq. (11-18) for a into Eqs, (11-14) leaves the direct
and quadrature-rotor port voltages as

I'~ 1-" sin a = II~o


= '\1'2 v;
= - 0 v cos a = v;o
(1l-19)
I
which are eonstants. If the left-hand sides of Eqs. (11-17) are constants,
then in the steady state the responding currents must also be constants.
This reduces Eqs. (11-17) to

(11-20)

where the ('onstant electrical angular frequency win is again substituted


for the rotor angular velocity w',
In order to examine the energy-conversion properties of the syn­
chronous machine operating under the imposed constraints, let us cal­
culate the torque of electrical origin. Before we proceed, a few con­
venient definitions will tend to simplify the notation. For this reason
the following three quantities are defined:

(11-21)

(11-22)

(11-23)

The quantities X 9 and X d , defined by Eqs. (11-21) and (11-22), are com­
monly given the names quadrature and direct-axis synchronous reactance,
respeetivcly. For identical rotor windings on the primitive-machine
model, Eqs. 00-.')1) and (l0-.''j2) give the rotational inductances involved
in the synchronous reactances as being

n(L' - L.) G";a = nL~ n(U + L.} (11-24)


Th. synchronous machine 481

where Lr and L. are defined by (10-53) and (10-54). When Eqs. (11-24)
are substituted into Eqs. (11-21) and (11-22), we have
Xd = IUl(l] + L.) (11~25)
For a nonsalient-pole machine the air gap between the stator and
rotor is uniform and L. = O. If this is the case, then the quadrature­
and direct-axis synchronous reactances are equal and
XII. = X. nonsalient machine (11-26)
and X. is simply called the synchronous reactance of the machine.
In terms of the three defined quantities given by Eqs. (11-21) to
(11-23), Eqs. (11-20) become
(11-27)
The voltage terms involving R~ and R; al'e usually much smaller than
the other terms in these equations. Thus, without introducing a sig­
nificant error in a practical calculation, these terms can be neglected.
Eliminating the ohmic terms, Eqs. (11-27) are now simply
(11-28)

upon subst.itution of expressions (11-19) for the left-hand sides. The


two rotor eurrents ('arl now easily be found, and they are given by

. = -------;X"""d---
~dO
V2 V cos 6 - E I (11-29)

The d-q roordinates ran he used in a per-phase calculation of the


torque of electrical origin for the actuu.l machine because the d-q to a-b
transformation is' power-invariant. This torque is computed from the
portion of the electric power entering the electrical ports and not going
into electrical loss or electric-energy storage. In the model of the
synchronous machine we are considering, all ohmic loss is being neg­
lected. Also, since the model is operating in the d-c steady state, no
mutual- or self-inductive power flow is present. Therefore, all of the
electrieal-port power goes into conversion and is included in the p. term
of Eq. (11-3). Owing to the power-invariant nature of the three-phase
to two-phase coordinate transforlllation, p. for the model must equal
p. for the actual machine on a per-phase basis.
By using the result given by Eq. (11-4) and the definitions of (11-21)
to (1l-2!l), the total torque of eieetriral origin under steady-state condi­
tions is found to be

(11-30)
482 Principle. of electromechonieol-energy conver.ion

Motor ---1
Generator
operation operation I
TotoI torque curve

nVEf .
V2wX Sl~ 8
d

Fig. 11·7 Synchronous torque of electrical origin per phase for a salient·pole
machine aa a function of the torque angle.

By substituting Eqs. (11-29) for the two currents ~~o and ~~o and dividing
by 2, this torque expression can be manipulated into the following form:

n [VEl'
T .. = w V2 Xd Sill "
V2
+ 2XdX (X
d -
X)'
q Sill
2]
"
(11 31)
­
q

The model machine used for this analysis has two phases, so dividing
the total torque, given by (11-30), by 2 makes Eq. (11-31) the synchro­
nous torque of electrical origin per phase. Equation (11-31) multiplied
by 3 would be the total generated torque for a three-phase machine.
Figure 11-7 shows a typical plot of the per-phase torque of electrical
origin T.. as a function of the angle". From Eqs. (11-25) the quantity
Xd - X q is equal to 3wL, and is positive. This assumption is made in
the plot of Fig. 11-7. For a nonsalient-pole machine, Eq. (11-31)
simplifies to

T.. = :;'VEI sin" (11-32)


2wX.
and the torque of electrical origin is a simple sine function of the angle fl.
The positive reference direction for T•• is taken to be clockwise.
This means that if T .. is a positive quantity, the torque of electrical
origin is in the clockwise direction. Equation (11-16) has set the machine
The ,ynchronou, machine 483

speed to be constant at win rad/sec in the positive, or counterclockwise,


direction. Thus, if T"" is clockwise and w' is counterclockwise, the
machine is acting as a generator converting mechanical power to electric
power. In Fig. 11-7 the region of posit.ive T.~ is labeled generator opera­
tion. Similarly, where T .. is negative, or in the same direction as the
motor speed, the machine is acting as a motor converting electric energy
to mechanical energy.
The angle 0 has been carefully introduced into the analysis by
Eq. (11-18) in such a manner that positive values of 0 correspond to
generator operation and negative values of 0 correspond to motor opera­
tion. In Sec. 11-8 t.he physical sil/:nificance of 0 is discussed. Expres­
sions (11-31) and (11-32) giving the torque of electrical origin per phase
for the nonsalient and salient machines, respectively, are functions of
the angle o. For this reason, the angle 0 is commonly given the name
torque angle.

Example J J·2
The 10-hp 230-volt four-pole 6O-cps three-phase wye-connected 8ynchronou8 motor
in Example 11-1 hll.8 its armature excited at rated voltage, 11.8 discuesed in
Example 11·1. The d-c field current is I!O adju8ted that E, .. 192 volts. The
windage coefficient at 8ynchronou8 speed i8 D .. 0.01 newton-m-sec/rad. Assume
the armature resistance to be negligible. The machine hll.8 II. nonsalient structure
with II. synchronous reactance X ... 1.5 ohms, and it is delivering rated power to a
constant load torque.
a. Determine the total power converted from electrieal to mechanical form.
All!O find the total electric power input.
b. Find the synchronous torque of electrical origin per phll.8e.
c. Determine the torque angle &.
d. Calculate the'magnitude of the line currents and the power factor of the
machine.
8olutmn: a. The total converted power is the output power plus the windage
losses; thus we have
p. == '(10 hp X 746) + D(",')'
p. ""-1460 + 0.01 X (.lP)t "" 1816 watts
Since there are no electrical losses, the total electric power input is also 7816 watts.
b. The total synchronous torque of electrical origin is the converted power
divided by the speed; therefore
P. 18\6
T • .. -;;. == - 377/2 == -41.5 newton-m

Since the machine hll.8 three phll.8C8, the synchronous torque per phll.8e is simply

T.. .. ~ .. -13.8 newton-m (per phase)

The torque of electrical origin can also be expreesed in units of synchrOfllJU4 watts.
These units 8imply use the corresponding power quantities 11.8 torque, without dividing
484 Principles of electromechanical-energy conversion

hy the synrhronous speed. Thull the total gen!'rat!'d torque would he 7H16 syn­
chronous watts, and the torque of I'le('tri('al origin per phase is 2605 syn('hronous watts.
c. The torque angle 0 for the nonsalil'nt machine can he det..rmin..rl from Eq.
(11-32). Hepeating (1l-:~2) for convenience, we hav..

nt'E, .
r.d> ----sm 0
V2wX.
as the g!'nerated torque per phase for a nonsalient machine. The voltage quantity V
is the rms phase voltage applied to the armature of the two-phas(, model; however,
the three-phase to two-phase transformation maintains the same phase voltage.
Thus V 2aO/ y:l and we have
X a77 X 1.5 X (­
2 X l:J:-l X 192
= -0.216
o -12.5 0
I
as the torque angle o. Since the maehine is operating as a motor, 0 should he a
negative angle according to our sign convention.
fl. The magnitude nf the line ('urrents can he determined hy first calculating the
transformed d..fJ rotor currents. From Eqs. (l1-2!1) we have
V sin 0
X.
for a nonsalient machine where X d X. = X.. Huhstituting numerical values
gives us
X l:.l:J cos (-12.5) 192
-5.:.13
1.5

= -27.2

From ScI'. to-a the d-q to a-b transformation for a four-pole mac,hine (i.e., two pole
pairs) gives UK for th(, a phase

Thus, the rms magnitude of i~ is simply

I' = V(i~/O)2 + (i;~ 19.6 amp


• V2
Sin('e the a phase of the three-phase armature has coordinates identical with the
a phase of the two-phase armature, I; is the magnitude of the actual line currents of
the thr..e-phaRe machine.
The power factor can he determined from the total electrkal input power, which
is equivalent to
PH = y3 VI.h cos e
where V L and II. are the respeetive thr('e-phaae lin .. quantities. Thull we have
7816
Power factor = cos e= = 1.0
X 2;~0 X 19.6
The marhine is operating at unity power factor, whi('h means that the line currentll
are in phase with the line-to-neutral or ph~e ooUages or the wye-eonneetl'd armature.
The qnchroneus machine .485

11·5 SINOR DIAGRAM OF THE SYNCHRONOUS MACHINE


Equations (11-8) impose constraints on t.he stator-port, voltages of the
actual synchronous machine. As we have seen, these constraints are
equivalent, under balanced excitation and constant speed, to certain
d-c voltages applied to appropriate windings of our d-q primitive machine.
The resulting primitive-machine port currents can be found, and by an
inverse transformation the stator-port currents of the actual synchronous
maehine are available. This process can be made more systematic by
developing a c1assieal sinor diagram which accomplishes the inverse
transformation. The sinor diagram will show how the primitive-machine
variables are combined in forming the actual-machine port variables.
The inverse transformation from rotor d-q voltage variables to rotor
a-b voltage variables is given by the matrix equation
(11-33)
which can be expanded as

1/:]
"r [
('o~ na
- Bill na
sin na] ll~]
cos na I';
(11-34)
b

as discussed in Sec. 10-3. Therefore, the a-b rotor voltages in terms of


the d-q prilllitive-ma('hine rotor voltages are given by
v: "~ cos na + II; sin ncr (11-35)
I); = -II~ sin na + I); cOS ncr
The steady-slale equilihrium equal ions for the primitive-machine model,
given hy (11-27), are repeated for convcnienC'c as
"d = I)~O R~t1o + X qt~o

v; = ";0 = - E, - Xdt~o + R;i;o (11-36)

The subscript zero indicates constant or d-c values which result from
the spccial constraints being C'onsidered.
By substitution of the constant values for the direct- and quadrature­
rotor voltages, t.he a-b rotor voltages of Eqs. (11-3;») can be expressed as

/!~ = lidO sin (ncr + ;) + v;o sin ncr


(11-37)
,.~ = -VAo sin na + ";0 sin (ncr + ;)
In order to simplify these cquations, sinor notation is to be introduced
in the following form. Let x(a) be a port variable defined by

x(a) V2 X sin (ncr + '1') (11-38)


486 Principle, of electromechanical-energy conversion

The sinor quantity representing x(a) is therefore given by


X Xd"' = XI", = [(X cos "') + j(X sin "')] (11-39)

in the three conventional forms. Notice that X is the rms magnitude


of the sinusoidal quantity X and the factor y'2 must be inserted In
changing the sinor quantity back into its sinusoidal functional form.
By using this sinor notation, Eqs. (11-37) can be written as

V'
"
(11-40) [
Multiplying the first of Eqs. (11-40) by the operator j gives us
·V) _ VdO + . v;o (11-41)
J"--y'2 Jy'2

which we see to be equal to V;;, and thus


(11-42)
This means that setting up a sinor diagram for the a-phase quantities
is sufficient to define the b-phase quantities completely. A simple
angular shift of the a-phase diagram by 11"12 rad yields the b-phase dia­
gram, according to the result of Eq. (11-42). By comparing Eqs. (11-5)
and (11-7) we can also sec that the a-phase variables are identical with
the a-phase quantities of the original three-phase excitation. Thus
when the a-phase variables have been found in terms of the d-q variables,
the b-phase variables are obtained by a 11"12 rad shift of all sinors. The
a-phase variables are identical with the a-phase variables, and the fj­
and 'Y-phase variables are obtained by appropriate 211'13 rad shifts on
the a-phase sinors.
Substituting the steady-state equations (11-36) for the synchronous­
machine d-q model into the first of Eqs. (11-40) gives us

(11-43)

as a complex equation for the sinor value of the a-phase rotor-port volt­
age. The original constraint on v:
is given by the first of Eqs. (11-8) "
and for convenience is repeated as
v: = y'2 V cos wt (11-44)

upon substituting the simplifying assumption of balanced excita.tion


given by Eq. (I 1-13). The angle a between the two coordinate frames

,
being considered is defined by Eq. (11-18), and upon slight rearrangement
The synchronous machine 481

Fig. 11-8 Complete sinor diagram for the a phase of the two-phase salient-pole
eynchronoUil machine.

it can be written as

wi = no: +!2 - 0 (11-45)

The a-phase voltage constraint, upon substitution of Eq. (11-45), becomes

v~ = v'2 V cos (no: + ; - 0)


or
v~ = v'2 V sin (no: + r - 0) (11-46)
In the sinor notation, defined by Eqs. (11-38) and (11-39), this port
constraint can be written as
v: = VIr - 0 (11-47)
Equations (11-33) and (11-47) can be represented very conveniently by
the sinor diagram of Fig. 11-8. The angle between the sinor E/lv'2
and sinor V is the torque angle 0 in accordance with Eq. (11-47). We
have adopted the convention of defining 0 to be positive if EJlv'2
leads V.
The a-b to d-q coordinate transCormation Cor voltage quantities
applies equally well t.o the corresponding port currents. Following
identical arguments for the port. eurrents, the following sinor equations
Cor the a-b rotor phase currents in terms of the d-q primitive-machine
rotor currents can be written:

(11-48)

As with the voltage quant.ities, t.he b-phase port current is ident.ical with
the a-phase current upon rotating :t by r/2 rad. The a-phase current
is also plotted in Fig. 11-8 according to the first of Eqs. (11-48). With
488 Principles of electromechanical-energy conversion

E, ~ 8
V2 Projection of ~
I onto V
Fig. 11-9 Sinor diagram for the IOB8less salient-pole synchronous machine.

both phase currents equal in magnitude the symbol I is used to represent


the rmll magnitude of I: in Fig. 11-8. The angle between V and I,
indicated by the symbol e, is the power-factor angle of either stator phase.
Notice in Fig. 11-8 that e appears to be a second-quadrant angle, greater
than 1r/2 rad. Thus cos e is a negative quantity and the steady-state
electrical-port power V I cos e is negative, indicating energy flow out
of the electrical ports and hence generator action. Thus, for a positive
torque angle ~ (i.e., sinor E,/y2 leading V) the machine operates as a
generator. This is consistent with the discussion at the close of Sec. 11-4.
By using the diagram of Fig. 11-8, the complete operation of the
synchronous machine under steady-state conditions can be described.
Before proceeding with this description, let us again make the simplifying
assumption that the ohmic voltages are negligible. Figure 11-9 is an
approximate sinor diagram, neglecting the ohmic terms. Directly from
Fig. 11-9, the folJowing expressions can be written;

V cos • _
(J -
E,
y2
+ XdidO
y2 (11-49)

which are exactly equivalent to Eqs. (11-28) previously derived. Solu­


tions for t~o and t~o are easily obtained as

. y2 V cos ~ - E, . y2 V sin ~
(11-50)
tdo =
Xd t;o = Xq

The per-phase electrical power into each of the two stator phases
is given directly by

PH_ = VI cos e (11-51)

where PH_ indicates electrical power input per phase. Now the term
I cos e represents the projection of I onto V taken positive if this projec­
tion is in the same direction as V, meaning e is an angle in the first or
fourth quadrants. In terms of its two components the projection of I
The rynchronOUfi machine 489

upon V is given by

I cos 8 = :2 sin 0 - .5~ cos 0 (11-52)

as is seen by examining Fig. 11-9. On substitution of Eqs. (11-50) into


(11-52) this projection becomes

I cos 8 = .!::2 (~
Xd
- ~)
X"
sin 28 - ---!L-
y2X
sin 0 (11-53)
d
after the trigonometric identit.y for the sin 20 is used. On substitution
of Eq. (11-53) into (ll-fil), the electric power intput per phase becomes

P H. = - V2(I - - - 1 )sin
. 2'
(I - ----
VE,.
sm 0 (11-54)
2 Xd X" y2Xd
With two phases on the actual machine the total electric power input
would be simply t.wice this amount, or

PH = - [Y~~E' sin 8 + X~~" (X d - Xq) sin 28] (11-55)

where PH is the total electric power input.


With three phases on the actual synchronous machine, the power
input pcr phase given by (11-M) must be mUltiplied by 3 to obtain the
tot.al electric power input. The total torque of electrical origin would
also be 3 t.imes thc per-phasc elcctrical torque.
Since the ohmic losses have been neglected, all of the electric input
power must go into the conversion process and

P E• = p •• (11-56)

where P•• is the portion of the electric power converted to mechanical


power per phase. The torque of electrical origin per phase is

T.. = - p ..
r
(11-57)
iN

and thus

T •• = ~ [ V~, sin 0 + . .~q (Xd - Xq) sin 28] (11-58)


w y2Xd 2Xd X

with the machine speed w' held fixed at win, the excitation frequency
divided by the number of pole pairs. Equation (11-.')8) is exactly
equivalent to the previous expression (11-31) developed with the d-q
primitive machine.
490 Principles of e/ecfromechanical-energy conversion

Example 11-3
A l3,200-volt 3000-kva eight-pole 6O-<:ps three-phase wye-<:onnected salient-pole alter­
nator is delivering rated kilovolt-amperes at 0.8 power factor leading and at rated voltage
to a balanced power system. The direct-axis synchronous reactance is X d - 66 ohms,
and the quadrature-axis synchronous reactance is X. - 50 ohms. Windage losses are
90 kw at synchronous speed. Armature resistance can be neglected.
o. Determine the magnitude of the three-phase line currents.
b. Find the total torque required to drive the alternator.
c. Find the torque angle 0 and the voltage E I induced by the field in the armature.
d. Draw a sinor diagram for these operating conditions.
SoluliDn: a. With the machine delivering rated kilovolt-amperes to a balanced
three-phase bus at rated voltage, we can write

V3 f L1 L "" rated kva


Therefore
I _ 3 X 10'
131 amp
L - V3 X 13,200
b. The total mechanical power supplied to the rotor shaft is equal to the con­
verted power plus the mechanical losses. Since the armature resistance is being
neglected, the converted power is equal to the electric output power. Therefore the
total mechanical input power is
PM = V3 VL1 L cos e + windage losses

= (3000 X 0.8 + 90) X 10'

= 2490 kw

The total torque required to drive the alternator is equal to the total mechanical
power divided by the synchronous speed of the machine. Thus we have
'2.49 X 10'
T' - 377/4 - 26,400 newton-m

The machine has eight poles or four pole pairs; therefore, synchronous speed is ",/4.
c. By using Fig. 11-9 as a guide, the following two current equations can be
written:
"r
.!!!. = I cos (180 - e- &) -'40 = I .
SIn (180 - IJ - 6)
viz viz
The current I is the magnitude of the two-phase armature currents, which are equal
to the three-phase line currents. The power-factor angle /} is equal to 1800 - 37",
since cos e must be 0.8. The angle e is less than 1800 because the power factor is
leading, which means that -1 must lead Vor, equivalently, -V must lag 1. t Substi­
tuting IJ = 143" into the two current equations leads to

.:/2 - 104.8 cos 0 + 78.6 sin 0


if'

.:;; = 78.6 cos 0 - 104.8 sin 6

using I - 131 amp.


t rn Sec. 11-6 the conventions for leading and lagging power factors for generator
and motor operation are discussed in greater detail.
The .ynchronou. machine 491

Also from Fig. 11-9, with R; = Rd = 0, we can write the following voltage
expressions:

X.l~.
- -= V'
sma
V2
corresponding to Eqll. (11-49). Substituting the exprell8ion for i;Ol V2 in the first
voltage equations gives us

50(104.8 cos /; + 78.6 sm. 8)


13,200 .
= v'3 sm /;

with V = 13,2001 v'3 as the phase voltage. By solving for the tan /; co (sin 8)1
(cos 8), we find
a .. 54.7·
By 8ub8tituting /; = 54.7° into the current equation for idol v"2 and using the
second voltage equation, we can obtain
E, . 13,200
V2 + 66(78.6 cos 54.7 - 104.8 sm 54.7) = V2 cos 54.7

which gives us

E,
V2 = 7050 volta E, = 9970 volta

d. The sinor diagram for the alternator is shown in Fig. 11-10. The current
idol v"2 is equal to -40 amp and therefore must be drawn downward in Fig. 11-10
opposite to the &ll8umed positive direction shown in Fig. 11-9. The sinor labeled
X .ti~oI V2 is also reversed in Fig. 11-10. Remember that the sinor diagram is valid
for the phases or both the two-phase machine and the actual three-phase machine.

jXq 6220

~ =7050

jXI/~=2640 idO =40


..n.
Fig. 11-10 Per-phase sinor diagram for a 3000-kva alternator delivering rated
kilovolt-amperes at 0.8 power factor leading.
492 Principle, 0' electromechanical-energy conver,ion
The time form for each of the sinor quantities must, of course, have an appropriate
phase angle depending on the phase being considered.
As an additional exercise, use Eq. (11-54·) and the values of E, and a found in
part c to obtain the converted power per phase. The total converted power per
phase is 2400/3 kw, and this additional calculation serves simply as a check.

11·6 CONSTANT-POWER OPERATION

Quite frequently, the synchronous machine is called upon to handle a


certain fixed amount of power in either driving a constant load torque
at synchronous speed or delivering a required quantity of electric power
to an electrical supply system. The sinor diagram of the preceding
sectioll p~ovides a very convenient tool for investigating how the machine
operates under the constant-power constraint. In order to simplify
the discussioll, let us assume that we have a nonsalient-pole lossless
machine. With a uniform air gap, the quadrature- and direct-axis syn­
ehrollous reaetam:es are equal; thus X q = Xd = X. and X. is termed
simply the synchronous reactance of the machine.
Inserting this simplification into the sinor diagram of Fig. 11-9 gives
us Fig. 11-11 for the nonsalient-pole machine. The sinor diagram shown
in Fig. 11-11 has been rotated 6 radians from that of Fig. 11-9 strictly
for ('onvenience in the following analysis. Notice that the voltage tri­
angle abc is formed by multiplying the current triangle a'b'c' by the
synchronous impedance jX.. Observe also that the current sinor I lags
the voltage sinor jX.I by 1r/2 rad. As before, the angle 8 by which the
phase voltage V leads the phase current I is the power-factor angle of
the corresponding phase.

Fig. 11-11 Sinor diagram for the 108111ess nonsalient-pole synchronoUll machine
rotated through an angle 8.
The synchronou. machimt 493

We now wish to examine the steady-state operation of the synchro­


nous machine on a per-phase basis under the constant-power constraint.
Substituting the nonsalient-pole condition (that is, Xq = Xd = X.) into
the per-phase electric-power equation (11-54) gives the total electric
power per phase as
VE,.
P E,. =- V2 X. sm 0 (11-59)

If the machine has two stator windings, or two phases, the total electric
power input is simply

Pr. = 21
)K,. V2 VE, Sill
= - -~X---:--
.
0 (11-60)

Since ohmic losses are being neglected, the expression for the total electric
input power per phase, given by (11-59), is equal to the total converted
power per phase P... Thus, the torque of electrical origin per phase
is readily derived as being
p.. nVE,.
T .. = - - = - - - sm 0 (11-61)
w' V2 wX.
With the machine revolving at synchronous speed, making w' = win,
the torque of electrical origin is related to the electric power input by
the factor -n/w. In many instances, the electrical t,orque is given in
units of synchronous watts, where the quantity p. is used as the torque,
with the understanding that a division by the synchronous speed yields
the magnitude of the actual electrical torque.
The ,constant-power constraint can readily be imposed by examining
the right-hand sitle of Eq. (11-.1)9). Since the machine is operating at a
constant terminal voltage, V is a fixed quantity. Similarly, X., the
synchronous reactance, is a parameter of the particular machine being
used, and it also is a fixed factor. Therefore, for constant power we
can write
E, .
V2 sm 0 = const (11-62)

where E" defined by Eq. (11-23), is directly proportional t.o the direct
field current and 05 is the angle by which the sinor EdV2leads the sinor V
in Fig. 11-11. The constant-power constraint, given by expression
(11-62), can be illustrated geometrically, as shown in Fig. 11-12. The
voltage triangle abo indicates a typical operating condition with the
machine handling rated power. The quantity (Edv2) sin 0 is equal to
the dista.nce be, and therefore, according to Eq. (11-62), it must be held
"9" Principler 0' e/ec:tromechaniea/-enerr1Y eOlfYerrion

constant if the machine continues to handle full rated power_ If the direct
field current excitation is changed, thus changing the magnitude of E" the
point b must move along the horizontal rated power line. Reducing
E, from its value in the abo triangle moves point b to position d and the
machine operation is given by the ado triangle. The angle 8, now given
by the angle Md, has increased in value and is maintaining the relation­
ship of expression (11-62).
If E, is further reduced to the point where the machine operation
is described by the triangle !leO, then the angle ~ is equal to 1('/2 rad.
This is the minimum value that E, can have with the machine still
delivering rated power. Increasing E, to a value such that EJ!-v'2 is
greater than V causes the operating condition to shift to a point such as f.
For leach of these operating conditions the sinor phase current I
must lag the voltage sinor jX.I by 1('/2 rad. Thus the current sinon!
corresponding to the operating points b to g in Fig. 11-12 are labeled
with primes. Since the machine is delivering constant power, the
projection of the sinor I on the sinor V must be held constant. A line
drawn perpendicular to sinor V, therefore, traverses the tips of each

e'

ower line

1.0 Rated power line


I
I. 5 Rated power line

Fig. 11-12 Geometrical interpretation of the eonltant-power constraint for a non­


Il8.lient-pole IOSllless a.lternator.
The qnchronou. madline 495

Fig. 11-13 V curves for the nonsalient-pole lossless alternator.

of these currents. The lagging power factor of the phase is defined as


the cosine of the angle by which sinor I lags sinor V. However, since
both I and V were defined positive if power flow is into the phase, the
definition of power factor lagging must be modified to be the cosine of
the angle by which either minus I lags plus V or plus I lags minus V.
For example, in Fig. 11-12, operating point b gives rise to the current
sinor drawn from 0 to b'. Taking the negative of this sinor indicates a
leading power-factor angle, since minus I leads V. Likewise, plus I
leads minus V by the same angle. For values of E,/y''2 smaller than
the value at point g the machine is operating at a leading power factor.
For values of E,/y'2 greater than the value at point g the power factor
is lagging. Unity power factor occurs at point g.
The various operating conditions along the 1.0 rated power line can
be summarized as shown in Fig. 11-13. The magnitude of the phase
current I is plotted as a function of the quantity E,/y'2, which is pro­
portional to the direct field current. The 1.0 rated power line in Fig.
11-13 has points b, I, and g from }<'ig. 11-12, as indicated. Also, three
additional curves for 0, 0.5, and 1.5 rated power are shown in Fig. 11-13.
496 Prfncip/es of e/edromechon;co/.energy conversion

Points of constant power factor on each of these four constant-power


curves are connected for the particular values of 0.8 leading and lagging
and unity power factor. Also, the points of minimum EtlV2 for each
value of power are joined to form the 8tability line in Fig. 11-13. If
EtlV2 is reduced below this line and the indicated constant power is
maintained into the mechanical port of the alternator, the machine
cannot continue to run at the indicated synchronous speed win. An
increased speed and loss of synchronism will result and produce an
unstable condition with large pulsations in the stator phase currents and
an entirely unsatisfactory mode of operation. Because of their general
shape, the curves in Fig. 11-13 are commonly called V curve8.
From Eq. (11-59) the power supplied to the electrical ports of the
armatur, is negative if the torque angle 6 lies between zero and +1'/2 rad.
Since the port variables are so defined that positive electric power cor­
responds to energy flow into the port, the negative power is interpreted
as energy flow out of the port. Thus, the operation indicated by Figs.
11-8 to 11-12, where sinor EtlV2leads sinor V and 8 has a positive value,
corresponds to generator or alternator operation. Let us now consider
the resulting sinor diagram if sinor EtlV2 lags sinor V, thus making 8
negative and yielding motor operation. Figure 11-14 shows such a con­
dition. Notice that the voltage triangle abc is similar to the current
triangle a'b'c' with each current sinor multiplied by jX., thus rotating
the sinors '1'/2 rad. The power-factor angle 8 is now less than '1'/2 rad
and thus VI cos 8 is positive, confirming the supposition that energy
flow-is into the electrical ports.
As in the case for the alternator, a constant-power line can be drawn
parallel to sinor V, as shown in Fig. 11-15. If the field current is reduced
such that Eli V2 is decreased from its original length ob to a new value
00, then the operating point on the constant-power line moves from b to d.
The torque angle 6 increases in the negative sense from angle boa to angle
doa. For clarity in drawing Fig. 11-15, the origin for all current sinora
has been shifted from point 0 to point 0'. For each operating position
on the constant-power line the corresponding current sinor is given by

Fig. 11-14 Sinor diagram for the lossless nonsalient-pole synchronous motor.
The qnchl'onoull machi,.. 497

(I'

f Constant power line

Fig. 11-15 Geometrical interpretation of the constant-power constraint for a non­


sa.lient-pole lossless motor.

the same lowercase letter with a prime notation. Thus current sinor
o'b' eorresponds to operation at point band (J'd' eorresponds to operation
at point d. Notice that reducing the field excitation causes an increase
in the magnitude of the phase eUrrC'nt, in addition to increasing the power
factor in the lagging direction.
Point g ill lng. 11-15 corresponds to unity-power-factor operation
and the minimulll magnitude for t.he sinor phase current. Increasing
Edv2 beyond point, g to point. f gives rise to the current sinor o'f' at a
leading power-factor angle. As in the ease of the alternator, the syn­
chronous motor ('an supply a ('onstant torque at cOllstant speed to a
load while extra('ting power from an infinite bus at an adjustable power
factor. Adjustment of the dire(~t field current controls EI/V2 and
thus the power factor. If EdV2 is reduced below point e, correspond­
ing to ~ = -1f/2, then the machine ean no longer supply the required
load power and will lose synchronism. This will cause large phase­
current pulsations, and the machine will eventually stalL
V curves similar to those shown in Fig. 11-13 can be drawn for the
motor operation. A succession of constant-power lines at varying
distances from the sinor V could be constructed on Fig. 11-15. The
phase-current magnitude I could then be plotted as a function of the
field excitation Edv2 for each of these constant-power lines. Curves
498 Principles of e/ectromechanical·energy conversion

identical with t hose for the alternator would result wit h the simple
interchange of the words leading and lagging ill Fig. 11-13. For values
of Etlv'2 greater than the unity-power-fador eonditioll the alternator
delivers power at a lagging power-factor angle, whereas the motor would
receive power at a leading power-factor angle. The remainder of Fig.
11-13 is the same for alternator or motor operation.

Example 11.4 1
A 5O-hp 230-volt d-e shunt motor is coupled mechanically to a 5O-hp 230-volt six­
pole 60-<,ps three-phase wye-connected nonsalient synchronous machine, as shown in
Fig. 11-16. The field of the do(' machinfl is connected in series with a rheostat R
acroSll a 230-volt doC infinite bus. The a-(' RynchronoUH machine i" connected to a
230-volt tbree-phll.lle a-c infinite bus, and it is running at llyn{·hronou8 speed. The
IIwitch K, in scries with the d-c machinfl armature, is open. The machine parameters
are 8.8 follows:
For the doC maehine:
Field resistance R, = 20 ohms
Rotational inductanee
Armature resistance fl.
G.,
= 0.22 h
0.1 ohm
For the a-e machine:
:;;ynehronous reactance X. = 0.8 ohm
Armature resistance is negligible
The windage loSlles of each machine are 3 kw when the machine is running at syn­
chronous speed.
a. With switeh K open the field of the synchronous machine is so adjusted that
the a-e line currents are a minimum. Determine the magnitude of the line currents
I, the torque an!!:Je 0, and the induced voltage Etlv'2.
b. While the a-e machine field current is held constant, the d-e field rheostat
Il is so adiusted that, when switch K ill {'Iosed, the direct armature current remains
at zero. Calculate the required value of Il.
c. While the a-e machine field current is kept at the value set in part a, the d-c
field rheostat is reduced 2.0 ohms from the value set in part b. Calculate the new
I This example makes a very interesting laboratory experiment.

230 1----<> 230 volts


volts 3<1>
d-c

Fig. 11-16 A synchronous machine mechanically coupled to a d-e shunt machine.


The synchronous machine 499

steady-state direct armature current J., torque angle, alternating line current, and
power factor. Which way is energy flowing?
d. With R held at the value set in part cdetermine E,/V2, 0, and J for conditions
of unity power factor and a power factor of lUi leading.
Solution: a. The synchronous machine is acting as a motor supplying the required
windage losses of its own rotor and the armature rotor of the d-c machine. Since
the armature electrical losses of the a-c machine are zero, the total converted power
is equal to the power drawn from the three-phase bus. Therefore, we have
v'3 V Ll cos 6 '= 3000 + 3000
where the phase current I equals the line current, since the machine is wye-eonnected.
For minimum line currents the a-c machine must operate at unity power factor; thus
6000

I - v'3 X 230 - 15.1 amp

The per-phnse sinor diagram corresponding to unity-power-factor operation for a


motor is shown in Fig. 11-17a. Notice that I is in phase with V, indicating energy
flow into the electrical port. The voltage sinor jX.I leads sinor I by ."./2 rad and
therefore is perpendicular to V. Using V as reference, we have from Fig. 11-17a

~ - V -jX.I
V2

E,
fi
iXI/~
Vz230/"tj /=15.1

tal !h)

r
112volts

LiX,1
~V~z-1-33----JL---------~

(e)

Fig. 11-17 Sinor diagrams for Example 11-4.


500 Principles of electromechanical-energy conversion

which reduces to
Ej 2ao . I
~2 v'3 - }O.S X 15.1 = la3..=§.~

Therefore, the torque angle Ii _6.2° and the induced voltage E,I..;'i "" 133 volts.
Since Ii is less than 5.7°, the sin Ii = tan Ii and E,I..;'i "" V = 13:J volts.
b. In order for the d-e armature current I G to remain zero when the switch K
is closed, the back emf of the d-e machine must equal the d-c line voltage. I Thus
we must have
Go,..,' T, 230 volts
On solving for the field current, we have
230 .
I, 0.22 X (3n/a) = 8..32 amp
The total\ field l'ireuit resistance consists of the rheostnt R in series with the field
resistance R, = 20 ohms. The required value II is ohtained from
2ao
]{~+ /{,
By solving for R, we have
2ao
]{
X.:12 - 20 7.fi ohmtl

c. When R is reduced hy 2.0 ohms, the direet field current rises to


2ao
" = 27.6 -2.-0 = 9.0 amp

The doc machine armature circuit equation is given by


G.,w'I, + /l.I. 2aO
Solving for T. leads us to

- - - --;;-'c;----
9.0 = -
2 00 amp

Since T. is negative, the doc machine must he acting as a generator. Power is bcing
drawn from the a-e line!!, !'onverted to mechanical energy hy the synchronous machine,
transmitted through the shaft to the doc machin/!, rec,onverted to electrical form, and
supplied to the doc bUB.
The total power converted hy the a-c machine is equal to the windage of both
machines pillS the power converted hy the doc machine. Thus we have for the a-c
converted pow('r per phase
P,q, = i(total windage power + G.,..,'l,T.)
p,. i(30()() + 3000 + 0.22 X 126 X 9.0 X 2(0)
P •.,. 1101.7 kw (per phase)
Sim'!' the alternating field current is held ('qual to the current set in part a, the
valu(' of Et! v'2 remains at 133 volts. Thus the per-phase sinor diagram has the
inter{'sting form shown in Fig. 11-17b with E,l V2 and V forming two sides of an
, Capital letters are heing used for the constant d-c variables ra.ther than the more
('umbersome subscript zero used previously for constant values.
Th. Iync:hronoul moehirHt 501

isoscelm. triangle with jX.I as the base. Since the sinor current I is nonnal to jX.I,
we have sinor I bisecting the torque angle a, and thus the power-factor angle (J = '/2.
By using Eq. (11-59), we have

sin 0 X.P••
VE,/v'2
and so
, = -57.7°
The minussign indicates that Ed V2 should lag V as shown in Fig. 11-17& for motoring
operation. The power-factor angle is simply

(J ". ~ .. 28.9° lagging

Hinre the a-c machine has no elm:trical 10l1l'i1',8, the power input from the a-c Jines
equals the power convcrtNI. Thus on a per-phase basis, we have
Vlcoll(J=P..

