3_Control System Components
3_Control System Components
aMinmnlnMinttilK
McORAW«HIU SERIES IN
mtro
.jumniuusiiimHHJnmimiWHtmsmi'ims
621.8 G35c
Gibson & Tuteur 1058249
PUBLIC LIBRARY
FORT WAYNE AND ALLEN COUNTY, IND.
ACPL ITEM
DISCARDED
CONTROL SYSTEM COMPONENTS
McGraw-Hill Series in Control Systems Engineering
John R. Ragazzini and William E. Vannah, Consulting Editors
1958
CONTROL SYSTEM COMPONENTS
Preface. v
Damper 8.6 Gears 8.7 Design of Gear Trains for Minimum Inertia 8.8
Gear Ratio for Load Matching 8.9 Backlash in Gears 8.10 Ball-screw
Actuator Problems
Index 481
n
CHAPTER 1
10 ohms 33 ohms
12 39
15 47
18 56
22 68
27 82
and powers of 10 thereof
Figure or Figure or
■ /First significant figure number of number of
Color zeros Color zeros
~ ^—Second significant figure Black 0 Green 5
Brown 1 Blue 6
■ '^"Multiplier Red 2 Violet 7
. Orange 3 Gray 8
Tolerance Yellow 4 White 9
Color % tolerance
None 20
Silver 10
Fig. 1.1. Resistor color code. Gold 5
given in Table 1.1. The values are chosen so that the percentage change
in resistance between each of the values is approximately constant. Values
at one-half the standard percentage step are also produced but are not so
widely stocked and may be somewhat more expensive.
Wire-wound resistors may be obtained in any resistance value and
wattage rating. Standard RMA values in the 5- and 10-watt ratings
are commonly available from stock. Several companies specialize in
supplying precision resistors to any accuracy required.
1.3. Variable Resistors. Sometimes the term “variable resistor” is
applied to a device which has only two terminals. One terminal goes to one
end of the resistor, while the other terminal is connected to the wiper arm.
The symbol is given in Fig. 1.2a.
O
Fig. 1.3. Cutaway of precision variable resistors: (a) single-turn; (6) helical-wound
resistor; (c) multiturn slide-wire. (Courtesy G. M. Giannini and Co.)
form, which is then bent into a circle. The wiper bears on the edge of the
card and moves from turn to turn (see Fig. 1.3).
Accuracy of a variable resistor, or “linearity,” is commonly defined in
two ways. So-called “best linearity,” “independent linearity,” or nor¬
mal linearity is the maximum departure of the actual resistance curve from
the best straight line which can be drawn through the actual curve, this
departure being expressed as a percentage of the total resistance of the
4 CONTROL SYSTEM COMPONENTS [Chap. 1
device (see Fig. 1.4). Note that the straight line is not drawn through
the origin, and thus a better figure for linearity is obtained than may be
obtained if the zero-based linearity is used. Zero-based linearity is the
maximum departure of the actual resistance curve from the best straight
line drawn through the origin. This is the more natural definition, but
precision-resistor manufacturers have adopted normal linearity almost
universally. A single-turn wire-wound resistor may be obtained from
stock with 0.1 per cent normal linearity for resistance values of 5,000 ohms
and above. The body diameter of this device is of the order of 3 in. If
either the resistance or the body diameter is smaller, the number of turns
of wire must be reduced, and thus the linearity is reduced.
The resolution of such a device is also important. Resolution of a varia¬
ble resistor is defined as the smallest change in resistance which can be
detected, expressed as a percentage of the total resistance of the device.
,y dr / Straight line
approximating curve
/ A of potentiometer
/
Total /
/
/ but passing through
Total
zero point
travel / travel
T~o
End voltage or Travel of contact Endvoltage or Travel of contact
resistance of resistance of
potentiometer potentiometer
(o) Normal linearity (£) Zero-based linearity
_1_
Resolution = X 100% (1.1)
total number of turns of wire
0-3)
cCir-f) + i = ° <'-4>
1 Gardner and Barnes, “Transients in Linear Systems,” John Wiley & Sons, Inc.,
New York, 1942, pp. 38-43.
2 Readers not familiar with the Laplace transform method for analyzing networks
are referred to Gardner and Barnes, ibid., and also to Chestnut and Mayer, “Servo¬
mechanisms and Regulating System Design,” John Wiley & Sons, Inc., New York,
1951, vol. 1, pp. 66-96. For a mathematical treatment of the Laplace transformation,
see Doetsch, “Theorie und Anwendung der Laplace-transformation,” Dover Pub¬
lications, New York, 1943,
12 CONTROL SYSTEM COMPONENTS [Chap. 1
We use the shorthand notation e = E(s); hence Eq. (1.5) may be rewritten
in the form
e0 _ RCs
(1.7)
% ~ RCs + 1
It may be shown1 that the quantity RC has the dimensions of time, It is,
therefore, referred to as a time constant and given the symbol T. With
this notation, Eq. (1.7) can be written in the equivalent form
K _ Ts
(1.8)
Ci Ts + 1
Equations (1.7) and (1.8) are in the time-constant form. This form is
identified by the fact that the constant term is unity. The time-constant
form is one of two standard forms in which transfer functions are com¬
monly expressed. The other common form, which may be referred to as
the root form, is given for the simple network discussed here by
= S = S (1 9)
Ad s + (1/22C) « + (I/?7) V '
charge
RC = time (sec)
charge/time (sec)
SEC. 1.11] ELECTRIC NETWORKS FOR CONTROL SYSTEMS 13
Fig. 1.7. (a) Magnitude and (6) phase angle of the transfer function for the circuit
shown in Fig. 1.6.
have been no limitations placed on the voltage that is applied to the cir¬
cuit except that it exists only for positive time. If we are willing, how¬
ever, to restrict ourselves to sinusoidal voltages and to steady-state oper¬
ation, we may solve for the response of any circuit to any frequency given
merely the transfer function of the circuit. It is only necessary to sub¬
stitute yco for s (j = — 1) and then to let co be any value of frequency
in which we are interested.1 The magnitude of the resulting complex
number is the amplitude ratio of output to input, and the angle is the
phase difference between output and input. For the circuit considered
in Fig. 1.6 this information is available in Eq. (1.11). The complex num¬
bers in the numerator and denominator are converted to polar form and
divided. The resultant magnitude is shown in Fig. 1.7a, while the angle
between input and output is shown in Fig. 1.76.
1 See Chestnut and Mayer, op. tit., pp. 99-103.
14 CONTROL SYSTEM COMPONENTS [Chap. 1
Thus the asymptote for low frequencies is the log-co axis, while for high
frequencies the asymptote is a straight line with a positive slope of 1,
intersecting the log-co axis at the point co = 1/T. These two asymptotes
are shown in Fig. 1.8 as the solid lines. The actual characteristic
log |1 + jwT\ is shown dotted. Note that the actual curve does not
depart very far from the asymptotes at any point. For real T the maxi¬
mum departure occurs at the point where the two asymptotes meet.
This is the so-called “break point,” for which co = 1/T. At this point
the value of the actual characteristic is \/2; hence the departure of the
asymptotes from the true curve is of the order of 30 per cent. Since the
asymptotes represent the actual curve so closely, it is common practice to
draw only the asymptotes to represent the amplitude characteristic of a
transfer function. The resulting curve is the asymptotic amplitude
diagram.
* The bar indicates that ja> has been substituted for s.
Sec. 1.12] electric networks for control systems 15
For the simple network of Fig. 1.6 we obtain the asymptotic representa¬
tion by taking the logarithm of the magnitude of the frequency ratio
[Eq. (1.8) in which s = ju>]. We obtain
Thus there are three components. The asymptotic characteristics for the
three components are shown in Fig. 1.9a, while the total characteristic for
the rate network of Fig. 1.6 is shown in Fig. 1.96. The dotted line again
represents the actual curve.
the asymptotic curve near the break point. The reader is referred to
texts on servomechanism theory for a complete discussion of this problem.1
The procedure is best illustrated by an example. Thus, consider the
transfer function
Ks(T2s -f* 1)
1 > Tl > T2 > T, > Ta > Tb
(Tis + l)(Tbs + 1 )(T4s + l)(Tbs -f- 1)
(1.15)
'out
db = 20 log (1.16)
Asymptotic amplitude
Physical Transfer
Names characteristic vs.
configuration function
frequency
Reset net¬ So 1
R work; inte¬ 6i RCs + 1
O—vWA
gral con¬
i troller; low-
pass filter
Proportional 50 R2 RiCis 1
51 R1 + R2 / R1R2 \
jii *
plus rate
network;
lead net¬
\Ri + R2J
C is + 1
work
■j£l
plus rate 6i l(Rl + R2)C2S + l] (g^j-Cu + l)
plus inte¬
R, (approximate; see Sec. 1.21)
gral con¬
troller ; lag-
c2
x lead net¬
work
network is not loaded. The transfer function for the last network given
is approximate and has been obtained by a technique explained in Sec.
. .