Solving for the line or phase current 1 leads us to


18.7 X 10'
1 133 X <'.os 28.\1 .. 161 amp

Energy How is from the a-e bus to the d-c bUll.


d. If R ill held at the same value II.S in part e, the d-c machine wiII continue to
supply the same power to the doC bUB. The a-c machine must therefore continue
converting 18.7 kw per phase in order to supply the total windage losses plus the
d-c converted power. From Eq. (11-59) we have

.!lL sin II .. _ X.p.. = 0.8 X IH.7 X 10'


---13-3-- ­
-112 volts
v'2 V
The minus sign means that a is a negative angle, which indicates that Ell V2 lags V.
Figure H-l7e illustrates the unity-powcr-factor condition, and so we have
X.I ." 112 or
112
1 .. 0.8 = 141 amp

AI80 from Fig. 11-17e we have

a .. tan-' ~I .. tan-' tH .. 40.3°


In addition, from Fig. 11-17e we can write

.. 175 volts

The 0.8 leading-power-factor case is illustrated in Fig. 11-17d. Since the sinor
jX.I is perpendicular to the sinor circuit I, the angle between jX.I and a vertical
line is the power-factor angle (J = 37°. Thus we have X.l cos (J .. 112, and 80
112 112

1 .. X.OO8 , = 0.8 cos 37 = 176 amp

502 Principles of efecfromechanic;ol-energy conversion

The torque angle is given by

-I 112 t -I 112 2730


• = tan V + X.I sin 8 = an 133 + 0.8 X 176 sin 37 - .

Also, from Fig. 11-17d we have

E, _ ~12 = .JJ:2_ = 245 volts

V2 Sin' Sin 27.3

Notice how the adjustment of the field current of the a-c machine alters the power
factor of the synchronous motor from lagging in Fig. 11-17b to unity in Fig. 11-17c
and on to leading in Fig. 11-17d. The converted power in all three cases has not
changed.

11-7 PER-PHASE EQUIVALENT CIRCUIT FOR THE


NONSALIENT-POLE MACHINE

The complete sinor diagram for the salient-pole synchronous machine


is shown in Fig. 11-8. For a nonsalient machine including armature
resistance in each phase, the sinor diagram reduces to Fig. 11-18. Here
we have set

(11-63)

where R' is the resistance of each armature phase and X. is the synchro­
nous reactance. The diagram shown in Fig. 11-18 has been rotated
counterclockwise by an angle 6 from that of Fig. 11-8 strictly for conven­

/
I
I
I
I
I
/
/
, .,fl I
, rr I
"/2~/
, I
, I
'v
Fig. 11-18 Complete sinor diagram for one phase of a non salient-pole synchro-
DOUB machine.
The synchronous machine 503

• Energy flow for motor

Energy flow for generotor

R' jX.

Fig. 11-19 Equivalent circuit Cor the ~+{~}+ 0 I 0 V


nonsalient-pole synchronous machine.

ience. The four voltage sinors drawn in heavy lines represent the equation

v= E, + (R' + jX.)I (11-64)


v'2
The sinor labeled jX.I leads the phase current I by ..,/2 rad, as indicated
by thej operator in (11-64). For the nonsalient machine the synchronous
reactance voltage leads the phase current by ..,/2 rad and the sinor labeled
R"I is in phase with the current I.
Equation (11-64) can be represented by the simple equivalent circuit
shown in Fig. 11-19. The sinor V is the phase'voltage, and sinor I is the
phase current. The angle 8 between V and I is the power-factor angle for
each phase of the machine. If the quantity V I cos 8 is positive, then
energy flow is from right to left in Fig. 11-19 and the machine is drawing
power from the electrical port and running as a motor. If VI cos 8 is
negative, then energy flow is from left to right. The machine is acting as
a generator delivering power to the electrical terminals. The ohmic losses
per phase in the armature are simply PR'. The power absorbed by the
left-hand port of the circuit shOWJl in Fig. 11-19 is the converted power
per phase. Thus we can write

p .. = '::21 cos (8 + 15) (11-65)

From Fig. 11-18 notice that 8 + 15 is the angle between Edv'2 and I.
For a cylindrio&l-air-gap, or nonsalient, machine the equivalent cir­
cuit shown in Fig. 11-19 is very convenient. For a salient-pole machine
this equivalent circuit gives a reasonable approximation to actual results,
particularly when the machine is running near full-load conditions, by
use of the average value of the direct- and quadrature-axis synchronous
reactances as X,.

Exomple 11-5
For the alternator deecribed in Example 11-3 IIOlve Cor the torque angle 8 and the
induced armature voltace E, by using the nonsalient equivalent circuit of Fig. 11-19.
504 Principle. of electromechanical.energy conversion

Use a synchronous reactance value which is the average of the direct- and quadrature­
axi!! reactan!'es.
Solution: The sinor phase voltage V for the wye-eonnected machine can be
specified as

V '" !~!200 /0
V3 . . . ..

For convenience we are taking the phase voltage as reference with an angle of zero.
Since the machine is acting as an alternator at a leading power factor of 0.8, let us
define a sinor phase current I' = -I. Thus from part a of Example 11...3 we have
I' 131/37°

The synchronous reactance is approximated by

In terms of I' = - I in Fif!:. 11·19, we have the following sinor loop equation:

~ =. V +jX,I'
V2
with Rr = O. 1'iubstituting the values for V and l' gives us

~ = !3,2C!Q /0 + (.')8/90)031/37)
V2 V3 - .~ --­
which reduces to

7620 + 7590/127
By solving for E" we have, finally,
E, = 9640L6:l.2
The torque angle is defined as the angle by which E, leads V, and thus .$ '" 63.2°,
The magnitude of E, is 9640 volts. The correct values for a and E, as found in
Example 11-3 are 54.7 0 and 9970 volts, The results of this simple nonsalient analy­
sis are a reasonable approximation to the correct results.

11-8 PHYSICAL SIGNIFICANCE OF THE TORQUE ANGLE


In the synchronous-machine analysis presented in the preceding sections
the torque angle 0 plays an outstanding role in predict,ing response. We
have seen how the sign of 0 dictates motor or generator act.ion, Also, 8.
critical maximum value for 0, equal to ±r/2 rad for a nonsalient-po!e
machine, gives the limits for astable steady-state operation of the machine.
In this section the physical significa.nce of t,his angle is to be studied and
the reason for its prominent role explained.
The synchronous machine 505

The angle l> entered into the analysis in Eq. (11-18), which, for con­
venien('e, is repeated as follows:
W
a(t) = - t -
n
11'
-
2n
+ n-l> (11-66)

where w is the electrical excitation frequency and a is the instantaneous


angle measured from the d axis to the a-rotor axis or, equivalently, from
the a-stator axis to the a-rotor axis. In other words, a is the angle
between the d-q axis frame and the a-b axis frame as originally defined in
Fig. 10-2.
The steady-state operation of the synchronous machine has been
examined under the condition of operation froll! an infinite bus. This
constraint fixes the stator terminal voltages regardless of the port currents.
Arbitrarily, the stator voltages for the balanced two-phase synchronous
machine were so chosen t.hat l
v: = y2 V cos wl v~ = -y2 V sin wt (11-67)
N ow let us pose the problem of finding the magnetic air-gap field linking
the stator coils in such a manner that the stator currents are zero. In
other words, we are looking for the rotating field that induces precisely
the port voltages specified by Eqs. (11-67). Then upon connection to an
infinite bus of the salHe voltages, no currents would flow in the armature
windings.
From the matrix (10-67) the equilibrium equation for the a-axis stator
winding, shown in Fig. 11-3, is given by
v'a (11-68)
where i; is the direct field ('urrent and lIf:~(a) is the mutual inductance
between the a-stator and a-rotor windings. The factor of i is inserted
as discussed in See. 10-8. The mutual inductance is a function of a and
is defined by the first of Eqs. (10-71) as being
(11-69)
where M~T is a constant. The direct field current has been designated as
i~o in the preceding seetions; therefore, we have

(11-70)
On substit.uting Eqs. (11-69) and (11-70) into (11-68) we have for ,the
a-axis stator equilibrium equation
v'a (11-71)
IEquations (J I-R) speeify these armature voltages as being applied to rotor windings.
Rempmber that the synchronous machine is analyzed by inverting the stator and
rotor and that the actual two-phase equivalent voltages are applied to the stator.
506 Principles of electromechanical-energy conversion

Substituting Eq. (11-66) for a(t) leads to


1': = (R~ + !L:p)i~ + wM~'i~o cos (wt + Ii) (11-72)
By using the expression (11-23) for H" we have, finally,
v: "" (R~ + !L!p)i! + E, cos (wi + Ii) (11-73)
Equation (11-73) can be interpreted as follows. The voltage v!
applied to the winding terminals is opposed by a voltage H, cos (wt + 0),
which is induced by the rotating field winding. The difference between
the applied voltage I'! and the voltage induced by the revolving field
establishes an armature current i!. The voltage induced by this phase
current in the resistance and self-inductance of the armature winding
v:
makes ¥p the difference between and the field-induced voltage. If E,
is adjusted to be equal to .y2 V and Ii is equal to zero, then
E, cos (wt + 0) = .y2 V cos wi v! (11-74)
and i: will be zero. A similar derivation for the b-axis stator winding
would show that the same values for H, and 0 would make i: = O.
Under the conditions of Eq. (11-74) the synchronously revolving doc
field has been adjusted to establish an air-gap field inducing stator voltages
to exactly oppose the applied armature voltages. Both the magnitude
and spatial position of the doc field must be properly adjusted to achieve
this condition. From Eq. (11-66), at time t = 0 the doc field in Fig. 11-3
is located at -r/2n rad from the a-stator axis in the counterclockwise
direction, since

a(O) = - !!.. with 0 0 (11-75)


2n
The synchronous field axis in Fig. 11-20 shows the position of the doc field
at time t = 0 required to oppose the applied voltages exactly. If Ii were
not equal to zero, then the actual position of the doc field would be o/n
radians in the direction of rotation from the synchronous field axis. Thus
the torque angle 0 divided by the number of pole pairs n is the angle
between the synchronous field axis and the actual doc field axis.
Figure 11-7 shows that a positive torque angle corresponds to gen­
erator operation and a negative torque angle corresponds to motor opera­
tion. Suppose that we adjust E, to be equal to .y2 V by manipulating
the direct field current i~o. Then if Ii = 0, both the armature currents
are zero and no power is electromechanically converted. The rotor is
being driven at w' win rad/sec, and the armature is connected to the
two-phase bus voltages. If slightly more torque were applied to the
rotor, the angle SIn would increase and the rotor axis would move ahead
of the synchronous field axis. The added rotor-port power must be con­
The synchronous mochine 507

Synchronous
field oxis

Fig. 11-20 A schematic represen­


tation of the two-phase synchronous
machine showing the physical sig­
nificance of the torque angle I.

verted to electric power, since synchronous speed is maintained. The


torque angle is positive and the synehronous machine is operating as a
generator. If a load is pla(~ed on t.he rotor shaft, then the d-c field axis
will move behind t he synchronous field axis. Electric power is then
extracted from the two-phase bus, eOllverted to mechanical power, and
supplied to the shaft load. The torque angle is now negative and we have
the syn('hronous ml1('hine operating as a motor.
The (,hange ill t he torque angle can be observed in the laboratory by
using a stroboscopic lamp synchronized to flash at the frequency of the
supply voltage. A radial line insrribed on the end of the synchronous­
maehine shaft will appear as n radial lines, since the lamp blinks n times
per shaft revolution. Remember that synchronous speed is w/n. As a
load is placed on the shaft of the machine the inscribed lines will appear
to move opposite .to the direetion of rotation. The angle through which
the lines appear to move is - a/no Applying additional driving torque to
the shaft of the machine causes the lines to move in the direction of rota­
tion, giving a positive torque angle. Thus the torque angle {j is a phys­
ically observable coordinate.
The number of degrees through which the inscribed lines appear to
move can be mechanically measured, and they are often termed mechanical
degrees. The actual electrical torque angle a is the number of pole pairs
n t.imes the number of merhanical degrees. The units of a are ofben
termed electrical degrees. Thus we have
a (in electrical degrees) = n X a (in mechanical degrees) (11-76)
Remember that all of the expressions developed in this chapter have 6
expressed in electrical degrees (or, actually, in electrical radians), The
term "mechanical degrees" is used only for convenience to correspond to
the laboratory measurement.
508 Principles of electromechanical-energy conversion

11-9 DETERMINATION OF SYNCHRONOUS-MACHINE PARAMETERS


The analysis of synchronous-machine performanee relies on our ability to
determine the machine parameter values. The general steady-state
synchronous-machine equations derived from the two-phase d-q primitive­
machine model are given by (11-27), and for convenience they are repeated
as
(11-77)
Let us look at each of the parameter coeffieients and review their relation­
ship to the original three-phase synchronous machine. I The general
sinor diagram shown in Fig. 11-8 is drawn for the a phase of the two-phase
model-machine armature. The three-phase to two-phase transformation
we have1been using keeps the a phase of the three-phase system identical
with the a phase of the two-phase system. Thus, Fig. 11-8 also represents
the a phase of the three-phase armature.
The two-phase equivalent model was constructed with two armature
windings each identical with anyone of the three armature windings on
the original three-phase machine. The distrihution factors K: b of the
two-phase stator windings in Fig. 11-3 are identical with the distribution
factors K~fJ.. of the three-phase stator windings in Fig. 11-2. As discussed
in Chap. 10, the equivalent d-q model machine has the same distribution
factors as the a-b machine. By inverting the a-b model as shown in Fig.
11-4, an equivalent d-q primitive machine shown in Fig. 11-5 is constructed.
After all these manipulations we have distribution factors for the t.wo d-q
rotor windings in Fig. 11-5 equal to the distribution factors for the original
three-phase stator windings shown in Fig. 11-2. Thus we can write
(11-78)

where K dq is an abbreviation for Kd = K; and K~(j.. is used for


K~ = Kp K~.
The two resistances Rd and R; in Eqs. (11-77) are equal to each other,
since the direct- and quadrature-rotor windings are identical windings.
For an n-pole-pair winding, Eqs. (6-105) and (6-106) give the doc resistance
values in terms of the physical dimensions of the machine and the
resistivity of the wire used. Since the three-phase distribution factors
are identical, as given by (11-78), the values for the resistances
Rd = R; = Rr can be obtained by simply measuring the resistance of any
of the three phases of the actual machine. If thc machine is wye-con­
nected, then each phase resistance is the line-to-line resistance divided

I The three-phase machine is by far the most widely used synchronous machine.
Therefore, all of our results are being carefully referred back to the three-phase
machine.
The synchronous machine 509

by 2. For a delta-connected armature t.he phase resistance is i times the


line-to-line resistance.
The first of Eqs. (11-65) shows the total electrical loss per phase to
be given by fIR', where f is the phase current. Practical machines have
electrical losses in addition to this d-c ohmic heating. Eddy currents
induced in the iron structure and in the winding conductors, and also
skin effect and resistance changes with temperature, contribute increased
electrical losses. By slightly increasing the value of R', these additional
losses can be included in our analysis. A new effective-resistance value,
which we shall call R~If' can be obtained by the following tests.
With the direct field current equal to zero, the rotor can be driven at
synchronous speed by a motor (often ealled a prime mover) coupled to the
shaft of the synchronous machine. The mechanical power required from
the prime mover represents the mechanical windage losses. The armature
is now short-circuited, and the field current is adjusted until rated current
flows in each phase. The mechanical power now required from the prime
mover minus the windage losses is the total electrical loss of the machine.
The effective phase resistance can now be calculated as
p~o ••
(11-79)
(l;.ted) 2
where p~.. is the total electrical loss at rated current. divided by the num­
ber of phases.
The quantity E, in the second of Eqs. (11-77) represents the voltage
induced in each of the armature phases owing to the direct current in the
field winding. Equation (11-23) defines E, as

(11-80)

where i do is the direct field current. Since the d-q primitive-machine


model has armature windings identical with the three-phase armature
windings and an identical field winding, the quantity Md' can readily be
calculated by using Eq. (6-94).
Experimentally the direct-axis mutual inductance Md can be deter­
mined by means of an Qpen-circuit test. The rotor of the synchronous
machine is driven at synchronous speed by a prime mover and with the
armature terminals open-circuited. The magnitude of the open-circuit
induced voltage per phase is then plotted as a function of the direet. field
current i do , as shown in Fig. 11-21. From the slope of the curve the value
of Md can be calculated.
With the a.rma.ture open-circuited the three-phase currents are
i~ = i~ = i~ = O. Therefore the equivalent two-phase currents i:
and ~
are also zero. The equivalent d-q machine currents are also zero. With
510 Principles of electromechanical-_rgy conversion

I
/
Air-gop line ...._/
/
/

Fig. 11-21 A typical open-circuit character­


Field current i ~o istic for a synchronous machine.

the a('tual armature open-circuited Eqs. (11-77) reduce to


(11-81 )
By using the a-b to d-q transformation, given by Eq. (10-:13), the cor­
responding a-phase voltage is obtained as
v~ = v~o cos na + v;o sin na (11-82)
The a-phase voltage is equivalent to the a-phase voltage of the three­
phase armature, so the open-circuit induced voltage becomes
-Ef sin na (11-83)
upon substitution of Eqs. (11-81) for v~o and v;o. When the machine is
driven at synchronous speed, eX = win; so we have

a(t) = ~ t + a(O) (11-84)


n
Upon substituting (1l-80) and (11-84) into Eq. (11-83), we can see
that the rms magnitude of the open-circuit induced phase voltage is 1
Ef wM~r .
V~ = 0 = 0 t~o (H-85)

Therefore, the slope of the curve shown in Fig. 11-21 is wM:/10 and the
mutual inductance M d' can be obtained. The straight-line portion of
Fig. 11-21, below the knee of the curve, is called the air-gap line. If we
use the air-gap line as an approximation to the actual open-circuit char­
acteristic, M:{ is a constant independent of the field current.
I With both i do and I~O equal to zero the sinor diagram of Fig. 11-8 shows us directly

that the quantity E tI v'2 equals the magnitude of the phase voltage.
The synchronous machine 511

The direct- and quadrature-axis synchronous reactances Xd and X q


in Eqs. (11-77) are the last two parameters to be examined. These
reactances were originally defined by Eqs. (11-21) and (11-22) as
(11-86)
The self-inductanres I~d and L; are given in terms of the machine dimen­
sions by Eqs. (6-92) and (G-93), where the distribution factors Kd and K;
are identical with the three-phase winding distribution factors.
Experimentally the direct-axis synchronous reactance Xd can be
determined oy the short-rircuit test previously described in this sec­
tion. With the three-phase armature short-circuited the equivalent
voltages on the d-q model Illarhine arc zero. Assuming the armature
resistances to be small, Eqs. (11-77) reduce to
(11-87)
From the first of Eqs. (11-87) we have 1.~o = O. Directly from Fig.
11-8, we can see that the magnitude of the short-circuit phase current I
is equal to 1.'do/V2. From the second of Eqs. (11-87) the direct-axis
synchronous reaetam'e is given by

Xd = lil =J~d~? (11-88)


ido I
Usually the d-(~ field eurrent i~o is adjusted to make the rms phase current
I equal to its ral<~d vn.lue. If i~o is known, the corresponding value of
Et/V2 ean be read directly from the open-circuit characteristic shown in
Fig. 11-21. If the machine is nOllsalient, then the synchronous reactance
X. = Xd = Xq call be obtained from this short-circuit test.
The experimental determination of the quadrature-axis synchronous
reactance Xq is more involved than the determination of Xd. The best
method is the maximum-lagging-current test, which will be considered in
detail. l With no load on the rotor, rated armature voltage is applied to
the salient-pole synchronous machine. The torque of electrical origin per
phase is given by Eq. (11-31), and for convenience it is repeated as

T.. = n [ ---- . 0 + --'


V E , sm V% (Xd -
X)'
q sm 20
] (11-89)
w V2 Xd 2XdX q
With the windage torque as the only load 011 the rotor the machine can
run as a motor at synchronous speed even with the d-c excitation on the
I See lL V. Hhepherd and C. E. Kilhourne, The Quadrature Synchronous Reactance

of Salient-pole Machines, Trans. AlEE, 62: 6R4-689 (1943). ,Four tests tor deter­
mining X. are discussed in this paper, induding the slip test, the maximum-Iagging­
current test, and two other less popular tests known as the reluctance-torque and
negative-excitation tests.
512 Principle, of electromechanical·energy conversion

field reduced to zero. With E, 0 the second term in Eq. (11-89), often
called the reluctance torque, is sufficient to supply the required windage
torque.
Assuming the windage torque to be very small, the torque angle 0
will also be small, particularly with rated voltage on each phase. There­
fore, for small 0 we have sin 0 "" 0 and sin 20 "" 20, and Eq. (11-89)
becomes

(11-90)

For motoring operation 0 is a negative angle and 1'.. is a negative quan­


tity. Hemember that a negative torque of electrical origin corresponds
to a genbrated torque in the direction the machine is turning, thus giving
us motoring operation. Suppose now that the direct field current is
reduced to zero and then slowly increased ill the negal ive direction. The
bracketed term in (11-90) therefore becomes slllu.ller. Since '1"<1> is a
constant small negative quantity, the torque angle 0 must increase
negatively. When H, is large enough (in the negat.ive sense) to make the
bracketed term negative, the machine can no longer generate the required
motoring torque and the rotor tends to slow down and the machine loses
synchronous speed. After the machine slips one pole, the value of H,
effectively becomes positive and synchronism will most likely be regained.
The maximum negative value of Etlv'2 for stable operation can be
obtained by setting the bracketed term in (11-90) equal to zero. Thus,
w., have

(11-91)

with a larger value of field current causing unstable operation and a


momentary loss of synchronism.
With 0 very nearly equal to zero the direct- and quadrature-axis
currents given by Eqs. (11-29) reduce to

(11-92)

and so the phase current I has a magnitude given by

I '( t'do..)2 + (t~o)'


'J v'2 V2
"" V - EtlV2
Xd
(11-93)

for very small values of o. The value of I just before the machine pulls
out of step, as the field current is increased in the negative sense, is
The $,nchronou$ machine 513

given by
V - Ej.... /y'2 V
I pO = ·~~-X-;;--- = Xq (11-94)

upon substitution of Eq. (l1-m) for E,nRx/y'2. The quadrature-axis


rend nuee is 1hen-fore given by
V
(11-95)
11'0

wh{'r(~ V is I he phase volt a~() and I pO is the armature phase ellrr{'nt. just
prc('('(lillj:/; t he point wh{,l'e the llllll·hin{' loses sym,hronislll as }<;/ is Illude
i n('I'{'asi ngly ne~a t i ve,
The various parameter tests described in this sectioll are briefly sum­
lIlul'ized as follows:

1. The open-cireuit test gives liS the relationship between the field
indueed volt a~{' }<;/ and t he field cUI'l'cn\' i d(},
2. The short-eil'('uit test provides t he lIla~nit IIde of the direet-axis
synehronous real'lluH'e X d • If t he windage losses are known, then an
effect ive annal ure rcsist tlnt'e pel' phase R~rr ('an also be obtained frolll I he
short-eir(,lIil test.
:~, The IIInxillllllll-laggillg-('U1'1'{,llt test gives the value for the quad­
rature-axis sYIH'hroIlO\lS n~:wt IlIH'e X q.

11.10 TRANSIENT TORQUE-ANGLE CHARACTERISTICS


The sYlwhronolls-hHH'hine analysis pn's{,llted in the preceding sections
hIts been based Oil I he nssHlllpt ion that the Ill!whine is revolving at a ('on­
stant sYlwhrollOlls sppcd. Thlls, the rotor :tllglllar velo('ity has been
sp{'('i fied as

(11-96)

where w is the fre<lllelwy of the e\{'driea\ excitation and n is the number


. of pole pairs ('st ahlis\wd by the llllu·hille windings. Figmc 11-7 illustrates
how Ihe torque ang;le or t h(' sYlleh\'ollollS ll1!1ehille is adjusted to compcn­
sale for various \ondillg (·ollditiOIlS. In tho ease of nlOtorillg action the
lorquo lluI!;10 b('('onl('s ilH'\'onsillg\y Il('gnt ivc ItS I ho Ill{'('hallien\ load torque
is ilH'J"('as{'(1. For allprnatOl' opNalioll Ihe torque :tngle bc('omes positive
as the e1ce\rieallond is im'l'easetl. III this scction we shall examine the
prceise manllcr in w hi('h I he t Ol'qllo all~lo ehangcs rot' a sudden load ('hange,
S 14 Principle, of electromechanical-energy conversion

We are interested in the torque angle 6 as a fUllction of time after the load
on the synchronous machine is changed.
The mechJt.nieal-port equation is given by (I 1-2), which for conven­
ience is repeated as
T' = (Jp + D)w' + T. (11-97)
where J is the total inertia of the rotating system, D is the total viscous
friction coefficient, T. is the total torque of electrical origin, and T' is the
total externally applied torque. Usually the inertia of the rotat ing system
J is sufficiently large that any variation in speed, about synchronous
speed, occurs slowly enough that the elcrtrieal-port varil~bles are always
at their steady-state values. If the torque angle is to ehange in value,
then the speed of the machine must differ slight ly from synchronous
speed. However, t.hese speed fluctuations are sufiieiently slow that. the
torque of electrical origin given by Eq. 01-31) is still valid. Hepeating
Eq_ (11-31), we have for the electrical torque per phase for a salient-pole
machine

T ...(6) (11-98)

If 6 is a function of time, then T.~ is also a fUlletioll of time according to


(11-98).
The instantaneous angular position of the synehronous machine rotor
is given by Eq. (11-18) as being
w 11"
aCt) = n t - 2n + "it6 (11-99)

where a is the angle between the a-stator and a-rotor axes on the a-b
primitive-machine modeL The angular velocity of the rotor w' is there­
fore given by

w" - a. = -nw + -a (11-100)


n
assuming that the torque angle could also be a fUlletioll of time. Also,
the angular acceleration becomes

pw' =& n
(11-101)

On substituting Eqs. (11-100) and (11-101) into the mechanical-port equa­


tion (11-97), we have

Tr = ( -Jn p2 + nf»)
p 6 + D w + vT.~(a)
n
(11-102)
The synchronous machine 515

where v is the totalnumher of phases on the maehine. For a three-phase


lll!whine v = :3 and the total el('('tri('al torque 1'. = :3T.-p, Equation
(11-102) represents a seeond-order nonlinear differential equat.ion whose
solution gives us the torque-angle variation as n function of tillle,
Let us fir"t exallline Eq, (11-102) by the linearized incremental
analysis pl'('scnted in Chap, ii. Assume that t he externally applied torque
is equal to a ('onslan! d-e value plus a slllall time-dependent increment;
thus
1"(t) = 1'~ + 1'~ (t) (11-103)
whcre t he suhs(~ript 0 is used for t.he quies('ent portion and t he subscript
1 is used for the tilll<'-depcllden! inel'ellll'lltal pOl'lion of t he applied torque,
The I'esponse of Eq, (11-10:2) is also assumed to have the same form
ns I he f()1'(·i Ill!; funel ion; I hercfore, we ean wri te

(11-104)
Similarly, all nonlinear terms in Eq. (11-102) are expanded in a Taylor
sCl·jes ahout Ihe d-(~ value of th(' torque angle. The tOlal torque of
c1eetl'ienl origin is I herehy approximated by

(1l-1Q.t»

The ('oeffic:ient of 01 has units of reeiproeal eompliance, and so the symbol


1/K will he used, where

-1 = dT.(a) I vn [
=-
VEl
-",--cosoo+
V2 (Xd-Xo)cos2a ] ,
o
K d6 A, w y2 Xd XdXq
newton-m/rad (11-106)
Notice that the eompliallce is IargN in the ease of H. nOllsalient maehine.
The saliency, in effect, gives us a stiffer spring constant or smaller COlll­
plil1.1H'e fador.
Upon substitution of Eqs. (11-103) to (11-10;')) into Eq. (11-102) the
mechanicl1.l-port equilibrium equation becomes

(11-107)

\VhCll all time-dependent increments nre sct equal to zero, the steady­
state equilibriulll cqul1.tion is given by

To = D ~ + T.(oo) (11-108)
n

RYI1(·h\,ollous specd is w/n, so the quanl ity Dw/n is the steady-state


\,'indage torque. For all alternator, the constant applied torque T~ minus
516 Principles 01 electromechanical-energy conversion

the windage torque is equal to the steady-state electrical torque. Thus,


00 assumes some corresponding positive value making T.(oo) a positive
quantity. For motoring operation T~ is negative and opposes shaft
rotation. Therefore, 00 assumes a negative value, making T.(oo) negative.
The generated torque is thus equal to the load torque plus the windage
torque.
By sUbtra.cting Eq. (11-108) frOln (11-107), the incremental linear
equilibrium equation can be put into the form

(11-109)

where

(11-110)

(11-111)

Notice that as long as K is positive (Le., the slope of the T. versus 0 curve
is positive) the solutions of Eq. (11-109) are stable. Therefore all oper­
ating values of 00 between the first positive and negative peaks in Fig.
11-7 are stable operating points. !
For a step change in the applied torque Ti(t), the incremental torque
angle Ol(t) has a form similar to one of the curves shown in Fig. 11-22. I
H the damping ratio r is less than unity, the torque angle has a damped I.

~
oscillatory motion about its new final value. These oscillations of the
torque angle are often referred to as hunting.
In order to minimize this undesirable hunting, the damping ratio
must be increased. For a given machine at a particular operating point !
00, the inertia J, number of pole pairs n, and effective compliance K are
all fixed parameters. The only remaining factor influencing the damping
ratio r is the viscous coefficient D. However, if D is increased, then the
steady-state windage losses of the machine will also increase. A more
desirable way to inc-rease the damping ratio is by means of additional
rotor circuits known as damper windings or amortisseur windings. These
circuits consist of conducting bars imbedded axially in the pole faces of the
rotor or around the rotor periphery in a nonsalient machine. At each end
of the rotor a shorting ring connects the ends of all the damper bars.
H the rotor revolves at synchronous speed, then the bars revolve at
the same speed as the rotating air-gap field. However, if the rotor goes
into a torsional oscillation, then currents will be induced in the shorted
dampcr circuits in such a direction that t,hey oppose the oscillations. For
small torsional oscillations the additional damping torque is proportional

The synchronous machine 517

!< T~

2 3 4 5 6 7 8 9 10 11 12 WItt

FiJI;. 11-22 In('reml'ntal change in thl' torque an!l;le for a small step change in the
applied rotor torque, plotted as n function of the normalized time ",.t, for three par­
ticular damping ratios.

to the time rate of change of the incremental torque angle; thus we can
write .
(11-112)

where Tn is the additional damping torque and D' is the damping coeffi­
cient due to damper windings. Adding this additional damping
term into gq. (11-107) increases f) in expression (11-111) for the
damping ratio by an amount D' without changing the steady-state
windage torque Dw/n in Eq. (11-108). The damper windings are also
used as starting windings for synchronous motors, as will be explained in
Chap. 12.
If the increlllent al driving torque T~ (t) should contain a sinusoidal
cOlllponent, then «h(/) will be fOl'ced to oscillate at this same driving fre­
quency. The Bode plots described in Chap. 4 are a convenient represen­
tation fOl' t he sinusoidal torque-angle response as a functioll of the driving
frequell('Y. For small damping ratios, a forcing frequency near the
natural frequen(,y w. of the rotor as given by (11-110) can cause large
torsional oscillations that result in extrellle shaft torques and lIlechanical
failure.of the system. Pulsa! ing loads such as air compressors or pulsating
prime movers such as internal-combustion engines must never contain a
518 Principle, of electromechanical-energy conversion

torque term having a frequency component near the resonant frequency


of the rotating system.
The torque-angle transient characteristics have been studied in this
section by considering linear incremental motion about a steady-state
operating condition. Suppose we have a synchronous motor running
with no load on its shaft. Under steady-state conditions, if we neglect
all damping torques, Eq. (11-102) shows us that the electrical torque
must be zero, which means that 0 is also zero. If a sizable load torque is
now suddenly applied to the rotor, the torque angle will seek an appro­
priate negative value at which the steady-state torque of electrical origin
will just balance the applied load torque. The ehange in 0 is not incre­
mental, so the previous analysis cannot be used to determine 0 as a func­
tion of time. In general an analog- or digital-computer simulation of Eq.
(11-102) is the most convenient way to solve for 0(0. Very often, how­
ever, our interest centers upon whether the system is stable and settles
down to a constant torque angle or whether the system is unstable and
loses synchronism or pulls out of step. A sudden change in the electrical
load on an alternator is equivalent to a step change in the prime-mover
torque applied to the shaft. Again we are interested in the conditions
for a stable system.
If all damping torques are neglected, the mechanical-port equilibrium
equation (11-102) reduces to

(11-113)

Now with Tr = 0 the torque angle 8 is also equal to zero. Suddenly


applying a step torque Tr(l) = T~u(t) causes the torque angle to change.'
The energy introduced into the system as the angle changes by an amount
d8 is given by

TO do = J (p 2 8) do
n
+ vT.~(8) d8 (11-114)

The first term on the right-hand side of (11-114) represents the portion
of the supplied energy going into storage in the form of kinetic energy in
the rotating system. If the torque angle changes from zero to a value
8..., the total change in kinetic energy is given by

,.. J
KE change =
to n - (p 2 8) do (11-115)

" Remember that a positive value of T' means an applied torque in the direction or
rotation correllponding to alternator operation. For a load torque on a synchronous
motor, rr would be negative.
The synchronous machine 519

(a)

(b)

(c)
Fig. 11-2:J A graphical determination for the stability of the synchronous machine
for a large suddenly applied rotor torqueTo.
520 Principles of electromechanical-energy conversion

Figure 11-7 is a general plot for the torque of electrical origin per phase
T,.,. By IIlultiplying T<4> by the nUll1ber of phases II, the last integral in
01-11;'») can be interpreted as ill Fig. 11-23a. The curve T~ - liT'4>(0)
is plot1cd as a funetion of o. The integral frolll 0 = 0 to 0 = 0",
is simply the total area between the eurve and the axis, as is indieated by
the two shaded portions labeled A and B. When the shaft torque T~ is
suddenly applied, the torque angle 0 increases until a lIlaximulll value 0..
is rettehed where area B is equal to ar('a A. At this torque angle the
kinetic energy in the system is zero, whi{'h means that 5 = 0 and the
torque angle has swung to its peak value. After a series of these oseilhL­
tions the damping whieh is aduully present in the system will eause the
rotor to seUle at Of in Fig. 11-2aa, where the e1erlri(~al torque equals the
applieq torque.
Suppose that the applied torque T~ were sufficiently large that area
B could not equal area A, as is the ease in Fig. 11-2!ib. The kinet ie energy
would never go to zero and a value for 0.. could not be found. The rotor
would be out of step with the synchronous field and stable synehronous­
machine operation would not result. After slipping a numher of poles,
the damping might {:ause synchronism to reo{'('ur and 0 would settle to a
value Of. However, such a mode of operation is usually undesirable.
The p('ak torque which can he suddenly applied wit hout loss of synchro­
nism is obtained hy having area ,1 just equal to the complete area B in
Fig. 11-2:{b. A sudden torque larger than this critical value causes loss
of synchronism.
If the initial torque angle just prC('cding sudden application of the
driving torque were 00 instead of zero, then all lower limits in (11-115)
would change to 00. Consider the case when the machine is initially
operating as a synchronous motor at, a torque angle - 00, as shown in Fig.
11-23c. A driving torque T~ is now suddenly applied, eausing the machine
to switch to generator operation. Stahility is maintained as long as area
B is larger than the complete area A shown in Fig. 11-2:k Assuming the
system is stable, the final operating value will he a torque angle Of as
shown in the figure. I

11-11 ELECTRICAL RESPONSE TO A THREE-PHASE SHORT CIRCUIT

The pr('('eding section presents an analysis of the mechanical motion of


the synchronous machine based on the assumption that the eleetrical
portions of the maehine are always operating under steady-state condi-

I For a more complete dis('ussion se«' J. J. Stoker, "Nonlinear Vibrations," pp. 70-S0,

IntcTscience Publish('Ts, Inc., New York, 1950.


The synchronous machine 521

tious. All the eic('tri('nl tranl'ients ('over It tillle duralion which IS so


1I11H'h short PI' than the lime dumtion of the lIleehani('ai transients that
for all pm!'! i!'al pUl'pOS!'R only t II(' I'll eady-1'1 ale ('Ie('(,rieal response influelu'es
1he I1lP('llani('al respOllse. I II this set't ion we shall exam inc t he nat \Ire of
I hp I I'aIlSiPlI1 (,\('('t ri('al j'('SPOIlSP to a lmddpnly applied short. (,in'uit· 011 al\
tluw' phases of a sYIl('hronolls Illaehill(,. In HOI'lllal operation a single­
phas(' short eir(,lIit is Illore ('onllllonly elH'OIllII<'rcd, hut Ihe more severe
threp-phase short (,in'lIit illustrates the :tnalyti(· tedllliques in a simpler
fOI'Ill.1
'I'll(' nUH,hine is hcing drivcn at sYlwhronolls speed hy a pI'ime mover,
and, dllrill~ thp tilllc interval requirt'd for th!' phase currcnts to I'eaph their
st ('ndy-sl ate short -(,in'lIit vailH's, hot h i1H' tOl'q\le nll~l() and speed of the
IIIIH,hi ne r(,lIlHin ('0111'11 nil t . Al!;td n \Y(' al'(~ s<'para t inp; I he pled I'ieul nnd
Illet'\lHlli('ld J'('SPOIlS(,S h('(':IlIs(' of t he wid!' dilTt'r('!l('c in their l'('sponsc t i llI,es,
For silllpli('ily til(' tllI'('(' phase!' are !l!;SIIIIICd to be opl'lJ-('ir('uited
pr('('('(lill~ til(' t imp ( = 0 wll('1l t hI' t hn'('-ph!ls(' shOl'l (,il'('lIit is applied.
TIll' <1-(' fipld i:o; exeitcd by a ('onSlanl-l'o/tngc SOIlI'('(~, From It pl'aeti('al
standpoint, lIIally of HI(' illlportanl tnmsi('nt dUI!';U'tel'isties wOllld be
lost if wc IIsed a ('onstant-('IIl'}'!'l\t SOIlI'('(' for the field ex(,itat iOIl. The
equivalen t d-q pri lllit ive-Ill:wili 11(1 model for the :H't ual three-phase
sYllehronolls mll(,hille is shoWIl in Fi~. ll-ii. Carefully review See. 11-2
to sec how this d-q model rclates to the :wt lIal IlHwhine, whieh is shown
s('helllut ically in Fig, 11-2.
Placing a short eirt:uit 011 the three arlllat urI' phases of the aet ual
machine means that

!'~ = 0 fOl' all t > 0 (11-116)


By IIsin~ thc thr<\e-phasc to t\\'o-pha!'1c transformation givcn by (11-6),
'an cquival('nt SI'I of a-b ('x(,italiollS al'(~ nlso fOl!lld to he z(,ro; thus
for all t> 0 (11-117)
By invert illg til(' sl at or :wd rot or of t he equivalent t wo-phase Illa(~hillc
and using the a-b to d-q tl'allsfol"lIlatioll, givclI by Eq. (11-10), the cor­
responding ('Ol\st raints on t he armature willdings in 1he d-q model of Fig,
11-;") arc found 10 he

v~ = I'~ =0 for all t > 0 (11-118)


Thlls, pl:willl!; a !'1~'llllllf:'t ri('al t hn'('-pliase shOl'l ('in'lIi! 011 1he original
nlaehilw is eqllivalent to a short ('in'llit 011 both armatlll'e windings of the
d-q lllodl'\.

I For ('olll[ll(,t!' dis('ussions of short-{'inuit r('spoII!<{>S 1>1'1': (1) n. Adkins, "The General

Throry of EiI'I'trie :'IlllI'hilll's," John Will'Y & :-:ons, 1111'., ~e\\' York, HIS!). (2)
C. Coneoruia, ":-:Yllchrono\ls :'Iim:hin('s," .John Wil"y & Nns, Ine" New York, 1951.
522 Principles of electromechanical-energy conversion

Tlu, oril!;inlll equilibriulll I'l!lwtiom; for the ('Iel'tri('al ports of the d-q
Illmll'l art' gi\'('11 hy Eqs. (11-1). The spc!'d of llw machine is assulllcd to
hI' {'!HIH! ant at sYJII"hrOIlOlls speed; t hils
w
w' =- -- (11-119)
n
Also, from Eq8. (10-;")0) to (10-;")2) we have
,rr
( rqd nJ.J~ (11-120)
ASSllll!il!1!; the t hrl'(, anllat 111"(' willdiuw> of the oril!;inal machine to be
idelltieul wit II resiHI uJ!{'es equal to H', we have
R~ R; = If' (11-121)
131use of the silllplifi{'ations givcn hy Eqs. (11-118) t.o (11-121), the
clcdrieal-porl t'quilihriulll equations redUl'c 10
P/d'P
IT' +JJ.J~p 01-122)
-~L~w

Wit 11 the fic!!l {~x{'it I'd hy a eo list ant -volt :lI!;C souree, we have sct v~ = VdO
ill (lJ-122).
Figure 11-24 shows a eompict e equivalellt eir('uit for the d-q model
of the syn(·hrolloIlH machine. The rotor speed is assullled <'onslant at
synchronolls value, flO a d-e SOUf('C w' win is eOlllledcd to the rotor
port. A d-I' sour!'e I'~o is eonneeted to the field port. The two armature
ports are simul1lllleollsly short-I'irt'uited at time t 0, as indi('ated hy the
t \Yo-pole swit('h.
The solutions for i; and i; will involve t hi I'd-order differential equa­
tions; t.herefore, three initial ('ondilions are required for eaeh unknown
variable. The initial values of i~ and i; are
o (11-123)
Binee the armature is open-eircuited for t < O. The initial first derivative
of i~ is ohtained hy evaluating the first two of Eqs. (11-122) at tinw
t = 0+, thus getting
IJ~pi~(O+) + iM~'pi~(O+) (11-124)
M::pi~(O+) + ~/J~pi;(O+)
The quantity t'~" - R,ji~(O+) is equal to zero; therefore, both
pid(O+) o and pi;(O+) = 0 (11-12i)
provided

L~TJ~ - (M'J)2 F- ° (11-126)


The synchronous mochine 523

• nL IIr l·r
'Zlq


'- ...,
I
I
I
I
.. l·r I
ztq I
I
t=O I
'­ I
...... ...J

Ideol

Fig. 11-24 Equiv/lknt C'in'uit for til(' d-q model of the synchronous machine with a
symmetrical short rircuit applied at t .., O.

The parameters L d, Va, and lII~r arc defined in terms of t.he machine
dimensions hy Eqs. (6-90), (H-!)2), and (6-94). If these relationships were
substituted into (1l-120), the expression would be equal to zero. Our
prirnitive-maehine model IlSsumes that. we have unity coupling between
all mutually coupled coils. In pnwtieal synchronous machines unity
coupling is not achieved (owing to fields around the end connections and
slot effeels which have been neglected), and so we shall assume that
(11-120) is not equal to zero.
By differentiating both of the first two of Eqs. (11-122) with respect
to t.ime and again evaluating the results at t = 0+, we have
(11-127)

From the third of 1':qs. (11-122), at t = O+, we have


'r(o+) = lIf~rwi~(O+) = _E, (11-128)
p't q aIr
V Jq
3Ir
"2 J q

and differentiating this equation and evaluating 11t t = 0+ leads to

(11-129)
524 Principles of electromechanical-energy conversion

The quantity E, = Mdrwid(O+) is the peak of the induced sinusoidal arma­


ture volt.age at the switching instant.
By s;)lving the first of Eqs. (11-122) for the field current in operator
form, we have

Let us examine the first term on the right side of (11-130). The resistance
of the field circuit is usually small, so this term can be approximated by
V~O l'ao

Rd, + L~p "" L~p

Howevtr, Eq. (11-131) is physically unrealistic, since the integral of a


constant would go to infinity as the t.ime t became infinite. We have this
result because the application of t'do to the field circuit neglecting the field
resistance would give us an infinite steady-state field current. Let us
assume that for t < 0 the field current established by the constant­
volt.age source had a value idO(O-). The voltage source has zero internal
impedance and the resistance of the field circuit is being negleeted, so the
current idO(O-) is circulating in a circuit having no resistance. In such a
circuit the total flux linkages must remain constant. l
The total flux linkages in the field cireuit for t < 0 would be
(11-132)
since i,i(O-) equals zero wit.h the armature open-circuited. The total flux
linkages in the field cireuit for any time t > 0 are given by

"d(t) (11-133)

From Eq. (11-130), neglecting Rd, we have

(11-134)

using (11-132) and the approximation that the field flux linkages remain
constant. Substituting (11-134) into Eq. (11-130) gives us

(11-135)

I Be~8.use a ronstant-voltal!:e SQUT('e vdO would I!:l'nerate an in finite field current ir


lert connected to the fi(lld windinl!:. wo nr(l efTectivrly makinl!: the Ilpproximntion that
the Rouree iR ronnected long enough to estahlish the "Ilm'nt i:/o(O-) and il'! then removl'Il
and replaced by a short circuit with i;'o(O-) still flowing.
Tile synchronous machine 525

By substituting Eq. (ll-lafl) into the second and third of Eqs.


(11-122), we have, aftel' rearranging terms,

0= {R: + [I d' - i!
,. ,
.<~!~o.~e.]
/(j + l/'dP p} iT!
'~'.
+ 3[ 'wi'
q
(11-136)
2
_ a [['d' - ( AJ d'L. p_] wid' + (Rr + [Ip)i r
ltd + Ljp (11-137)
.3
~ of. q

For convenience let us define the quantity

(11-138)

Solving Eqs. (11-1:~(i) lllld (11-1:l7) for i~ and i;, by using determinants,
leads to

A(p)i~ = -il.,;wEj (11-139)


A(p)i; = RrE! (11-140)

where

A(p) = tL~'(p)I.;p2 + ~R'["~/(p) + [.;]p


+ {w2J,~/(p)I,; + (R')2 (11-141)

Sinee the armature resislanee Rr is a small quantity, the (Rr)2 in (11-141)


can be lH'gl('(~ted. Also, the right-hand side of Eq. (11-140) eUll be tukell
equal to zero. The quantit.y A(p) ean t·hererore be put into the form

(11-142)

where the quadrat\lre-axis synchronous reaetalH'e X. = iw"~ and the


quantity X~(p) has been defined as

(11-143)

H the field resistanee [{~ is negleeted in (11-143), then X~(p) is no longer


a function of p and is given by

X'd --~w
a [I'
'd -
(Al~r)2]
--r;;- (11-144)

This reactanee quantity X~ is knowll as the direct-a.ris transient reactance,


and the induf'iall('e quantity ;[Ld - (M;n2/1,~J is termed the direct-axis
tran~iwt indllcian('('. B(,l'atlS(, both f(~ and H' are slllall, the quantity X~
is \fset! in the llIiddle lel'lll in (11-142) instead of X~(p).
526 Principle. of electromechanical-energy conver.ion

We now substitute Eq. (11-142) into (11-139) and (11-140), and, after
some manipulation, we have

(T~p + 1) (~+ 2;: + 1) id = -(T~p + 1) ~~ (11-145)

(T~p + 1) (~: + 2! p + 1) i; = 0 (11-146)

where we have defined


L~
T~ - seconds (11-147)
R~
L'd Lrd - (,lr or )2 X'd . seconds
T' -
d - R~Ld
d _
- Xd Td
(11-148)

r I Rr (J.,Xd + ..!.)
X

q
(11-149)
Xd ~wL~ the direct-axis synchronous reactance (11-150)
The field circuit time constant Td is often called the direct-axis transient
open-circuit time constant, and the quantity T~ is called the direct-axis
transient short-circuit time constant.
The solution to Eq. (11-145), using i~(O+) = pi:i(O+) = p2i:i(0+) = 0,
after some manipulation, can be put into the form

i:i(t) = - E,
Xd
E~ e-i.. ' cos wt -
+ Xd E, (J.,Xd - ..!.)
Xd
e-*" (11-1.'51)

Similarly, the solution to Eq. (11-146), using t~(O+) p2i;(0+) = 0 and


pi; (0+) Et!!L;, can be put into the form
(11-152)

In arriving at the solutions (11-11)1) and (1l-lfi2) the following approxi­


mations based upon the armature and field resistances being small are
used:
WTd» 1 WT~ » 1 r« 1 (11-153)
The d-q to a-b transformation given by (10-33) can be used to obtain
the a-b currents equivalent to the d-q currents of Eqs. (ll-liH) and
(11-1.52). Since the IX phase of the three-phase system is identical with
the a phase of the two-phase system, we have
i~ = i~ cos nIX + t~ sin nIX (11-154)
for one of the three short-circuit currents. The machine is revolving at
synchronous speed a win; thus
nIX = wi + 1J0 (11-155)
The synchronous machine 527

where 80 is the value of na at the instant the short circuit is applied.


Substitut.ing Eq. (11-155) into (11-154) gives us

i! = t~ cos (wl + 80) + ~~ sin (wl + 80) (11-156)

as one phase of the short-circuit currents. The other two phases i~ and
i~are obtained from (11-156) by adding -2'lr/3 and 2'lr/3 to the arguments
of the sine and cosine terms.
Equations (11-151) and (11-152) are now substituted into Eq.
(11-156), and, after some algebraic manipulation, the a-phase short-circuit
current can be put into the form 1

..
i' =

(11-157)

N"otiee that at l = 0, the short-circuit phase current is zero, as it should be.


The quantit.y l/X; - l/X q , often ealled the transient saliency, is
usually quite small, so the second-harmonic term in Eq. (11-157) can be
neglected. The second term ill (11-157) is a simple exponential decay
whose magnitude depends on the point in the a-c generated sine wave
where the short cireuit is applied. The first term in (11-157) represents
the steady-state short-circuit current, as is seen by referring to Eq.
(11-88). Remember that Eli Xd is the peak short-circuit current, whereas
(11-88) is in terms of the rms ShOl·t-circuit current.
Figure 11-2f>a is a sketch of the a-phase short-circuit current assuming
that 80 = 'lr/2. The second-harmonic term in Eq. (11-157) is being
'lr
neglected. Since 80 = 12, the d-c term in (11-157) is zero and the wave­
form is symmetrical about the time axis. Because of the last term in
(11-157) the magnitude of the short-circuit current is exponentially decay­
ing, with a time constant of T~ seconds. The exponential envelope starts
at Ell X; and decays to the steady-state short-circuit value of Eli X d •
The .8-phase current is obtained from Eq. (11-157) by substituting
80 - 2'lr/3 for 80 . The d-c term for the .8 phase is therefore given by

i~ (d-c term) = EI
2
(1,Xd + ~)
Xq
e-t"'t cos (8 0 _ 2'3lr) (11-158)

I The rollowing two trigonometric identities are used in this reduction:

cos w' cos (w' + (0 ) = l[coe 80 + cos (2...t + 9.) 1


sin ...t sin (wt + 80 ) = i[cas 8. cos (2...t + 8.)1
528 Principles 01 electromechanical.energy conversion

.S
1/1

(b)

·s
Iy

(c)
Fig. 11-25 Phase currents resulting from a three-phase symmetrical short circuit
applied when a(O) ,.. (Join = ."./2n.

For (Jo = 1f"/2 the term is positive, and at t = 0 it raises the exponential
envelope to

E, E, ( 1 1) (1f") (11-159)
X~ + 2" X~ + X q cos - 6
The i3-phase short-circuit current is shown in Fig. 11-25b, and it is similar
to i! except for a - 21f" /3 phase shift and the exponential deeay of the d-c
level. The"f phase is obtained by letting 00 in Eq. (ll-Vm become
00 + 21f"/3. The d-c term now shifts the envelope in the negative direc­
tion, as shown in Fig. 11-25c, to a magnitude identical with (11-159).
The syrn;hronous machine 529

11-12 SUMMARY
A model for the three-phas(> synchronous ma('hine is formulated in terms
of an equivalent, two-phase a-b llIaehillC'. Hinee t,he aet.ua.! machine often
has salient. rotor poles, the a-I) model Inust, be inverted with the salient
field winding ('onstrtleled on the slator. The windillgs on the a-a model
are identi('al with the windings on the three-phase maehille. The two­
phase excitations are equivalent in the sense that, I he total a-b surface­
current. distributiolls are two-thirds the m:\gnitude of the total a-{1-'Y
surface-current. distributions. However, all power quant.ities calculated
per phase for t he two-phase model nrC' ident ieal wit h the pC'r-phase power
quantit ies for the aetual maehine. The synchronous-machine equations
can be put in a more (·onvenient. f01"1II hy first transforming the a-b model
into an equivalent d-q model. Remembering that the a-phase variables
of the two-phase system are equivalent to the a-phase variables of the
three-phase syst em, a simple d-q to a-I> t ransformat ion changes the d-q
variables hack to aet.ual-machine variables for the a phnsc. The {1 and
'Y phases are obtained by phase-shifting the a-phase quantities hy -271-/3
and 211"/3 md, respeetively.
The most. commonly elH'ollntered opcl'llting mode for the synchronous
ma(:hilH~ is to have the armature eOlllw(·ted 10 a ('onstnnt-rolyphase-volt­
age supply known as an infinile hus. Under sleady-state eonditiolls the
excitations for the d-q model IIl1whine are shown to reduee to simple doc
constraints. The d-q Iluwhine ('urrrnts al'C also ('onsl ants, and the equilib­
rium equations are rcdu('ed to simple algebraic (nondifferent ial) equations.
Expressions for the eonverted power Hnd the torque of eleetrical origin are
ohtained for the d-q model. Hilllply dividing the torque and power expres­
sions hy 2 redu('es these quantities to ft. per-phase hasis. For the actual
three-phft.Re lIul.('hine, a lIlultipli('ation by 3 yields the total ('onverted
power ft.nd the tolltl gellernted torque.
On examining the d-q to a-b transformation for constant-speed opera­
tion, we ('an sec that, the d-q variables appear as coefficients of constant­
fr('queney sinusoidal quanl it ies. Therefore, t he I ransforlllatioll from d-g
10 a-b variables ('an most easily he :l('('olllplished by using a sinor diagram
with the d-q variahles appearing as sinor quanlities. We need only draw
the sinO!' diap;ralll for 01\(' phnsp of Ihe lwo-phnse a-Ii llIode'l, silH'e the see­
01H1 phase is !'limply di!'lpi:t('('d hy 7r 12 rad. The sinor dia!l;mlll for the a
phase also repr(~R('nls t II<' sinor diagmm for I he' a phase of I he Ihree-phase
syslem. Thus, hy drawing the a-phns(' diagram, we automatieally have
the a-phase' diap;ralll. The' fJ and 'Y phases are' ohtained by phase shifts of
-27r/;~ and 211"/:~ Tad, r('sp<'l'IiV<'ly, wllilp til(> II phase requires ft. 11"/2 rad
phase' shirl. By using the' sinor diagram. I he ('olllplete steady-st.ate
synehrollous-l1Ia('hine operation ('IHl easily he ohtained. Under constant­
530 Principles of electromechanical-energy conversion

power operation, either as a motor or alternator, the power factor of the


synchronous machine can be selected by adjusting the d-c field excitation.
Thus the synchronous motor can represent. a leading power-factor load,
and it is therefore often used for power-factor correction.
In See. 11-9 a series of experimental tests for det.ermining the machine
parameters are discussed. These tests are summarized at the conclusion
of the section. The analytic expressions for the primitive-machine
parameters, as developed in Chap. 6, are based on numerous assumptions
and approximations. For practical large-scale synchronous machines,
the primitive-machine approximations give operating results which for
most purposes are satisfactory. If, however, the parameter values for the
actual machine of interest are experimentally determined, the calculated
operati1g response will be in closer agreement with the observed response.
The transient operation of the synchronous machine is examined by
using the approximation that the electrical transients occur very much
faster than the mechanical transients. Therefore, as far as the mechanical
portion of the machine is concerned, the electrical ports have always
reached their steady-state responses. For an incremental change in the
externally applied shaft torque, the rotor of the machine will experience
damped torsional oscillations about the new steady-state torque angle.
The addition of damper, or amortisseur, windings in the rotor pole faces
increases the rate at which these oscillations arc damped. For!l. large
change in the shaft, torque, Sec. 11-10 presentR a graphical teehnique for
determining the stability of the synehronouR machine.
The electrical transients are examined by assuming the speed of the
machine to remain constant. For simplicity the case of a three-phase
symmetrical short circuit is considered. Suddenly applying a short cir­
cuit to a synchronous machine will certainly cause the shaft speed to at
least dip slightly. By the time the dip in Rpeed oc<'urs, the eicwtrical
transients are completed and all the electrical variahles have reac:hed their
steady-state values; therefore, we can assume the speed throughout the
transient interval to remain constant. The short-circuit currents depend
on the exact time when the short is applied, and they can reach extreme
peaks if the most inopportune time is choseH. The analysis of the sym­
metrical short-circuit current is developed by neglecting the damper bars
and assuming the field and armature resistances to be small. For a more
complete discussion of electrical transient responses, see the references
listed in Sec. 11-11.

PROBLEMS

ll-1 Find the speed, in radians per second and rpm, for the following syn­
chronous machines:
The $ynchronou$ machine 531

a. 60-cps three-phase 8-pole


b. 6()-cps two-phase 8-pole
c. 25-eps three-phase 12-pole
d. SO-cps six-phase 28-polc pairs

11-2 A 1Il0-hp 220-volt six-pole GO-eps three-phase nonsalient wye-eonneeted


syn(·hronou8 motor is c\(·livcrinll: full-load m('chanieal power. Windage and friction
losses arc 4 per rcnt of the 100.<1 powcr, and the armature resistance can be neglected.
a. Find the load torque, in newton-meters.
b. If the field current is adjusted for minimum line currents, find the line cur­
rents and power factor.

) 1-3 A 500-hp 2300-volt GO-cps unity-power-factor two-phase six-pole synchro­


nous motor has a full-load efficiency of O.95.
a. Find the full-load torque supplied by the motor.
b. Determine the full-load armature current in each phase.
c. Hepcat parts a and b for a three-phase wye-connected machine with the same
terminal ratings.

II-" a. The three-phase synchronous machine, shown in Fig. 11-2, has three
identical armature windings. How arc each of the two armature windings in the
inverted fl-q model, shown in Fig. 11-5, related to the three original windings?
b. How are the field windings in Figs. 11-2 and 11-5 related?
c. With the fi('ld structure of the three-phase machine in a suitable position the
maximum s('lf-inductan~e of one armature phasc is 1.2 mho After rotation of the
field hy nn appropriate angle the minimum self-inductance is found to be 0.8 mho
Is the rotor salient or nonsalient? \)('termine the quadrature- and direct-axis syn­
chronous reactanc('s for a rated frequefl!'y of GO cps.
d. Through how many mechanical dcgr('cs is the field in part c rotated if the
machine has a rated speed of gOO rpm?
N 11-5 A 'lIIlient-polc three-phase GO-cps four-pole 220-volt 30-hp delta-connected
\1Ynchronous machine is opernting in the steady state such as to have a per-phase
sinor diagram ns shown in Fig. Pl1-5. The direct- and quadrature-axis synchronous
reaetanres per phase are respectively given by X d = 2.0 ohms and X. = 1.5 ohms.
The nrmnturc rellistanrc is R' = 0.10 ohm pcr phase.
a. Calculate the rotor angular velocity in rpm and radians per second.
b. Find the power factor and the magnitude of the phase and line currents.
c. Is the machine operating as a motor or as a generator? Prove your answer.
d. Find the torque angle (I and the induced voltage E,/V2.
e. Find the total converted power and the total torque of electrical origin .

•r
~ =30omp
.[i

.r
~ =40omp
.fi

Fig. PII-5
532 Prillciples of electromechallical-ellergy COli version

11-6 A three-phase six-pole wyc-connecterl synchronous motor is rated at


500o-hp, 6600 volts, and 60 cps. The direct-axis synchronous reactance is X d 1.'>.0
ohms \':Ier phase; the quadrature-axis synchronous reactance is X q = 11.0 ohms per
phase; and the armature resistance is negligible. The motor is operating at rated
voltage and frequency with the d-c field excitation 80 adjusted that Btl viz 5600
volts and the torque angle 0 36°.
a. Draw a per-phase sinor diagram for the machine.
b. Calculate the line current and power factor.
c. Find the total developed power and the total power input.
d. Find the total developed synchronous torque expressed in synchronous watts
and newton-meters.
11-7 A 3000-kw three-phase wye-connected 4600-volt 32-pole 60-cps alternator
has negligible armature resistance. The direct- and quadrature-axis synchronous
reactances are respectively given by X d = 5.4 ohms and X q = 3.1 ohms. With the
machine 4elivering rated power and voltage at 0.8 power factor leading:
a. Determine the stator phase voltage induced by the field Etl viz.
b. Determine the torque angle o.
c. Determine the line current.
d. Draw the per-phase sinor diagram for this operating condition.
11-8 The voltage regula/ion of a synchronous alternator is defined as
Vol - VII

Vol

where V ftl = theno-load terminal voltage


VII the full-load terminal voltage
R = regulation, usually expressl'd in per cent
Thus a generator having a small rcgulation supplics a nearly constant voltage inde­
pendent of the load current. A nonsalient 300-kva three-phase 60-cps S60-volt
wye-connected alternator has a synchronous reactance X. 0.71 ohm. The d-c
field excitation is so adjusted that the machine generates mtcd voltage and rated
frequcncy with no external electrical load. Calculate the per cent voltage regulation
by. taking full load to be rated current at O.R power factor lagging. Draw sinor
diagrams for the no-load and full-load conditions.
II-9 Hepeat Prob. U.s for a salient-pole alternator having the same ratings
with a direct-axis synchronous reactance X d = 0.87 ohm and a quadrature-axis
synchronous reactance X. = 0.65 ohm.
11-10 A 500-hp 2300-volt 60-cps wye-connected three-phase nonsalient syn­
chronous motor has negligible armature resistance and a synehronous reactanee of
13 ohms per phase. With a direct field current of 94 amp the machine generates rated
open-circuit voltage when driven at synchronous speed: Assume a linear opcn­
circuit characteristic. The maximum available field current is 175 amp. With the
motor delivering rated power at rated voltage:
G. Calculate the maximum leading power-factor angle.
b. Calculate the armature current under the conditions of part a.
c. Calculate the smallest possible field current without losing synchronism.
d. Calculate the line current and power factor under the conditions of part c.
e. Calculate the dircct field current and armature current when the motor is
operating at unity power factor.
f. In one sinor diagram similar to Fig. 11-15 show these three cases.
The synchronous mochine 533

rent
\
~~.1
~
J A thrf'c-phnsf' nonsalif'nt syn('hronous alt('rnator has a diref't field .eur­
nmp. At rntt'd spf'(>{l tht' mn('hine generates nn open-('ircuit terlkinal
voltnge of IH6 volts and a short-('ircuit terminal ('urrent of on amp with all· three
tcrminals short-('ircuit('d. Cal('ulate the syndlronous rf'u('tnnt·f' r, if:
a. The Mtntor il'! wYf'-eonneeted.

b The stator is (Ielta-f'onnf'cted.

J 1"12 A 1.50-kva 22nO-volt 6{J-('ps three-phase wye-f'onneeted nonsulient-pole


synchronous alternator gives the following test data:

Armature open-circuited
Direct field curr!'nt, amp 10 20 40
Armature terminal voltng(', volts 131!l 22!}() 3770
Prime-mover driving pow('r, watts JO,lOO (const)

Armature sllort-circuited
Direct field current, amp 30.5
Armature current, amp 39.4 (rat('d current)
Prime-mover driving power, watts 17,100
Use Eq. (11-79) to ('alculate the effective phn!!e rcsistnnc(' for the alternator.
11.
b. Use the slope of the uir-gap line and Eq. (ll-RI;) to !'alculate the mutual­
inductanNl parametf'r Atri'.
c. From the short-cireuit test ('ompute the per-phase synchronous rea('tance X •.
d. Draw the per-phl!.8c equivalent ('ir('uit for the mlU'hine.

<E--- IJ"J3 A three-phase wye-conneet('d 1200-volt six-pole 60-!'ps salient-pole


synchronous machine, when driven at rat('d speed, gives the following test data:
Armature open-circuited
Direct field current, amp 5 10 15 20 25

Armature terminal voltaJ1:", voltI'< 51ir. !/x2 1242 1424 1575

Prime-mover driving power, watts 101)0 (const)

Armature 8hor/..circuit£d ,
Direct field current, amp 6.2

Armature current, amp 124

Prime-mover driving pow('r, watts 2!lOO

M azimu m-lag(1ing-rurrent t£st


The machine is t'onnected to n halant'('d 116()-volt three-phase infinite bus, with
no external load on the rotor. Tlw field eurr('nt is slowly int'reRsed in the negative
sense, and the shaft is obst'rvNI with a strohost'opie lnmp flashing at line frequent'y.
The line <'unent wlwn th(' motor Hlip;; n pole is found to he :H7 amp.
a. Calculate the effcctive armature rcsistance p('r phase nnd the windage coeffi­
cient D.
I,. Use the slope of the air-gap lin(' to enlculatc the mutual-inductance parameter
M:;.
c. Compute the dir('ct- and qundrature-axis synehronous reactan('cs per pha.se.
d. At rated voltng(' nnd frNlu('nt'y, l'nl<'ulnte the maximum reluctance torque
which the machine ('an generate.
534 Principles of electromechanical-energy conversion

e. By usinp; the as!!umptions leading to E'l. (l1-!)!), ('ftkulll.te the nep;ative ,lirt'('t
fielcl ('urrent when the motor slips a pol .. in the maximum-llLll:ll:inp;-ellrr('nt t ..st.
f. Write expresRions for the Ilelf-indudanc('s of the three Rtator phusel< as fune­
I ionR of I Ill' rotor anp;ular posit inn e.
II-J,i A pertain Hix-pole !lO-epl< Ryne'hronous nltt·mator haH n totnl shaft inr·rtill.
of 75 kp;-m'. The synt·hronous torque' of elt'l'tril'al orip;in genernted P('f flulian change
in torque anp:le is !HOO newton-m. OV('f the nornllLI operating rep;ion ('onsi<ier the
T, VB. 0 characteristic to he linpar. A 1~()-kw electri('al load is suddenly applied to
the machine. All windage, friction, and rotational damping are pra<'ti('ally nt'gligihle.
a. Find t he final 5t eady-state torque anp;le.
b. If the shaft is Iwing observed with a strnhoscnpi(' lamp flashinjl; at line fre­
quency, by how many degrees will the shaft appmr to move when tIl<' loftll is applied?
In which direction with respect to the dire(,tion of rotation?
c. Will the rotor oscillate about the final torque angle? With very Il1l1all damp­
ing find the frequency of these oscillations if they occur, as obs<'fved hy uHing the
Btroboseo~ic lamp in part h.
d. Find the maximum torque angle reached by the rotor after the electrical
load is applied.
e. If the motor were driving a periodically pulsating load, what would happen
if a component frequency in the load matched the frequl'ney found in part c1
11-15 An amort.isseur winding is now included on the rotor of the alternator
discussed in Proh. 11-14. The effective damping cocfncient for the o.TTlortisReur wind­
ing is 2000 newton-m-sec/rad. The 180-kw load is again applied at time t = o.
a, Find the final steady-state torque angle after applying the ISO-kw load.
/" Hketch the torque anp;le vs. time charactcril!ti" for all I > n.
c. Comment on the damping ratio I achievel! with this amorti!lSeur winding
design.
6 A 2000-hp 12-pole three-phase 6fJ-cps wye-conneded synchronous motor
is co ecte a 6600-volt (line-to-line) supply. The direct field current is so adjusted
that the per-phase induced armature voltajl;e E ,I v'2 = :i560 volts. The direct- and
quadrature-axis synchronous reactances arc respeetivcly p;iven by X d 10 ohms per
phase and X. = 7 ohms per' phase. The armature resistance can be neglected.
Windage, friction, and damping effects are also very small. The motor is initially
running with no shaft load, and full-load torque is suddenly applied to the shaft.
a. By using the equal-area method discussed in Sec. 11-10 determine the maxi­
mum torque angle reached by the motor on the first swing after the load is applied.
b. With full load on the motor, find the steady-state torque angle.
c. How much additional load torque can suddenly be applied without the motor
slipping a pole?
Il-17 This problem is a review of the more complicated steps in the develop­
ment in Sec. 11-11.
a. Starting with Eqs. (11-136) and (11-137), show that Eqs. (11-145) and (11-146)
are valid.
b. Carry out the solutions of the third-order differential equations in part a, and
arrange the results in the form of (11-151) and (11-152).
c. Derive the a-phase short-circuit current in the form of (11-157).
d. Write expressions for i~ and i; that are valid for all t ~ O.
11-18 1 A three-phase IO,OOO-kva wye-connected 60-cps 12,OOO-volt synchronous
machine has the following parameters:
I Answers to this problem are listed at the end of the book.
The synchronous machine 535

Sclf-induetllnce of !'neh armature phase varies between 6 and 10 mh as the rotor


revolves
Maximum value of the mutual inductance hetween the field winding and any
01\(' of til" armatur(' phaRe wirulingR = :u, m h
S('lf·iluitlCtalH'(· "f II", li.. ld wiluJinK = tl.22 h
!J-e rpsistuIH'c between any two stator terminals = 0.29 ohm
j)-e resistance of the field winrling = 1.10 ohms
a. Calculate tlw direct- and quadrature-axis synchronous reactances X d and X o'
I}. Find the dired-axis transient reactance X~ defined by Eq. (11-144).
r. Find the (lirect-axis transient open- and short-circuit timc constants r= and
r:t, respedivcly.
d. Compute the damping ratio r defined by (11-149).

11-19 The 1O,OOO-kva machine described in Prob. 11-18 is so excited that its
steariy-state Hhort-circuit current is 5 times rated current. A symmetrical three-phue
°
short circuit is applied to the armaturc of the machine at time t = when the field
axis is in line with the a stator axis.
a. Find the magnitude of the stator induced voltage E,/0.
h. }I'ind tho a-stator short-circuit current for all time t ~ 0,
c. Prepare a careful plot of i~ as a function of time similar to Fig. 11-25.
11-20 A nonsalient-pole Hynchronous motor is being supplied by a suitable
nonsalient-pole synehronous generator being driven by a prime mover at synchronous
speed. The synehronous reactanc-es of the motor and generator are respectively given
hy X ... and .Y.. Armature re!!istanl'c in hoth 1lIILehines can bc neglected.
a. Draw a pcr-phnse equivalent cireuit for the two machines. Use E,./0
and E,.. /V2 as the stator indm'ed voltages per phase for the generator and motor,
resp<'ctivcly, and IT and I as the phlk~c terminal voltage and current.
h. Draw a sinor diagram for the unity-power-factor case.
c. Show that the developed power p<'r phase by the motor is given by

p BloB,... (
,<f, =2(X.-+X;;') Sill ~. + ~..)
where ~. and ~ .. are .the torque an!l;les of the generator and motor, respectively.
d, For constant ficld excitations on hoth machines draw a sinor diagram for the
condition whereby the 1lI0tor is developing maximum torque.
chaptn

THE INDUCTION MACHINE

12

The induction machine represents one of the most


useful forms of a-e-powered electromeohanical
rotating machines. This device can be constructed
with no physi(~al connedions to the rotor circuits.
The rotor currents are generated beeause of the
magnetic coupling between the stator and rotor, and
the maohine becomes extremely rugged, dependa­
ble, and, relatively speaking, maintenance-free.
Large induction machines are usually operated
from a balanced three-phase supply system.
Therefore, the analysis presented in this chapter is
formulated for a machine having a three-phase
stator winding and a three-phase slip-ring rotor
winding. By using the three-phase to two-phase
transformation presented in Sees. 10-7 and lO-8, an
equivalent two-phase a-b model of the three-phase
machine is derived. As with the synchronous
maehine discussed in Chap. 11, the two-phase a-b
model can most easily be analyzed in terms of a d-q
primitive machine obtained by using the a-b to d-q
transformation. By using inverse transformations,
the results of the d-q analysis can be related to the
two-phase a-b model and also to the three-phase
machine.
The servomotor represents a very practical
two-phase induction machine commonly used in
positioning control systems and servo-instrumenta­
tion systems. The a-b primitive machine is a
straightforward model of the two-phase servomotor.
The a-b machine is an intermediate step in the
complete three-phase analysis; so by considering
the three-phase case, we are also obtaining results
applicable to the two-phase machine. The single­
phase induction machine, used in many consumer
536 products such as furnace blowers, refrigerators,
The induction machine 537

garbage disposal units, and dishwashers, can also be readily studied by


starting with the a-b model machine. Here again, the results of the
three-phase analysis ('an also be applied to the single-phase machine.
In order to give a simple qualitative discussion of the operation of the
indUe! ion ma('hine, t he squirrel-cage rotor structure will be introduced
first. The squirl'el-('age rotor is different ill eonstruction from the d-q
primitive-machine brush-('011lIllutator rotor or the a-b primitive-machine
brush-slip-ring rotor st ructure. A se('ond form of rotor construction,
known as a wouud rotor, is eonsidered next. This cOllstr~tion is similar
to rotor structures in previously (~onsidered Iluwhines, and t.he analysis is
therefore to be based on the woulld-rotor ('onstruetion. The analytic
simi luri ties hel weell I he sqllilTPI-eage Hlld wound-rotor constructions can
be shown, and thus the presellted analysis includes both types of induction
machine.