1 21
1.14. The Synthesis of RC Networks. Introduction. The problem
of synthesizing a physical network to meet a prescribed transfer function
is one that has engaged the attention of a long and distinguished list of
investigators. In general, the problem is complicated, and the methods
for solving it are correspondingly involved. Fortunately, a solution to
the general network-synthesis problem is not normally required in control
applications. One reason for this is that the transfer functions that are
encountered in control systems are usually quite simple, and ordinarily
18 CONTROL SYSTEM COMPONENTS [Chap. 1
Input Output
The subject of network analysis and synthesis is much too large and com¬
plex to be covered adequately in a few paragraphs, and the following
statements can serve at best as only a brief summary. For more exten¬
sive treatments the reader is referred to standard works1 on the subject.
Networks may be classified according to the number of terminals by
which they are connected to other circuits. Although a network may
have as many terminals as it has nodal points, most of the networks in
common use are two-, three-, and four-terminal networks, as shown in
Fig. 1.11. Note that a three-terminal network is a special type of four-
1 Bode, “Network Analysis and Feedback Amplifier Design,” D. Van Nostrand
Company, Inc., Princeton, N.J., 1945. Guillemin, “Communications Networks,”
John Wiley & Sons, Inc., New York, 1948, vol. 2.
Sec. 1.16] electric networks for control systems 19
terminal network in which both input and output have a common ground
connection. Networks with more than one set of input terminals are
occasionally encountered in control systems, particularly in places where
a number of signals must be added or subtracted. This type of network
will not, however, be discussed in any detail in this chapter. We shall be
concerned primarily with three-terminal networks, since this is the type
most commonly used in control systems.
A three- or four-terminal network may contain a large number of com¬
ponents, but it is usually possible and convenient to think of it as being
made up of a relatively small number of two-terminal networks. For
example, the ladder network shown
in Fig. 1.12 is made up of the two-
terminal networks Zi, Z2, Z3, Z4, etc.,
each one of which may contain a
number of resistors and capacitors.
The properties of the larger structure
are related to those of the two-termi¬ Fig. 1.12. Ladder network.
nal networks of which it is composed,
and it is, therefore, appropriate to consider first the properties of two-
terminal networks.
1.16. Properties of Two-terminal RC Networks. A two-terminal net¬
work is characterized by a single function, its driving-'point impedance,
defined as the ratio of the voltage applied to the two terminals to the
resulting current. In common with all other network functions, the
driving-point impedance is a function of frequency or, more generally, of
the complex variable s. The most important property of a driving-point
impedance of any passive network made up of lumped parameters such as
resistors, capacitors, or inductances is that it can be expressed as the ratio
of two polynomials in s as follows:
ansn + an^is"-1 + + no
z= (1.17)
+ 6m_ is™-1 + +
This may be demonstrated by writing all the Kirchhoff mesh equations
for the network and solving for the current as a function of the voltage at
any two terminals.1 Equation (1.17) may be written in the equivalent
factored form:
/
y _ K {s ~ ai)(s — a2) • ’ ' (s — an)
(1.18)
(« - 0l)(« -02)•••(«- Pm)
In accordance with standard convention the a s and p’s are referred to as
the zeros and poles, respectively, of the network function.
Guillemin2 has shown that the poles and zeros of RC driving-point
1 Guillemin, ibid., p. 208.
2 Ibid., p. 212.
20 CONTROL SYSTEM COMPONENTS [Chap. 1
impedances must be real and negative and that they must obey the
separation principle as follows:
This implies that the zeros and poles must be simple; i.e., factors of the
form (s — a)n cannot occur in the transfer function unless n = 1. It fol¬
lows that an asymptotic representation of an RC driving-point impedance
must take the form shown in Fig. 1.13 and that the actual driving-point
impedance must satisfy the relation
en = e<2, ci (1.21)
ZB
However ei = (1.22)
Za + Zb
Za
and e2 =
Za + Zb
A Fig. 1.14. Symmetrical bridge, or lattice,
e_o . Za Zb
Hence A
(1.23) network.
e: Za + Zb
Zo ? = H = (1.24)
6i
e* Zi +
We make the standard assumptions
Fig. 1.15 L-section type of RC
that the source impedance of the input
network. voltage source is zero and that the load
impedance is infinite at all frequencies.
The properties of the network are then easily deduced from the properties
of driving-point impedances derived previously. The restrictions on the
transfer functions obtainable with L sections may be stated as follows:
1. The poles and zeros must be negative and real.
2. The poles must be simple.
3. There can be no more than two successive zeros (or poles) without an
intervening pole (or zero).
4. The number of zeros cannot exceed the number of poles.
5. There cannot be a pole at the origin.
1 J. T. Fleck and P. F. Ordung, Realization of a Transfer Ratio by Means of an
R-C Ladder Network, Proc. IRE, vol. 39, pp. 1069-1074, 1951. Bower and Ordung,
op. cit.
Sec. 1.18] electric networks for control systems 23
^ d{log \H\) ^
(1.25)
~ d(logco) -
Equation (1.24) indicates that the poles and zeros of the transfer func¬
tion are produced by the poles or zeros of Z2 or Zx + Z2. Since these are
both driving-point impedances, the poles and zeros must be real and
negative, as noted in point 1 of the list above. Poles of the transfer func¬
tion result only from zeros of Zi + Z2, since a pole in Z2 is also a pole of
Z\ + Z2 and will therefore cancel out. Since zeros of driving-point
impedances must be simple, restriction 2 follows. Note, however, that
zeros may occur in pairs, because they are produced by either poles of
Zi + Z2 or zeros of Z2. Restriction 3 follows from the separation prop¬
erty of the poles and zeros of driving-point impedances. Thus suppose
that H has a pole due to a zero of Zi + Z2. The next singularity of
Zi + Z2 must be a pole; however, if this is a pole of Z2, no singularity in
H results. Zi + Z2 may now again have a zero, resulting in two succes¬
sive poles in H. The next singularity in H must now, however, be caused
by either a pole in Zi or a zero in Z2. Either of these singularities results
in a zero in H. Thus there can be no more than two successive poles. A
similar argument may be used to show that no more than two successive
zeros may occur. Restrictions 4 and 5 are restrictions on RC-network
transfer functions in general and need not be discussed again. Restric¬
tion 6 is related to inequality (1.20). The maximum rate of decrease of
the transfer function occurs when the numerator decreases at its maximum
rate and when the denominator is constant; the maximum rate of increase
occurs when the numerator is constant and when the denominator
increases at its maximum rate. Thus restriction 6 follows. Restriction
7 is almost obvious, since the network contains no amplifier, but it may
be shown more rigorously to be a consequence of the limitation on the
phase angles of Zx and Z2. Thus, suppose Eq. (1.24) to be rewritten as
fdllows:
Z2 1
H = (1.27)
Zi T Z2 (Zi/Z2) + 1
Since the phase angles of Z2 and Zi are restricted to the fourth quadrant,
24 CONTROL SYSTEM COMPONENTS [Chap. 1
we may write
-90° < < 90° (1.28)
Hence § + i>i
■^2
(1.29)
€\ = e%Hx (1.30)
Z,iy
ex = e,H2 Z OX “f" Ziy
(1.31)
i
T““l
x11 y nr ry
^ ox ”1 Aiy
A
to HxHy
or A (1.33)
ei 1 + ( Z0X/ Ziy )
It is clear that, if
Zox
«1 (1.34)
^7iy
e0
then A - HXHy (1.35)
ei
The maximum permissible value of \ZoxIZiy\ depends on the accuracy
desired, but a commonly accepted maximum is of the order of 0.1. Nor¬
mally only a very rough estimate of \Z0X/Ziy\ is required. If more than
two L sections are to be cascaded, the above argument is easily extended,
and inequality (1.34) must be met at each network junction.
Sec. 1.20] electric networks for control systems 25
dE
T 2 ~77 ~b E — 0 (1.44)
dt
Assuming that the voltage has some initial value Eq, the solution for E as
a function of time is
E = E0e-t/T> (1.45)
Similarly, we may solve for the current that flows when the two terminals
of the impedance are short-circuited together. Under this condition
E = 0, and Eq. (1.43) becomes
Tl d4 + I = 0 (1.46)
dt
or I = he-t/T i
(1.47)
28 CONTltOL SYSTEM COMPONENTS [Chap. 1
From Eq. (1.40) we see that l/7\ and 1 /772 are, respectively, the zero and
the pole of Z. Thus we arrive at the following conclusion:
When a driving-point impedance contains only one capacitor, it has
only one pole and one zero. The pole is the reciprocal of the discharge
time constant obtained on the open-circuited impedance, and the zero is
the reciprocal of the discharge time constant of the impedance when it is
short-circuited.
*2
-WVW-
if This rule enables us to write down
the driving-point-impedance func-
*1
o—AW/ -° tions by inspection of the circuit.
*3
Thus, consider the impedance shown
WWW
in Fig. 1.18. At low frequencies the
Fig. 1.18. RC impedance. capacitor is an open circuit; hence
the resistance is i?i + Rz. The pole is the reciprocal of (j?2 + Rz)C.