12-1 AN ELEMENTARY VIEW OF THE INDUCTION MACHINE


The squirrel-euge rotor structure, shown in Fig. 12-1, consists of copper
or aluminum bars plneed axially around the periphery of a laminated iron
cylinder. The lamination of the iron is necessary to force the rotor
currents to flow in the ('ondueting bars. At both ends or'the rotor cyiin~­
der, ('ondud ing <end rings arc IIsed to ele('\ rically connect the ends of the
rotor hal'S. Quitp obviously, thiH rotor iH n vpry n1gged, durable mecha­
nism and t h<'rc is lit t Ie cluuwl! of any meehllnical or electrical failure,
Around Ihe roIOl', It uniform stator structure is constructed, A
three-phase st atOl' winding is now asscmbled and is cOllnected to a balanced
three-phase voltage supply. The stator assembly of the three-phase
induclion l1la<'hine is very similar to Ihat of the three-phase synchronous
machine. Bnlaneed three-phase st alor currents in the three sinusoidally
distributed stator windings establish a surface-current density' on the
inner periphery of the stator rotating at synchronous speed win, ~here w
is the radian frequency of the excitation and n is the number of pole pairs

Fig. 12-1 The hasic squirrel-cage


rotor structure.
538 Principles of electromechanical-energy conversion

Stotor uniform field


direction inside rotor
dS =dSQ(J

Assumed positive direction

tor induced voltage

Fig. 12-2 The induced voltage in a single loop of thc stationary squirrcl-caRc rotor
structure.

established by each of the stator windings. For a nonsalient, or uniform­


air-gap machine, the magnetic field 'in the air gap is radially direeteg and
sinusciidally distrihuted. with a period of 27r/n radians, and it rotates at
synchronous speed. In the case of a two-pole field, n = 1, t he magnetic
field inside the rotor cylinder is uniformly directed, and it also rotates at.
synchronous speed.
Let us now consider what the currents induced in the squirrel-cage
rotor bars would be, assuming initially t hat the rotor is held stationary.
The entire squirrel cage can be considered to be I:onstructed from a num­
ber of Rhortcd loops which are joincd together at each end. One such
loop is shown in Fig. 12-2. The field dUI! to t he stator ex(·itat ion is uni­
form within the rotor, and it is assumed to he pointing along the direct
axis at time l = 0, as shown in Fig. 12-2. In cylindrical coordinates the
uniform two-pole field inside the rotor can he expressed as
B = B",[cos (wt - 8) a, + sin (wt - 8) ae] (12-1)

where Bm is the mllgnitude of the uniform field. Equation (12-1)


represents a two-pole field rotating in the positive angular direction. By
using Faraday's law, the voltage induced in this single loop, made up of
two rotor bars, can be found. Since the rotor is stationary, Faraday's
Jaw reduces to

e·In d = -
Jrs ~~.
at
dS (12-2)

Referring to Fig. 12-2, Eq. (12-2) gives


(12-3)
The induction momine 539

where a and l are the radius and length of the rotor cylinder, respectively,
and thus 2afis the cross-sectional area of the loop formed by the two rotor
bars. The angle th is the pia~e angle for the kth loop, as shown in Fig.
12-2.
The result expressed by Eq. (12-3) can be summarized for all of the
rotor bars, as shown in Fig. 12-3. The small dots and crosses indicate the
direction of the induced voltage in each of the rotor bars. For the
particular bar considered in Fig. 12-2, fh appears to be approximately
1f/3 rad. Thus at time t = 0 the induced voltage is negative, or opposite
to the assumed positivc dire(·tion for the indueed voltage shown in Fig.
12-2. Thus, the dots and crosses are as shown in Fig. 12-3 for this par­
ticular loop. The rule that a voltage is so induced as to cause a current
in a'direction to (~ancel the d\ll.llging flux linkages can aid in checking
these results. For the loop at 1f/3 rad the flux linkages are directed to the
right in Fig. 12-3 and nre decreasing at time t = O. Thus the induced
voltage has a direetion which would cause a current creating a field also
directed to the right to maintain these flux linkages, The rotor bar at
1f/3 has current coming out of the paper, while the bar at 1f + 1f/3 has
current, going into the paper. By the right-hand rule, flux is directed to
the righ t as desired.
The induced voltages, given by Eq. (12-3), are sinusoidal and have
the same angular frequency as the applied stator currents. With the
rotor stationary each of these rotor loops acts as a secondary of an ordinary
transformer; thus, the secondary and primary frequencies should be
identicl:l!. In the steady state the current-in each oft-he shorted loops
depends on the nature of the (~olllplex irnpedulwc of the loop's circuit.
This illlpedulH'c has a resistive eOlllpolwllt, and lUI illduetive reactive
eomponcnt. III general, the induced rotor currcnts in each loop, made
up of two bars Oli opposite sides of the rotor, can be expressed in sinor

Zero induced
voltage axis

Fi~. 12-:~ Induced voltt\~(,R in the rotor


hars of a stationary squirrf·l-"!l.~e rotor.
540 Principles of electromechanical-energy conversion

Currents log vOltoges


by this angle
I Zero induced
current oxis

Fig. 12-4 Resulting induced cur­


rents in the rotor bars of the stationary
squirrel-caKe rotor.

form bt
R lIr ~.=_.9k
(12-4)
~R+J~L·

where E = -B.. (2al)w/V2 from Eq. (12-3) and Rand D are the resist­
ance and self-inductance of each of the rotor loops. Since the rotor bars
are relatively large in cross-sectional area, their resistance is small com­
pared with the inductive reactance wD. Consequently, the induced sinor
currents will lag the induced voltage by an angle approaching '11"/2 rad.
Figure 12-4 shows the direction of the resulting currents in each of the
rotor bars at time t = O. Notice that t he axis of zero indueed currents
Jags the zero-induced-voltage axis, shown in Fig. 12-3, by an angle
approaching '11"/2 rad.
The force on each of these current-carrying conducting bars can be
determined from
F=lixB
where 1 is the length of each bar, i is thf1 inst,antaneous-current vector,
and B is the magnetic-field vector at the conduding bar. The directions
of these forces are up for currents coming out of t.he paper and down for
{'urrents going into the paper, as shown in Fig. 12-4. The net sum of
these forcf1s causes It counterelockwise torque on the rotor strudure. The
rnagnit ude of this torque of eleetrieal origin can be shown to be propor­
tional to sin 0, where 0 is the angle between the direetion of the stator
field and the axis of zero induced currents, as shown in Fig. 12-4. As
the stator field rotates, the axis of zero induced currents also rotates and
maintains a ('onstant dlsplacement angle 0 and thus a constant torque in
the sallle direction as the direction of stator field rotat ion.
If the rotor turns at an angular velocity less than that of the rotating
stator field, the voltage induced in the rotor bars has the same form as
The induction machine 541

Eq. (12-3), except that the frequency will be the difference between the
field and rotor angular velocities. The angle a will still remain constant
at some value between zero and 11"/2 rad; thus a constant torque is still
generated. Since a lower-frequency voltage is induced in the rotor, the
impedance of the rotor loops becomes smaller and more resistive in char­
acter. The inductive reactance wI- is reduced because w is decreased.
Thus, the rotor currents are more in phase with the induced rotor voltages.
Referring to Fig. 12-4, this means that a would increase, approaching
11"/2 rad. The sin a also increases as the value 11"/2 is reached. However,
the magnitude of the induced voltage decreases and the generated torque
may actually increase or decrease depending on the exact values of the
parameters. In a later section these considerations are discussed
analytically in great detail.
If the rotor revolves faster than the stator rotating field, induction
generator action results and energy flows out of the stator ports. The
wound-rotor construction is introduced in Sec. 12-2, and the analysis
proceeds from this mechanistic model. This initial section is intended
only as an introduction to give some physical feeling for how the induc­
tion machine operates.

12·2 PRIMITIVE-MACHINE MODEL OF THE THREE-PHASE


INDUCTION MACHINE
The induction machine is usually constructed with a uniform air gap;
thus both rotor and stator are nonsalicnt structures. The wound-rotor
three-phase induction machine is shown schematically in Fig. 12-5.
The three stator windings are sinusoidally distributed windings with
equal distribution' factors of K' turns/m. The three rotor windings are
also sinusoidally distributed with equal distribution factors of Kr turns/m.
In general, K' ."r= Kr. The axes of the stator windings are separated by
211" /3n radians, where n is the number of pole pairs developed by the
three·phase windings. The axes of the rotor windings are also separated
by 211" /3n radians.1 Slip rings are used to make external connections to
the rotor windings. Very often, the three rotor phases are wve-connected,
and therefore only three slip rings are required.
The three-phase stator is usually connected to a balanced three­
phase sinusoidal voltage supply known commonly as a three-phase
infinite bus. The rotor windings either are externally short-circuited
through the slip rings or are connected to a balanced three-phase load,
which is usually resistive. We can always assume the rotor windings to
1 For a complete description of a sinusoidally distributed multipole winding see

Sec. 6-8.
542 Principles of electromechanical-energy conversion

be short-circuited and account for any external resistance load by simply


increasing the value of the rotor winding resistance parameter.
The resistance of each of the three stator windings is designated as
R', and the total resistance of the three identical rotor windings, including
externally added resiBtance, is designated as Hr. The self-inductance of
the three stator windings is to be designated as L~/ty, where the three sub­
scripts simply mean that this symbol reprcsents the self-inductance for
all three windings. Similarly, the self-inductance of the three rotor
windings is given by L;/ty. From Eqs. (10-71), we can see that the mutual
inductance between the stator and rotor windings is a sinusoidal function
of the angle a between the stator and rotor axes. Equations (10-71) give
the general form of the mutual inductance between a stator and a slip-ring
rotor j\vinding. Let us designate t~e peak value {)f the stator-rotor
mutual inductance as M::/ty. The resistance and inductance parameters
of the three-phase induetion machine are summarized as follows:
Stator phase resistances R! = R~ = R; = R'
Total rotor phase resistances R~ = RJ, = R; = R'
Stator phase self-inductances I,! = I-~ = 1-; = L!/t-r
Rotor phase self-inductances V:. = Lit = IJ; = IJ~/t.,
Peak stator-rotor mutual inductances
(12-6)
The three-phase machine pictured in Fig. 12-!i can most easily be
analyzed by transforming to an equivalent two-phase a-b model machine,

y'

Fig. 12-5 The three-phase wound-rotor induction machine.


The induction machine' 543

Fig. 12-6 Equivalent two-phase a-b model of the induction machine.


as shown in Fi~. 12-6. 1 The a-II machine has two identical stator
windings sinusoidally distributep with 11 distribution factor of K' turns/m.
Slip rings are used 10 make exlerll:\! {'onneetions to t.he two ident.ical rotor
windings. The rotor windings are also sinusoidally distributed, but with
a distribution factor of K' turns/ill. Notice that the stator windings on
the a-I> modeillre identical with the three stator windings on the original
three-phase machine. Also, the two rotor windings have exactly the
same construction as the rotor windings on the three-phase machine.
The stator and rotor structures 011 t he three-phase and two-phase model
have exactly the ~me physieal dimensions.
The three-phase variables can be transformed to an equivalent set of
two-phase variables by using the a-fJ--y to a-b transformation given by Eq.
(10-119). Since three-phase quantities appear on both the stator and
rotor of the actual induction machine, the transformation must be
expanded to indude both stator and rotor variables. Thus the three­
phase currents are transformed according to the matrix equation
(12-7)
where
i~
iiJ
.,,. i; (12-8)
l ../I'Y = {r
..
ij.

t';,

IFor a detailed discussion of the three-phase to two-phase transformation review


Secs. 10-7 and 10-8.
544 Pr;nc:ip/es of e/ectromechonico/.energy conversion

and the complete transformation matrix is given by

1 -i -j 0 0 0
0 va Va
-2 0 0 0
2
(l~ ...~y = i 0 0 0 1 -i _t
(12-9)
11"

0 0 0 0
va
--2 2
va
The voltage variables are also transformed hy a similar matrix equation;
thus
(12-10)
where I
v~

v~

r v~ (12-11 )
v!~y = V'
a
v~

v~

The factor of j is introduced in the transformation of Eq. (12-9) in


order that the per-phase powers for balanced excitations be invariant
under this transformation. Section 10-8 shows us that the impedance
matrix of the a-b model machine must be slightly modified to compensate
for the introduction of the factor t. Each of the inductance parameters
appearing as coefficients of transformed currents in the equilibrium equa­
tions must be multiplied by a factor of~. If we were to write a 6 X 6
impedance matrix for the three-phase machine and properly use the
a-fJ--y to a-b transformation matrix, according to Eq. (10-140), this
factor of ~ would appear before each of the appropriate inductance
parameters. Therefore the parameters of the equivalent a-b model
machine can be summarized as follows:
Stator phase resistances R! = R: = R'
Total rotor phase resistances R~ = R; = R'
Stator phase self-inductance L~ = L:' = fL~/ly = L'
Rotor phase self-inductance L~ = L; = IL~y =U
Peak stator-rotor mutual inductances
(12-12)
The three-phase a-{3--y parameters are summarized in (12-6).
As in the case of the synchronous maehine, the analysis of the indue­
tion machine is greatly simplified if we further transform the a-b model to
The induction machine 54S
q .,
...!.!

Fig. 12-7 Equivalent d-q model of


the induction machine.

an equivalent d-q model, as shown in Fig. 12-7. The equivalent d-q


variables are obtained from the a-b variables according to the matrix
equations

i~ = (l~~ •..J:l (12-13)


Vd~ = (l~~ •."V:b (12-14)
where

(12-15)

and the a-b to d-q. transformation is dcfined as

_.[~l_ ~
a"
Iif.ob - 0 0 cos na
o 0 sin na
-'i~ no]
cos na
(12-16)

The d-q and a-b stator quantities are identical. However, the rotor
quantities are transformed in such a fashion that the d-q machine using a
commutator-and-brush mechanism has a total rotor-surface-current den­
sity that is identical with the total rotor-surface-current density of the
a-b machine using slip-ring connections.
In construction the stator and rotor windings of the d-q machine are
respectively idcntical with the stator and rotor windings of the a-b
machine (which in t urn are rcspcetivcly idcnt ieal with those of the original
three-phase machine). The equilibrium equations for the d-q primitive­
S46 Principle. of electromechanical-energy conver.ion

machine model are given in general by

'l] [Rl + /,l.


v'q
v~
= ' 0
M~'p
R;
0
+ JJ~p
G~~wr R~
lIfd'P

+
0
L~p

v; -G;dw' M:;p -G;~w'

The mechanical-port equation is given by

T' = (Jp + D)w' + 1', (12-18)

with the torque of electrical origin derived in the conventional fashion,


describfd in Sec. 7-3, as

T tI -- - P.
wr

(12-19)

The positive sense of T. is opposite to the assumed positive direction for


the rotor speed w', For T. positive, the electriralIy generated torque in
Fig. 12-7 would be in the dockwise direction opposing positive speed w'.
The relationships between the a-b parameters and the equivalent d-q
parameters are sUlllmarized by Eqs. (10-70) to (10-7..t). Based on these
relationships, the following sUlllmary of resistanees and self-induetances is
formulated:!

L dq = L!. = ~L~tI'Y (12-20)


L~q = L:. = FJ:tI'Y
The mutual inductances have the relationships

(12-21)

By llsing Eqs. 00-49) to 00-fi2). the rotational inductances can be written


in terms of the self- and mutual indll(~ta!H'es as follows:

G~; nM;' nM" G~~ = nL; = nU 02-22)


G;d = nMd' = nM" G;'d = nL; = nlI
Substituting the simplifications of Eqs. (12-20) to (12-22) into the five

J The rrlationllhips giV('n by 02-20) are interpreted by using only one of the suh­

scripts. Thus, the term Uj. is a shorthand notation for Uj, = U; and U'ap-y stands
for fl~ = flo R;.
The induction machine 547

port equilibrium equations given by (12-17) and (12-18) leads us t.o

Vti]
I"
g =
[HI +0
Up
H'
o
+ Up
A/"p
o
l'~ A/"p nAI"w' +
Fl' Up
l'~ -nAl"w" A/"p -nlJ'w'

where the expression for 7'. from (12-19) has heen substituted into the
rotor-port torque cquation.
Equations (12-2a) and (12-24) represent a set of nonlinear simul­
tancous differential equut ions describing t he opera! ion of the d-q model of
the induction machinc. \Ve huv(> been very careful in the formulation of
this model to keep traek of thc relationship between t,he d-q parameters and
the paramcters of hoth the two-phase a-II model and the three-pha,se a-{J-"{
induction machine.

12·3 CONSTRAINT EQUATIONS FOR STEADY-STATE OPERATION


ON A BALANCED INFINITE BUS
The rotor circuits of the induet ion mlwhine are to be eit her short-circuited
or terminated by sOllie external \'('sistall{'e, If this extemal terminating
resist ance is lumped with t he resistance of the winding, then in general the
port voltages can he set equal to Z(,I'O; thus

I!~ = !I~ = v~ = 0 (12-2:»

By substituting (1'2-2.') into the three-phase to two-pha,se voltage trans­


formation given hy Eq. (12-10), we obtain for the equivalent a-b rotor
voltages

(12-26)

Fiimilarly hy use of the a-I) to d-q transformation equat.ion (12-14), t,he


equivalent d-fl rotor volt:tges are

(12-27)

for ordinary indudion-maehille operation. In certain special applications


the rotor eireuits are not short-!'ircllited, hut these (,a,ses are omitted from
t his present at ion,
Th(' !It alor ports of the t hr(>e-phase llllwhine are very often connected
to a balanced three-phase infinite-hus voltage source. The three stator
548 Principles of e/ectromechanico/·energy <;onver$ion

voltages can therefore be specified as

II; = y2 V' I~OS wt v~ y2 V' cos (wt + 2;)


v; = y2 V' (~Os (Wi - ~f) (12-28)

where V' is the rms value of the applied phase voltages. By using the
stator portion of the a-{3-'Y to a-b transformation given hy (12-9), the
equivalent two-phase stator excitations are found to be
lJ~ = V2 1" cos wi l'~ = y2 V' sin wt (12-29)
The stator portion of the a-b to d-q transformation, given hy (12-16),
shows 1l' that
t'~ = 1': = y2 P cos wi (12-30)

If the induction machine is operating from a halanced three-phase supply,


then the equh'alent two-phase excitations are also halanced. The single­
phase inch!!'t ion motor anl1 the two-phase servomotor do not, in general,
have balanc'ed excitations. However, these llla(~hilleS lire special cases
of the two-phase machine rather than the three-phase machine. We shall
not he concerned with unba\aneed three-phase operation.
Equations (12-27) and (12-30) represent four external constraints on
the d-q model of the induction machine. The fifth constraint arises from
a specification of some relationship coneerning the mechanical port. This
relationship eould be in the form of a passive mechanical load if the
machine is being used as a motor. If generator operation is the case, the
manner in which mechanical energy is being supplied into the mechanical
port would be given. These five constraints, together with the five equi­
librium equations (12-23) and (12-24), are sufficient to completely deter­
mine all 10 port variables.
As a practical consideration, we shall be interested particularly in
the steady-state response to the sinusoidal excitations spec'ified by Eqs.
(12-28) to (12-30). Under steady-state ('onditions, with the speed being
constant, all time-dependent d-q variables are sinusoidal functions of the
excitation frequeney w. By using sinor notation, all p = did! operations
can rigorously be replaced by the jw operator as discussed in Hec. 4-6.
The four electrical-port equations therefore become

V~] _ o
V'9
o -
[R< +0 jwL'
jwM"
R' + jwV
n,~J"wr
jwM"
o
RT + jwJI
o ]
jWM".
n11w' I;;
1'I~ ]
o -nJ~I"w' jwl\J" -nI1w' R' + jwlI I;
(12-31)
The induction machine 549

The si nor rcprcsen (at ions for the sl at or excitations, as given by Eqs.
(12-30), arC' 10 he taken as
Vd = jVd (12-32)
where v~ and VZ a\"(' tim rlltH vahl!'!' of thl' t wo-phai\(' sl al or-port voltages.
If a balalH'cu t.wo-phase supply is used, then \'d = ~.; = V'.
If balanccd phase voltagcs, in the forl\l of (12-30), arc applied to the
two-phasc stator windings, then Ihe stator ('urrcnts will also form a bal­
am'ed two-phase set with the salllc frcquency of w rad/scc. In Sec. 10-2
hal anced stational'y C'urrcnts of fl'cqueney ware shown to produce R
!Ilagncti(~ field rot at ing in t h(' air ~ap at all allgular frcquency equal to
win, where n is the nUllIhcr of pole pail'S estahlished by the windings.
The angular spced of the rot at illg Illngnetit, field win is termed the syn­
chronous speed of thc mat'hinc. Let us define this synchronous speed as
w
w, = (12-33)
n
In Icrms of the sYfH'hronolls ~p('('d tilt' ,~lilJ of the mlwhine is defined as

(12-34)

whcrc w' is the aetual rotor angular vcloeity and 8 is the dimensionless
quantily tcrmcd "slip." Notil'e that if the rotor angular velocity equals
w., thcn thc slip equals zcro. Also, if the 1'01.01' isstandingstill, then the slip
cquals unity. Thc rotor angular veloeity is given in terms of the slip by

w' = (1 - S)w, = (1 _ S) w (12-35)


n
Equations (12-31) can now be rewritten by substituting Eq. (12-35)
for the rotor angular vclo{'ity, thus giving us

V~]
V'q
[ R' +0 jwl,' R'
o
+ jwl,'
jwilf"
o o
jwM" ]I~l
I:
o -- jwllf" wM"(1 - 8) R' + jwU wU(1 - S) I~
o -wl\f"(l - 8) .iwM" -wU(1 - S) Fl' + jwU I;
(12-36)

as the sinusoidal stcady-stMe equilibrium cquations for t.he four electrical


ports. The pro('edurc for solving Eqs. (12-;)6) (~all be greatly simplified
hy introducing a new or t·ransforlllf'd scI. of elect,rieal coordinates. For­
mlllly. 1his pw('ed1ll'e is kllown as a symmetrical-component transformation.
B('forc int rodu('ing t h(' sYllllllet rieal-{'olllponent transformation, let
us examine thc torque of ele('\rit'al origin T. undcr the sinusoidal steady­
550 Principle. of e/ec'romechanical·energy conversion

state operating mode. Equation (12-24) gives the instantaneous value of


t he elect rieally generated torque as

(12-37)

If all eurrents are sinusoidal time funetiolls, then the average value of T.
is given by

(12-38)

where He indieat!'s thai only Ih!' real part of the ('ollJplex expression is
retained and th!' • III ("a liS to conjugate the iudi('atcd siuor-eurrenl term.
Rpuwlllbpr that Eq. (l2-:~8) g;ivl's ouly th(" aI'pra(le I'af/l(~ of Ihp torque of
clef'tri('a! origin, Pulsating {'olllpon("nls with zero llveral1:e vllluc {'ould
very ~\'ell exist, hut t h("y nr(" not ine\ tided ill expression (12-:18).

12·4 TWO-PHASE SYMMETRICAL-COMPONENT TRANSFORMATION


Thc I wo-pha!!c sYl\llllcl rieal-I'omponcnt I rallsforlllal ion for the !!\.ILlor
voltages is given hy the following matrix cquation:

(12-39)

The vollag(" matriecs in Eq. (12-:m) am givrnt individually by

v.
dq =
Vd]
V; (12-40)

V.

= V~]
V'­ (12-41)

Notice that both thed-q voltage matrix and the (+ -) voltage matrix are
composcd of sinor quant ities. The transformation is defined in terms of
complex quantities and is ('ommonly callcd a complex transformation.
The a-b to d-q transformation, defined in See, 10-3, involves real-time
functions for the variables, and hence it is called a real transformation.
The symmetricai-romponent transformation matrix for the stator
quantities is defined by·

.
«+-.d9 = 1[11

V'2 (12-42)

1When used with the d-q variable!!, thill transformation is often ealled 8. jQrvlard­
backward (I-b) ·t.ransformation.
SCI' D. C, White and H. H. Woodson, "Electro­
mechanical Energy Conversion," ('hap. 4, John Wiley & Sons, Inc., ~cw York, 1959.
The induc:tion machine 551

The complex power supplied to the sl alor ports of the d-q 1ll11ehine, in
terms of 1he si nor quail t Hies, is j!;iven by
P~q = [(Idq)t]*Vaq (12-43)
where V dq is defined by (12-40) and
[(I~q) 'J * = (I~) * (I;) * (12-44)
1--_

The superscript. t indicates t.hal t he (~olulJln Iaq matrix should be trans­


posed int.o a row Illal rix, and the * indi('ates that, I he matrix is to be con­
jugated. In terllls of I he (+ -) variables the power supplied is given by
P~_ =
[(I~_)tl*V~_ (12-45)
where V~_ is defined by (12-41) and
[(I~_)tl* = (I~)* (I~)* (12-46)
I I

For both complex power expressions, (12-43) and (12-45), the real part
represents the watts and the illlnginnry part t,he l'ars being supplied to the
stator ports.
When 1·:q. (12-39) is subst.ituted into Eq. (12-4!i), t,he (+ -) power
beeomes
(12-47)
The eurrents undergo I ransforll1!1.t.ioll in the sallie fUHhioll as the volt.nges;
thus
(12-48)
and b y transposin~, we have
(I~j = (a~_.dqIdq)1 = (Idq)t(a~_.,fq)t (12-49)
where we have used the mat.rix identit.y for the transpose of t,he produet of
two matriees. By conjugation of Eq. (12-49) and substit,ution into
Eq. (12-47) the (+ -) complex power is given, finally, by
P'+_ = [(Idq)ll*[(a~_.dq)'J*a~_.dqV~q (12-50)
By comparing Eq. (12-50) with Eq. (12-43), we ean see that the power
will be invariant under tmnsforlllnt ion, meaning P'+_ = Pdq, if
(12-51)
where 'U is the unit matrix with l's on the main diagonal and O's in all
other positions in the matrix. For the ease of a real-transformation
matrix, such as the a-b to d-q transformal ion, all terms are real and tJ:e
conjugate operation is not required. Thus, for a real transformation,
552 Principles of electromechanical-energy conversion

Eq. (12-51) reduces to the form of Eq. (10-101), which does not have the
conjllgale operation.
Let us now t esl the symmetrical-eolllponent. transformation III at rix,
defined by Eq. (12-42), for power invariance.

,.
[(a:_.&q)'] a~_.dq =
1
y2 [1_j 1]
j X
1
y2 [1
1

·1[1+1 j-j]
= 2 -j+j -p-p
= [~ ~J = ~
Thus the symlllet ri("al-('olllponent I ransforlllal ion III at rix keeps the lotal
complex power invariant under I nlnsforlllat iOIl.
Given I he (+ -) voltage quant ities, the d-q voltage quantit.ies can
readily be found simply by inverting Eq. (12-39), which yields

V~. (12-52)

where

(12-53)

The d-q {'urrents {'an be determined ill a similar fashion if the (+ -)


('urrents are known.
The Rignifi('allce of the sYllllllctri!'al-eomponent t.ransformation <:an
be illustrated hy its applicatioll 10 the direet- and quadrature-stator-port
voltages given in sinor form by Eqs. (12-32), which for convenience are
repeated as

V~ = jVd (12-54)

Expanding the transformation given by matrix (12-39) gives us the stator


(+ -) vo\t.ages as
V' = ....~ (V' + J!V') V'- ~ (V'd -
= y2 J!V') (12-55)
+ V2 d •
q

Substituting the direcl.- and quadrature-stator sinor voltages of Eqs.


(12-.54) into (12-05) gives us

V·- = ~
V2 (V'd - V!)

(12-56)
The induction machine 553

If the lJIagnitudes of the sinusoidal voltages should happen to be


equal or balaneed, meaning V~ = V; = V', then

V~ = j y2 V· and V'-- = 0 (12-57)

Balanced eX{'itations give rise to a uniform revolving magnetic field in the


positive-ungulnr dircdion. For this rellson the + voltage variable is
termed the p()sitil'e-sequerlcc voltage, and similarly the - voltage variable
is e:tlled the negatil'e-sequctlce volf llj!;e. As seen from Eqs. (12-57), for a
halalwed excitation resulting ill 1\ uniform positively revolvillg field,
ollly positive-sequPIH'P jPrIlH! will rpslIl1. On the ot her hand, if

thml

V~ = 0 (12-58)

and only negativc-sequem'e terms will he nonzero. This later stator


ex('itation corresponds to 11 ulliforlll field revolving in t,he negative­
angular direction. For unball\need stator exeitations, where V,~ ;r6. V:,
both positive- I\nd nCj!;at.ive-sequcn('c terms exist. The result.illg field
is made up, in genentl, of two unequa.l uniforlll fields. One field revolves
in the positive diredion, corresponding to the 'positive-sequence quan­
tities, and the other revolves in the negative direction, corresponding
to the nega.tive-sequence terms. Usc of the transformation to det.ermine
induetioll-nllwhine operation will ('Iarify these concepts.

12·5 SINUSOIDAL STEADY-STATE EQUIVALENT CIRCUIT


The sinusoidal steady-state equilibrium equations for the four electrical
ports, given by Eqs. (12-36), arc for convenience repeated as

V~]
V'
q =
[ R' +0jwV o
R' + jwL!
jw.U"
o o
jwM" ] I:Id]
V~ jwM" wMor(l - S) Rr +jwU wII(1 - S) I,j
V; -wM"(l - 8) jwM" -wl./(l - S) R' + jwU I;
(12-59)

where V.j and V; have been reinserted for darity in the following develop­
ment. Let us now apply t.he sYlllllletri('al-component transformation,
introduced ill SC('. 12-4, to Eqs. (12-!'iH). The two-phase symmetrical­
component transformatioll matrix for the two stator windings is given by
554 Principles of electromechanical-energy conversion

(12-42). The same transformation CI\n also be applied to t.he two rotor
windings; and so, in general, the complete two-phase symmetrical-com­
ponent transformation for both stator and rotor windings is given by

.,.
a+_. dq =
1
-\12 ['
1
~
j

-j

0 1
0 1
lJl j
-j
(12-60)

Notice the two superscripts used to designate the complete 4 X 4 trans­


formation matrix.
Equations (12-59) can be written in matrix form as

(12--61a)

By using the complete symmetri<,al-component transformation, given by


(12-60), we can put Eq. (12-61a) into the form

(12-61b)

where

V':_ = a':_.dq V~~ (12-62)


1':_ = a':_.dqI~~ (12-63)
and
(12-64)

Rather than just write the results of Eqs. (12-61) to (12-64), let us carry
out these matrix operations in a more detailed manner. By expanding
Eq. (12-62), we have

:v~ ~2 (Vd + jV~) V·- =~


v'2 (V'd ­ jV')
q
(12-65)

1 1
V"+ v'2 (V~ + jV;) V: = V'2 (V~ - jV;) (12-66)

Substituting the equations for Vd and V; from (12-.')9) into Eqs. (12-65)
gives us, after some slight rearrangement of terms,

(12-67)
The induction machine 555

Expanding the transformation of the currents as given by (12-63) leads


us to

1~ =
1
V2 (14 + .11;) 1~ = J2 (14 - .11:) (12-68)

1~ = J2 (1~ + .11;) 1~ = J2 (1~ - .11;) (12-69)

Substituting these p08itive- and negative-sequence currents into Eqs.


(12-67) gives us, finally,

V~ = (Ro + jwl")I~ + jwM"I~ (12-70)


I V~ = (R' + jwU)I~ + jwlvf"l~
as the t.ransformed stator equatiolls.
The rotor e(luations are transformed in a similar fashion. Beginning
with Eqs. (12-66) we substitute the equations for V~ and V; from (12-59)
and thus get

(12-71)

By using the definitiolls for the positive- and negative-sequence currents,


given by Eqs. (12-68) and (12-69), and recognizing that the last two terms
in each of Eqs. (12-71) involve - j times the p08itive-sequence currents
and +j times the negative-sequence currents, these rotor equations can
be reduced to

./ V~ = jwM"SI'+ + (R' + jwUS)I~ (12-72)


V~ = jwM"(2 - S)I~ + IR' + jwU(2 - S)ll~
If the posi tive-sequence e(lua! ion is divided by S and the negative-sequence
equation by 2 - S, the rotor equations are given finally by

(12-73)
556 Principles of e/ectromechanical.energy conversion

The forcing fundiollsllre zero ill both cuses, since short circuits are assumed
on the rotor ports of the induetion machine aI'! originally proposed in Sec.
12-3.
By c'olllhining Eqll. (12-70) and (12-7a) into matrix form, we have
Eq. (I2-Hlb), where

n' + jwV 0 jwM" 0


0 n' + jwV 0 jwllf"
Z"\ ­ jwM" 0 H' + JW. (" 0
J(r .
0 jwM" 0 2 ":",,) + JWU
(12-74)
I
The t mnsforllled (+ -) impedanC'c III at rix is IIYllllllCt rical ahout its prin­
cipul diagonal, and therefore I hc posi t ive- and negat ive-scqllenc'e equilih­
riulll equatiom; !'an be repf(~lIell1<·d by a passive ('i)'(·uit. This four-loop
('ire'"il i!' shown in Fig. 12-8, as is easily verified hy simply writing the
loop equations for the equivalenl ('ireuit. For a given set of stator volt­
all:C'H, V: aJ\d V~ ('all he dC'tcmllined. The steady-stale analysis of the
induc·tion maehine now redu('cs to simply solving for the positive- and
n('gativ<~-s('quen('e ('urrents in the ('in'uit of Fig. 12-8. By suitahle inverse
t mnsfol'lIlal ions, t he ad ua! !'t al 0)' and 1'0101' curren Is ('an t hen he fou nd.
Not ic'e, however, I hal two )'('!'ist ive clelllenls in Fig. 12-8 are functions of
the slip 8. Thus, ill order to detcfllIine the c'irc'uitall'csponse, somc infor­
malion c'OIwel'llillg t he rotor angular velo('ity lllust be known. By rc\uting
I he lorquC! of pic!'! rieal origill to I he pOf;il ive- and Ilegat ive-scqllc!lce cur­
rent!' and spP('ifyi!lg Ihe cxterual shafl load, Ihis rotor speed information
('all hp dC'1 C'rllli!lcd. Thc nexl sc('l ion c'onsiders I he lorquf' relationships.