The zero is the reciprocal of [R1Rs/(Ri -f Rs) + R^C. Hence the imped¬
ance function is
Z = (7?i + /?3)
[R 2 + (RiR*)/(Ri + R-i)]Cs + 1
(1.48)
(R2 Rd)Cs + 1
In an L-section network like that shown in Fig. 1.15, a pole in the trans¬
fer function is produced only by a zero of Z\ + Z2 that is not simultane¬
ously a zero of Z2. Hence for a network containing only one active
capacitor a pole in the transfer function is always the reciprocal of a dis¬
charge time constant of the entire short-circuited L section. Zeros in the
transfer function are produced by zeros of Z2 or poles of Zi, since a pole of
Zi corresponds to a pole of Zi + Z2. Thus zeros are given either by the
discharge time constant of Z2 short-circuited, or of Zi open-circuited.
To synthesize an L-section network, we first draw up the network
qualitatively according to the following rules:
1. If the asymptotic representation of the transfer function has a zero
slope, the network has no active capacitors in it.
2. If the asymptotic representation of the transfer function has a +1
slope, a capacitor in Z\ is active, and if it has a — 1 slope, a capacitor in Z2
is active.
3. A zero in the transfer function at zero frequency requires a series
capacitor, not bridged by any resistance in Zx. A zero at infinite fre¬
quency requires a shunt capacitor across Z2.
The procedure is best illustrated by means of a few typical examples.
Consider the transfer function whose asymptotic diagram is shown in
Fig. 1.19. The network must contain two capacitors, one active in the
interval coi ^ co ^ co2 and the other active m the interval co3 < CO < C04.
We consider these two intervals separately. A network to synthesize the
transfer function to the left of co3 is shown in Fig. 1.20a. Note that at low
Sec. 1.21] electric networks for control systems 29
frequencies the capacitor acts as an open circuit; hence the network gain
is unity. At intermediate frequencies the capacitor becomes active and
reduces the gain, and at high frequencies the capacitor is short-circuited,
and the network gain remains con¬
stant at a reduced level. For fre¬
quencies above w3 the network gain
must increase. By rule 2 this re¬
quires a capacitor in Zi. If the net¬
work gain is to be less than unity
at high frequencies, however, the
Fig. 1.19. Example of transfer function.
capacitor must short out only part #
of R\. Thus the complete network shown in Fig. 1.206 results.
To complete the design quantitatively, we note that the break at cox
represents a pole in the transfer function of magnitude coi. This pole is
caused by a zero of the input driving-point impedance of the circuit of
o-AA/VA wvw-
*1 Rz
O-
Fig. 1.20a and is therefore the reciprocal of the discharge time constant
of that circuit when the input is short-circuited. Thus
— = RzCi (1.50)
co2
The next zero, corresponding to co3, is due to a pole of Z\. It is, therefore,
the reciprocal of the discharge time constant of Zi when it is open-
circuited;
— = R2Cz (1.51)
Finally, l/o>4 must be equal to the discharge time constant of the entire
30 CONTROL SYSTEM COMPONENTS [Chap. 1
network with the input terminals connected together and with C1 replaced
by a short circuit.
1 _ C^R^iRi Rz)
604 R1 -f- R2 T R3
Although other equations may be written about the circuit of Fig. 1.206,
it will be found that they are not independent from the four given above.
Thus we have only four independent equations. There are, however, five
unknown parameters in the circuit: two capacitors and three resistors.
This situation is encountered in all methods of transfer-function synthesis,
since the impedance level of the net¬
work is not specified by the process.
The value of any one component of it
the network may therefore be set :/?i
arbitrarily, depending on the require¬
ments of the remainder of the circuit.
ia)
Thus, suppose the circuit to be driv¬
ing the grid of a vacuum tube. Then
the maximum resistance between
grid and ground might be specified as Rz .Cs
o-MW-
less than 1 megohm, so that we would
have the additional requirement that
Rz + R\ < 106 ohms. If then we
arbitrarily let R\ -f- R2 = }/\2 megohm,
R. 'T''Co
ic)
Fig. 1.21. Network transfer function. Fig. 1.22. Steps in the design (see text).
we have five equations, which may be solved to find the values of the
five parameters. Since most of the equations are, in general, quite
simple, like Eqs. (1.50) and (1.51), it is usually possible by a judicious
choice of the arbitrary parameter to simplify the solution very consider¬
ably. Thus, in the foregoing example, if we assign a value to Ci, Eq.
(1.50) may be solved immediately for R3, and Eq. (1.49) for Ri + R2 + R3.
Then division of Eqs. (1.51) and (1.52) yields an equation having only R1
as the unknown, so that all the resistances may be obtained easily. C2
is then obtained from Eq. (1.51).
As a second example, consider the transfer function shown in Fig. 1.21
which has a zero both for zero and for infinite frequencies. The +1 slope
extending down to zero frequency requires a series capacitor in Z\ by rule
Sec. 1.21] electric networks for control systems 31
3. Hence the transfer function to the left of co2 is synthesized by the net¬
work shown in Fig. 1.22a. At the frequency a>i the capacitor Ci becomes
inactive. In order to obtain the — 1 slope between co2 and o>3, a capacitor
in Z2 must become active in this interval; however, unless there is a series
resistance in Z i, such a capacitor cannot have any effect on the transfer
function. Thus the network must take the form shown in Fig. 1.225.
The final break at oh must be produced by another capacitor in Z2, so that
the network of Fig. 1.22c follows. Note that a capacitance is shunted
across Z2 as required by rule 3.
To evaluate the R’s and C’s, we proceed as in the previous example and
obtain the following equations:
We have four equations but six unknowns. One more equation may be
obtained, as in the last example, by assigning an arbitrary value to one of
the components, but this still leaves one additional equation to be deter¬
mined. This additional equation is required to specify the gain level of
the network. Thus, suppose that the gain of the transfer function
between the frequencies 001 and w2 is set at g, where g < 1. Then since Ci
is assumed short-circuited and C2 and C3 open-circuited in this frequency
interval, we have
_ Rz 4 Ra
~
(1-57)
R2 4 Rz 4 Ra
~ ~
Thus we have the necessary six equations, which may now be solved for
the six components.
It should be noted that the question of gain level did not appear in the
first example because the level was tacitly assumed to be unity at fre¬
quencies below coi. Had a level of less than unity been desired, then an
additional resistor (probably across C\ in Fig. 1.205) would have been
placed in the circuit. An additional equation would then have been
needed to specify this resistance.
Normally one is interested in as high a gain level as possible, since low
gain must be made up with amplifying circuits. The question naturally
arises, therefore, as to how large g may be made. For a transfer function
of the form shown in Fig. 1.19 the answer is obvious: the maximum gain
is unity. However, for network functions of the type shown in Fig. 1.21,
32 CONTROL SYSTEM COMPONENTS [Chap. 1
the answer is not obvious, and it cannot be obtained from the approximate
synthesis procedure described here. All that can be said at this point is
that g < 1, since, by Eq. (1.57), Ro = 0 if g = 1.
The essential difference between the transfer functions of Figs. 1.19 and
1.21 is that the latter contains two successive poles, while the former does
not. It will be shown in connection with the exact synthesis method to
be described subsequently (Sec.
1.23) that this question of gain level
arises generally only when two suc¬
cessive poles are specified for a
transfer function. An approximate
idea of the maximum value of g per¬
Fig. 1.23. Transfer function with two missible in such cases may, there¬
successive poles.
fore, be obtained by considering a
transfer function having only two poles. The asymptotic diagram of
such a transfer function must have the form shown in Fig. 1.23, and an
analytic expression of H is
__ _ Zo _gcc2s_
H = (1.58)
Zi Z2 (s + o?i)(s -f- C02)
This may be rewritten as
Z\ + Z 2 _ 1 _|_ Z\ _ (s + COi)(s + 0)2)
(1.59)
Z2 Z2 geo 2S
Z1 S2 -T (cO 1 + CO2)s “b COiO>2 — ge02S
so that (1.60)
Z2 ge^2S
The roots of the numerator must be real and negative. By a short
algebraic manipulation it may be shown that this requirement leads to
9 < 1 -
S
A plot of this function is shown in
(1.61)
_R 3 + (1/6*1$)_
R 3 + (1/Cis) + /2i + [Rz/iRzCzs + 1)]
_(RzCis T 1)(R2C2S -j- 1)_
(R\ + RzjRzC2Cis2 -f- (R1C1 + R3C1 -f* R2C1 T I£2(72)s 4- 1
34 CONTROL SYSTEM COMPONENTS [Chap. 1
R2C2 = —
co3
(1.64)
flsCl = -
C02
Substituting these values in Eq. (1.62) and equating Eqs. (1.62) and
(1.63), we obtain
1 1 1 / 1
CO1C04
+ — H-Js + 1 = — ( — + R\C\ ] s2
CO 1 CO4/ CO3 \co2
R&i = (1.66)
CO16O4 C02
Substituting this value for R\C\ in Eq. (1.65) and equating the coefficients
of s, we obtain finally
= —
CO 1
+ --—
CO4 CO1CO4
- —
CO3
(1.67)
Equations (1.64), (1.66), and (1.67), along with the specification of one of
the R’s or C’s to provide the desired impedance level for the network,
constitute a set of five equations, from which the five network parameters
are easily obtained.