R'
R' jw(L'-M'") 5 jw(L'-M")

jwM"

jwM'"

R' jw(L'-M"j R'


2-5
Fi~. I:.!-S COIIlJllet" HillUl'loirial IItf'adY-Rtatf> "'lllivalent "ircllit for the induction
maC'hine.
The induction machine 551

12-6 TORQUE RELATIONS


Equation (12-38) expresses the total average torque of electrical origin in
terms of the direct and quadrature sinor <mrrents (or the two-phase model
machine. For <!onvcnicllce this c<tu:~tioll is repeated as

(12-75)

where Rc means to retain only the real part. of the complex expression in
square brackets. By using the inverse symmetrical-component trans­
(ormatioIl, given by

(12-76)

we can eliminate the direct and quadrature currents in favor of the posi­
tive- and negative-se<tuel1(~e currents. Expanding Eq. (12-76) for stator
and rotor quantities gives us

I~ = ~2 (I~ + I~) (12-77)

r=~I--=(r+r)
d V2 + ­ (12-78)

Let us now form the two produ('.ts ill Eq. (12-75),

_~-- (I~ + I~) _ ~- (J1~* - J1'-*)


v2 v2 (12-79)
I:I~*
. =~ V2 ('-J1'+ + J1~) -~-
V2 (1'*
+
+ I:")
Notice that in conjugating a direct- or quadrature-current expression, the
sign preceding all j's is (~hanged and all positive- and negative-sequence
currents are also eonjugated. Upon multiplying the binomial expressions
in (12-79) and forming the differell{'e of t.he two products, we have

(12-80)

By substituting expression (12-80) into Eq. (12-75), the total average


electrical torque is found \'0 be

(12-81)

in terms of the positive- and negative-sequence currents,


558 Principles of e/edromechanical-energy conversion

From Fig. 12-8, depicting t he equivalent eireuit for the induetion


maf'hine, we can write the following two relationships:
jwM"
1+ = - -.--.~----- I' (12-82)
N:IS +iwU +

iwM"

- ~-'--'-'--'--'----- I'
1-:.. (12-83)
R'I (2 - 8) + jwl.! ­
When E<)s. (12-82) and (l2-8~) arc solved for 1't. and I:' and the results are
substituted into (12-81), the total average cleetrieal torque becomes

Tw. He { -n (.~~ + iT,) (1+)2


+ n[(;T!!::'{)~ +i U ] (I~Y} (12-84)

where 1'(1')* equals the magnitude of the ('orrcsponding rotor current


sinor s~;;-d, or (/'p.
Thus by retalliini only the real part of Eq. (12-84), we have

T•.• v • (12-8;"})

Notice that the total average torque of eleetrieal origin is made up of two
terllls. The first term depends on the positive-scquenee rotor curren t,
and t he second term depends on t he negative-sequence rotor eurrent.
Equations (12-;')7) and (12-;)8) give the sequenee voltages for balanced
st at or ex!·it at ions on t he induction Illa!'hine. These equat ions show I,hal
the negative-sequence 8t ator voltage is zero if a UllifOl"lII posit ively rotating
field result s frOlll the d-q stat or voltages. Referring to the eif(~uit of Fig.
12-8, if V'- = 0, then I~ O. Thus, only the first term ill Eq. (12-8.'»
remains and T ...... is a negative quantity. By the sign eonventions we
have been using, a negative eleetrical torque means that the machine is
generat iug a torque in the positive-angular direetion. Thus, the positive­
sequence portion of the circuit generates a torque in the po.~itive-angular
direction. An entirely similar argument follows for the negative-sequence
portion of the equivalent. cireuit, wit h the se!'ond term in (12-85) generating
a torque in the negative-angular direction.

12·7 PER-PHASE QUANTITIES AND THE PER-PHASE


EQUIVALENT CIRCUIT

Equation (12-81») expresses the total average torque of elertrical origin


generated by'bot h phases of the d-q two-phase lllodei. RilH'e t he machine
is eonne<'ted to a balanced two-phase infinite bus, we can divide (12-85)
The indudion machine 559

by a factor of 2 and obtain the average electrical torque per phase as


being

(12-86)

Both the symmetrical-component transformation and the a-b to d-q


transformation are power-invariant, so Eq. (12-86) is the generated
electrical torque per phase for both the d-q and a-b models of the induc­
tion machine. The three-phase to t.wo-phase, or a-~-'Y to a-b, trans­
formation is arranged to maintain the per-phase torque and power
quantities invariant. Thus (12-86) also represents the per-phase torque of
electrical origin for the original three-phase machine.
The per-phase converted power for the machine is given by
(oJ
P •• = - T •• (1 - S) - (12-87)
n
since the rotor angular velocit.y wr = (1 - S)w/n, according to Eq.
(12-35). Therefore, on substitution of Eq. (12-86) into (12-87), the
converted or developed power per phase becomes
P = /l"(1 - S) (Ir )2 _ Rr(l - S) (I')! (12-88)
.. 2S + 2(2 - S) ­
Each of the two resistive elements ill t.he rotor portion of the equivalent
circuit of Fig. 12-8 can be split. into two parts as follows:

and (12-89)

The equivalent, !'ireuit, ('all 1I0W he I'cdrnwn, as shown in Fig. 12-9, with
these resistors divided as indi('ated hy Eqs. (12-89). If the positive­
and negative-sequence stator exeit,ations V+ and V'- in Fig. 12-8 are
H'

r l-S
-H ­2-S

H'

Fig. 12-9 Complete per-phase sinusoidal steady-state equivalent circuit for the
induction machine.
560 Principles of electromechonical·energy conversion

divided by V2, the four loop currents will also be reduced by a factor of
I/v2, as shown in Fig. 12-9. With the equivalent circuit in this form,
the power dissipated in the two resistors, which are functions of the slip S,
is preeisely the eonverted power per phase, as given by Eq. (12-88). For
balanced stator excitations, Eq. (12-!>7) or Eq. (12-.'i8) indicates that
V~/V2 or V'-/V2 has a magnitude equal to the actual voltage v- applied
to eaeh stator phase for both the two- and three-phase model machines.
Referring to the original sinusoidal steady-state relations for the
induction maehine, given by Eqs. (12-36) in terms of t.hc d-q model, we
can identify the per-phase average power lost in ohmic heating of the
stator willdings as being given by
Pert-phase power loss in stator = !R'(I~I~* + 1;1;*) (12-90)

This POWN quantily is usually given the name stator copper loss. The
real-part operator fie l:3.n be omitted in (12-90), since no imaginary terms
appear. This results from the fa(~t that any sinor times its conjugate is a
real number equal to the magnitude of the sinor squared. If Eqs. (12-77)
are substituted for the direct- and quadrature-stator currents in Eq.
(12-90), the stator loss becomes

Per-phase stator copper loss


= lR- [il~+_I'-) (I~* +_e) + (-~1~ ~Jl'-) (Jl~* -Jl~)]
~ V2 V2 V 2 · V2
= !R'(I~I~* + I,-e)
or finally

Per-phase stator copper loss = R' /' )' + R' (V-2


(V) /' )2 (l2-!H)

In other words, referring t{) the equivalent circuit of Fig. 12-9, the machine
stator copper loss per phase is simply the power dissipated in the two R­
resistors by the positive- and negative-sequence stator currents. In an
identical fashion the per-phase rotor copper loss can be shown to be given
by

Per-phase rotor copper loss = Rr 1')' + R' (1')2


(;2 :y-z (12-92)

or simply the power dissipated by the positive- and negative-sequence


rotor currents in the two Rr resistors of Fig. 12-9.
In summary, the per-phase equivalent circuit of Fig. 12-9 gives the
following information concerning t he steady-state operation of the induc­
tion machine:
Tne induction macnine 561

1. The power dissipated by the positive- and negative-sequen('e rotor


currents, I:/y'2 and I'jy'2, ill the resistors W(l - 8)/8 and -Rr(l ­
8)/(2 - S), respeet ively, is t he developed or ('ollverted power per phase
for the nllwhille.
2. The powcr dissipatrd by t hc posi1.ive- and negative-sequence
stator currents, I~/y'2 and I'Jy'2, in the two U' resistors represents the
per-phase st aior ('opper loss fot' I he Illaehine.
:3. The power dissipat rd by t he posit ive- and negal ive-sequenee rotor
currents, 1:/ y'2 and I~jy'2, in the two W resistors is the per-phase rotor
copper loss for the ma(·hille.

Example 12-1
A 23lJ1l-volt 75-hp four-pole flH-cps thrl'e-phase wye-connected wound-rotor induction
motor ill <lrivinl/: Il ('Olllltant lond torque at It spe{'(\ of 1741l rpm. Windage losses are
5 per <'ent of tI", lotal input 1){)\V('r. Th" parltlneter~ of the three-phase machine are
as follows:

Htator resiRtan('e per phase U' 2.IlO ohms

Hotor resistance per phaRe N' = 2.22 ohms

Stator 8elf-indU!,tance per phase I'~/l~ = O.5il7 h

Hotor ..elf-indu('tan('e p('r phase I';'/l~ = 11.51'7 h

Peak stlltor-rotor mutulil indudan('(' M;;I1~ = 0.57:3 h

Find:
a. The machine slip
h. The complete per-phase equivalent circuit
c. The stator line currents and power fal'lor
d. The total stat or nnd rotor ('opper IOllsell
c. The tolal ('onverted po\V('r and the value of th" load torque
f. The eflieiency of th!' motor
Solution: a. Th!' synchronous SPI'C(] of !I. four-pol!' (or two-pole-pn.ir) machine is

w, = ~ = .;!~~ = IS!) rad/sec = lS()() rpm


2

Therefore the slip, at a speed of w' 174fl rpm, is given hy

b. Thl' complete per-phase equivn.lent cireuit for the induction motor is shown
in Fig. 12-9. The excitations applied to this cir!'uit are the positive- and negative­
sequence stator voJtap;es divided hy 0, as shown in Fig. 12-9. With balanced three­
phallI' voltages applied to the stator, the equivalent-circuit excitations, according to
Eqs. (l2-57), reduce to

IV'I
~ = V' "" vf:i
2300 = 1330 volts I~\ =0
562 Principles of electromechanical· energy conversion

2.90 j7.54 2.22 j7.54

~
1330[9 j324 .f2 71.B

Fill;. 12-10 Per-phase equivalent circuit for thl' machine in Example 12-1.

Since the negative-sl'qu€ ' nce excitation is zero, for halanced ex('itations, only the
upper half, or positive-sequence half, of Fill:. 12-9 needs to he considered.
The inductance qunntities in the per-phase I'quivalent eircuit are J times the
inductan('e quantities for the three-phase machine. Thus we have

T,' 1 ~L~/I'y ~. X 0.5i!7 = O.SRO h

],r =. il'~/lY
= ~ X 0.5"7 = O.RRO h

M" =~M::/lY = .~ X 0.573 = O.i!60 h

For a slip S = 0.03, the output resistor has a value of

1 - S J - 0.03 _
R, ";<," = 2.22 X-1fli:f'- = ,1.8 ohms

Thus the ('omplete per-phase equivalent cir('uit is as shown in Fill:. 12-10. For con­
velli"Il('I', tlu, phuHIl afll/:le of V·,! 0 i!l tILkr'fi all 7,I'TO.
r. The actual Htator line eurrent is equal to 1-:,1 V2, sirll'e the rwgativc-sequf'nr'e
stator ('urn'nt I ' j 0 is zero. Let us review why this is true. Equations (12-77)
show us that

I~

with I:' = O. Hince the d-q and a-b ma"hine stator quantitil's are id("ntical, we also
have

I~ I:' = I;

No\\' tho:> a phase of the three-phase machine is identi('al with the a phas!' of the t\\'o­
phase equivnlcnt machino:>; thus

I'
I'a = I'a = ....±..
y'2

and so the actual stator phase curTo:>nts, or lin!' ('UTrents for the wye connection,
(!'1 ual 1'.1 0.
'I'll" "I1Trents in the o:>quivaient dn'uit or Fig. 12-10 ('an he ohtll.in(>d by a loop
formulation, hut a simpler proredure is to solve for the impc;lan('e seen hy the \'oltage
source. This drivinll:-point impedanee is
The induction machine 563

The stator currents are therefore

The line ~urrentl'! nrc 17.4 nmp at a power fadOT of cos 2:1.1 = (U)2 lagging.
d. The per-phase (·opper los.~el! arc simply the power dissipated in the 2.90- and
2.22-ohm r('sistors in Fig. 12-10. The positive-sequence rotor current is given by
r =_
._J~~
r+
V2 (71.8 + 2.22) + j(7.54 + ;)24) ~2
Thus we have

Total stator copper losses =:1 ('0)'/' U· ax (17.4)' x 2.90 = 2630 watts

Total rotor copper losses = a( r )2 Ii'


.0 ax (16.6)2 x 2.22 = 1836 watts

e. The converted power per phllse is the power dissipated in the output resistor;
therefore, the total converted power is

p=a
• ( r)' ) s
v'2
+ fl'
S
:\ X (lG.6)' X 71." 5!l,aOO wntts

The windage lo...~('l' ar(' I} p('r ('('nt of til(' lotal input power. The total three-phase
input IlOwer i..
P F. = v'3 V 1.1 I, ('Os (J = v':l x 2:100 X 17.4 X 0.92 = sa,800 watts
where VI, nnd h ure the rcspc(,tive line quantities. The total input power can also
be obtuincd by adding the total stator and rotor copper losses to the converted power,
thus giving
PI-' 26ao + IHall + 50,:100 = Ga,766 watts

Taking 5 per cent o( the input power, we have for the windage losses

P .'nd = 0.05 X 63,RO() 3190 watts

and so the net output power absorbed by the constant load torque is

Plo.d = P. - P.ind = 59,aOO 3190 = 56,110 watts


The actual speed of the machine is given as

",' = 1746 rpm = IRa rad/sec

and so the majl;nituoc of tho load torque is

f. The overall cffidency of the machine is

PI...d 56,110 8

P; = 63;800 - 8 .0 per cent


564 Principles of electromechanical-energy conversion

In EXfllllpic 12-1, th(' POSiliw-s('qll(,ll<'(, mlor enrr(,llt jR I!;iv('n hy


r
Y; = -HHi/-lO.;j (l2-H:3)

RiIH'(' 1h(, nf'galiv<'-s('qU(,lH'P mtol' ('llfl"(,llt ]S 7,('1'0, ttl(' ('quiva\ent d-q


rotor ('\I!TPlltS, using (12-78), arc given by
1',
I~ Hi.fi/ -J O.;j
y2
(12-94)
r
rq "h = jHUi/ 10.:,

TI1<' !illr fllll!'lionR ('orr('spondin!!; to the d-I! sinor quanli!ips arc

i~(t) = - V2 X }(j.n sin (wI - 10.;')


(12-H.')
i;(1) = V2 X lO.n COI'l (wl - 10.;')
By Ul'(' of H](' d-" to a-I! Ir::uH,forlllatiol\ the ('orresponding a-phase
ro tor ('\lrren t is

i~ ('OS na + i; sin na (12-96)

The rotor speed w' = a, and therefore, hy suhstitulinl!; i~ and i;. we have
- V2 X 16.6 sin (wi nw"t + '1') (12-97)

where I{J ill an llppropriat e phal"e angle dependinl!; on how t o is ehosel\.


Frolll Eq. (I 2-:~;;) defining slip wc (:ItJl show that,

Sw = (w - nw') (12-98)

Thus I he a-phase rotor rurrelll, which is ident j<'al wit h t hc a-phase rotor
('11 rren t on the three-phase machine, is

i~(t) = - y2 X 16.n sin (Swi + I{J) (12-99)

X ot ice I hat the [l('t lIal fre<jueney of the indtl('cd rotor currents is
Sw, whieh is comll1only known as slip frequency. At sl andstill the slip
S = 1.0 and the rotor frequeney equals the sl ator applied frequency.
As the rolor speed approa('hes synchrollous speed the frequency of the
:trtllal rotor ('urrents heeomes smaller and approaehes zero as S goes to
zero. The d-q primitive machine provides a ('ollvellient simplificution
hecause. for the d-q model, both the rotor and stator currents are Itt the
Ralllp ('xc·it at ion frPCIIICIl('Y w, t hUB allowinl!; t he usc of simple sinor notation
ill O(l(' ('olllplete p<jllivalelli. (:ir('lIit. If Wp worked direetly from the a-I)
model. we would have' 10 deal with hoth the exeitatioll frequency w und
the rotor slip frequen('y Sw.
The induction machine 565

12-8 THE LEAKAGE-MUTUAL EQUIVALENT CIRCUIT


Th(, per-phase equivalrlli ('in'Hit developrd ill the pr('('cding sections and
showlI ill Fig. 12-!J is not t he most ilH'llIsivr efluivalent cireuit for the
imhu,t ion madlillp, A port iotl of t he input. power is lost in hysteresis
alld (·<ldY-('UlTPllt efTerls in t h~ 81 at or :lIId rotor it'OIl sl rudllres. By
Wilkin!!; It llJodification of the I'quivalcll\, eireuil of Fig. 12-9, I,hese addi­
t iOllal {'led ri('al losses whi('h hav(' \)('('n lI('glcd~'d ('all be illduded in our
analysis ill at \('!tst 1Ul approxillJ:l.tl' manll(,l'. TllP illdu!'tioll maehine has
rertaill Rillli\urities to nil ordinary trallsfol'lll('r wit h It short-dreuited
secondary winding, so leI liS IWl!;ill Ihis 1II0dific-al ion by bl'iefiy studying
the t \Vo-winding lmnsfol'lllCr,
The equililll'illllJ equatiolls for the 1wo mutually coupled coils shown
in Fig. 12-11a ar(' given ill matrix fOl'1ll by

[HI .up
+ IjiP
R2
JlIp
+LP 2
]il] '~]
i2
(12-100)

Rj N,: N2 R2

v,

EJ~ (al

R, L,-II( R2 L z -II(

v,

9'~F~l
(bl

R, L,-aM a2Ra a 2(L2- If)

v,
E;{~'M~
(el
j.. ,,1
'3 1

Fill:. 12-11 (a) Two IlHll!:nf'ticnlly coupl('d ('oils. (I.) Self-inductance-mutual­


inductance equivalent cir(,uit. (c) Another equivalent circuit involving the trans­
formation ratio a.
S66 Principles of electromechanical-energy conversion

'1>11

Fig. 12-12 Component equivalent fluxes for the two couplcd coils.

Fignre'12-11b shows an equival('nt. eir('uit having the Sl1lll(' equilihriulll


equatiolls. The ('ire'llit showll ill Fig. 12-11" is very similar to I,he posi­
t ive- or ncgal ivc-sequclH'e pori iOlls of Fig. 12-!J. FiJ?;lII'e 12-11c shows
a second poR,<>ihlc cfjuivalcnt I'ir{'uit having ('qllilihriulll cquations ident,i('al
with Eqs. (12-100). The faetor a is a ('onslant knowll as the transforma­
tion ratio. For our purposes we shall aSSUIlI(! I he t.ransforlll:tt ion mt io to
be equal to I he turns ratio of the two eoupled {~oil8 of Fig. 12-11; thus wc
have
Nt
a=- (12-101)
lV,
Let us now review the definitions for the self- and mutual inductances
shown in Fig. 12-11. Figure 12-12 shows the same two coupled coils
wound on an iron core in order to show clearly the sense or direction in
which each coil is wound. For simplicity we shall assullle all inductances
to be linear. The equivalent magnetic flux lines shOwn in Fii 12-12 are
defined as
'follows: l

<f!ll = flux due to current i l linking only ('oil 1


<f!l' = flux due to current i2 linking only coil 2
<f!12 = flux due to current i l linking both coil 1 and ('oil 2
<f!21 flux due to curren t i2 linking bot h coil 1 and coil 2

For linear inductances, the total flux linkages in coils 1 and 2 are
given respectively by

(12-102)

I The term "equivalent flux" means that if the flux quantity is multiplied by the
number of turns which it links, the correct flux linkages are obtained.
The induction machine 567

Therefore, by using the equivalent flux definitions, \Ye have


I'! = ~ i
11 :it-O
= !V_I (<p~" ±_<p_I~)
1.1
(l2-1O~)

1'2 = ~21 = N 2 (<P12,+ <P21) (12-104)


12 ,-,-0 12

and for M we have 1\\'0 equivalent defining rebt ionships

M =~I = Nl;~1 (12-10.'5)


t'2 il=O 'h

lIf = ~I = N 2,<P12 (12-106)


11 ;.-0 t\

From Fig. 12-11c let us now ('xnmille ! he 1hree indtH't nll('e quantities. \ (,"
By usillJ.!: Eqs. (12-IO;~) and (12-100), \\'(' lmv('
l P"
J. (
1'1 - aM =
N (
1 <Pt!,
+ <P12) _____
N N
! 2.<p12 =
N
~<P11
\ (Jv<\'-:"~
(12-107)
t\ N2 t\ t\

The induetmwe quantity d('lincd hy (12-107) results frolll flux lines whieh
do 110\ link hoth ('oils. These Ilonlilikitlp; lines are (·all(:\rit.~i.:~g~}l.:ux
lines, and 1he ('OlT('spondillp; indu(,t !\Il('C is known as 1he Iral,age in(/.uctance,
which will be designated by n low('f('nse I. Thus the leakagc'Tildtj('tance
of ('oil 1 is defined as

II = N I,<PII = 1'1 - allf (12-108)


tl

Similarly, by using 1<:<)s. (12-104) and (12-IO;i), we have for the leakage
inductance of ('oil 2

(12-109)

Substituling 1<:qs. (12-10:{) nnd (12-105) into the first of Eqs.


(12-100) leads to

(12-110)

upon using the leakage indu('tall('e definition of (12-108), The sum of the
two equivalent fluxes ill the last IeI'm of Eq, (12-110), ('omlllonly termed
I he mutual flux, is denoted by
~-.~-~

<P... = <P12 + <P21 M "tv. .. f +I v 't- (12-111)

From Fig. 12-12 we ('an see that the Illutual flux <Pm is the total equivalent
flux in the iron ('ore aetually linking hoth eoils. The last term in Eq.
'-(12-1 iO), which can be written as e = N 1 d<Pm/dt, is the voltage induced in
568 Principles of electromechanical-energy conversion

coil 1 by the total flux in the eore_ This voltage p, appears aeross the
inductance aJl in Fig. 12-11c. The current 111 the inductance a:l! is
equal to i l + i 2/a, which ('an be written as
. . ( ai2)
to:' = h - - r.... "') f'\ ~ .1! J
16. )-- - (
(r' / ( : - (12-112)

where im is known as the magnetizing current.


The current -i2
is the current in a load which could he cOllnected to
the secondary winding, with the minus sign in('hHied for a passive load.
When -i2 is divided by the transformatioll ratio a, we say that the load
current is referred to the primary. Thus Eq. (12-112) indieates that the
primary current minus the secondary load c~urr<'nt ~ef(>rrpd to the primary
is equf.1 to the magnetizing current. The magnetizing ('urrellt im
establishes the equivalent mutual flux IPm in the iron core. The induct­
ance aM is usually known as the "!l_aunetizing inductance.
The <;.ore-Ioss effects composed of hysteresis and eddy currents are
related to the time rate of change of the mutual flux IPm. Thus the eqre
losses are direeUy related to the voltage e = N1(dlPm/dl) and ('all be
taken intu ac'count by a resistance placed in parallel with the magnetizing
indw·tatH'e aM, as shown in Fig. 12-13. The suhscript h + e is used to
dcnote hysteresis plus eddy-current losses. If the value of R"H is
properly adjusted, the power dissipat.ed can he made to correspond to the
power lost in tore heating. The resistance a 2 R2 and leakage indlletance
a 2 12 arc termed Ihe secondary resistance and leakage inductance referred to
the primary. The entire circuit shown in Fig. 12-13 is the complete
equivalent circuit for the transformer referred to the primary. 1
The complet.e per-phase equivalent ('inmil for the induction machine
shown in Fig. 12-9 can now be put into an equivalent form, as shown in
Fig. 12-14. The eore-Ioss resistor Rh+. is adjusted in sueh a fashion that
the total power dissipated in the two core-loss resistors in Fig. 12-14 is the

1 For a more extensive discussion of this form of the transformer equivalent circuit
see, (or I"xample, A. S. Langsdorf, "Theory o( Altcrnating-current Machinery," 2d
ed., chap. I, McGraw-Hili Book Company, New York, HI5S. ,
-,

v aM

Fig. 12-13 LcakaRe-inductance-mutual-Rux equivalent circuit for the two coupled


coils including core-loss effects referred to the primary.
The induction machine 569

R' jwl'

V!
J2

R' jwl S a 2R' jwa 2 l'


Filt. 12-14 Leakaltc-indurtanN'-mutual-flux per-phase equivalent circuit for the
indurtion lllllf'hine referred to the IItator.

tot al machine ('ore loss per phos". Heartanre quantities are shown for
all of the iuJu!"! ivc clcnwnts. Tlw ,'e:\('tall('('s wZ', wZ', and wall!" are
~enerally known :\s t he stator and rotor leakage reactances and the mag­
nf~tizi7lg rl'lldau('(', I"{'spe('t ivC'ly. Xot i{'C' I hal all rotor parameters ('ontain
the t ransfol"llHlt ion ml io squa.reu and :\I"e I hus referred to the stator. Since
th(' positive- and negalive-sequC'IH"C rotor ('urrents are divided by a, all
rolor pOWN!'! nrC' the salllc a,,, with the Ol'igillal drruit of Fig. 12-9.
In !'!UIII nllU'y , t he per-phase <,quivalcnt cireuit. of Fig. 12-14 is more
useful than the dr('uit shown ill Fig. 12-H because the core-loss effects
eall be takC'1l illto ac('oullt. Pl'lwtieally nil of the literature discussing the
induct ion nllwhine uses the leakage-indu(·tancc-lIIutual-ftux equivalent
('ircmit shown in Fig. 12-14 and not the sC'lf-induetance-mutunl-inductance
circuit of Fig. 12-9.

12-9 THE APPROXIMATE EQUIVALENT CIRCUIT AND THE


EXPERIMENTAL DETERMINATION OF MACHINE PARAMETERS
In a practical induction machine, a certain portion of the electric power is
absorbed in iron-core losses. These losses rcsult from hysteresis and
eddy-currcnt cffects in the ('orc material as a consequence of the periodic
magnetic fields existing in thc machine. As an engineering approxima­
tion thcsc ('orc-Ioss effects arc taken inlo Il,Ceollnt, hy placing an additional
resistalH'e in Uw equivalent rir('uil, as discussed in Sec. 12-8. This
resistalH'c, whi('h is designated RH • for hysteresis and eddy currents, is
placed in parallel with the magnetizing inductance aM". The parallel
combination of Rh+. and aM" is commonly termed the no-lQad braTl!=A
570 Principles of electromechanical.energy conversion

A sN'ond v('ry useful approximation is ohtained hy realizing that in a


pradi('almachinc the voltage drops arross R' and jwl' arc generally small
enough that t he volt age arl'OSS t he no-load braneh i8 very nearly thc­
posit iv(~- or neg;al ive-sequell('(' volt age. Thel'(~f()rc, relatively i nsignifi('un t.
errors will, ill lliost pnwt ieul (~IlSeS, resull frolll \lsi ng I he (.j r<~ui I shown in
Fig. 12-};'). This eireui t is t erllled t he approximate per-phase e'luil'aient
circuit for the indll('\ion lllll('hine. The following l'eaetanee quantities arc
used in Fig. 12-15:

Stator leakagp reaetanee x' = wl'


Hotor leakage rea(,t unee x' = wi'
:"IIagnetizing rNl,danec Xm wa.V"
I
The f'omponenl of slalor ('Urrellt gOIng into the Ilo-Ioad hranrh is ('om­
monly ('alled I he no-load current. Thercforp, the supers(,!,jpl n is used 10
designate these ('urrent eomponents. The name "no-load hraneh" stems
from I he faet t hat if I~ and 1'- were zero, then 110 power would he ('on­
verted and the ent ire st alor ('urrellt would go t hrollgh only t his parallel
hranrh. By using t he approximate equivalent eireuit we ('an see how
the panl.lllet<'fs of the induction nmehille ('an hc determincd by two rela­
tively ~illlple tests.

The running-light or no-load test

Hated voltage is applied to the stator of the ma(:hinc. With no loa<;,toll


the rotor the-maehine will ope~atc at a value of slip very ('loSclo-~zcro.
If halaneed voltagcs arc applied to the stator, the negutivc-seqll-~;~·e

R$ jx$ a 2R' ja2x'

.1.L
120.
V! jX". a
2R,I-S
S
Jf'

jx'
Fig. J2-J!j Approximate per-phase equivaJl'nt ('ireuit for the induction machine
referred to the stator.
The induction machine 571

Fig. 12-16 (al Per­


phase no-load equivalent
circuit. (bl Per-phase
blocked-rotor equivalent
circuit.

quantities are zero, and only the upper half of Fig. 12-15 needs to be con­
§idered. If ~r.; .. ""0, the output resistor R'(l - 8)/8 becomes very large.
Effectively, l~ is very nearly zel'O, and nil of the st ator current goes
through the no-load branch. The per-phase equivalent circuit therefore
reduces to Fig. 12-lUa.
By measuring t he phase voltage V., phase current I., and per-phase
input power p., t he elements of the no-load branch are computed as
follows. The input power under no-load conditions is mainly absorbed
by windage and fridion and by hysteresis and eddy-current losses; thus
we have
(12-113)
where p.,+! is the combined per-phase windage and friction loss and Ph+< is
the hysteresis and eddy-current loss per phase. l The sinor phase current
I. lags the sinor phase voltage by the angle (J, where

(J
Ph+<
= cos- l _ ,- (12-114)
l.I.
The driving-point admittance of t he no-load equivalent circuit IS given by

(12-115)

I For a simple technique for separating these losses see A. F. Puchstein, T. C. Lloyd,
and A. G. Conrad, "Alternating Current Machinery," 3d ed., chap. 22, John Wiley
.& Sons, Inc., New York, 1954.
The inc/uction machin. 573

No-load lest
Stator voltage (Iine-to-Iine) = 220 volts
No-load stator line current = 16.6 amp
Total power input = 964 watts
Windage and friction losses at synl'hronous speed = 450 watts
Rlocked-rotor test
Stator voltage (Iine-to-line) = 35 volts
Rtator line current = 61.2 amp
Total power input '" 1638 watts
D-c resistance (at rated temperature) oetween any two stator terminals =
0.126 ohm
With rated voltage applied to the stator the op!'n-circuit induced rotor voltage is
measured as 110 volts line to lin!'. Determine the per-phasc approximate equivalent
circuit for the machine. Also, finrl the actual resistance, self-inductance, and the
stator-rotor peak mutual inrluctance for t.he windings of the actual three-phase
machine.
Solution: The pcr-phaRc ('quival{'nt circ'uit for the no-load t{,.8t is shown in Fig.
12-1&. The per-phase power absoroed hy the core-loss resistor RAH is given by
P A+. = 1(964 450) = 171 watts
where we have used Eq. (12-11;». The power-factor angle for Fig. 12-1&, accordill'g
to (12-114), is
171 •
II - coS-I---...---~-- ... 90
(220/0)(16.6)

By using Eqs. (12-116) and (12-117), we have

(220/0)'
RAH ­ --171- '" 94.3 ohms
and for the magnetizing reactance

220/0
X .. - ",aM" = ---.-~--
16.6 sm 90
= 7.65 ohms
The per-phase equivalent circuit for the blocked-rotor test is shown in Fig.
12-16b. The power absorbed oy the circuit per phase is given by

1638
p. - --- = 546 watts
3

By using (l2-11S), the power-factor angle is found to be

6 - cos-' ----~~-- = 63.S"


(35/0) (61.2)

From Eq. (12-119) the tota.l resistance per phase is

R' + a'R' '"' (6~:)S = 0.146 ohm


572 Principle, of electromechanical-energy conversion

which reduces to
1 I. Ph+e
Rh+e = V. cos (J = (V.)2 (12-116)

.l.. = _1_ = ~sin (J (12-117)


X.. waM" V.
where X", = waM" is defined as the !.rlagnelizinJL!.~t;L!lc..e.. By using
Eqs. (12-116) and (12-117), the no-load--branch can be determined at
rated voltage by this simple no-load test.

The blocked-rofor fesf

If the ,otor of the machine is blocked or restrained from rotating, the slip
is equal to unity. Under blocked-rotor conditions, the output resistor in
the positive-sequence half of Fig. 12-15 is equal to zero. The stator and
rotor resistances and leakage reactances have a much lower impedance
than the no-load branch, so under blocked-rotor conditions practically
none of the stator current goes into the no-load branch. The positive­
sequence equivalent circuit thus reduces to Fig. 12-16b, where x' and x'
are the stator and rotor leakage reactances equal to wl' and wl', respec­
tively. By using a reduced phase voltage adjusted to obtain rated phase
current, we again record the phase voltage V~, phase current I., and per­
phase input power p.. The parameters are then obtained-as follows.
The angle by whiclii. lags V. is given by

(J = cos- 1 p. (12-118)
V.I.
From the driving-point impedance of Fig. 12-16b we have

R' + aIR' = V. cos


I.
(J = . p.
(1.)2
(12-119)

X' + a2x' = ~: sin (J (12-120)

By using a direct current, R' can easily be measured, and aIR' can be
obtained by using (12-119). The two leakage reactances x' and a 2x'
appear only in series in the approximate equivalent circuit, so we do not
have to know how the total leakage reactance is actually distributed__
between the stator and rotor.

Example J 2.2
A 30-hp 22()..volt four-pole three-phase 6O-cps 1725-rpm wye-connected wound-rotor
induction machine gives the following no-load and blocked-rotor test data:
The induction machine 515

for the two-phase model machine. Thus we have


L' = i' + aM" = O.3ll:l + (2 X 10.2) ~ 20.8 mit

Mor 102

L' = I' + a = (l.OllS + - 2'- = 1).2 mh


The actual per-phase inductances for the three-phase machine are therefore

l'~/IY= jV =i X 20.S = 1:{.5 mh

l'~/IY= jU = j X 5.2 = a.5 mh

M::/Iy = jM" = i X 10.2 = 6.1'! mh

The coefficient of coupling hetw(len the stator and rotor when the windings are coaxial
is defined a8
M" 6.8
k - _ r-' - _I - .. 1111.0 per cent
v UTI v 13.5 X a.5
and we can scc that the stator and rotor circuits are very closely coupled.

12-10 THE CIRCLE DIAGRAM FOR THE INDUCTION MACHINE


Figure 12-1.') shows t.he complete per-phase approximate equivalent circuit
for the induction ma('hine. The stator excitations V~/v'2 and V':..Jv'2
are obtained from the two-phase excitations by using the symmetrical­
component. t.ransformation defined by Eq. (12-39). If the actual induc­
tion machine of intefCst is a three-phase machine, then an equivalent
two-phase set of variables is obtained before using t.he symmetrical-com­
ponent transformation. Very often the lllaehine is operated from an
infinite bus, so the positive- and negative-sequence stator voltage sinors
are fixed quantities. Consequently, t.he steady-state electrical response
can be obtained by' solving for the stat,or and rotor sinor currents in the
equivalent circuit of Fig. 12-1.'), Bahmred excitations on the stator of the
induction machine result in either the positive- or negative-sequence quan­
tities being zero, depending on whether the resulting uniform magnetic
field revolves in t.he negative- or positive-angular direction, respectively.
Therefore, for the balanced case, only half the circuit of Fig, 12-1.') needs to
be considered, silwe the other half will have zero excitation.
A very convenient graphical technique for solving eit.her half of the
eircuit of Fig. 12-1!} ('IHI he used. Let us begin this development, by con­
sidering the simple H-L cireuit shown in Fig. 12-18. The sinusoidal
jX

v R
Fig. 12-11'1 An R-I, circuit in th!' sinusoidal steady
state.
574 Principle, of electromechanical-energy conver5ion

)0.148 0.083 1'0.148

0.083 '~S

Fig. 12-Ji' Approximate ppr-phase equivalent ~ircuit r!'ferred to the sta.tor for
Exampln 12-2.

For a wye-('onn!'<'ted stator, the stator resistance per phase is simply

N'I= !!:;26 = n.063 ohm I

Thus the rotor resistance referred to the stator is


aiR" = 0.146 - 0.063 = 0.083 ohm
Similarly, hy use of Eq. (12-120) the total leakage reactance per phase is
3f)/V:~
x' + a'x' = ---
61.2
sin 63.R = 0.2!!6 ohm

As an approximation, let us divide the total leakage reactance equally between the
stator and rotor; thus
x' = 0.148 ohm a'x' = 1l.148 ohm
The approxima.te per-phase balanced excitation equivalent circuit is shown in Fig.
12-17, where all resistances and reactances are in ohms.
Assuming that the wound rotor has the same number of poles as the stator, the
transformation ratio is very nearly given by the ratio of the applied stator voltage to
the open-circuit rotor voltage; thus
220
a = ~.~ = 2.0
IlO
Since the excitation frequency is OJ = 377 rad/sec, the stator-rotor peak mutual
inductance for the two-phase model is

,\for = X.. = _~~ = 10.2 mh


OJa 377 X 2

The stator and rotor leakage inductances for the two-phase model machine are
respectively given by

l' = ~ "" Q:148 = 0.393 mh


OJ 377
a'x' 0148
l' = ;ai = 3ff-x 4 = 0.098 mh

From Eqs. (12-108) and (12-109) we can lICe how to convert the mutual inductanee
/If'" and the leakage inductances t, and I' into the stator and rotor self-inductances
576 Principles of electromechanical-energy conversion

steady-state current is to be found as the resistance R is varied from ­ 0()


to + o(). Consider the negative-resistance values as being at least mathe­
matieally definable. The sinor eurrent is given by
V
1= R +jX (12-121)

A simple graphical technique for solving this equation is shown in Fig.


12-19. The sinor voltage V is taken as reference and is drawn to a con­
venient voltage scale in the zero-angle direction, horizontal and to the
right. A circle of diameter l' /X is drawn as shown on the negative
imaginary axis to a eonvenient amperage seale. For each value of the
resistanee R from minus to plus infinity, the tip of the sinor I must be on
this circle. Therefore, values of R are shown as a parameter with arrows
indieati\lg the direction of in(~reasing resistance. Once the sinor I is
drawn, the voltage across the resistance RI in phase with the current and
the voltage across the reactaneejXI leading the current by 7r/2 rad can be
drawn, vectorially summing to give V as shown.
The angle e by which I Jags V is the power-factor angle of the circuit.
The average power input is given by V I cos e. For all values of Riess
than zero, elies between -7r /2 and -11" rad, thus making V I cos enegative.
In this region of operation energy flow is out of the circuit and energy is
being absorbed by the sinusoidal voltage source. Since the power factor
is usually taken as a positive quantity, the sense of I is conventionally
reversed, thus indicating a leading-power-factor condition throughout the

Increosing

R=-X

Fig. 12-19 A sinor diagram for an R-I, circuit &II R is varied from -.. to + GO.
The indtx:tion machine 577

lmoginorv
axis

v! Reol
-J2 oxis
J

Fig. 12-20 Positive-sequence pt>r-phase eirel!' diagram for t,he indu!'tion machine.

negative-resistance l·egioll. Noti<'<l t hat. the eireuit. absorbs power only at.
a lagging power factor and supplies power only at a leading power factor.
The steady-state approximat.e per-phase equivalent eireuit. for the
induction maehine, shown in Fig. 12-1!i, can he analyzed by a similar
procedure. The sinor cirde diagram of Fig. 12-20 represents the positive
sequence or upper half of the circuit of Fig. 12-11), The positive-sequence
stator voltage is taken as referenee, and it is drawn to a convenient voltage
scale as sinor OJ. The no-load sinor current. I~, drawn to a convenient
current scale, is shown as OA, lagging V~/V2 by angle AOJ. The com­
ponent OK, in phase with V~/V2, is the current through Rh+.; the com­
ponent 01... at 1f/2 rad behind V~/V2 is the eurrent through jX.. in Fig.
12-15. The remaining portion of the cir<,uit in Fig. 12-15 is similar to the
example circuit shown in Fig. 12-18.
As the slip S varies from minus to plus infinity, the total resistance
will take on all values from minus to plus infinit.y. Thus, the tip of the
positive-sequence rotor current I~ must traverse a circle with t.he slip as a
parameter. The diameter of this drde equals V~/V2 (x' + a'xr) and is
578 Principles of electromechonicol-energy conversion

given hy AN. Since the positive-sequence stator current is given hy

(12-122)

moving the circle to the tip of I~ allows the addition of currents to be made
in a convenient fashion. Point B represents a particular value of slip
between zero and positive unity. Sinor -I~/v2 a is given by AB, and
the stator current I~/V2 is given by OB. The power-factor angle is
BOJ and is lagging, since I~/V2lags VVV2.
When the slip equals zero, operation is at point A. Values of slip
between 0 and + 1 result in operation on the sector ABH of the circle. For
values of slip hetween 1 and 00 operation takes place between Hand M.
Thus, nil positive values of slip are Joeated 011 t.he sedor ABHM. All
negative values for the slip are located Oil the remainder of the circle,
eoverillg the sedor A P!If going frolll 0 to - 00. Once the slip has been
determined, the solution for the equivalent-cireuit currents is a simple
mat ter of using the circle diagram.
A good deal more information than just t.he current solutions is
available from the circle diagram of Fig. 12-20. The component of the
no-load current which is in phase with V~/V2 is given by OK. When
this component is multiplied hy the magnitude of the stator voltage
1!~/v2, the product equals the watts dissipated in the positive-sequence
resistor Rh+< in Fig. 12-1.1. We could, therefore, set up a wattage scale
which, in effect, would perform the multiplication by the constant magni­
tude V~/V2. Applying this scale to the in-phase components (i.e., in
phase with the voltage V~/V2) would yield various powers directly.
As an example, suppose the rotor is blocked or restrained from turn­
ing. Therefore, operation is at point H, where S 1.0. The sinor OH
represents the stator current IVV2. The in-phase component of OH,
given by FH, is proportional to the per-phase power supplied to the
positive-sequence portion of the equivalent eircuit. ~\feasuring the length
FH with our wattage ruler would give this per-phase power directly in
watts. Since w' = 0, none of this power can be coming out of the mechani­
cal port. Also, FG = OK is the power absorbed by Rh+<. Therefore GH
represents the per-phase power absorbed by R' and a 2R'. The power
dissipated in R' and a 2R' is the per-phase positive-sequence contribution
to the stator and rotor copper losses, respectively. Thus, GH is the
positive-sequence contribution to the per-phase losses in the stator and
rotor" windings due to ohmic heating. If point Q is so positioned on GH
that
GQ R' (12-123)
Gil = R' + a2R'
The induction machine 579

then GQ represents the stator copper loss and QH would be the rotor
copper 1088. This statement is readily proved, since

GH ex: ( vi;l' a)2 (R. + a Rr)


2
(12-124)

Then

GQ ex: ( vi;l' )2 R­
a (12-12.'j)

by use of Eq. (12-123).


When operat.ion is at the value of slip given by point H, the in-phase
component of the stator eurrent CH reprcst'nts the pt'r-phase powt'r into
the positive-seqllenee (·ircuit. Tht' portion CD is t.he power tnken by
NH ,. The remainder of CB, namely, DH, must be proportioned into
three eomponents. These are per-phase positive-sequen('e eontributions
to (1) stator {'opper losses, (2) rotor copper losses, (:J) converted power
Pt"" The component DE will be shown to make up portions 1 and 2,
thereby leaving EH as the positive-sequell!'e ('onverted power P:",.
Rin('e triangle ADE is similar to triangle AGH, we ('an write
DE AD
(12-126)
GH = AG
The length AG is the component. of the rotor current at ."./2 rad to the
applied posit ive-sequence voltage when S = 1.0. Therefore,

(12-127)

where
Zl = [(R- + a2Rr) + {x' + a 2.rr)2Jh
2

Similarly, the lengt.h AD is the out-of-phase ('omponent of the rotor cur­


rent, which can be shown to he

(12-128)

with Sn being the value of slip at point. H. Now t.he length GH is


proportional to the stator plus rotor copper losses with S = 1.0; therefore

(12-129)
580 Principles of electromechanical-energy conversion

Substituting relations O:l-1:l7) to (12-12\) illlo Eq. (12-12(\) ~IVP8 the


lengt h DE as

(l2-1:l0)

Thus, the ]Pllgth /)l~' is proportional to the stator pIllS rotor eoppcr losscs.
II owpvpr, si II!'€'

DY DE
(12-1:H)
flQ (/H
we ('un writ€', using Eq. (12-12:;),
I
/)}' OQ U'
(12-1:~2)
f) E (; Jl = /(7+" oHF
Ttwfefore, D Y rcprespnts 1he posit ive-sequcrwe per-phase contrihut ion
1081 utor I'opppr 10001S(,S and Y E f(>preRenls thc <,ontrihul ion to rotor copper
lOHHl's.
For t hc opcrat ion at I hc valuc of slip represented hy point B on the
eire-Ie diagram of Fig. 12-20, the foIlowing positive-sequenee quantities
{'an he summarized. All lengths 011 the diagram are to he measured with
a suitahle scale calihratpd in volts, ampcres, or watts depending on what
the lellgt h represents.

OJ = stator phase voltagc


= no-load current
OA
AB
= negative of the positive-sequence rotor phase current
referred to the st.ator (-I~/v'2 a)
OB = slator phase current
Angle BOJ sl ator power-fad Of angle
CB per-phase eledric power input to positive-sequence
circuit
CD = power in Rh+<J positive-sequence per-phase component
of the core losses
DY = stator copper loss, positive-sequence per-phase con­
tribution
Y E = rotor copper loss, positive-sequence per-phase contri­
hution
BB positive-sequenee per-phase component of the COIl­
verted power
('BlOB = power fa(~lor of the positive-sequenee circuit.
YEIl'B = slip at point B(SB)
The induction machine 581

The laRi ratio ('all I)(! prov('d quite simply, Rill('P

n~' 0: rotor ('opper loss = U:/V2FJlr


YB 0: (rotor ('opper loss + convcrted power) = U~/V2)2(RT/SR)

Th!' quantity Y B is often terllled til(' power acro,ss the air gap.
Let us consid('1' one other operatinl!; eondition on the diagram of Fig.
12-20. Suppose the machine is opcmtinl!; at a negative slip such as at
point P. Rememh('r I hat a Il('gativ(' slip occurs whcn CUT > CU" or the
rotor is turni ng at a gl'CI\1 er anl!;ular velori t y than t he synchronous speed
for the presrribed cxdtation. The positive-sequence stator current. is
now given hy 01". Angle PO.! is t he positive-sequence powcr-fact.or
an/l:le, and for this value of slip it is grel\ter than1f'/2 rad. Thus, energy
flow is out of the posilive-scqllcnee st alor t.erminals. The magnitude
of Ihis genernt('d output po\\'('r is giv(,H hy I'll, the eOlllponent of the
stator ('urrellt in phase with the stator voltage. An induction machine
/l:enemt i ng eled ric power is known as an induction generator.
The power fador at which \.hi8 power is heinl!; delivered ('un he deter­
mill('d hy taking I he lllagnit.\Ide of IIIC! eosine of angle POJ. However,
since the power factor is llsually a positive (!llantity, t.he sense of either
I~/V2 or V~/V2 !lIlIst he reversed. In either ease, -I~/V2 leads
V~/V2 or I~/V2 I('ads -V~/V2 hy angle POX. The cosine of angle
POX cquals Illinus the ('osinc of anglc POJ and represents the leading
power fador at whi('h power PH is delivered to the infinite bus. Froll1
the diagralll of Fig. 12-20, we see 1hal POWN ('an he delivered only at a
leading power faetor.
The distance NT represents t he power absorbed by RH •. Equation
(12-130) shows, for operat.ion nt point. B 011 the circle, Ihat the distance
measured 011 a line p'amllcl to t hc applied voltage V~/V2 from t.he diam­
eter A N to t hc 8 = 1.0 line A Il is equal to the st.ator-plus-rotor eopper
losscs. The proof leading 10 thc result of Eq. (12-1:m) is entirely general
for allY ot.her opemting poin1 011 I he eirde. Thus, for the induction
generator ('ase we arc 1I0W ('ollsidcring, with operation at point P, the
dist ance 'I'll', measured hy the wal t age scale, gives t.he stator-plus-rotor
('opper losscs. By again using Eq. (12-12::\), the following relation can be
developed:
TlI all R' (12-133)
1'W = il/J iV+ a2 f{r
Thereby, the copper losses ar.:' divid!'d with Tll being the stator copper
loss and F I\' t he rotor loss.
A get\('ml rule in lIsing 1he drde diagram ean be lIlade corwerning the
S = 1.0 line, given hy A II. The dis1 :1nc'(' along a Jille parallel to V~/V'2
582 Principles of electromechanical-energy conversion

from All to any operating point on the cirde represents t,he posiHve­
sequence converted power per phase P:",. For values of slip between
zero and unity, corresponding to rotor angular velocities from synchrollous
speed to standstill, the circle lies to the right of the line All and the con­
v"rteu power iH froll! deetrieal to Illechanir~al. Vor all other vl~lues of
slip the eirele lies to t he left of line A H and the converted power is from
mechanical to elce! rieal. For operat ing conditions 011 the portion of the
circle below the horizontal line FH the ('opper losses can still be divided
between the stator and rotor by extending the lines AQ and All. The
intersection of these extended lines with a horizontal line drawn through
the operating slip on the eirde gives these eopper losses as distances
measured with the wattage seale.
~Il currents derived by using the circle diagram of Fig. 12-20 are for
the positive-sequence portion of the equivalent eircuit of Fig. 12-15.
The power quantit ies taken from this circle diagram represent positive­
sequCIH'e eontributions to the per-phase quantity. For example, the

Imaginary
axis

Reol
oxis
J'

Direction of
increa~inq slip

Fig. 12-21 Negative-sequence per-phase ('ircle diagram for the induction machine.
The induction machine 583

distance DY is Ihe positive-sequenee eontribut,ion to the per-phase


stator ('opper loss with the induetion machine operating at the slip given
by point B. In effect, DY is the first term on the right-hand side of Eq.
(12-91). The total per-phase stator copper loss ean be found by adding
the ncgntiv<'-HcqIlCIH'e (:ontrihutioll al this SIlIllC opcrating slip.
The negative-sequence pOl'tion of the equivalent circuit has the same
elements as the positive-sequelwe portion with the simple substitution of
2 - S for S. Therefore, a ('ire-Ie diagram identkal with the positive-.
sequenee diagram ('an be drawn for the negative-sequeIH'e (~ircuit with the
substitution of 2 - S for every S appearing in Fig. 12-20. Figure 12-21
shows the results of this substitution, with primes used to designate all
lettered points eOJ'responding to Fig. 12-20. Thus the point Q' is so
positioned th!tt
G'Q' GQ R'
(12-134)
G'H' = Gli = k' + ii.in-"
corresponding to Eq. (12-123). In general, V'.JY2 and VVy2 are
different in magnitude, making the two diagrams different in size,
assuming the same voltage, eurrent, and wattage scales are to be used.
Operation at point 8 in the positive-sequence diagram of Fig, 12-20
would correspond to operation at point B' in Fig. 12-21. Therefore, the
following negative-sequence quantities can be listed:

0' J' = stator phase voltage


0'A' = no-load current
A' 8' = negative of the negal.ive-sequenee rotor phase current
referred to the stator (- I~I y2 a)
0'8' = stator phase current
C' 8' = per-phase electric power input to the negative-sequence
circuit
C'D' = power in Rh+o, negative-sequence per-phase component
of the core losses
D' Y' = stator copper loss, negative-sequence per-phase con­
tribution
Y' E' = rotor copper loss, negative-sequence per-phase contri­
bution
B'E' = negative-sequence per-phase con tribution to the con­
verted power (frolll mechanical to electrical)
C'B' 10' B' = power factor of the negative-sequence circuit
Y' E'I Y'B' = 2 - S, where S is the slip at point 8

Notice that all negative-sequenrc power is consumed in R H • and


copper losses. A port.ion of these losses, C'B', comes from the electrical
584 Principles of electromechonical-energy conversion

stator t.erminals in the equivalcnt (·in·uit. The remainder of these losses,


B' E', eOllles from trw meeh:lllieal port.
Total per-phase powers for operation at a slip equal to SJj ean now
he eomputed hy adding the ('omponents from the two diagrams_ Thus,
DY + D'Y' is the total per-phase stator eopper loss for the machine.
The tot al per-phase ('onverted power from electrical to mechanical is given
by EB - E'B'. I\otiee that the negative-sequenee eircuit is capturing
some of the meehanical power converted by the positive-sequenee circuit.
For balaneed stator voltages either V:h!2 or V'-/y2 will be zero and
only one of the Iwo dia!1;rams needs to be used. Rather than a(·tually
const ruet two ('irdc diagrams, I he voltage, (mrrent, and wattage rulers
can be multiplied hy a fador '" V~ and used on the positive-sequenee
diagralll to give negative-sequpnee rcsults. Of <'OllrRe, rcmember the
2 - k for S suhstitution when determining the negative-sequence results.

12-11 STEADY-STATE PERFORMANCE CHARACTERISTICS


In Ree. 12-;') a sct of steady-slate eonstrainis are imposed on the indudion
machine. The two constraints, involving I he shorl-<~ircuiting of the
rotor ports, have been incorporated in the equations leading to the equiv­
alent eireuit of the ma<,hine. The voltage constraints on the stator ports
of the equivalent two-phase machine are given by

I'~ = y2 Vd eos wi I!~ = y2 V~ sin wi (12-135)

In sinor notation these constraints become

Vd = jVd v; = V; (12-136)

By using the SYI1l111ct rieal-('omponent transformation introdu('ed in See.


12-4, the positive- and negalive-sequence voltages are derived to be

V·- =L
y2 ('"d - V')
If
(12-137)

The per-phase torque of ele<:trical origin is found from the relationship

T <jJ = - -~P.~~ (12-138)


• (1 - 8)w

where the per-phase developed power P •• is obtained from the approxi­


mate equivalent eireuit of Fig_ 12-1;') as being the power dissipated in the
resistors a2Rr(1 - 8)/8 and -a 2Rr(l - 8)/(2 - 8). Thus the devel­
The induction machine 585

oped power per phase is

Pt~ = W(l ,;8) (JzY - ~!r~i>;J (~~Y (12-139)

By substituting Eq. (12-139) in Eq. (12-138), the generated electrical


torque per phase is found to be

T,~ = - Sw 1'2)2 +
nRr ( J' nRr
(2 - S)w
( J' )'
y72 (12-140)

as originally given by Eq. (12-8(j). If the ('Iectrical torque T.. is positive,


the torque is generated ill the negat ive angular direction; if T.~ is negat.ive,
the torque is generated in the positive angular direct.ion.
From the approximate equivalent circuit, shown in Fig. 12-15, the
positive- and negative-sequence rotor currents are found to be
r
--~ = ~.----~-----~--~~.~~-----...~..--­
y2 a (R' + a R'IS) + j(;r,. + a 2x')
2
(12-141)
I':. -V'jy"2
~2~ = [n;-+-;i2Il;/(2"':S")]+j(:r' + a 2.x;r)
Upon substitution of the square of the magnitudes of these positive- and
negative-sequence rotor currents, given by Eqs. (12-141), into Eq.
(12-140), the per-phase average electrical torque becomes
n(a2RrI8)(V~Y

T.~ =
2w[ (R- + a R' I S)
2 --:+---T(x'-c:-·-+:---a7"2x---")C;:;2]
2

n[a 2Rr1(2 - S)l( V~'y


+ 2wTflF+a2R;/(2-'=-S)]2 n~' + -<1"2./:;)21 (12-142)
By using Eqs. (12-1~n), the llIagnitudes of the positive- and negative­
sequence slntOl' voltages are substituted into Eq. (12-142), which gives
the eledri('ully gellerntcd torque per phase as

T.. = - T:. + T;. (12-143)


where

(12-144)

and

(12-145)

Equation (12-144), for T:•.


gives the average value of the per-phase
torque of electri('al origin attempting to drive the rotor in the positive­
586 Principles of electromechanical-energy conversion

allgular diredion. Similarly, Eq. (12-14;), for T-;~, represents the per­
phaf;e portion trying to drive the rotor in the negative-angular direction.
Let us begin our investigation into the chll.raeteristics of these gen­
erated torques by assuming a balanced stator excitation producing a
positively revolving uniform lIJagnetic field. Sueh all excitatioll corre­
sponds to the statement that
,..
q (12-146)
Substituting this balanced condition into Eqs. (12-144) and (12-145)
gives us
na 2 R'(V·)2jSw
(12-147)
(H'+ a 2H' ls) 2 + (x' +~x~')2
1J;~ = 0 (12-148)
Thus, for balaneed excitation in the positive direction an electrical
torque only in the positive direction results. A typical plot of Eq.
(12-147) is shown ill Fig. 12-22, where rt~ is plotted as a funetion of the
slip.
For values of slip between zero and unity the rotor angular velocity
is positive and less than the sYIH'hrollous speed w,. Since r:..
also is
positivr~ for these values of slip, the machine is operating as an induetion
lIlotor eonverting eleetriC' energy to meehanical energy, as indicated in
Fig. 12-22. If the slip is greater than unity, the rotor angular velocity
must be negative. The electrical torque, however, is still positive;
hence power is being supplied to the mechanical port. An examination of

- - - Generotor ---.....;.>-- Motor ---r- Broke-­


I
I

1.0
s

Fig. 12-22 Positive-sequence per-phase contrihution to the torque of electrical


origin as a function of slip.
The induc!ion machine 587

the circle diagram of Fig. 12-20 shows t.hat with V~ = 0 and t he slip
greater than unity, power is also being supplied to the eleetrieal ports.
Thus, with energy flowing into all ports, the maehine is operating asa
brake (~onverting both 1lIC1'luUlieal and elccl rical energy into hCltt ene~gy.
For negative values of slip the rotor angular velocity is positive and greater
than synehronous speed. With the electrieal t.orque negative, power is
again being supplied to the mechanical port. However, with the slip
being negative, energy is flowing out of the eleetrical ports and" the
machine is operating as an indu<:tioH generator. The circle diagram of
Fig. 12-20 shows that for values ~f slIp just siightly negative no power
will flow out of the machine terminals owing to absorption by the core­
loss resistor. Also, for negative slips beyond point Z in Fig. 12-20, no
power aetually flows out the electrieal ports. However, conversion is
still from mechanical to electrical form.
With the balaneed positive exeitation, proposed by Eq. (12-146), Q1~.
starting torque ('an easily he found silllply hy setting S = 1.0 in Eq.
(12-"147), which gives us

G f )~ 10 (12-149)

Since T;. is zero for the positive balan('ed excitation, Eq. (12-149)
represents the complete per-phase average starting torque generated by
the indudion ma(,hine.
The maxilllum torque the llIaehine ('an generate can be found by
first findiJ;ith~ "Slip ;t" ~hich this nHtXillllllll torque OC(~urs. This slip (,an
he dcic~;;~illed i)y setting the slope of the torque-slip ('urve equal to zero
and solving for the slip that satisfies the zero-slope ('oudition. Thus,
upon setting dT:./dS = 0, the slip at whieh maximum torque O('ClirS is
found to he

(12-150)

Equation (12-HiO) indi('ates two valu('s, one positive and the other
negative, for t he slip Itt whieh IllaxillHlIll torque is developed. These two
peaks are shown ill Fig. 12-22 at !"qual distauees from the torque axis.
Upon suhstituting the two values for the slip at maximum torque into
Rq. (12-147), we have as thE' vahl(,s for the maximulI\ torques

(12-151)
588 Principle, of e/edromec:hanical·energy conversion

Notice that the negative peak is larger than the positive peak, since the
denominator of the second expression is smaller than that of the first
expression.
Substituting 8 m ., from Eq. (12-1.'i0), into Eqs. (12-151) gives the two
torque peaks as

r:. (positive max)


(12-152)

Since Eqs. (12-1!i2) do not contain the rotor resistance HT, the peak torque
must be independent of this quantity. Equation (12-1.'>0) shows that the
faetor b,2HT/S.., must remain constant for fixed values of H', X', and xTi
therefore, varying the rotor resistance can only cause the slip at which the
maximum torque occurs to change. However, th~y:aluELQft!!~~rl!\x}nw~
torque remains constant. Figure 12-23 shows a set of torque-slip curves
for various values of rotor resistance. The magnitudes of the two peaks
are constant" but the slip at whi(~h they occur increases in magnitude as
the rotor resistance increases.
For eurve c the peak occurs at a slip of 1.0 or at starting. Thus, for
one partieular value of rotor resistance the machine generates maximum
torque at start. Selting 8 m,1.0 in Eq. (12-1.')0) gives this particular
rotor resistance as

for max torque at start (12-1.53)

iw'=o
i

a
b
c
s l

Fig. 12·2:! (lcm'rated "Iectrical torque for various values of the rotor rcsistance 1\8
a function of slip.
The indudion mcxhine 589

One advantage of the wound-rotor induction machine is the facility for


changing the rotor resistance by adding external resistors into the rotor
circuit.. When the machine is used as a motor, the total value of the per­
phase rotor resistance can be adjusted to the critical value given by Eq.
(12-1.')3) in order to have a large value of starting torque. Then, as the
machine accelerates, the rotor resistance can be reduced to have a running
characteristic which is more nearly at constant speed. In fact, if the
wound-rotor machine is operating wit.h a constant torque applied to its
mechanical port" some degree of speed adjustment can be obtained by
adjusting Rr. Consider, for example, the load torque TL in Fig. 12-23.
As the rotor resistance is redueed from curve d t.o curve a, the operating
condition moves from point 1 to point 2 and thereby decreases the slip or
increases the rotor angular velocity.
Suppose now that a balanced excitation is applied to the stator,
producing a uniform magnetic field revolving at synchronous frequency
w. in the negative-angular direction. Simply reversing two line connec­
tions on the three-phase machine or reversing the connections on either
of the two st ator ports of the two-phase machine would result in a nega­
tively revolving field; thus for example we might have

v; = - V: = V· (12-154)
Referring to Eqs. (12-144) and (12-141}), the electrically generated torque
per phase becomes

T:. = 0 (12-155)
T _ == _....~-~..._---_._,,_.--
na 2Rr(V.p/w(2
.....----~-.-~-
S)
(12-156)
.. [fl· + a 2 Rr/(2 - 8)]' + (x' + a 2;cr)2
.
Now the positive-sequence component. of t.he torque is zero and the nega­
tive-sequence component is, in general, nonzero. Remember that posi­
tive values of the negative-sequence torque are directed in the negative­
angular direction. Equation (12-1;)6) is identical with Eq. (12-147) for
the balanced positively revolving condit.ion if 2 - S is substituted for S.
By the definition of slip, given by (12-34),

w. - (-w r )
2 - S = ---~-- (12-157)
w.

Thus the examinat.ion of the negative case will give results identical with
those for the positively revolving case if t.he sense of the rotor angular
velocity wr is reversed. Equation (12-157) shows us that reversing the
assumed direetion of wr would require the substitution of 2 - S for S in
all of the previous rfsults. ConsequenUy, all conclusions for the balanced­
positive case carry over for the balanced-negative case.
590 'rinciplea of electromechanical·energy conyersion

Exomple 12-3
Thf' ~O-hp 220-volt four-polf' t,hree-phal!f' 6O-(:ps 1725-rpm wye-connected wound­
rotnr inrludion machine rliscuRiled in Exampl" 12-2 has th" following per-phase
"'luivlll"nt-rirruit pararnet"rs all r.,r"rr"rl to the Rtator:

x~ = 7.6.'> ohmll a'll' O.O1{3 ohm


R' H !14.a ohm" a'z' = 0.148 ohm
R' == 0.063 ohm a = 2.0
z' = 0.148 ohm
The rotor is short-cirruited and the stator is conne<'ted to an infinite hus at rated
voltage.
a. Calculate the starting inrush current and the starting power factor.
b. Determine the total starting torque.
t. Fihd the maximum torque the machine can generate and the speed at which
this torque o<'eurs.
d. Cal<'ulate the additional rotor resistance required to hlwe the maximum
torque availahle for starting the machinf'.
Solution: a. At the instant of starting, the slip S = 1.0 in the equivalent circuit
of Fig. 12-17. The rotor phase current referred to the stator is calculated as

which reduces to

(220/V%LQ

~Di6:f"-f-(j.lls:$f+ R6~I~fif+-o~f4S)

The no-load hranch current is given hy

(220/'\1'3)/0 (22Ul'\.i:i)/U

I"+ 94.3 ,- + -77.65 .~ = US - j16.6

Therefore, the total stator inrush current in each phase, which iM also the line current
in a wye connection, is simply the 8um of these two currents, as is seen from Fig.
12-17. Thus the inrush current at start is

I~ = I';. - I. = 172 j361 = 400{ -64.fi,


The minus sip;n hefore I~ results from the assumed direction of the rotor sequence
current. Thus the starting inrush <'urrent is 400 amp at a power factor of C08 64.5 =
0.431 lap;ginp;.
b. The torque of electrical orip;in per phase generated hy the machine at any
slip is the power clis.'lipated in the output resistance of the equivalent circuit divided
hy the rotor speed ",' = (1 - S)",/n. Thus the generated torque at start, where
S I, is

n(l;»a'R'
T~
Sw
na'/l'( l~)' 2 X U.Ol~:l X (;J.'!5)'
05.2 newton-Ill /phMC
Sw 1 X 377
The induction machine 591

The total starting torque for three phases is


T. (at start) = 3 X T:. = 196 newton-m
since 7';;. = 0 for balanced poliitive excitation.
c. From Eq. (12-1 ;,cl) the slip at which maximum torque O('curs is

S O.OR:J 0275

.., = [(O.(63)' + (0.148 + 0.14R)'I\~ = . •

The speed corresponding to this value of slip is

"':., = (l - S ..,) ~ = (I - O.27!\).i!.F = la7 rad/sec = law rpm


n
The value of the maximum torque ean he found hy simply subatituting parameter
values into th(' first of Eqs. (12-152). However, Il!t us determine T"",. by first
finding the positive-sequence rotor current when S = S .., = 0.275. From the cir­
cuit shown in Fig. 12-17 we have

-I. =< (ifti6:i +


(2'..W/V:J)/O
fl.08:JjO.:i75)T}(0:14R

The torque generated per phase is ther!'fore given hy


-Fo.l48) = 270L=~9.0

'J'+ na·/l'(l.) , 2 X O.O!~:J X (270), 117 newton-m/phase


.. = S",·· =1i.m-X37-7- =

or the total maximum torque is simply


T. = 3 X 7':' = a!)l newton-m

d. Using Eq. (12-IS:J) the maximum torque of a51 newton-m will occur at start
provided

a'R' = [(O.06:J)' + (0.148 + 0.14R)'II' = 0.302

Thus the total rotor r<"l'istance in each rotor phase must he

U' = o.;~ = n.0755 ohm for T max at start

This means that we must add in each phase a resistance equal to


00R3
R. dd = 0.0755 - T = 0.0548 ohm
If three 0.0548-Qhm resistances are joined in a wye connection and connected to the
three slip-ring rotor terminals, T .... will occur when the machine is started.

12·12 TRANSIENT PERFORMANCE CHARACTERISTICS


In the pre(~edillg section we have considered the steady-state characteris­
t.ies of the polyph,ll.se indUdion IlIlU'hine with balanccd and unbalanced
excitl\tions. III this section we shall examine the machine's transient
592 Principles of e/edromechanicol·energy conversion

response for incremental motion around a normal operating condition.


We then present a graphical technique to show how nonincremental
transient operation can be predicted. The transient analysis can be
greatly simplified by assuming that the electrical transients are very much
faster than the mechanical transients. Thus, the electrical-port variables
have always reached their steady-state values before any mechanical
changes can take place. This same assumption was used in Sec. 11-10
for the synchronous machine. The sillusoidal steady-state equivalent
circuit shown in Fig. 12-15 can therefore be used to determine currents,
powers, and torques as functions of time. If the speed of the machine is
changing as a function of time, then the slip 8 is also changing with time.
The output resistors in Fig. 12-15 are therefore funetiOlls of time, sinee
these resistors are dependent on the slip. At each time instant the circuit
paramet!rs are known and t,he magnitude of the eurrents can be calculated.
Since the eleetrical portion of the induction maehine is always in the
steady state, the torque of eledrical origin, given by Eq. (12-143). is
exerted 011 the mechanical portion of the machine during the transient
interval. Under balanced excitation the positive-sequence per-phase
generated torque is given by (12-147). When the numerator and
denominator are multiplied by 8 2 , the total per-phase torque of electrical
origin becomes
na 2 R r 8(V')2/w
1'.. = - T;. + T;;. = - ~~R'+ a2Rr)1 + 82(x' + a2xr)2 (12-158)

with T;. = O. The normal operating range for the induction machine is
usually at small values of slip, or near synchronous speed. On examining
Fig. 12-22, we can see that for small positive slip values the machine is
very nearly a (~onstant-speed device over a large range of generated torque.
For small values of slip Eq. (12-l!iS) is given approximately by

T .. = n(Vep
wa 2 Rr
S = -k8 for small slip values (12-159)

where all terms in t.he denominator involving 8 have been neglected.


Thus, the torque of el('drical origin is linearly related to the slip in the
normal operating region.
The mechanical-port equilibrium equation is given in general by
(Jp + D)w' + "T•• = Tr (12-160)
where" is the number of phases on the machine, J is the total shaft
inertia, D is the total windage eoefficient, and Tr is the externally applied
shaft torque. The speed in terms of the slip, from Eq. (12-35), is

wr = (1 - 8) w (12-161 )
n
The induction machine 593

and so
W
pw r = - -n p8 (12-162)

By substituting Eqs. (12-159), (12-161), and (12-162) into Eq. (12-160),


we have

- ~ .IpS
n
+ ~n D(1 - S) - IlkS = Tr (12-163)

Equation (12-163) is the me(~hanic!l,l-equilibrium equation in terms of the


slip, valid in the neighborhood of the normal small-slip operating region.
The inl'rementnl operation of lhe Iluwhine call be examined very
eonveniently by assuming each variable to have a <:ol1stant steady-state
portion plus a slllall lillie-dependent portion. Thus, assuming a passive
load torque which opposes posilive speed leads to
Tr(t) = - T~ - TW) (12-164)

The slip can also be expanded in a similar form, thus giving us


(12-165)
By sUbstituting (12-164) and (12-165) into Eq. (12-163), we obtain the
steady-state equilibrium equation defining the operating point and an
incremental differential equation defining the mot,ion about the operating
point. After some slight manipulation, we have

~
n D(1 - So) + To = "kSo (12-166)

[;i~/P + (;i 1) + Ilk)] 8 1 = TI(t) (12-167)

The steady-state equation (12-166) simply equates the steady-state


windage torque plus the load torque to the steady-state torque of elec­
trical origin. Equation (12-167) is a first-order linear differential equa­
tion describing how the slip changes about the operating-point slip So.
Suppose that the load torque were suddenly increased by an amount T~
at time t = 0; thus T~(t) = Tiu(t). From Eq. (12-167), the increment
in the slip is found to be

(12-168)

where
.I
l' = -~~-- seconds (12-169)
D + nllk/w
594 Principles of e/eclromechonical·energy conversion

..

(a)

I~
V2
.',­ - -,--r-­
-n-rr-n-TTII£'

-'­ -­
(bl
Fill:. 12-24 (al Slip as a function of time following an incremental step increase in
the load torque on an induction machine. (b) The stator phase current following the
step increase in load torque.

Figure 12~24a shows the tDtal slip as a function of time. For l < 0
the slip 8 = 8 0 • After the increased load torque is applied the slip
exponentially increases to a slightly greater value. From the positive­
sequence portion of the equivalent circuit shown in Fig. 12-15, we can see
that as 8 increases, the output resistor a 2R'(1 - 8)/8 decreases in value.
The value of the stator~phase current IVv'2 will therefore increase
exponentially, as shown in Fig. 12~24b. Since the slip is a known function
of time, the magnitude of the currcnts in the equivalent circuit can be
calculated at each time instant. The frequency of the current in Fig.
12·24b has been rcdu('cd for clarity ill drawing the figure. Actually, the
current would go through many more eydes during the transient time
intervaL
Equation (12·11)9) is a linear approximation to the tDrque-slip char­
acteristic valid in the normal operating region near zero slip. If the
machine is operating in a different incremental region (i.e., where the slip
is not close to zero), then the constant kin (12-11)9) can be suitably altered
to correspond to the slope of the torque-slip {'urve around the new oper­
The incluction machine 595

ating point, and linearized incremental transient motion can still be


studied.
Onc of the most important transient problems, not involving incre­
ment.al speed fllH~tuatiolls, is ('on('erned with starLing the machine.
Again assuming that the eledrical transients do not influence the mechan­
ical transients, the steady-state torque characteristic can be used.
Neglecting windage, Eq. (12-160) can be put into the form
dw'
J di = T~(w') - TL (w') = ~ T(w') (12-170)

where T; is the total generated torque tending to drive the machine in the
direction of positive speed and Ti. is a passive load torque. The quantity
~ T is therefore the difference between the generated torque and the torque
absorbed by the load. Figure 12-25a shows the generated tcrque plotted

--('
Load
torque
(T[)

o
(a)

w'
(b)

w'
wr(<o) -----­ ?-'---­
r
W,

I<'ig. 12-25 (a) Torque-speed curve for


I\n induction motor. (0) Inverse of the
torque availaole for accelerl\tion as 1\
function of speed. (c) Machine speed a8 t,
a function of time. ((;)
596 Principles of electromechanical-energy conversion

as a fun('tioll of t he machine speed. The T~(w') curve is obtained by


• taking th{' torque-slip curve in the motoring region of Fig. 12-22 and
inverting the slip axis. Thus when S = 1.0, the speed w' = 0; and when
the slip S 0, we have w' w.. The load curve T'L(w') is also plotted
in Fig. 12-2;'")a. If t he windage torque is 1I0t actually negligible, it could
be induded as part of the load torque. The difference between the two
('urves is t:.T(w'). Tbe steady-state speed w'( 00) is the intersection of the
two curves.
Equation (12-170) call be lIIanipulated into the integral form

(12-t71)

Figure 1212.% shows a plot of the quantity ./ / t:.T(w') as a function of w'.