This method may be used for other simple networks having only a few
poles and zeros. As the networks become more complicated, considerable
ingenuity may have to be exercised in particular cases to make the sub¬
stitutions indicated above. Also the method does not optimize the gain
level for transfer functions having two successive poles, like that shown
in Fig. 1.21. Such optimization would have to be a trial-and-error proc¬
ess with larger and larger values of gain assumed until an unrealizable
situation (a negative value for an RC product, for instance) appears.
Despite its shortcomings, the method has considerable merit where appli¬
cable, since it is considerably simpler than the general exact synthesis
method to be described in the next section.
1.23. The Exact Synthesis of L Sections. General Method.1 The
L section considered in this section is that shown in Fig. 1.15. Suppose
1 J. L. Bower, P. F. Ordung, and J. T. Fleck, “The Synthesis of Resistor-Capacitor
Networks,” Yale \Jniv..Dept. Elec. Eng. Tech. Rept., August, 1948, pp. 21-24.
Sec. 1.23] electric networks for control systems 35
that the desired transfer function meeting all the restrictions for L sec¬
tions discussed above has the form
Z\ + Z2 _ , Z\ _ 1 (s d~ T~ 02)(s H~ ^3)
(1.69)
Z2 Z2 h (s + ai)(s + «2)(s + 0:3)
Defining
(s + + &2)(s + ^3)
F(s) A (1.70)
(s + ai)(s + 0:2) (s + ol 3)
we find that
Zi 1
(1.71)
h [F(s) ~ h]
From this equation it is clear that poles of Z1 or zeros of Z2 result from
poles of F(s), while zeros in Z\ or poles in Z2 result whenever F(s) — h = 0.
The a’s and 7*s are known, but K1 and K% are not yet determined. The
ratio K1/K2 is found most easily by using Eq. (1.71) evaluated at a con¬
venient value of s. Thus, for s = — 00 ? Eq. (1.71) becomes
Zi Kl _ 1 _ 7N
(1.74)
Z2 k2 h(1 h)
Note that the partial-fraction expansion is valid only for proper fractions;
hence the impedance function of Eq. (1.73) must first be converted by
division into the sum of a constant and a
proper fraction, as shown in Eq. (1.75). -WAV-
The constants i£3 and K± are found by
standard techniques. Equation (1.76)
suggests that Z2 consists of three com¬ C
1L
ponents in series. The first component
may be identified as a resistance K2, while Fig. 1.27. Resistor and capacitor
in parallel.
the other two both consist of a parallel
combination of a resistor and a capacitor, as shown in Fig. 1.27. This
may be shown by evaluating the impedance of this combination:
R/Cs 1/C
Z = (1.77)
R + (1 /Cs) s + (1 /RC)
K-
3/7i
Kt4//?
-vWW- -vWW-
K*
1L _1 (_
13 !/K,
• /‘4
Fig. 1.28. Synthesis of Z2 (first method).
S ($ + 7i)(s + 72)
K2 $($ + <22) ($ + 0:3)
A1
+ +
D,
K2 \ S
131 H-j-1-j-
$
B2S
I
S -f 0:2
0:2
B3s
S + 0C3
$4-0:3,
(1.78)
where B1 = A1/K2, etc. The form of Eq. (1.78) suggests that the admit¬
tance Y2 consists of three components in parallel. The first of these is the
conductance Bh and the other two both consist of a capacitor and resistor
in series, since the admittance of such a combination is given by
Cs/R
(1.79)
Cs + (1 /R)
Hence by the second method the impedance is synthesized in the form
shown in Fig. 1.29, with resistance and capacitance values again given in
ohms and farads.
We may now complete the design OCy
of the transfer function given in Eq.
(1.68). Using the first method of
driving-point-impedance synthesis,
we obtain the network of Fig. 1.30,
in which the values of resistances
and capacitances are given in ohms
and farads, and where K1/K2 is
given by Eq. (1.74).
Fig. 1.29. Synthesis of Z2 (second Fig. 1.30. Network having transfer func¬
method). tion of Eq. (1.68).
work. In Fig. 1.316 and c are shown the a diagrams of two L sections,
which together will realize the transfer function of Eq. (1.80). The two
component transfer functions are
(s + 20) (s + 50)
x x (s + l)(s + 200)(s + 500)
(1.81)
(s + 10) (s + 50)
v v {s + 2)(« + 200)
In making the distribution of poles and zeros between the two component
network sections, we keep in mind that the gain level between two succes¬
sive poles has an upper bound for realizable transfer functions (see Fig.
Fig. 1.31. a diagrams of over-all transfer function and component transfer functions.
networks of poles and zeros that do not exist in the over-all transfer func¬
tion. The reader is referred to some of the problems at the end of this
chapter which illustrate these principles (see, for instance, Prob. 1.14a).
We proceed by first synthesizing Hx. A sketch of F(s) for the transfer
function Hx is shown in Fig. 1.32, with the horizontal line for the maximum
s F(s)
-0.1515 85.67
-304 85.62
-308 85.67
-310 85.65
-320 85.06
258 \
Zlx = 0.01167AT,
s + 50/
( 0.06447 0.935 \
Z2x = K 2x
\s + 0.1515 + s + 308/
Sec. 1.25] electric networks for control systems 41
The complete design of Hx is shown in Fig. 1.33, in which the values are
expressed in ohms and farads. K2x is left arbitrary to permit adjustment
of the impedance level.
The design of Hy proceeds in a completely analogous manner. A sketch
of F(s) is shown in Fig. 1.34. The maximum value of hm&x in this case is
F(0), because larger values of h would require zeros or poles in one of the
impedances for positive value of s. This is ruled out by the restrictions
on realizable impedance functions. The intersection of the horizontal
line for hmax with F(s) at s = —770
0.0602K2x
can be easily determined in this ex¬
ample by a direct algebraic solution.
The distribution of zeros and poles
K ij/ 1
(1 - 0.8) = 0.25 (1.85)
K 2y 08
720
Hence Z\y = 0.257^22/ (1 +
8 + 50,
10 A 2 V
Z 2y = K^y +
The final design of Hy is shown in Fig. 1.35, with resistances and capaci¬
tances in ohms and farads.
The final step in the design is to determine the ratio of K2x to K2y so
that, when Hx and Hy are connected, the second L section will not load
42 CONTROL SYSTEM COMPONENTS [Chap. 1
the first one excessively (see Sec. 1.19). In general, this step calls for a
calculation of the ratio of the impedances of the two sections, as required
by Eq. (1.34). Since one order of connection may have advantages in
the size of capacitors required, it is good practice to compute the circuit
for both orders of connection.
The impedance ratio ZoxIZiy for the case when network Hx precedes Hy
is most easily found from Eq. (1.38).
Zix z2x n
^ 0 K2y (s + 0.1515)(s + 10)(s + 50)(s + 308)(s + 770) , Q .
K2x (s + 1 ){s + 2)(s + 200)2(s + 500) 1 ;
The maximum value again occurs for 1 < o> < 2 and has a value of less
than 40K2y/K2x- Thus, if the ratio of the impedances Zoy/Zix is to be
less than 0.1, K2x/K2y > 400. The
3.6 KZy resulting network is shown in Fig.
-VWW"—
0.25 K^y 1.37.
-VWW— The final decision as to which cir¬
cuit is better depends now on prac¬
-1(-
0.00555'
Kty
tical circuit considerations. Thus,
K,2y
if the network were to drive the
.OJ_
grid of a vacuum tube, a maximum
K2y d-c resistance between grid and
ground of 1 megohm might be speci¬
Fig. 1.35. Final form of Hy. fied. This specification would be
met in the circuit of Fig. 1.36 with
a A of about 3 X 106, while in Fig. 1.37 K would be about 3 X 104.
(It is assumed that the network driving source has zero d-c resistance.)
Under these conditions the largest capacitance in Fig. 1.36 would be about
5.1 ijlf, while in Fig. 1.37 it would be only 3.3 gf. Since large capacitors
are bulky, the network of Fig. 1.37 has a slight advantage. In general,
it will be found that the size of the largest capacitors is minimized by
Sec. 1.26] ELECTRIC NETWORKS FOR CONTROL SYSTEMS 43
0.0602 K 0.216 K
—A/VW— -WWA-
0.01167 K 0.015 K
■wwv— t-\A/VW-< -o
Input Output
-l(- Hf—
0.332 0.0926
K K
0.06 K
0.426K
1.67
K
0.00304 K
3.6 K 24.1 K
placing the network section with the lowest frequency poles toward the
end of the cascade.