The area uuder the ('urve in Fig. 12-251>, up to any plutieular speed w~, is
equal to the tillle tl required to reach that speed. Thus the integral on
the right Ride of Eq. (12-171) ean be evaluated graphically, and the speed
vs. time ('urve showll ill Fig. 12-2;}c can be obtained. If the speed is
known as a function of tillie, the steady~state equivalent (~ir<~uit ('an be
used to obtain the lI1agllil ude of the sinor currents at each Hme instant,.

12·13 THE TWO-PHASE SERVOMOTOR


One of the 1lI0flt ('ollllllonly IIsed ele(~t rOllle(~halli(,1l1 posi 1iOllillg d(~vi('es fOl'
low-power applicatiOlls ,is the two-phase indu('t ion serVOlllotor shown
s('hel1lat ically in Fig. 12-26. The 1I0nsalien t rot or of the l\1a(~hine is
usually ('Ollstru('ted in the forlll of a high-resist arwc squirrel cage as
pictured in Fig. 12-1. Therefore, 110 eled ri('al ('0 III Ie !'\ lOllS to t he rotor
arc required, and t hI' lIla('hlne is a very ruggcd, dCpi~lIdahl(' devi('c requir­
ing a minimulll of maintellau('e. Th(' nOllsalicnt stator has two identical
phase windings plal'cd Tr /2 clel'l ril'al radians apart. Thus il elwh of the
stator windings establishes an air-gap field haviug n pole pairs, the axes
of the two stator wiudings are separated by Tr/2n radians.

FiR. 12-26 i'chl'nlntic rlin!l:ram of thfl two-phase


Sflrvomotor.
The induction machine 597

The stator winding on lhe dired or horizontal axis in Fig. 12-26 is


designated as the reference phase. The reference phase is excited by a
fixed sinusoidal voltage source

I'R(I) = v'2 F Il cos wt (12-172)

The stator winding on the quadrature or vertical axis is known as the


control phase. The voltage exritatiolJ for the control phase is displaced
from the referell('e-phase t"xcilalion hy 7r/2 rad. Thus the control-phase
terminal voltage ('an he given by

vc(t) = v'2 V c sin wt (12-173)

If r e = V Il. I hen we have a halanced hvo-phase excitation resulting in a


uniform positively rolaling ('ounler('\oekwise air-gap field. The servo­
motor' develops !l steady-slate torque jn l.he posHive-angular direction.
If Ve = - If H. then the sallle sl eady-sl ate torque is generated in the
negative dockwise angular direel ion. Thus, by controlling the magnitude
of "e, I he direction and Illagnil tide of I he servomotor torque can be con­
trolled; but the phase of !Ie is always held at, 7r/2 rad, eit,her leading or
lagging the reference-phase voltage I'/{.
The two-phase a-b induction maehine, shown in Fig. 12-6, can be
used as a model for t.he two-phnse s('rvomotor. The d-q equivalent
model, showll ill Fig. 12-7, can now he employed, ltlld lhe ent.ire steady­
state analysis of t.he preceding sections of this ('hapier applies to the
servolllolor. The PX(·it at,ions on I he stalor willdin\s of t.he a-b model are
identical with those on the stator of the d-q model; thus we have

II~ = I'~ = y'2 V c si II wt (12-174)

These voltage constraints are identical with Eqs. (12-135). Following


the sallle development as used in Sec. 12-11, the per-phase torque of
electrical origin is given by Eq. (12-143). Sillce the servomotor always
has two phases, the total electrieal torque can be put into the form

na 2RrS(V R + V cP/2w
T. = - T; + T; = - (SF+' a2Rr)2+ S2(x'+a%xr)2
na 2 R'(2 - S)(V R - VeF/2w
+ [(2 - sTJ{' + a2R'p +(2 ":::-syt(X~~+-~(i2x'2) (12-175)

The first term in (12-175) is the positive-sequenee torque tending to


drive the rotor in the positive direction, and t.he second term is the neg­
ative-sequen('e c1eetrieal t.orque tending to drive the rotor in the reverse
direction. Figure 12-22 shows the general positive-sequence torque
598 Principles of electromechanical-energy conversion

-r. = r;- r.
Positive -sequence torque ( r;J

Negative-sequence torque (-r.)

Fill:. 12-27 Total servomotor torque of electrical orill:in as a function of slip for a
positive value of the rontrol-phas(, voltall:c less than the reference-phase voltage.

charaet crist iI· plot ted as a funel ion of slip. Thc IIcgal ivc-sequcnee
chara<'1eristie is idclltieal ill form with two exccptions:

1. Its magnilude is in general different.


2. Thc quantity 2 - ,"; is suhstituled for 8.

Figure 12-27 Hhows a typi('al plol of Ihc tol al posit ivc- and negative­
scqllmll'e clc('tri('altol'Clll<!R for a se'rVO!ll{)tor. The' rolor resishuwe is large
enough 10 haVf~ Ihe' peak positivc-seqllelwe torquc o('(~ur at a slip greater
t han tinily. The reason for slH'h a de'sign wiII he darified latcr in I.his
sel'lioll. The' ('\Irve'!, in Fig. 12-27 arc drawn on t he assumption I hal
the ('onl rol-phasc volt age V c is less I han the referclH'c-phasc voltage V H.
Thus at ,wro spI'cd, 01' ,"; = 1.0, II\(' posit ive-seqllcl\('c torquc is greater
than the ncgative-sequell!'c torq\le and the' lotal steady-state torqlle is
ill I he posit ive-allgular dire(~tioll.
The servomotor is usually used as H positioning devi('c, so the char­
acter of t he torque ('urves arOUlld zero Ilpef'd is of greatest interest. For
vahlf's of slip very Ileal' unity, the denominators of t.he two terms in
(12-17;i) are hoth very nearly given hy th(' per-phase blocked-rotor
impedance of the machine. From Fig. 12-161> the per-phase hlocked­
rotor il1lpedan('e has It Illagnitude equal to
(12-171)
For Inrgf' 1'0101' rl'sislalH'e'1'I a 2 U' is the· dominant term in the denominators
of (l2-17.'i); so as long as the slip is fairly near to unity, Eq. (12-176) is a.
The induct ion machine 599

reasonably good approxi mat iOIl for t hI' dl'nomillator expl·essiolls. By


8ubstHuting the rotor speed w' (I - S)w/n for t he slip, Eq. (12-17;))
ean now be puL into the forlll
(12-177)
where

nt'w! nn-IlH\('('/ rad (12-178)

newton-m/ vol t (12-179)

Figure 12-28 is a plot of Eq. (12-177) for vnriollS valu('s of ttl(' eontrol­
phage voltage V(~. The lleglLtive of the total torque of elC(~tri(·:tl Ql·jgin
'It.
- 7'. is ploH ed !I."I a flllletiOll of speed to ('orrespolld to Fig. 12-27. Remem­
ber that a negative torq\w of c\('("\ rit'al origin means that the machine is

-T"

FiR. 12-28 Servomotor torque-speed ",hara",teristies for various values of the con­
trol-phase volta~c, MIIullling the motor speed is much less than synchronouB speed, or
S ... 1.0.
600 Principles of electromechanical-energy conversion

gen('rat ing (t torque ill t he posit ive 8pc('d direct Ion. The torque generated
hy Ihe lIl(t('hine is linearly relotN} to the eontrol-phase voltage at w' O.
The 1110101' r:ieclricfll dmnpiny cor.flicicnl J)"" given hy (12-178), defining
t h(' slope of I h(' t orqu('-8pc('d ('urves, i8 prael.ieally independen t, of the
{'ontrol-phase voltage as long as ~·c« V R. However, for larger valll(,s of
Vc the slopes ilH'rease roughly as shown ill Fig. 12-28.
The trunsi('nt operation of the servomolor ('an he studied by using
I he aSstllllplioll I hat the eieetrieal transiellls arc very much faster than
Ihe lIl('('hanical t mllsients. Th(' rotor-port equalioll for t he servomotor
is given by

T' = (./p + j))w' + T. (12-180)


wh('rc '(r' i8 th(' total torque exterrmlly applied to the rotor shaft,./ is the
tot al rotor plus load inert in, and J) is t he viscous damping ('oefficient for
the rotor and load. For a constant passil'c load torque with a magnitude
of T~L newlOIl-1ll the exlprnal torque is given by

T' = - sgn w' T~L (12-181)

where
sgn w' +1 for w' > 0
(12-182)
sgn w' -1 for w' < 0
Substituting Eq. (12-177) for the st('ariy-state generated torque into
(12-180) gi ve8 IlS
(l2-18~)

Equation (I2-18~) has a slable response to inNel1lental changes in I.he


applied lorque or I he ('Ollt rol-phase voltage provided t he lot 11.1 damping
coefficient J) + J)", is positive. The quantity /)'" i8 positive provided th('
slope of the T. vprsus w r eHrve has a positive slope, or as shown in Fig.
12-27,the -1'. versus slip eurve near S = l.Olllust have a positive slope.
By examining Fig. 12-27, WI' find we have a positive slope al S = 1.0
provid('d the slip at maxinlUlII lorque $"" is greater than unit.y. The high
rotor resist allee thus makes t he servomotor have a stable operal ing
charact('risti{~ around zero speed .

__ ~pt

or
/{_........ ..
Tp+l -~.-~
....

Fig. 12-29 &lrvomotor block diagram.

The induction machine 601

Output
indicating
device

+
P. .=... E
S~ide
wire
L -_ _ _ _ _ _ . _ _ _ _ _ _ _ _ _

~
~
FiR. 12-30 Hchemntic repres('ntlttion for D. nnsic closed-loop servo-instrumentation
system.

Equation (12-183) can be represented 1Il block-diagram form as


shown in r'ip;. 12-29, where we have defined
K =_~1_ rad/newton-m-sec (12-184)
D+D",
J
'f" = D+D... seconds (12-185)

and or is the shaft angular position, in radians. For small values of the
control-phase voltage, the slight. dependence of D ... upon V c can be
neglected.
One of the most frequent uses for the two-phase servomotor is in the
system of Fig. 12-30. In the figure are shown the essential features of a
closed-loop serm)-instrumentation system. The output voltage of the
transducer I'. is relatcd 10 the physiea\ va.riablc being instrumented. The
potcnl iomct('r PI, whieh is usually known ItS a slide-wire, is connected to a
constant d-c voltap;e source B. The slidc-wire output voltage Vow is sub­
tracted from t he transducer output 111, and the difference is supplied to the
d-c to a-c (~Ollverter. This converter is a synchronously driven vibrator
or choppcr, which is csscntially a double-polc double-throw switch. The
choppcr output, is very nearly a square wave with a magnitude propor­
tional to its d-c input voltage. The frequency of the square wave
depcnds on the line frequency applied to the chopper driving coil, which
controls thc rate at which the switch vibrates.
The chopper output is supplied through a transformer to a high-gain
a-c amplifier. The amplifier output is ('onnected to the control phase of a
two-phase servomotor. The rcference phase of the motor is usually con­
nected in series with a capacitor across the same a-c line that drives the
The induction machi,.. 603

where
wn 2 = liQ~ . KKfh (rad/secp (12-187)
T

(12-188)

For a step change in the transducer voltage 1'1, the optimum syst.em
response is achieved with a damping ratio r = 0.6. The bandwidth of
t,he system is limited by the und:ullped natural frequClwy Wn. Notice
that increasing the amplifier gain K" improves t.he bandwidth but
decreases the damping ratio. For too small a damping ratio t.he system
has excessive overshoot and an oseillatory response.

12-14 THE SINGLE-PHASE INDUCTION MACHINE


One of t he major drawblwks to th{' polyphase induetion machine is the
requirement. that, a t.hree-phase or, more rarely, two-phase supply be
available. Industrial il1Rlltllntioll!<l usually huv!' thrl"e-phase power.
However, most hOllies and husilll"ss offi('es ('olllIHonly have only a single­
phase three-wire Edison supply. For this reason the single-phase induc­
tion machine plays 1\1\ important role in driving a wide variety of consumer
produ(,tH, usually requiring Icss than 1 hp output.. Adually, t,he single­
phase machine is simply a spc('ial easc of t he two-phase ma('hine with
one of the stator phases open-cirrll£fl'd. The rotor i!<l oftl'n ('onstnwted in
the form of a squinel ('!tl/:e, as shown in Fig. 12-1, so elcl'lri('al ('onnections
are made only to the single stator winding. The. air gap is uniform.
Figtlre 12-::l2 s('hcmat,j('ally shows the sinl/:le-phase induction ma('hille.
The two-phas<' a-b Illodel shown in Fig. 12-() ('an be used as a model
for t he single-phase IIllLl'hine. The d-q equivalent model, shown in Fig.
12-7, ('lUI now be substitutcd, and the steady-state analysis developed for
polyphl}Sc mll('hines ean be' used. Only t hI" st at or-porI volt age constraints
are SOllll"whllt differ('nt. ASStllllillp<; the single stator winding to be on the
a stator axis 01', (''luivull'nfly, the d stator axis of the two-phase model

Fig. ! 2-:i2 Schematic Tl'presentation of the


single-phase induction machine.
602 Principle. of e/ectromechonical·energy conversion

or

Fig. 12-31 Block diagra.m for the closed-loop servo-instrumentation system.

chopper. The capacitor shifts the reference-phase terminal voltage by


the required 11'/2 rad from the control-phase voltage. The motor shaft
is connected to the slider of the slide-wire through an appropriate mechan­
icallillkage. Now if the chopper input is positive (that is, VI > v... ), Va
leads V It by 11'/2 rad and the motor runs in a direction such as to drive the
elider up-scale and increase V'W' When v.w = VI, the chopper input is zero.
The control-phase voltage is also zero, and the motor stops. If Vt < V. w ,
then the chopper input is negative. The control-phase voltage V c now
lags V II hy 11'/2 rad and the motor drives the slider down-scale and
decreases v... until a balance point is again reached.
The shaft of the motor can also be mechanically linked to a wide
variety of indicating and controlling devices. For example, the motor
could drive a pen across chart paper moving at a uniform rate and thus
continuously record the variable of interest as a function of time. By
using a second servosystem to drive the chart paper, we have an X- Y
plotter whereby one variable can be plotted as a function of a second
variahle of interest. The motor shaft ean also he coupled to a digital­
dial indicator and thus provide a visual digital read-out. By having the
motor drive a cam, a microswitch can be actuated for controlling a process.
If the transducer is a thermocouple, the microswitch can be used to turn
a heater on and off in order to maintain a constant temperature. These
examples are only a few of the vast number of practical uses for the basic
system shown in Fig. 12-30.
Figure 12-31 shows the block diagram for the servo-instrumentation
system, assuming that the load torque on the motor shaft is negligible.
The amplifier gain K,. relates the magnitude of the control-phase voltage
Veto the chopper input voltage V, - V' The feedback constant K'b
W'

relates the slide-wire output voltage v.w to the radian position of the motor
shaft 9". The transfer function of the two-phase motor is taken from
Fig. 12-29 with Tr = O. The closed-loop transfer function, from Fig.
12-31, can be put into the form

(12-186)
VI
604 Principle, of elecfromechonical·energy conversion

machine, we can specify the applied voltage as

v~ = v: = V2V' cos wt (12-189)


The quadrature-stator winding is open-circuited, 80 the terminal voltage
cannot be specified. However, the winding current due to the open cir­
cuit must be zero; therefore,
(12-190)
Figure 12-11> shows the approximate per-phase steady-state equiv­
alent circuit for the induction machine. Let us see how the two single­
phase constraint equations (12-189) and (12-190) can be applied to this
equivalent circuil. The positive- and negative-sequence st.ator currents,
from the symmetrical-component transformation, are defined by Eqs.
(12-68) as

r+ = _1_ (1'd + 11')


V2 Jq
l' = _1_ (I' - 11')
- V2 d Jq
(12-191)

From the cO!l!!traint equation (12-190), we have I: 0, and therefore

I ,+ =
I'
- = VI~2 (12-192)

Figure 12-33a is the same per-phase equivalent circuit shown in Fig.


12-1.'> except that the negative-sequence stator voltage source V'-/V2 has
its direction reversed. Correspondingly, the negative-sequence stator
and rotor eurrentH ill Fig. 12-3:Ja are also Rhown revers(~d from Fig. 12-15.
The eonstraint equation (12-192) can he satiRfied by removing the con­
nection laheled A-A' ill Fig. 12-33a, therehy f()rcing IVv2 to equal
I'-/v2. The two sequence-voltage sources are now (~ollnected in series.
Using Eqs. (12-65), these two sources are added to give

(12-193)

Thus, the sum of the two sequence-stator sources is simply the sinor
voltage applied to the single stator winding of the actual single-phase
machine. The per-phase equivalent circuit of the single-phase machine
(built from a two-phase model) is shown in Fig. 12-33b. Using Eq.
(12-192), we have
I'+
(12-194)
v2
which is half of the actual current in the single-phase stator winding.
The induction machine 60S

R' lx'

v! (if ~
.f2 R,..• jX", J2a
.f2

v~ (if
.f2 Rh+# jX", ~
.f2a
.f2

R' jl(' a 2R' ja 2 x'


la)
R' jx' a 2R' ja 21(r

jX", ~
J2a

v;
jX",
~
J2a

R' jx'
(1) )

Fiji;. 12-:J:3 (al Per-phase equivalent circuit rOT the two-phlUle induction machine.
(II) Two-phR.se equivR.lent. dreuit ,",odifi"d to irwludl' thl' Miflll;ll'-phlUlC ('onstraint.

The equivalellt. ('in'uil shown ill Fig. 12-:3;~1) contains the eOllstraint.s
for 1he single-phase ·maehinc. However, we IllUst remember t hat, the
eir('uit was ('ollstrut'led from a I wo-phase model. Therefore, all powel'
(\uantities eal('ulatcd by using Fig. 12-a:~b arc still per phase and must be
multipli~d by 2 to obtain the total quantity for the actual siilgle-phase
machine. For example, the ele('\rie power supplied to the circuit of Fig.
12-33b is given hy

(12-195)

which is simply half of the !wt ual power supplied to t he single-phase


machine. Similarly. the core losses, stator and rotor copper losses, and
the cOllverted power as eal('u\ated from Fig. 12-33b must all he doubled
1.0 give t.he proper quant ity (or the ad ual single-phn.8e machine. If all
of the impedllnees in Fig. 12-:t~b were divided hy 2, then, for the same
excitation, all of the currents would he t.wiec as large. Powers calculated
The induction machine 607

The admittanre of the upper half of the circuit is given by


I
Y, = - - + - + - ---­+ 2.R + 53.2) +
50R j69 (1.97 + 2.3)

Y , = 0.024IL::37

Rimilarly, the admittanl'e of the lower half of the equivalent rinuit is


I
Y 2 = --
50/0\
-1 + (1.97
+ j69 + 2.R
I
"""--"--"--"------­
- 1.36) + + 2.3)
Y. = 0.IR7L -51l.~
Th., total irnp('(laorl' sel'n hy the lIS-volt sour,'e i.'l therl'fore

Zr = -YI , + -Y,
1 1
= ---------
O.024t/. -37
I
+-----------
(l.1R7L::-!iIl...:'>
= 46.,;/39.2

and 80 the stator-phase curr!>nt is given by

111\/11

2 X46.:'\L:~9~2 4.1I4/-:l1l.2

The magnitude of the stator C'urrent is 4.!)4 amp at a power fa!'tor eflual to ('Os 39.2
n.771l lagJl;ing.
h. The voltages aeross the upper an.1 lower halves of thl' efluivalent C'ircuit in
FiJI;. 12·34 arc givl'n re,spectively hy

V, =-y-;--:'y;IIIl/O = 102/ 1$

V, = __1::"'_._ U5/n = la.I/16.;


Y, + Y. - "

Thl'rdore, thl' two ,'urrent!! I, nn" I, "all he ('al"ulated 8.1',

The stator eopp!>r 10RSes per phase are therefore given by

Per-phase stator copper losses = 1,'R' + 1,'R' = [(1.76)' + 2.29)'1 X un

= 16.4 watts

The total stator ('opper IORSes arc ohtainC'd hy douhlinJl; the per.phase loases; thuR we

have

Total stator copper losses = 2 X 16.4 = :l2.R waUR

Similarly, the total rotor ('Opper 10ssl'II are

r. The per-phase 1'0nvI'I'ted powl'r is simply the power dissipated in the two
output resistors of the equivalent circuit shown in Fig. 12-34. The total converted
606 Principles of electromechanico/.energy converrion

with the modified circuit would then automatically be doubled. This


modified half-impedance equivalent drcuit is very often used for the single­
phase induction machine. I

Example 12-4
A t-hp lIS-volt four-pole single-phase 6O-cps induction motor is driving a constant
load torque at rated voltage with a slip S = 0.05. The per-phase equivalent-eircuit
parameters referred to the stator are as follows:
X .. = 69.0 ohms RA+. 508 ohms
R' = 1.97 ohms a'R' = 2.8 ohms
x' = 2.3 ohms atz' = 2.3 ohms
The tota~ friction and windage losses at this speed are 15 watts. Determine the fol­
lowing qJantities:
a. The fltator current and power factor
b. The total stator and rotor copper losses
c. The total converted power and the total torque of electrical origin
d. The total core loss
e. The output power and overall efficiency

Solution: a. At a slip S 0.05 the two output resistors in the equivalent circuit
of Fig. 12-33b are calculated to be
1 - S I - 0.05
a'R' - S - = 2.8 X 0.05 = 53.2 ohms

-aiR'; :=~ -2.8 X; := ~:~~ = -1.:~6 ohms


The complete per-phase equivalent circuit for the machine is "hewn in Fig. 12-34.

I See, for example, A. E. Fitzgerald and C. Kingsley, Jr., "Electric Machinery,"

2d ed., chap. ll, McGraw-Hili Book Company, New York, 1961.

1.97 j2.3 2.8 }2.3

r;*1
~
2 V 508 j69 53.2

I'
115LQ
T--­
Vz

I
508 j69
~ -1.36

1.97 j2.3 2.8 j2.3


Fig. 12-34 Per-phase equivalent circuit for the single-phase machine in Example
12-4.
608 Principles of electromechanical-energy conversion

power is twice the per-phll.Be value, thus giving us

1 -
p. = 2P•• - 2 [ l,'a'R'-- - II"a"R'-­
8
S
2-8

1 - S]
2[(1.76)1 X 53.2 - (2.29)1 X 1.36) = 316 watts

Notice that the negative-sequence converted power subtracts from the positive­
sequence converted power. The torque of electrical origin has a magnitude equal to
the converted power divided by the rotor speed, or
P 316

IT.I = (T=- 8)w/1I = (1 _ 0.05)(377/2) = 1.77 newton-m

d. The per-phase core loss is equal to the power dissipated in the two R~+.
resistors. Thus the total core 1088 is

Totat core lo&~


V,I
= 2 ( 508 + 508
V.I) = 41.8 watts

t. The output power delivered hy the motor is equal to the converted power
minus the windage and friction 10Mes. Thus we have

p ou • = p. - P"'+1 = 316 15 301 watts

The total power into the machine is equal to

Pi. = V.I. C08 (I = 111) X 4.94 X 0.7715 = 440 watts


and therefore the overall efficiency is

P ou • 301
'I - .. -~ - 68 5 per cent
Pi. 440 .

As a check, we can add the core loss, total copper los5, and the windage and friction
1088es to the output power and thus get

41.8 + 32.8 + 46.8 + 15 + 301 = 437 watts

which is within slide-rule accuracy of the 440 watts previously calculated as the
input power.

From Eq. (12-86), the torque of elect.rical origin generated per phase
is given by
nRr nRr
- 2Sw (I~->2 + 2(2 _ S)w (I~.Y (12-196)

At starting, the quantity 2 - S is equal to S,since the slip S 1.0. With


S = 1.0, the upper and lower halves of the equivalent circuit of the single­
phase maehine shown in Fig. 12-33b are identical, since both output
resistors are zero. Therefore, at starting we have I~ = I: and the average
generated torque given by (12-196) is zero. The basic single-phase induc­
tion machine has no starting torque. Once the machine is turning in either
The induction machine 6C9

direetion, so that the slip is no longer unity, the output resistances are no
longer equal and I: "F I:. A net. torque is generated and the machine
can operate successfully as a motor.
Single-phase motors can be made to develop a starting torque by
placing a second winding on the stator in quadrature with the main
single-phase winding. A machine of this type is known as a split-phase
motor. If the impedance of the auxiliary, or starting, winding is different
from that of the main winding, then the two winding currents will be out
of phase when the windings are placed across the same single-phase
supply. The phase displacement in the winding currents roughly sim­
ulates thc operation of a two-phasc mot.or connected to an unbalanced
two-phase supply. A starting t.orque is thereby generated, and after a
sufficient speed has been reached t.he starting winding can be open-cir­
cuited and the machine operates in a true single-phase mode. The
starting winding is usually disconnected by a centrifugal switch set to
operate when t.he rotor is revolving at nearly synchronous speed.
The phase displacement of the current in the starting winding can
also be aehieved by plaeing a eapacitor in series with the starting winding.
Such a ma<:hine is called !\ capacitor-start induction motor. Usually the
starting winding and series ('apadtor are open-circuited again by means
of a centrifugal switeh set to the proper speed. In some single-phase
machines t.he capacitor and starting winding are not open-circuited.
Such a machine is known as a permanent-split-capacitor induction motor.
For small, low-efficiency, inexpensive single-phase motors, such as
those used in small household fans, a salient-pole stator is used. A small
shorted copper loop is placed around a portion of the stator poles. The
resulting dissymmetry in the magnetic circuit gives the machine a small
starting t.orque. The copper loop is known as a shading coil, and so the
machine is called a ~haded-pole motor.l
All of these starting arrangements obtain a rotating field from a
single-phase supply either by unbalancing the winding impedances, as in
the split-field and capacitor-start machines, or by unbalancing the mag­
netic circuit as in the shaded-pole machine.

12-15 SUMMARY
A d-q primitive-machine model has been used as the basis for analyzing
the operation of the induct.ion machine. The d-q model relates to an.
equivalent a-b model representing a two-phase induction machine. With
I For a more extensive discussion of starting arrangements see A. F. Puchlltein,
F. C. Lloyd, and A. G. Conrad, "Alternating Current Machines," chap. 28, John
Wiley &: Sons, Inc., New York, 1954.
610 Principle. of electromechonical-energy convenion

the three-phase to two-phase transformation the model can represent the


very common three-phase induction machine. The two-phase servo­
motor is a direct copy of the two-phase model. The single-phase machine
can al80 be studied from the two-phase a-b model by imposing the proper
stator terminal constraints.
The analysis presented in this chapter is based on the assumption
that the electrical-port variables are in the steady state. Only the
mechanical transients are considered to be significant. Therefore, after
the equilibrium equations for the d-q model machine are written, sinusoidal
steady-state conditions are imposed. For stator voltage excitations at a
frequency of w rad/sec, the quantity jw is substituted for the differential
operator p. Sinor variables are used to replace corresponding time­
dependel,t variables. After a two-phase symmetrical-component trans­
formation on the complex sinor equations derived from the d-q model
machine is imposed, a per-phase equivalent circuit for the induction
machine is derived. A careful examination of the equivalent circuit
shows us how power and torque quantities in the original induction
machine can be calculated directly from the circuit without the necessity
of inverse transformations. For balanced polyphase voltage excitation
on the original induction machine, the per-phase equivalent-circuit excita­
tion reduees to the adual phase voltages, and the input current is the
actual phase current.
The equivalent circuit developed from the d-q model machine con­
tains self- and mutual-inductance parameters. By substituting another
equivalent circuit in terms of leakage and magnetizing inductance param­
eters, the hysteresis and eddy-current losses, or core losses, can be included
in our analysis. These core losses are represented by an additional linear
resistance in the leakag~mutual-inductance equivalent circuit. By con­
ducting blocked-rotor and no-load tests on the induction machine, the
equivalent-circuit parameters can be determined experimentally. These
parameters are related to the original primitive-machine parameters, and
80 they can be calculated in terms of the machine's physical dimensions.
The cireJe diagram is a very convenient, graphical technique for study­
ing the equivalent circuit of the induction machine. From the diagram
we can obtain the sinor values for all of the currents in the equivalent
circuit. By using an appropriate wattage scale, per-phase power quan­
tities ean be read directly from the diagram for a1\ operating speeds. The
complete circle diagram includes induction-generator operation where
the rotor is driven faster than synchronous speed and mechanical power
is converted to electrical power delivered at a leading power factor.
The steady-state torque of electrical origin is divided into positive­
and negative-sequence portions. The positive-sequence generated torque
tends to drive the rotor in the positive-angular direction and the negative­
The induction machine 611

sequence port.ion tries to drive the rotor in the negative-angular direction.


For balanced cxei talioll one of the sequen('e torques is zero. The max­
imum torque which a particular indue! ion ma(~hine can generate, for a.
specified stator excitation, is found to be independent, of the rotor resist­
ance. However. the slip at. whieh t his maximum torque occurs can be
controlled by adjusting the rotor resistalwe. In the wound-rotor machine
we have access to the rotor windings, and therefore the rotor resistance
can readily be adjusted. For a partieular rotor resistance the maximum
torque can be generated at starting to overcome st.arting friction and
minimize starting times. As the machine speed increases, the rotor
resistance can be reduced to obtain a lIlore desirable running torque-slip
characteristic.
The trnnsient operation of the induetion machine is studied by
assuming that no ele(~tri('al transients alfe!'t the mechanical transients.
Electrically, the machine variables are always in the steady state, and so
the steady-state torque of electrical origin can be used in the mechanical­
port equilibrium equation. Normal operation of the induction machine
occurs at a slip near zero, or at a speed near synehronous speed. By
linearizing the 100'(IUe-slip ehametcristic in the region near S = 0, an
analytic solution for the mlLehillc specd as a fUlletion of time can readily
be obt !lined. For llleeimni('ul transients not confined to the zero-slip
region, a graphical integrat,ion can be used to obtain w' as a function of
time. If the machine speed or slip is known as a function of time, the
steady-state equivalent eir(~uit. can be used to obtain the envelope of the
responding currents during the lIlechanical transient interval.
The two-phase servomotor is an indUe! ion machine commonly used
in low-power positioning applications. The normal operating region for
the servomotor is near zero speed, or a slip equal to unity. In order to
have a stable operating point around 8 = 1.0, the servomotor rotor
resistanl'e must, be increased until the slip at which maximum torque
occurs is greater than unity, usually about 1.5. For small values of the
control-ph~e voltage, a linearized transfer relationship for the servo­
motor, valid about S = 1.0, can be obtained. Control systems contain­
ing a servomotor can be studied by using this linear incremental transfer.
relationship.
The two-phase model ma('hine ean serve as a basis for analyzing the
single-phase nlllehine. By imposing an opC'n-cireuil constraint on one of
the I wo stator phases on the two-phase model, a modified per-phase
equivalent ('ir<'uit, valid for the single-phase maehine, is obtained. The
single-phase nmehine is annlytielllly equivalent to a particular set of
const mints on the two-phase model ma.chine, and it has a torque char­
aeteristic similar to that of I he polyphase induction machine operating
near zero slip. However, at starting, or S = 1.0, the single-phase machine
612 Principles of electromechanical-energy conversion

generates zero torque. Therefore, some auxiliary means must be used to


obtain a starting torque. The split-phase, eaparitor-start, and shaded­
pole single-phase induc-tion mot.ors are briefly discussed in See. 12-14 in
order to present the startiug techniques most commonly eneountered.

PROBLEMS

12-1 For the following in<ludinn machine~ fin,l the synchronou~ Ilpeed in
radianll per second and rpm:
a. I6-pole 60-cps three-phase
b. 16-pole GO-cps two-phase
c. 6-pble 50-cps three-phase
d. B-pole 25-cps three-phase

12-2 What ill the highest synchronous "peed that 25- and 6O-cps induction
machines have in common? How many poles must eaeh machine have?

12-3 A IO-hp 440-volt 6O-cpll eight-pole three-phase induction motor has a


full-load speed of R22 rpm.
a. Compute the full-load slip.
b. Find the frequency of the rotor currents at 822 rpm.

12-4 Perform the matrix transformation indicated by Eq. (12-64). The d-q
impedance matrix is given in 02-5Q), and the complete symmetri(~al-componcnt
transformation appears in 02-60). The (+ -) impedance matrix is given by
Eq. 02-74}.