1.26. Bridged-T and Twin-T Networks.1 In this section we discuss
two networks that are commonly used to obtain transfer functions with
complex zeros. The bridged T is the simpler of the two networks, but
the complex zeros produced by it can have only a relatively small imagi¬
nary part. The twin T is more complicated but also more flexible, per¬
mitting the synthesis of transfer functions with purely imaginary zeros
(i.e., complete rejection of one fre¬
._\(°1_ quency). The twin T also can be
designed with a nonminimum phase
o- >-VWW--VWVv—1 --o type of transfer function; i.e., it
/?, Ro
may have zeros with positive real
Input cz Output
parts.
The form of the bridged T is
shown in Fig. 1.38. The transfer
Fig. 1.38. Bridged-T network.
function can be found by applica¬
tion of standard techniques (the node method gives the result most
rapidly) and can be put into the form
1
(1.89)
" = = VRiRiCxC*
It is convenient, therefore, to normalize the frequency variable s by
defining
v = — = s
a>c
(1.90)
With these definitions the transfer function may then be put in the form
p2 + 2fp + 1
H(p)
a + l\ (1.92)
v2 + 2r + V + 1
2r )
Physical realizibility requires both f and a to be positive real numbers.
Thus f cannot be made equal to zero, and the network cannot completely
reject the frequency coc. In fact, it is seen from Eq. (1.92) that, as f and
therefore the minimum gain of the network are decreased, the two poles
Sec. 1.26] electric networks for control systems 45
of the transfer function move farther and farther apart, and therefore the
phase lag introduced by the network at frequencies well below coc becomes
larger. This is illustrated by the attenuation and phase-shift curves
0.1 0.2 0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 10.0
"/ft/c
(a)
given in Fig. 1.39. Since the complex zeros produced by the network are
quite often used to cancel a set of complex poles produced by some other
component of the loop, the large phase lag produced at low frequencies is
sometimes objectionable.
It should be mentioned that, although the bridged-T network of the
46 CONTROL SYSTEM COMPONENTS [CHAP. 1
form shown in Fig. 1.38 seems to be most commonly used, it is also possi¬
ble to construct the network with re¬
sistors and capacitors interchanged.
The basic form of the transfer func¬
tion as given by Eq. (1.92) is not
changed by this procedure.
The twin-T network is shown in
Fig. 1.40. Its transfer function is
Pig. 1.40. Twin-T network. again most easily found by use of
the node method and can be put into the form
I = Hw = m (L93)
where
N(s) = fli7?2fi3CiC2CV + (Ri + RJRzCiCzs* + 728(Ci + C2)s + 1 (1.94)
and
A(s) = R\R2R3C\C2C3s3 + [(Ri T RijR^C\C2 -f- (R2 -f- Rz)R\C2C3
R iR %C\C s]s2 + [Ri(C2 + C 3) + R2C2 + Rz{C\ + C2)]s + 1 (1.95)
N(s) =
uc2y
+ OJr + - + 1 (1.96)
7
where y is the third zero, which must be negative and real. Comparison
with Eq. (1.94) indicates that
1 _ RiR2R3C\C2C3
(1.98)
^7 “ -Ri(Ci + C2)
C3 _ R9(Ri + R2)
(1.99)
C1 -b C2 R1R2
where n is any real, positive number. It will be shown presently that n
has an optimum value depending on the ratio of Ri to R2 and that of C1 to
Sec. 1.26] electric networks for control systems 47
«! = «1 (1.101)
1 + j8
and using this result and Eq. (1.100) in Eq. (1.97), it is found that
RiC1 (1.102)
s
(1.103)
JT / 1 +
\ _ P“
(1.104)
~ H{v) ~ 1 + Kp +
n( 1 + a) + «(1 + ff)
where (1.105)
y/&(3n
Thus, for a given null frequency the circuit behavior is completely deter¬
mined by n, by the ratio of R2 to R1, and by the ratio of C2 to C\.
Normally it is desirable to have as sharp a null as possible at the rejec¬
tion frequency. The sharpness of the null may be defined in a number
of ways, but a convenient one is to consider the slope of the magnitude
ratio \H(ju/uc)\ at the null as a measure of sharpness. Accordingly we
48 CONTROL SYSTEM COMPONENTS [Chap. 1
define the optimum value of K as the value giving the steepest slope of
|H(jco/c0c) | for co = coc. By differentiation of the magnitude of Eq. (1.104)
with p set equal to jco/wc, it is found that the slope at co/coc = 1 is — 2/K.
Hence the optimum value of K is the smallest realizable value. In deter¬
mining this value, we must keep in mind that a, (3, and n must be positive
real numbers. It is convenient to express K as the sum of two terms:
K = (1.106)
Note that the two terms indicated by the parentheses are both positive
and that K is therefore minimized by minimizing each one separately.
The first term has the form (x + 1/x). By differentiation this function
is easily shown to be minimum for x = 1. Hence one condition for mini¬
mum K is
afi _ j
n (1.107)
or n = a(3
K = 2 + a + i (1.108)
where cos is the relatively low signal frequency at which the shaft of the
synchro is being oscillated, while coc is the carrier frequency, which is
100
-100
0.1 0.2 0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 10.0
[&)
showing that the signal consists of two sidebands, one at the frequency
o)e — cOs and the other at the frequency coc + cos. Similarly the output
signal may be expanded into the form
Tp
e0 = {cos [(<oe — ccs)t — 6] + cos [(coc + cos)t + 0]} (1.113)
Thus we see that the function of the network must be to provide phase lag
at frequencies below the carrier frequency and phase lead at frequencies
above the carrier frequency. Inspection of Figs. 1.39 and 1.41 indicates
that both the bridged T and the twin T will perform this function, pro¬
vided that they are tuned so that their null frequency coincides exactly
with the carrier frequency. Note, however, that they perform the oper¬
ation only approximately, since their characteristics are symmetrical with
respect to the logarithm of frequency, while the desired characteristic, as
shown by Eq. (1.113), should be symmetrical with respect to the frequency
itself. This is not, however, a serious objection if the signal frequencies
are low with respect to the carrier frequency. A much more serious diffi¬
culty with this type of series equalization is that the performance depends
very critically on the carrier frequency, and even relatively small shifts
from the nominal value result in a serious loss of performance. This is
due to the fact that the network characteristics depend on the absolute
frequency of the signal rather than on its ratio to the carrier frequency.
To examine this problem quantitatively, and also to indicate an
approach toward a design procedure, suppose that the signal frequency is
given by
ct>s = ccoc (1.114)
Then the upper- and lower-sideband frequencies are, respectively,
wc(l + c) and xoc(l — c). The discussion is illustrated by using the
Sec. 1.26] electric networks for control systems 51
bridged-T network but is easily extended to the twin T (see Prob. 1.19).
The transfer function is given in Eq. (1.92). In order to get the frequency
function, let p — ju = jco/coc so that the normalized upper- and lower-
sideband frequencies become simply j(l + c) andj(l — c), respectively.
Substituting into Eq. (1.92) and replacing 2f + (a + l)/2f simply by
2fd, we obtain, after some algebraic rearranging,
c(c + 2)
1 +j
H (ju) 1 2r(c + i) (1.115)
U c(c + 2)
i +j
2?d(c + 1)
for the upper sideband and
c(c — 2)
i - j
H(ju) 1 2r(c - i) (1.116)
td c(c - 2)
i - j
2Uc ~ 1)
for the lower sideband. When the network is to be used to provide phase
lead in a servo, is made very much larger than f. Hence, for small c the
denominator is approximately equal to unity, and the transfer function
becomes approximately
c(c + 2)
H(jco) « £ 1 ±3 (1.117)
2f(c ± 1).
With the same value of as before, the phase lead of the upper sideband
is about 63.5°, while the phase of the lower sideband is zero. By invert¬
ing the process used to derive Eq. (1.113), it can be shown that the signal-
frequency phase shift is approximately equal to one-half the difference
between the phase shifts of the sidebands, which in this case amounts to
about 31°. Thus a small shift in carrier frequency results in a large loss
of phase lead. A further effect is that the carrier phase is also shifted by
about 30°. As shown in Chap. 7, this results in reduction of torque gener¬
ated by the motor and excessive heating.
Although the commercial 60-cps supply frequency is usually maintained
very close to its nominal value, this is not true of most 400-cps-circuit
supplies. The frequency of currently available aircraft supplies cannot
be guaranteed to be regulated to better than ± 10 per cent, or ±40 cps.
Such a frequency variation will make networks of the type discussed here
quite useless. The regulation problem of aircraft supplies may be solved
in the next few years, thus making high-performance a-c servos practical
in aircraft.
1.27. The Use of Amplifiers with RC Networks.1 The variety of
transfer functions available with RC networks can be extended greatly by
using networks in conjunction with amplifiers. The subject is too large
to be dealt with completely here, and only some of the possibilities are
indicated by the following examples.
A common procedure is to use networks in the return path of a feedback
amplifier, as indicated in Fig. 1.42.
If the amplifier gain is G and if the
Amplifier
network transfer function is H,
then the transfer function through
e{
the entire circuit is
A
Network en G
1 + HG
Fig. 1.42. Feedback amplifier with RC
network in feedback path. 4 if HG » 1 (1.118)
Thus, provided the loop gain is sufficiently high, the approximate transfer
function is the reciprocal of the network transfer function.
This principle can be applied in the generation of transfer functions
having complex poles by making the network in the return path of the
feedback amplifier a twin T. Qualitatively such an amplifier will have
a gain of unity at very low and very high frequencies, but at the null fre¬
quency of the twin T, the gain is G, since the feedback is not operative
there. Clearly the transfer function is not the exact reciprocal of the
twin-T transfer function. Even though the transfer function of the twin T
1 Valley and Wallman, op. cit., pp. 391-408.
Sec. 1.27] electric networks for control systems 53
has imaginary zeros, that of the feedback amplifier cannot have imaginary
poles, since this would require infinite amplification at the null frequency.
The discrepancy can, however, be made as small as desired by making the
gain of the amplifier, G, large enough.
Another interesting application of the twin-T circuit in connection with
a feedback amplifier is shown in Fig. 1.43. This circuit is referred to as
the rejection amplifier, and it has the effect of sharpening the null of the
r (P2 + 1)
ea _ (p2 + Kp -f -1) v2 + i (1.119)
* ~ , g (p2 + i) K
V + 1
^ (P2 + Kp + i) «?+l)
where G is the amplifier gain. Comparison with Eq. (1.104) indicates that
the effect of the circuit is essentially to change the K in the denominator
of the transfer function of the twin T to K/{ 1 + G). In the discussion
of the previous section it was shown that the slope of the magnitude of the
frequency function at the null was —2/K for the twin T; hence it is clear
that the action of the amplifier is to make the slope steeper, and therefore
to sharpen the null.
In both the examples given, the transfer functions that were obtained
are of the type that could also be obtained by the use of inductances.
Thus, complex poles are commonly obtained by simple RLC resonant cir¬
cuits, while the sharp rejection characteristic of the second example may
be shown tQ be identical to that of a bridged-T network in which one of
the resistances is replaced by an inductance.1 This bears out the state¬
ment made in Sec. 1.14 that, by the use of RC networks with amplifiers,
practically any transfer function that can be realized with resistances,
capacitances, and inductances can be realized without actually having to
use inductances.
A final example illustrates the use of a difference amplifier (see Sec.
2.11) to generate a nonminimum phase transfer function. The circuit
is shown in Fig. 1.44.
1 Valley and Wallman, op. cit., p. 385.
54 CONTROL SYSTEM COMPONENTS [CHAP. 1
Fig. 1.44. Use of difference amplifier to produce nonminimum phase transfer function.
Designers must be encouraged not to design new circuits for each new
device but to use standard circuits that are mass-produced and under rigid
quality control.
The second cause of unreliable devices is the attempt by the designer
to wring every last bit of performance from electronic components. The
statistical life history of a batch of components or complete units, what¬
ever their type, usually falls into a pattern such as that shown in Fig. 1.46.
Fig. 1.45. Over-all reliability of a device as a function of the reliability of the individual
components.
The curve may be divided into three regions. Region A, or the infant
mortality region, consists of failures that occur in the break-in or shake-
down period. The Navy’s insistence on a shakedown cruise and specifica¬
tions by buyers for 10 to 20 hours operation of a unit at the factory are
both good procedures, as is seen from this chart. Next comes region B, the
operating region. If the unit is well designed, the failures that occur in
Fig. 1.46. Typical life history of a group of devices or components. Region A is the
break-in region, B represents the operating region, and C is the wear-out region.
this region are few and random in nature and time of occurrence. Finally,
region C, or the wear-out region, is a bell-shaped curve in a well-designed
piece of equipment.
It has become a vogue in the guided-missile field to attempt to design
equipment for short operating life. The argument goes that the unit
need operate for only an hour or two at the most; thus, it is wasteful to
Sec. 1.28] electric networks for control systems 57
design for more. We can see the fallacy of this argument from Fig. 1.46.
The break-in period blends into the wear-out region and there is essen¬
tially no period of trouble-free operation. We can categorically state
that it is impossible to build a reliable, short-lived piece of equipment,
and our missile program will be successful only when this is realized.
In general, the designer only shortens region B when he pushes compo¬
nents up to and beyond their design limits. It has been found, for exam¬
ple, that operating vacuum tubes at 120 per cent of their rated plate dissi¬
pation will result in about 200 per cent the failure rate of tubes operated
at their design value. Several manufacturers have recently set up as
standard-design procedure that vacuum tubes are to be operated at a
maximum of 50 per cent of their rated plate dissipation. This has
resulted in a 4:1 improvement in reliability.
The third cause of unreliability is operation under difficult environ¬
mental conditions. Extrapolation of present trends indicates that
environmental specifications, such as ambient temperature, shock, and
operation at low atmospheric pressures, will continue to become more
severe. Derating of conventional components will soon fail to provide
an operating margin. Extensive research on new materials and methods
for component manufacture is imperative.
Finally, unreliability is due to improper maintenance. This may be
the result of improperly trained technicians, lack of thought on the proper
layout for ease of maintenance, or simply, equipment complexity.
Usually all three causes are at work. One solution gaining favor is the
use of modular packaging. Modular packaging of subunits of an assem¬
bly reduces repair to simply finding the troublesome module and replacing
it. This is not only quicker and easier but is apparently more economical
in the long run.
PROBLEMS
1.1. By means of resistance padding, construct the best approximation to the square
law y = 2x2, where x is shaft position in radians and y is resistance in ohms. Let
1/max = 10,000 and use a maximum of three segments.
1.2. It is desired to construct by voltage padding the response function shown in
Fig. 1.47. Use four intermediate voltage taps.
Radians
58 CONTROL SYSTEM COMPONENTS [Chap. 1
y = x — x2 + 0.5rc3
fc(0.1s + 1) (0.05s + 1)
a‘ s(0.01s + 1)(0.00001s2 + 0.001s + 1)
, s(5s + l)(0.2s + 1)
(10s + l)(s + 1) (0.1s + 1) (0.02s + 1)
1.5. A network consists of two sections, H i and H2. The output impedance of
Hi is given by
Aq(0.5s + 1) (0.02s + 1)
(2s + l)(0.1s + 1) (0.01s + 1)
^ k2(s + 1) (0.05s + 1)
i2 s(0.25s + 1)
Find the ratio of k i to k2 such that section H2 does not load section Hi.
1.6. Two lead networks (see Table 1.2) are to be cascaded to produce an over-all
transfer function with a +2 slope in the asymptotic a diagram. The attenuation at
low frequencies is the same for the two networks. Show that the ratio of impedance
levels required to prevent loading between the two networks is a maximum when the
two lead networks have the same break frequencies coi and co2.
1.7. Write down by inspection the impedance functions of the networks shown in
Fig. 1.48.
R
WVvV-
R C
o-MW-|f
c
(o)
id)
Rz
WW—
—AW- —AW—4
Rz c
< >—v/VW—o
c ft*
y WWW
ic) id)
Fig. 1.48
1.8. For the driving-point impedances containing more than one capacitor shown
in Fig. 1.49, write the impedance function by inspection by assuming that for any
particular frequency only one capacitor is active. Also assume that the active zone
for C2 comes at a higher frequency than that of Ci, etc.
Problems] electric networks for control systems 59
Rz
R\
—AW—
C\
"](■
*3 Cz
■VWW-
■tt
[o)
R\ Rz
[—vWv^- -MW-
0.2s (0.05s + 1)
a. H
(s + 1) (0.01s + 1)
(0.5s + 1) (0.05s + l)2
b. H
(2s + 1) (0.2s + l)(0.01s + 1)
,_s(0-2s + 1)_
c. H (,h is to be as large as possible.)
(s + 1) (0.05s + 1) (0.01s + 1)
(0.1s + 1) (0.025s + 1)
d. H
(5s + 1) (0.01s + 1) (0.002s + 1)
1.10. Design the networks of Prob. 1.9a and b by the exact method described in
Sec. 1.22.
1.11. Use two L sections to synthesize the following transfer functions. Use the
approximate method described in Sec. 1.21 to obtain the L sections, and adjust the
ratio of impedance levels to prevent loading.
1.25s2(0.4s + l)2
a. H
(10s + 1) (2s + l)(0.1s + l)2
(0.05s + l)2(0.01s + 1)
b. H
(s + 1) (0.005s + l)2
(s + 1) (0.01s + 1)
c. H
(0.1s + l)2
(0.0143s + l)3
d. H
(s + l)(0.1s + 1) (0.001s + l)2
s2(0.005s + 1)
e. H (Design for maximum h.)
h (0.1s + l)2(0.05s + 1) (0.02s + 1)
1.12. Use the exact method of Sec. 1.22 to synthesize the individual L sections for
the transfer functions given in Prob. 1.11. Compare the resulting designs to those
obtained in Prob. 1.11.