12-rf(A I ~hp
IIO-volt two-phase two-pole wound-rotor induction motor has
two identl II. s tor windings and two identical rotor windings. The machine h8.8 a
smooth nonsali!' air gap. The winding parameters are as follows; R' 6 ohms,
I,' = 1.22 h, fl' = f# ohms, and 11 = 1.23 h. The mutual inductance between the
a-axis stator windings and a-axis rotor windings is 1.20 cos a henry, where a is the
angle between these two axes.
The two rotor windings are short..;)ircuited, and the stator windings are excited
by voltage sources as follows:

v: = 141 cos 4001 Vb = 141 sin 4001

a. Calculate V~ and V: and draw the complete self-inductance-mutual induct­


ance per-phase equivalent circuit, similar to Fig. 12-Q.
b. For the given balanced stator excitations calculate the steady-state developed
power and developed torque when the machine is driving a load at 367 rad/sec.
Also, compute the Iota I copper losses (rotor plus stator) for the machine under these
conditions.
c. The machine speed is reduced to 339 rad/sec and held at this value. The
D-Ilxis stlltor coil is suddenly removed from the voltage source (141 sin 400t) and is
then 8hort-circuited, making lIZ = O. Now calculate the total developed power and
The induction mochine 613

torque for the machine. Is the converted enerl/:y Howing out of or into the mechanical

rl
.
port? What are the total copper losses with this new ('xcitation?
12-6 A IS-hp 440-volt 1i0-eps four-pole three-phase wound-rotor wye-connected
'-'i~uetion ma!'hine at standstill has a line-to-Iine open-circuit rotor terminal voltage
which is twiC'e the applied line-to-line stat.or voltal/:e. With the rotor circuit open,
the rotor shaft is driven at 1600 rpm in the direction of the rotating field. Rated
voltage at rated frequency is being applied to the stator.
a. Write suitable expressions for the open-!'ircuit voltages appearing across each
pair of slip rings.
b. At what speed should the rotor he driven to deliver a rotor frequency of 80 cps?
c. If the rotor is driven at )R()() rplll in the direction of the rotating field, what
will be the magnitude and frequency of the line-to-Iine induced rotor voltage?
d. If the rotor is driven at ) ROO rpm in the direction opposite to that of the rotat­
ing field, what will be the map:nitude and frequency of the induced rotor voltage?
12-7 A 20-hp 60-cps three-phase four-pole induction motor has a full-load
speed of 1723 rpm. Assuming windal/:e and friction 10ssC8 are 5 per cent of the out­
put power, determine the total rotor copper losses.
12-8 A commercial power transformer has the following name-plate data:
1200 V /120 V (Primary/Secondary)

96KVA

60 cps

The components of the approximate equivalent circuit are as follows:

Primary resistance = 4 ohms

Primary leakage reactance = 3 ohms

Secondary resistance = 0.024 ohm

Secondary leakage reactance = 0.018 ohm

Core-loss equivalent resistance = 600 ohms

Equivalent magnetizing reactan('e = ROO ohms

a. With no load on .the secondary windinp: of the transformer find the magnitude
of the primary current with rated primary voltage applied.
b. Under the conditions of part a find the power snpplied to the primary.
c. On the basis of the approximate equivalent circuit, calculate the primary plu8
secondary ohmic losses for a secondary current of SUO amp.
d. On the basis of the approximate equivalent circuit, find the magnitude of the
primary voltage required to deliver 96 kva to an O.8-power-factor-lagging load, with'
rated secondary voltage across the load.
e. Calculate the self-inductance of the primary and secondary windings and the
mutual inductance between the two windings.

r.""i 12-9 A .')..hp 220-volt three-phase 60-cps six-pole wye-connected wound-rotor


i,ug~tion machine has the following no-load and blocked-rotor test data:

No-load test
Stator voltage (line to line) = 220 volts

No-load stator line current = 3.16 amp

Total power input = 590 watts

Windage and friction losses at synchronous speed = 312 watts

614 Principles of electromechanical-energy conversion

Blocked-Towr test
Stator voltage (line to line) "" 34.3 volts
Htator line ('urrl'nt 14.5 amp
Total power input = 710 wattll
At rated temperature, the d-e resistance between any two stator terminals is 0.48
ohm. With rated volta!!;e applied to the stator the open-circuit induced rotor voltage
is 220 volts line to line.
a. Determine the per-phase approximate equivalent circuit for the machine.
b. Find the resistance, self-inductance, and stator-rotor peak mutual inductance
for the windin!!;s of the actual three-phase machine.
c. Calculate the starting inrush current with rated voltage applied to the motor
and the rotor terminals short-circuited.
d. How much inductive reactance should be added to each stator phase to limit
the starting inrush current to 65 amp?
~2-1b A 220-volt three-phase 6O-cps wye-connected induction motor has a doc
resistance of o.:n ohm between any two stator terminals. With no external shaft
load the total power input ill 380 watts and the line current is 12 amp at rated stator
excitation. When the applied volta!!;e ill reduced to 40 volts, the total power input
decreases to 90 watts with a line current of 4.5 amp. At this reduced voltage, the
hysteresis and eddy-eurrent losses may be considered negligible. Compute the total
and per-phalle hyst!'r!'sis and I'ddy-current losses (or core losses) and the total windage
and friction losses.
12-11 A 5O-hp 440-volt 6O-cps four-pole three-phase induction motor draws
9.12 amp and 2134 watts at no load when operated at rated voltage and frequency.
With the stator windin!!;s wye-eonnected:
a. Calculate the elements in the no-load portion of the per-phase equivalent
circuit, assumin!!; a resistor and inductor in parallel.
b. Calculate the elements in the no-load branch of the per-phase equivalent cir­
cuit, assuming a resistor and inductor to be connected in series.
c. Repeat parts a and b for stator windings which are dclta-connected.
12-12 a. From Eq. (12-147) show that the per-phase developed power with
balanced~phase stator excitation can be put into the form
y~ a'W(l - S)(V'pIS

P,~ = [R,+"(i~R'+aq{'(C~-s)7sr'+-(x'-+-a'i;f'

b. Show that this expression for P,~ can be manipulated into the form

a'll'{k - S) =(T/')'y~'i-- 2R, {I ± ~;-= t(V.)~:~·'- 211;}


where R. ~ R' + a'R' and Z. "" IR.' + (x' + a'xr),l~i.
Note: The latter expression becomes quite convenient when the developed power
is known and the slip at which this power occurs is desired. The ± sign gives two
values of the slip, one of which is usually extraneous.
c. Show that the slip at which maximum power occurs is given by
a'R"
S.p = ii'R~-+Z.
12-13 A 5-hp 220-volt 6()-eps six-pole three-phase I5.3-amp wye-eonnected
induction machine was tested and the following data were obtained.
The induction machine 615

No-load test

Line voltage Line current Total watt.~

230 1).0 380


220 8.2 :J40
200 7.0 290
ISO 6.1 254
160 5.2 226
140 4.6 204
120 3.9 180
100 3.3 160
45 2.(1 120

Blocked-rotor test

Line voltage Line current Total Wl\tts

20 5.7 RO
40 11 .5 24U
6U 17.3 610
80 23.0 117U
lUll 2R.7 1881l

The stator resistance gives a voltage drop of 10 volts hetween terminals with 16
amp doc flowing.
a. From the no-load te.8t data plot the line current and total power input as
functions of the line voltage. By extending the watts ve. voltage curve to the zero­
voltage axis, the windage.and friction losses ('an be obtained as the watts input at
zero voltage. Find the windage and friction loss('8 and diSl'uss the reason why this
method is valid.
b. From the blocked-rotor test data plot the lin(' "urn'nt and total power input
as functions of the line voltage. CaJ.-ulate the power-fal·tor angle and also plot a
curve of 1 cos 8, the component of the line eurrent in phase with the line voltage, as
a function of the line voltage. Extrapolate th(' two ('urrent ('urves to the rated voltage
axis. The blocked-rotor power factor at rated voltage "an thus be obtained as the
ratio of the in-phase current to the total line current both taken at rated voltage.
c. By using the techniques described in parts a and b calculate all per-phase
approximate equivalent-circuit parameters.
d. Draw the complete circle diagram for the machine operating at rated voltage.
Show values of slip at equai intervals around the circle.
e. From the circle diagram plot a developed torque VB. slip characteristic curve.
f. For the value of slip where the machine is delivering maximum power oper­
ating as lln indu('tion generator, find the mal'hine spel'd in rpm, the total power
delivered, the hystl'r~sis and. eddy-current losses, the total stator and rotor copper
losses, the line current, the power factor, and the overall effi('iency. Use the circle
diagram, with current and wattage rulers, for obtaining these data.
616 Principles 01 eledrome<:honico/-energy conversion

12-14 A 4DO-hp 2300-volt 6O-cps six-pole three-phase wye-eonnected induction


motor draws a line current of 2.') amp and a total of 15,000 watts with no external
shaft load. The stator resistance is 0.42 ohm per phase, and the rotor resistance
referred to the stator is 0.23 ohm per phase. The total leakage reactance of the
stator and rotor is 1.90 ohms per phase. With rated voltage at rated frequency
applied, the motor is loaded until its slip is 3.0 per cent. For this condition find:
a. The rotor phase current referred to the stator
b. The stator line current and power factor
c. The total developed torque
d. The total developed power and efficiency
Use the approximate per-phase equivalent circuit in these calculations. Windage
and friction losses total 10,000 watts for speeds near synchronous speed.
12-15 The motor of Prob. 12-14 is operating at rated voltage and rated fre­
quency. With the machine delivering rated power compute:
a. T\le slip and speed, in rpm, closest to synchronous speed
b. Tile rotor phase current referred to the stator
c. The stator line current and power factor
d. The total developed torque

I!. The efficiency

Hint: See the results of Prob. 12-12b.

12-16 For the motor of Prob. 12-14 operating at rated voltage and frequency
"alculate:
a. The starting inrush current and power factor
b. The starting torque
c. The speed, in rpm, at which maximum torque occurs
d. The maximum torque
f!. The additional rotor resistance (referred to the stator) required to have maxi­
mum torque available for starting the machine
12-17 For the three-phase induction machine of Prob. 12-14, find the self­
induetanee per phase for the stator and rotor and the peak stator-rotor mutual induct­
ance. Find L:'/I.." L;'/I.." and M::/I,,' Draw the self- and mutual inductance per-phase
equivalent circuit similar to Fig. 12-9. Assume x' - a'x' and that the transfor­
mation ratio a ... 2.0. Can a hysteresis and eddy-eurrent 1088 resistor R' H be placed
in this form of the equivalent circuit? Explain your answer.
4
12-18 By using the approximate equivalent circuit for the induction machine,
neglecting the stator resistance, show that

where T is the developed torque at any value of slip 8, T ma. is the maximum torque
the machine can generate, and 8 ..., is the slip at which maximum torque occurs.
12-19 A 2300-volt 1000-hp 60-eps 12-pole wye-eonnected wound-rotor induction
motor has the following per-phase parameters: R' = 0.10 ohm, Rr = 0.034 ohm,
l' - 1.14 mh, and I' = 0.35 mho At rated voltage the no-load current is 75 amp and
the power drawn is 75 kw. The transformation ratio a - 2.0.
a. Select convenient. voltage, current, and power Bcales and draw a complete
circle diagram for the machine.
The induction machine 617

b. Locate thE'S - J.O point on the cir('le, and from the diagram determine the
blocked·rotor line current, power ftl.('tor, ('ore 10SII, stator cop))!'r loss, rotor cop))!'r
loss, and developed power.
c. Repeat part b for S = 0.5.
12-20 A polyphase induction motor has a total per-phase leakage reactance
X' + a'x' equal to 1.5 timE's th(' total ))!'r-ph!lll(, relli"tancE' Jl. + a'R'. The rotor
resistance per phase, referred to the stator, is equal to the stator rrsi~t.an('e prr phase.
At what slip does maximum torque occur?
12-21 A 20-hp 220-volt 60..,·ps eight-pole two-phase induction motor has the
following per-phase approximatt' equivalent-cireuit parameters all referred to the
stator:

R' = 0.25 ohm x.. .. 37 ohms

.£' - 1.00 ohm Rh • = 4.63 ohmK

aiR' "'" 0.20 ohm

alx' .. ] .20 ohmB

For each of the following connections draw the pl'r-phase approximate equivalent
rircuit. Rhow all parameter values and voltage exl'itations. Assume the motor
speed to hI' 840 rpm. Cal('ulate the line "urrent in rAch of the stator phaees, the total
dl'veloped pOWE'r, and the total devclo))!'d torque.
a. The two 8tator phascs are connl'cted to a halanced two-phtl.8l' infinite-hull
voltage Bupply delivering rated voltage at rated frequency.
b. One stator phase re('eives rated voltage at rated frl'qlll'n(1Y wit.h the ot.her
stator phasc short-('ireuited.
c. Same conne('tion as part II elWl'pt that th(' 8000nd phase is now open.."ircuited,
thus giving single-phll.8e operation.
12-22 The developed torque VII. slip curve for an induction motor running
near synchronous speed ('an he IlJ)proximat.rd hy a straight line. Thus,

T. '" -kS

where thl' usual sign convention for 1', is used. The ma('hine is driving a constant
load torque of 100 newton-m under steady-state conditions when, at t "'" 0, the load
is suddenly redueed to SO newton-m. The parameters are as follows: k = 200
newton-m, J = 2 kg-m' (shaft plus load inertia), D .. 0 (neglect windage), anfl
"'. = 500 rad/see. Find the slip as a function of time for all t ;:: o.

12-23 A 5-hp 440-volt 6O-cps four-pole three-phase induction motor has a


speed.torque characteristic as given in the following table:

Speed, rpm T{)rque, lb-/t Speed, rp ttl Torque, lb-/~

1800 I) 1200 39.8


1725 18.3 !lOO 37.1
1650 27.1 600 34.4
1575 :l3.2 300 32.4
1500 36.6 0 30.7
1350 39.4
618 Principles of electromechanical-energy conversion

a. Plot the torque-speed characteristic for the machine. Use mks units of
newton-meters and radians per second. respectively. (Consult Table A-I, Conver­
sion factors.)
b. The ma('hine is coupled to a constant load torque of 16.2 Ib-ft. The total
inertia of the load and the machine rotor is 50 Ib-fV. By using the graphical method
discusscd in Sel'. 12-12, obtain a plot of the speed as a function of time when the
motor is started.
c. Write an inl'remental linear differential equation similar to (12-167) that is
valid in the neighborhood of the steady-state operating speed of part b. Assume the
windage coefficient D to be zero. If the load torque is now increased by 2.0 newton-m
at t 0, derive the incremental slip S,(t) and write an expression for the total slip
as a function of time valid for all t ;::: O.
12-24 A 3-hp 440-volt 60-cps four-pole two-phase wye-conneeted wound-rotor
induction machine has the following per-phase equivalent-circuit parameters all
ref erred to fhe stator:
R' 2.2 ohms fl,,, '" 300 ohms
X' = 4.8 ohms X .. = 100 ohms
a'R' = 2.1 ohms a = 0.5
a'x' '" 4.2 ohms
The machine is to be used as a servomotor.
a. Compute thc additional rotor resistance per phase required to make the slip
at maximum torque S ... be equal to 1.5. Why must S .., he set at such a value?
b. For rotor speeds near zero, determine an approximate expression for the total
developed torque of this servomotor as a function of the shaft speed and the control­
phase voltage. Assume rated voltage on the reference phase, and the total rotor
resistance 1I~ determined in part a.
12-25 A IllJ-volt 400-cps two-phase servomotor has the torque-speed charac­
teristic shown in Fig. P12-25. The rotor moment of inertia is 3 X 10- 1 kg-m'. The

Speed. rpm
Fig. P12-25
The induction machine 619

servomotor is to be used in the position-control system shown in Fig. 12-31, whereby


an additional load of h. = 2 X Hl-t oz-in.' and D£ '" 4 X 10- 3 oz-in.-sec/rad is
coupled to the motor.
a. Linearize the torque-speed characteristic and find D", and k,. as defined in
Eq. (12-177), expressed in mks units.
b. Find the transfer relationship hctween the shaft angular position and the
magnitude of the control-field voltage.
c. Taking K,b in Fig. 12-31 to he 3 volts/rad. calculate the required amplifier
gain K. to have a closed-loop damping ra.tio equal to 0.6. With this value of K ••
find the undamped natural frequency w. in <,ycles per se("ond.
12-26 A i-hp IIO-volt eight-pole 6!)-cps single-phase induction machine is
driving a. constant load torque at ratl'd voltage at a speed of 864 rpm. The per­
phase equivalent-circuit parameters rpfcrrc(1 to thc stator are as follows:
x.. '" 33.5 ohms aIR' = 2.68 ohms
R.+t '" 478 ohms x' = 3.8S ohms
R' .. 2.51 ohms atx' = 3.8S ohms
Total windage and friction losses at this speed are 6.0 watts. Determine the following
quantities.
a. The stator current and power factor
b. The total stator and rotor copper losses
c. The total converted power and the total torque of electrical origin
d. The total core loss
11. The output power and overall C'fficiency

12-27 A single-phase induction motor is rated at t hp, 110 volts, 60 cps, 4.3
amp, and 3420 rpm. No-load and hlocked-rotor test data are as follows:
N o-loa,J. lesl

Stator voltage = 110 volts (rated value)


Stator current = 3.6 amp
Power input .. S8 watts
Blocked-rotor tellt .
Stator voltage = 19.2 volts
Stator current .. 4.3 amp (rated value)
Power input'" 46 watts
The d-c stator resistance is found to be 1.4 ohms.
Draw the per-phase approximate equivalent circuit shown in Fig. 12-33b. Give
all parartleter values and the applied voltage.
Hint: In making these calculations assume that the impedances of R.+. and X ..
are much larger than any noninfinite impedances connected in parallel.
&

appendix CONVERSION FACTORS AND


PHYSICAL CONSTANTS

TABLE A-I Conversion factors


Length (l) 1 m = 3.281 ft
39.37 in. (= 2.54 cm)
Mass (M) 1 kg 0.0685 slug
2.2051b
= 35.27 oz
Velocity (II) (translational) 1 m/sec = 3.281 ft/sec
= 2.237 mph
Velocity (w) (rotational)
1 rad/sec = 9.549 rpm
Force (I)
1 newton 0.225 Ib
= 3.597 oz
Torque (T)
1 newton-m = 0.738Ib-ft
= 141.6 oz-in.
8.851 Ib-in.
Energy (W) 1 joule = 0.738 ft-Ib
141.6 in.-oz
= 2.778 X 10- 7 kwh!'
= 9.478 X 10-' Btu
= 0.239 g-cal
Power (P) 1 watt 0.738 ft-Ib/sec
= 44.25 ft-Ib/min
= 141.6 in.-oz/sec
= 1.341 X 10- 3 hp
Moment of inertia (J)
1 kg-m2 = 0.738 slug-ft'
23.73 Ib-ft'
= 5.467 X 10' oz-in.2
Spring compliance (K) (translational)
1 m/newton 14.59 ft/lb
= 175.0 in./lb
= 10.93 in./oz
Spring compliance (KB) (rotational)
1 rad/newton-m = 1.356 rad/lb-ft
= 7.06 X 10- 3 rad/oz-in.
0.1130 rad/lb-in.
Damping constant (D)
1 newton-sec/m = 68.6 X 10- 3 Ib-sec/ft
= 5.72 X 10- 3 Ib-sec/in.
91.5 X 1O- 3 0z-scc/in.
Damping constant (DB)
1 newton-m-scc/rad = 0.738 Ib-ft-sec/rad
141.6oz-in.-sec/rad
= 8.851 lb-in.-sec/rad
Magnetic-flux density (B) 1 weber/m 2 = 10' gauss
= 64.52 kilolines/in.·
Magnetic intensity (H) 1 amp-turn/m = 0.0254 amp-turn/in.
= 0.01257 oersted
622 Principles of electromechanical-energy conversion

TABLE A-2 Conductivity of common metals

Metal Conductivity, mhos/ In


Silver 6.17 X liP
Copper 5.80 X 10 7
Aluminum 3.54 X 107
Brass 1.57 X 10 1
Steel 0.5-1.0 X 10 7

TABLE A·3 Density of common metals

.Metal Density, kg/tn'


Silver 10.47 X 10'
Copper 8.62 X 10 3
Aluminum 2.70 X 10 3
Brass 8.44 X 1O~
Steel 7.82 X 10 3

TABLE A-4 Physical constants

Permeability of free space Jl.u 4,... X 10- 7 him


Permittivity of free space EU 8.854 X 10-1' farad/m
Acceleration of gravity g 9.807 m/sec 2
&&

BIBLIOGRAPHY

Energy-state-function formulation
Goldstein, H.: "Classical Mechanics," Addison-Wesley
Publishing Company, Inc., Reading, Mass., 1950.
Guillemin, E. A.: "Introductory Circuit Theory," John
Wiley & Sons, Inc., New York, 1953.
McCuskey, S. W.: "An Introduction to Advanced
Dynamics," Addison-Wesley Publishing Company,
Inc., Reading, Mass., 1959.
White, D. C., and H. H. Woodson: "Electromechanical
Energy Conversion," John Wiley & Sons, Inc., New
York, 1959.
Whittaker, E. T.: "A Treatise on the Analytic Dynamics
of Particles and Rigid Bodies," 4th ed., Cambridge
University Press, London, 1937.

Introductory electromagnetic theory


Hayt, W. H., Jr.: "Engineering Electromagnetics,"
McGraw-Hill Book Company, New York, 1958.
Plonsey, R., a.nd R. E. Collin: "Principles a.nd Applica­
tions of Electromagnetic Fields," McGraw-Hill
Book Company, New York, 1961.
Reitz, J. R., and F. J. Milford: "Foundations of Elec­
tromagnetic Theory," Addison-Wesley Publishing
Company, Inc., Reading, Mass., 1960.
Seely, S.: "Introduction to Electromagnetic Fields,"
McGraw-Hill Book Company, New York, 1958.

Linear-8Yste'TfUJ theory
Cheng, D. K.: "Analysis of Linear Systems," Addison­
Wesley Publishing Company, Inc., Reading, Mass.,
1959.
Gardner, M. F., and J. L. Barnes: "Transients in Linear
Systems," John Wiley & Sons, Inc., New York, 1942.
LePage, W. R.: "Complex Variables and the Laplace
Transform for Engineers," McGraw-Hill Book
Company, New York, 1961.
Lynch, W. A., and J. G. Truxal: "Signals and Systems in
Electrical Engineering," McGraw-Hill Book Com­
623 pany, New York, 1962.
624 Principles of electromechanical-energy conversion

Seely, S.: "Dynamic Systems Analysis," Reinhold Publishing Corporation, New


York, 1964.
Reza, R. M., and S. Seely: "Modern Net.work Analysis," McGraw-Hill Book
Company, New York, 1959.
Van Valkenburg, M. E.: "Network Analysis," 2d ed., Prentice-Hall, Inc., Engle­
wood Cliffs, N.J., 1964.
I n.cremental-motion transducers
Lion, K. S.: "Instrumentation in Scientific Research: Electrical Input Trans­
ducers," McGraw-Hill Book Company, New York, 1959.
McLachlan, N. W.: "Loud Speakers: Theory, Perrol'mance, Testing, IUld Design,"
Dover Publications, Inc., New York, 1960.
Neubert, H ..K. P.: "rnstrument Transducers," Oxford University Press, London,
1963.
Aavant, C. J., Jr.: "Basic Feedback Control System Design," 2d ed., McGraw­
Hill Book Company, New York, 1964.
Truxal, J. G. (ed.): "Control Engineel"'S Handbook," McGraw-Hill Book COIll­
pany, New York, 1958.
Primitive-roochine theory
Adkins, B.: "The General Theory of Electrical Machines," John Wiley & Sons,
Inc., New York, 1951.
Bewley, 1. V.: "Tensor Analysis of Elect.ric Circuits and Machines," The Ronald
Press Company, New York, 1961.
Koenig, H. E., and W. A. Blackwell: "Electromechanical SYAtem Theory,"
McGraw-Hill Book Company, New York, 1961.
Kron, G.: "Application of Tensors to the Analysis of Rotating Electrical Machin­
ery," 2d ed., General Electric Review, Schenectady, N.Y., 1942.
Ku, Y. H.: "Electric Energy Conversion," The Ronald Press Company, New
York,1959.
Messerle, H. K.: "Dynamic Circuit Theory," Pergamon Press, New York,
1965.
Seely, S.: "Electromechanical Energy Conversion," McGraw-Hill Book Com­
pany, New York, 1962.
Takeuchi, T.: "Matrix Theory of Electrical Machinery," The Ohm-Sha, Ltd.,
Tokyo, 1958.
White, D. C., and H. H. Woodson: "Electromechanical Energy Conversion,"
John Wiley & Sons, Inc., New York, 1959.
D-c roochines
Fitzgerald, A. E., and C. Kingsley, Jr.: "Electric Machinery," 2d ed., McGraw­
Hill Book Company, New York, 1961.
Harwood, P. B.: "Control of Electric Motors," 3d ed., John Wiley & Sons, Inc.,
New York, 1952.
Langsdorf, A. S.: "Principles of Direct-current Machines," 6th ed., McGraw­
Hill Book Company, New York, 1959.
Bibliography 625

Pestarini, J. M.: "Metadyne Statics," The Technology Press of the Massachu­


setts Institute of Technology, Cambridge, Mass., and John Wiley & Sons,
Inc., New York, 1952;
Siskind, C. S.: "Direct-current Armature Windings," McGraw-Hili Book Com­
pany, New York, 1949.
Siskind, C. S.: "Direct-current Machinery," McGraw-Hill Book Company, New
York,1952. .
A-c machine8'
Alger, P. L.: "The Nature of Polypha.se Induction Machines," .John Wiley &
Sons, Inc., New York, 1951.
Concordia, C.: "Synchronous Machines: Theory and Performance," John Wiley
& Sons, Inc., New York, 195I.
Fitzgerald, A. E., and C. Kingsley, Jr.: "Electric Machinery," 2d ed., McGraw­
Hill Book Company, New York, 1961.
Langsdorf, A. S.: "Theory of Alternating-current Machinery," 2d ed., McGraw­
Hill Book Company, 1955.
Lyon, W. V.: "Transient Analysis of Alternating-current Machinery," The Tech­
nology Press of the Massachusetts Institute of Technology, Cambridge
Mass., and John Wiley & Sons, Inc., New York, 1954.
Puchstein, A. F., T. C. Lloyd, and A. G. Conrad: "Alternating-current Ma­
chines," 3d ed., John Wiley & Sons, Inc., New York, 1954.
Siskind, C. S.: "Alternating-current Armature Windings: Theory, Practice, and
Design," McGraw-Hill Book Company, New York, 1951.
Stoker, J. M.: "Nonlinear Vibrations in Mechanical and Electrical Systems,"
Interscience Publishers, Inc., New York, 1950. (Synchronous machine
stability.)
Veinott, C. G.: "Fractional Horsepower Electric Motors," 2d ed., McGraw­
Hill Book Company, New York, 1948.
_BEcm&i&q:ggg_W 4U: 1&45.. _ •. adask 41.

ANSWERS TO PROBLEMS

I-I a. Nonlinear. The g-v relationship is not a straight line


dg g",T. 7rV 7rV
c. C(v) - = -SIn-COS­
dv Vm 2v", 2v m

d. C ( 2vm) =
gmT

2v",' max value of C

a. Nonlinear, since L is a function of i; thus X-i curve is not straight


lo2
­ lA
2
1-5 b. T'(WA,W8) iJ 808: + ¥ AOA·, T(lA.lo) = ~-
2Js
+ 2J..

1-7 a. Nonlinear. b. g(v)

e. W:(v) = gm (-1
a
+ e-'" + av). d. W.(v) gm (1 _ e-'" - ave-.')
a

e. C = agme-"·, tCv" a:" v"e- a ., W. ;;c iCv' because C is a function of II

I-ll a. R 77.1 ohms. c. Rnew = 70.2 ohms

1-13 a. p = em. sin-l~.


c

2-1 a. .e tJIOl' + ¥:02' + ¥aOa 2 - (812~.

b. Q. T., Q. = -To, Qa = 0

2-3 UJ + (g - fit) sin 8 = 0, iit l(e sin 8 +0 2


cos 8) + ~K y, 0

2-5 a. T' tm(x' + il' + z·)

c. z l cos </>, y l sin </> sin 8, x = l sin </> cos 8


d. .e = iml·(~· +
02 sin' </»

2-7 a . .e jL 1q.2 + Mg g, + t L 2g.'


1 - :~~ - 2C.

[These are analog quantities)


{I

2-9 b. q. = ql, g. = -g2, gd = q, g., gb = g. - q. [Holonomic, since


these equations are integra.ble]
627
628 Principles of electromechanical-energy conversion

Aa 2 Ab·
2-11 a. We
I l'
TtClAe2
l'
+ TtC.Ad', W", = 2L. +
b. .c = W; - W... c. No. d. Ac A., Ad = A" ~a A. - ~2' 4 ~2
2-13 b.:J = 975 watts (approx), :J' 625 watts (approx)
c. D8 0.113 newton-m-sec/rad, iDI!W2 1390 watts

3-3 a. A

dA - 0). f.' ON 2hr{a


. .. .

b. L = - = , Imear smce L 16 mdependent of t


di
2g

c W' = J.Li2 = /Jo N2hr {a


- O)i', W
~ 11& 2 4g 0)

IJ" f.' ON2hr(a 0)i2 (0


3-5 a. .c Tt u 2 +----~------
4g
b. :J = ~Dg/J2 + iRi2
0 /JoN·hri·
c. J /j

+ Deu
A
+ K8 +---
4g

f.' ON2hr/Ji 2
- 2g
+ f.'oN hr(a
2g
- 0) di

-dt + Rr vet)

3-7 a. I. (Negative sign means force acts to


increase xl
b. Ie = i)W,(q,x) q'ln (b/a)

i)x

4"<0 G+ x r
A~·
3-II b. W'•. = ­ 6 (x.•

c. 'U

3aA~~ bA2

3-13 a. .c
4
+ -2 (x c)' + iM.t'

x-c
c. AU 2bA'(x c)3 + ~ + D.t -Mg
j-aA -%>, + b>,(x - 0)4 + 4bA(x - C)3.t + GA = i(t)

9 • Xl' . 3X.·X.'h2;:'
3-15 a. W. TtX.X. 2 A.' + x,
A.A. + 4

3X2 2;)..4 4x.'


- ;\.;\. -
9X.X24;)..~f.
----~--
b. I" 2 ~ 2
f.2
Answers fo problems 629

oW;(~,x)
c. lei = - , same results a.s for b
ox.

4-1 a, b, c, linear. d. Nonlinear


4-3 c. L. and L3 commute. d. AU are linear operators; L. and L3 are constant­
coefficient operators

X
4-11 a. = 2.5/ -50 0 or 7.94 db, -50 0 phase. b. 12.06 db, -50 0
I --­
c. -12.06 db, 130" or -230°. d. 12.06 db, -140°
4-13 a. get) (i + fe-at + J/'r 6e- ')u(t)
= U 5

b. get) = + O.62e-" cos (t + 120")Ju(t)


[T'\r" l
c. get) ( lrP ~\t + rlo e-" - Tf'o e- 21 ) ;0 u(t)
d. g(t) = (te-' sin 2t - 40e-' cos 2t)u(t)

10-15 a. All poles in LHP, stahle. b. Two poles in RHP, unstahle. c. Two
poles in RHP, unstable

5-1 a. Xo = 2, Yo = 4. b. a 108. c. b 108


d. 5i, +
8x, +
(3x021/0')X' (2x03yo)Y, Al (t)
+
12ih 20],it + (41/.)Xl +
(4XO)Yl = B,(I)
+
5-3 a. 1f,1 30i'. b. /' = 0.1 08 + a.6i,. r. Exart (0.1 H)()7), approximate
(0.1188). d. 20 rna

5-5 b. 00 = O,7r. c. Ml'tJ, + Del + Mgl ('OS 11011, 0


d. Houth's method lI:ives Mgl cos 00 > O. Only 110 = 0 is a stahle point
5-7 a. 0.0995 weber/m'. b. 0.1)91 kll:. r. 3.25 newton-s(>('/m
d. k 2.04 mv-sec/m, Wn aO.2 rad/sec, I = 0.501. e.:; dl> down at .5.25 t'PS
f. With AI, 3 dh down at n..55 cps; therefore eu is better
b
5-9 a.
I'
Xo -
K + N2puA.
xo%
l02 = H, R~. = Vo.
.
b. LS7 X 10-' m
c. 2,46 X lO-' m (stahle), 1.02 X lO-' m (unstahle)
5-11 b. M = .5.6\1 X lO-' kg, K = 3.18 X 10- 8 m/newton. d. a.:33 X 10-' m,
4300 volts. .f. .5.2 milliwatts. g. 10.4 watts, 0.05 per cent

6-1 (II) - g, +2 01 + 2(0' 7r- 01) [


cos n -
(?o) - l cos :3n(211) +
*
c. 9 -
cos 5n(211) - . . .J
360 poKai
c. 2 turns/III, O. d. B - 2g(lI) cos 211 a" 0
7r a
480 240 120
6-5 c. 0, - , -2 , 2
turns/m
1\'"'a 1I'" a 1I'" a
630 Principles of electromechanical-energy conversion

240
6-7 c. turns/m. d. K is independent of the rotor's angular position
,..2a
6-11 a. 0.446 ohm. b. 2.18 ohms. c. 974 turns/m. d. Lr 0.048 henry,
L' 0.573 henry
e. 0.143 henry. f. Gd~ G;d = 0.573 henry, G';d = G'J. = 0.048 henry. g. 0,
144 volts
7-1 a. Uniform at 45°. b. Same as part a. c. Fields cancel
d. Uniform, revolving CCW at l/n rad/sec. e. Same as part d

7-3 a. V~D = v~o 20 volts, VdO 124 volts, v~u -90 volts

b. p~. = 80 watts, into ohmic heating. c. '1', 24 newton-m, clockwise


d. Pdq = -44 watts. e. '1'; 26 newton-m, P~n = 104 watts

7-5 a. (1 e- 5t ,3)u(t) amp

b. v; = v~ 0, v:i = [21(28e-$·13 + 42 70e- 2.") + 5e- 5tI3 )u(t)

7-9 a. 5 voltages and 5 currents. b. 4. c. 1. e. 1 voltage, 4 currents

f. 1 voltage or 1 current

8-1 a. 119 volts. b. 0.202 henry. c. 109 volts, 77 volts. d. 6980 watts

8-3 a. 75 kw has 30.4 per cent, 150 kw has 69.6 pel' cent, 222 volts

8-5 b. 222 volts. e. 0.99 volt/volt, 21.3 watts/watt

8-7 b. R f = 24.4 ohms, rheostat 26.8 ohms. c. 48 amp. d. 0.268 henry,

0.122 ohm. e. 115 newton-m


8-9 a. 14.3 ohms. b. wr(t) = 1800 3600(1 - e-O.lH)tt(t) rpm
c. 6.3 sec. d. 75e- o· lltu(t) amp. e. -1800 rpm
8-11 a. 248 rad/sec, 17.3 amp. b. wr(t) = 248 + 160(1 e-U•10")u(t) rad/sec

8-13 :2 (~)I joules, wr( co)


Gi/0
= ~, so energy quantities are equal
Gt/o

8-15 a. 25Ib-ft. b. 162 rad/sec. c. 49 per cent

8-17 a. 5.43 ohms. b. 4.36 ohms. c. 3.60 ohms

8-19 a. 0.0248 newton-m-sec/rad, 0.0376 henry. b. 91.2 newton-m

c. 121 - 10(1 e- 7 .16')u(t) rad/sec

9-1 b. 2.95 amp, 78 newton-m

d. w~(t) 100 [10.4 + +


28.8e-4.671 sin (1.8t - 159°))u(t) rad/sec

9-3 a. Gal = 0.408 henry, Ga. = 0.545 millihenry. b. 125 volts (cum.),
109 volts (diff.). c. 92.2 per cent (cum.), 91.2 per cent (diff.)
9-5 b. 133 + 66.6[1 + 4.85(0.le-: 10 ' 0.306e- 3 · 21t )]u(t) volts
c. 10.2 amp. e. 5870: 1
9-6 b. 16.4 volts, change 0.027 rad/sec decrease

r1~1(8) 11.2

[Stable]
C. V Rl(8) = 1.38 X 10- 38 3 + 0.04898 2 + 0.4518 + 12.2

9-7 a. 1930 rpm. b. 0.474 henry

Answers to problems 631

9-9 a. 53 ohms. c. 221 volts, 1630 amp


9-11 a. 0.833 X 10-- 3 sec. h. 2.71 volts. c. J)i~continuity \)G.l amp
d. 0.0262 weber 1m"

10-3 12.7 turns/m, 18.0 turns/m, 14.3 turns/m

10-5 b. L~ L; = t henry

10-7 a. 10 volts. b. Vb = 0, v~ -IHOn sin 3a, vI; -lHOU eos ;30:

c. Vd 10, v; Vd = 0, v; = -180n volts. d. 10 newton-Ill


10-9 a. ~t. b. v~ = Vb = o. c. V:I = 70.7 volts, v~ = () (short cil't'uit),
iri i: = 0 (open circuit). d. Vd = v; = n. e. v: 141 cos 7541,
Vb = 141 sin 754t
ll-l a. 900 rpm. b. 900 rpm. c. 250 rpm. d. 107 rpm

11-3 a. 2960 newton-m. b. 85.4 amp. c. 2960 newton-m, ~)8.5 l1mp


ll-5 a. 1800 rpm, 189 rad/sec. b. Phase 50 amp, line 86.6 amp
c. Generator. d. 12.6", 299 volts. e. 25,860 watts, 137 newton-Ill

11-7 a. 2053 volts. h. 33.1". c. 470 amp

11-9 74.6 per cent

11-11 a. 1.89 ohms. b. 5.66 ohms

11-]3 a. 0.0282 ohm, 0.101 newton-m-sec/rad. b. 0.24 houry

c. Xu 3.21 ohms, X. = 1.93 ohms. d. 1176 newton-ln. e. 7.1 amp


f. L~ = (3.03 + 0.75 cos 66) millihenry
11-15 a. 9°. b. 0.157[1 + 1.41e- 13 .$I sin (13.5t - 135°)]1£(1) m.1. c. r = n.70ti
11-18 a. X d = 5.65 ohms, X. = 3.a!) ohms. b. X:/ = 1.94 ohms
c. l'ri = 0.2 aoc, 1'~ = 0.0688 sec

] 1-19 a. 13,600 volts

h. iri -3410 cos 377t + 7800e- 12 . 91 + 2120e- 12 ... ('.06 7541 ­


6510e- I <.51 cos 377t
12-1 a. 450 rpm. b. 450 rpm. c. 1000 rpm. d. 375 rpIll
12-3 a. 0.0867. b. 5.21 cps
12-5 a. V~ j141, V'... = o. b. 280 watts, 0.765 newton-Ill, 42.8 watts
c. 29.8 watts, 0.088 newton-m, out the mechanical port, 323 watts
12-7 704 watts
12-9 a. R' = 0.24 ohm, Rr = 0.89 ohm, x' + x' 0.767 ohm, X.. 41.3
ohms, R,,+< 174 ohms. b. L~fl'" = L~lh 0.0737 h, 1If;;/h = 0.0731 h. c. H3 amp
d. 1.45 mh

12-11 a. R k .;.. = 90.7 ohms, X", = 29.2 ohms. b. R h +< 8.55 ohms, X", =

26.6 ohms. c. RA+o = 272 ohms, X .. = 87.6 ohms (parallel), R,,;. 25.7 ohms,
79.8 ohms (series) ,
12-13 a. 102 watts. c. RII+< = 203 ohms, X,n = 15.5 ohms, R' = 0.31 ohm,
a~Rr = 0.43 ohm, x' + aZxr = 1.89 ohms
632 Principles of eleclromechanical-energy conversion

'12-15 a. 0.0145, 1183 rpm. b. 81.1 amp. c. 85.3 amp, OJI56 lagging
d. 2480 newton-m. e. 91.7 per cent

0.02:3H h, lIf:;ll-, = 0.0471 h, no Rh+<


12-19 b. 1380 amp, 0.35 lagging, core = 75 kw, stator 798 kw, rotor = 1100
kw, P. = O. c. 1180 amp, 0.63 lagging, core = 75 kw, stator 5'27 kw, rotor = 188
kw, P. = 1475 kw
12-21 a. 98.8 amp, 16,600 watts, 189 newton-Ill. b. 101 amp, 32.6 amp (in
short), 7900 watts, 89.7 newton-m. c. 116 amp, 5332 watts, 60.5 newton-Ill
12-23 c. Set) = 0.0345 + 0.0051(1 - e-··~at)u(l)

12-25 a. D.. 2.7 X 10-' newton-m-sec/rad, k,. = 1.5 X }()-4 newton-m/volt


c. K. 8.45 volts/volt, 12 cps
12-27 a 2R' = 1.09 ohms, x' + a 2x' 8.72 ohms, R A+. = 439 ohms, X",
57.8 ohms

Anda mungkin juga menyukai