1.13. Design a two-terminal impedance to have the following impedance function:
(0.02s + 1) (0.0025s + 1)
Z = 10,000
(0.05s + 1) (0.01s + 1)
1.14. Use the exact synthesis method described in Secs. 1.23 to 1.25 to obtain the
following transfer functions. Use the minimum number of L sections. In all cases
design for maximum h.
1.16. Find the input impedance and output impedance of the bridged-T network
of Fig. 1.38. Assume the source impedance is zero and the load impedance infinite.
1.16. Repeat Prob. 1.15 for the twin-T network of Fig. 1.40, in which Ri = R2 =
YzRz and Ci = C2 = 2C3.
1.17. A twin-T network for which a = (3 = n — 1 is to be cascaded with a simple
low-pass network (Table 1.2, second row). The pole in the low-pass network is to
coincide with the null of the twin-T network. Find the ratio of impedance levels
required to prevent appreciable loading between the two networks. Assume that the
low-pass filter follows the twin T.
1.18. By the use of a bridged-T network together with one other L section, design
a cascaded network to synthesize the following transfer function:
Check that the two sections do not load each other appreciably.
1.19. Show that a twin-T network tuned to the carrier frequency of an a-c servo
system has an effect on the carrier envelope that is similar to the effect of a perfect
differentiating circuit on a low-frequency signal.
1.20. Design a bridged-T network to provide a 30° phase lead at a signal frequency
of 6 cps for a servo using a 400-cps carrier. What is the effect of a 10 per cent shift
in carrier frequency on the network characteristic?
1.21. An improperly tuned bridged-T network produces a phase shift of du in the
upper sideband and a phase shift 6i in the lower sideband. Show that the effective
phase lead of the envelope frequency is approximately A (6U — 6i). Also show that
the phase shift of the carrier is approximately A(0u + 0j).
CHAPTER 2
D-C AMPLIFIERS
2.2. Drift in D-C Amplifiers. There are three major factors that cause
drift in a d-c amplifier. These are:
1. Variations of power-supply voltages
2. Changes of components with temperature and age
3. Effects of heater-voltage variation
Only the third of these is assumed to be unfamiliar. Drift is caused by
heater-voltage variations because changes in cathode temperature result
in changes of the initial velocities of the electrons emitted, so that the
electrode voltages required to maintain a given electron flow must change.
It has been found that, if the plate current is small compared to the total
emission from the cathode, this effect is essentially independent of the
plate current. The heater-voltage effect is best expressed in terms of the
amount by which the cathode voltage must change relative to the other
electrode potentials in order to hold the plate current constant. For
oxide-coated cathodes, this amounts to about 0.1 volt for a 10 per cent
change in heater voltage about the normal value, and it is relatively
independent of the tube type.1
Various methods are used to counteract these effects. An obvious
remedy is to regulate carefully all power supplies including heater sup¬
plies. High-quality components, wire-wound resistors, etc., will tend to
minimize the effect of component variations. A number of cancellation
methods have been devised to minimize the effect of cathode-temperature
variation, and some of these will be described in detail below. These
methods usually make use of the fact that the difference between the
voltage changes required in two tubes to keep the plate current constant
is much less than the change re¬
quired in either tube alone. When
b -~7^°.
_ 60 +^ they are applicable, push-pull cir¬
(a)
cuits are very effective in reducing
drift from all three causes, particu¬
larly if matched tubes and compo¬
nents are used. Negative feedback
is usually ineffective, since it does
not ordinarily affect the signal-to-
[t>)
noise ratio of an amplifier. This is
Fig. 2.1. Block diagrams of circuits used
for the calculation of the effect of nega¬
demonstrated below.
tive feedback on drift. 2.3. Effect of Negative Feedback
on Drift. Suppose that the ampli¬
fier shown in Fig. 2.1a has a gain G and generates an internal noise having
the value v at the output. (Drift may be thought of as noise of very low
e0 — Gei + v (2.1)
C\ @e0 (2.2)
Therefore e0 = G\Ge\ + Gv i + v
— GiGci — GiG(3e0 Gvi T v
_ G\G . Gv i v
(2.3)
1 T GiG/3 1 1 T GiG(3 1 T GiG/3
Gffl » 1 (2.4)
~+ i1,1 + oh ” (2-5)
In order to have a valid comparison between the feedback amplifier and
the original amplifier, we make the gain of both amplifiers the same; i.e.,
1//3 = G. Then
We see that, although the original noise v has been reduced by the gain
of the preamplifier, Gi, an additional noise term (G/Gi)vi has been added.
If we assume that the amount of noise added by an amplifier is propor¬
tional to the amplifier gain, i.e., if v\ = (Gi/G)v, then we see that the feed¬
back amplifier has about the same noise output as the original amplifier.
The above argument does indicate, however, that if the preamplifier
were noise-free, the noise output of the feedback amplifier would be con¬
siderably reduced. It also indicates that, in the design of a feedback
amplifier, all care should be taken to make the first stages as noise-free as
possible.
Getting back to the problem of drift, we shall find that it is possible to
construct d-c amplifiers with zero drift and very high d-c gain. The gain
does, however, drop sharply as the frequency is increased to values as
high as 1 or 2 cps. If such an amplifier is used as a preamplifier in a feed¬
back circuit like that of Fig. 2.16, the drift introduced by the main ampli-
64 CONTROL SYSTEM COMPONENTS [Chap. 2
fier may be reduced to as small a value as is desired. The fact that the
gain of the amplifier drops at higher frequencies is of no consequence, for
as long as the inequality (2.4) is satisfied, the gain is given by 1/(3 and is
essentially independent of the gain of the forward path of the feedback
loop. This principle is employed in the chopper-stabilized d-c amplifier1
described in detail later.
2.4. Analysis of Simple Triode Circuits. Before considering the opera¬
tion and design of complete amplifiers, we must examine the behavior of
single stages rather completely, with particular emphasis on the fact that
they are to be used in d-c amplifiers. The quantitative effect on drift of
the three factors listed in the last section, as well as questions of gain and
terminal impedances, will be discussed.
where the voltages epk and egk are the plate and grid voltages, respectively,
measured relative to the cathode. The tube may therefore be represented
egk
T
ek
by the equivalent circuit shown in Fig. 2.3, where the arrows on the
voltages represent the assumed positive sense.
This circuit is accurate only for normal cathode temperature, since it
does not provide for the insertion of the cathode-temperature-variation
effect. The initial velocity of the electrons results, approximately, in a
“virtual” cathode voltage differing from the cathode voltage measured
at the terminal. The effect of heater-voltage
variation may, therefore, be most easily repre¬
sented in the equivalent circuit by a voltage source
en inserted between the cathode terminal and the
cathode, as shown in Fig. 2.4. Hence the equiv¬
alent circuit of Fig. 2.3 may be generalized to
include heater-voltage variation by considering
the terminal marked “cathode” as the “virtual”
cathode inside the tube and by connecting a eg
voltage source eh between this terminal and the
outside connection. The voltage eh may be de¬
fined as T
Fig. 2.5. Triode equiv¬
dek alent circuit with heater-
eh Ae/
de/ iP = c voltage effect.
where Aef is the change in filament voltage. This leads to the equivalent
circuit of Fig. 2.5, which applies to all triodes. For this circuit the equa¬
tion for the plate current becomes
66 CONTROL SYSTEM COMPONENTS [Chap. 2
where egk is the voltage between the grid and the virtual cathode, as
shown in Fig. 2.4.
2.5. The Cathode Follower. Cathode followers are used primarily as
impedance changers, since they offer a very high input impedance and
very low output impedance. The voltage gain is slightly less than unity.
The actual circuit is shown in Fig. 2.6a and its equivalent in Fig. 2.66.
The output voltage is ek, measured between cathode and ground. The
current iQ is the load current and is included for generality. It will be
noted that the cathode is not returned to ground but to a negative
voltage Ekk- This is necessary if the output voltage ek of the cathode fol¬
lower is to swing through negative as well as positive values of voltage.
The complete equation for this circuit may be derived by use of Kirch-
hoff’s current law. The expression for the current leaving the node ek is
+ Ch l -—b —) ~■ i0 — 0 (2.11)
'Tp ^PI
Rk rp
_ uRkeg + Rk(Ebb ~ Ex) rpEkk + + Rkrpi0 — (1 + fi)RkCh
(2.12)
Rk(u 1 + )
Sec. 2.6] D-C AMPLIFIERS 67
dek fj.Rk
G = 1 (2.13)
deg (n + 1 )Rk + rp M 4" 1
The approximations made to obtain the final result are that (/x + l)Rk ^>rp
and n^>> 1. For a typical triode (6SL7) common values are /x = 70,
Rk = 200,000 ohms, rp = 50,000 ohms; hence these approximations are
very good. Similarly we may find the output impedance
Common values for the output impedance of a cathode follower are from
500 to 1,000 ohms, corresponding to a gm of from 1,000 to 2,000 nmhos.
Note that the value of Rk does not appear in the approximate expression
for the output impedance.
The drift due to changes in Ebb or Ekk is found from Eq. (2.12) to be
dek 1 _ 1
(2.15)
1
dEbb (m + 1 )Rk + rP jU + 1 H
dek rp TP ^ 1
and (2.16)
dEkk (m T l)Rk + Tp (/x + 1 )Rk gmRk
Note that drift from these two sources is rather small, especially if Rk is
large. The drift due to heater variations is, however, large; it is given by
dek _ — (1 + ii)Rk
(2.17)
deh (1 + n)Rk + rp
input impedance. Since the two leakage paths from the grid are in
parallel, it is more convenient to consider the input admittance. We
define leakage admittances between grid and plate and between grid and
cathode as shown in Fig. 2.7. These admittances are assumed to be
measured on a dead tube, i.e., on a tube for which \x = 0.
The current ig must be equal to
dig
— Ygp + ( 1
deg (
— Ygp + 1 - Y gk
(m + 1 )Rk + tP/
\r* + rP)Y gk Y gk
= Ygp + Ygp “F (2.19)
(/J, + 1 )Rk + Tp +1
where the assumption has been made that Rk » rp, so that Rk + rp « Rk.
Thus we see that, although the cathode fol¬
lower has no effect on the grid-to-plate
admittance (which is the same as the grid-
to-ground admittance, since the plate is at
“signal ground7’), the grid-to-cathode ad¬
mittance, measured when /x — 0, is reduced
approximately by the factor 1/(1 + /i)
when the tube is operating. This result
should be.compared with that obtained with
a conventional grounded-cathode amplifier,
where the input admittance is equal to the
Fig. 2.7. Input-admittance
paths in cathode follower.
sum of the grid-to-cathode admittance plus
the grid-to-plate admittance multiplied by
the gain of the stage.1 This comparison shows the cathode follower to
have a much lower input admittance than an amplifying stage.
2.7. Drift due to Component Variation. Equation (2.12) can also be
used for computing the drift caused by component variations. Thus, for
instance,
d@k Rjc[(Rk H- Rk:(Ebb Ex) EkTpto Vp(Ekk "I- 6h)\ /.-) <)i
dJI [(A* + 1 )Rk + rj2 1 j
These relations and similar ones for other component variations are
useful, not only for the computation of drift, but also to determine the
maximum range of deviation of the output voltage as a result of compo¬
nent tolerances. If such a range is known, it becomes a relatively simple
matter to specify the range of balancing potentiometers required in the
amplifier. Note that, in the application of formulas for drift due to com¬
ponent variations, all voltages and currents involved in the circuit must
in general be known.
2.8. Cathode Follower with Plate-load Resistor. Occasionally, a
plate-load resistor RL is inserted between the plate and the positive supply
voltage of a cathode follower. This occurs particularly in situations
where the tube operates simultaneously as an amplifier stage and a
cathode follower. The analysis derived thus far applies to cases of this
sort almost without change; all that is required is the replacement of rv in
all the relations by rv + Rl. This does, however, have the result, par¬
ticularly if Rl is large, that a number of the approximations made to
simplify the final expressions are no longer completely valid.
2.9. Equivalent Circuit for the Cathode Follower. In the analysis of
complicated circuits involving cathode followers, it is sometimes con¬
venient to make use of an equivalent circuit. Such a circuit may, of
course, take any number of forms. It is merely required that the expres¬
sions for gain, drift, and impedance derived from the equivalent circuit
correspond to those derived previously. The circuit shown in Fig. 2.8 is
suggested for situations where only the gain, output impedance, and
drift rate due to Ebb, Ekk, and eh are of interest. This circuit does not give
the correct answer in problems involving component variations. The
boxes marked n, fi + 1, and l/(/x + 1) are perfect amplifiers with zero
output impedance. The input and output voltages, as well as the output
current, are to be thought of as voltage and current variations about some
fixed, unspecified quiescent value. For this reason it is not necessary to
70 CONTROL SYSTEM COMPONENTS [Chap. 2
determine the effect of the cathode resistor. Thus, for the circuit shown
in Fig. 2.9, the plate current is
This could have been found also by inspection of Fig. 2.10, since all
voltage sources in that figure are assumed to have zero impedance.
The drift due to Ehb is
de p r:
dEbb Rl + En
rP T (/x -f~ 1 )Rk
(2.30)
Rl + rp -j- (n + l)Ri
dei (m + 1 )Rl
(2.31)'
deh Rl + tp -j- (ix -f- 1 )Rji
Here we find again, as with the cath¬
ode follower, that filament-voltage Fig. 2.10. Equivalent circuit of triode
variation causes a very large amount amplifier with cathode resistor.
of drift, since the effective cathode
voltage en acts on the tube in approximately the same way as the grid
voltage, which is the desired signal.
Drifts due to variations of the components may be obtained by differ¬
entiating Eq. (2.27) with respect to the desired component. Specific
computations are left to the reader.
72 CONTROL SYSTEM COMPONENTS [Chap. 2
dep _ (m + 1 )Rl
(2.32)
de'k Rl + rp + (yu + l)Rk
Comparison of this equation with (2.28) shows that the gain for the two
methods of signal introduction is about the same, the gain for cathode
input being (yu + l)//x times that for grid input. More important is the
fact that the sign change associated with amplification when the signal is
applied to the grid does not occur when the signal is applied to the cath¬
ode. This fact is of importance when it is desired to apply negative feed-
back around an amplifier.
One disadvantage of applying the
signal to the cathode is that the input
impedance is very low. The input
impedance may be deduced from one
of the results obtained for the cath¬
ode follower; i.e., it is the sum of Rk
and the output impedance of a cath¬
ode follower with a plate-load resistor
Rl and an infinite cathode-load re¬
sistor. Figure 2.9a shows the cir¬
cuit; the impedance in question is
that “seen” by ek. Hence
Z, = Rk + r* +,R,l (2.33)
Fig. 2.11. The difference amplifier. M + 1
This is usually much less than the input impedance at the grid.
2.11. The Difference Amplifier. We have seen that one of the most
serious causes of drift is the variation of heater voltage, with its influence
on the effective cathode voltage. A number of methods have been
devised to counteract this effect, one of the most useful being the differ¬
ence amplifier, shown in Fig. 2.11. The two triodes may be in separate
envelopes;however, the two sections in a single envelope are usually better
matched and will, therefore, give better results than two triodes picked at
random.
To analyze the circuit, we suppose that the cathode of the left-hand
section is disconnected at the point marked x in the diagram. The right
half of the tube is then a standard cathode follower, and if for the moment
we consider only the- effect of the grid voltage on the output, we have
Sec. 2.11] D-C AMPLIFIERS 73
_CgZlJ-zR’k_
(2.34)
0*2 + 1 )Rk + T p2
The subscript 2’s refer to the right-hand triode. Also the impedance to
ground seen by the left-half triode is the output impedance of the cathode
follower. It is given by
Z = (2.35)
(1 + M2)Rk + rP2
Hence the circuit may be redrawn in the equiv¬
alent form of Fig. 2.12. The output voltage
e0 for this circuit as a function of eg\ and eg2 is
given by the combination of Eqs. (2.28) and
(2.32), with Rk in both equations replaced by Z
of Eq. (2.35). Thus,
— niRLCgi T- On T 1 )R
L (/*2 + 1 )Rk + VV2
T p2Rk
Rl + Tp\ + Oil + 1)
(m2 T" 1 )Rk + Tp2
(2.36) for difference amplifier.
not, in general, reduce the drift from this cause completely to zero.
Since no two tubes have exactly the same characteristics, a circuit which
is to be insensitive to heater-voltage variation must contain an adjust¬
ment to accommodate different tubes. Such a circuit was described by
Miller1 and takes the form shown in Fig. 2.13.
The analysis proceeds as for the difference amplifier. We assume the
existence of the heater-effect voltages em and ek2 in the two sections of the
tube and, for the moment, assume the left-hand section disconnected at
point x. The right-hand section may then be considered as a cathode
Ri _ R\
6g2 p i p &k p' (2.39)
III "T II2 /t/c
(see Fig. 2.135). The voltage at the break x may then be deduced by the
superposition of Eqs. (2.13) and (2.17). It is necessary, however, to
replace rv by rp2 + Rl2 in accordance with the discussion in Sec. 2.8. Also
eQ in Eq. (2.13) must be replaced by Eq. (2.39). The result is
ek =
— (^2 + l)Rkeh2 + /JL2Rk(Rl/Rk)€k
(m 2 + 1 )Rk
Ch1 — 6h2
Rk(ll 2 + 1) + Tp2 + R L2 — II2R1
(2.41)
Cm = ac^2 (2.42)
we can solve for the value of R1 that results in zero drift. The result is
the circuit [see Eq. (2.49) below]; hence it should not be larger than
necessary.
In practice it is undesirable to make the cathode resistor, Ri R2 = Rk,
constant, as implied in the previous analysis, since any adjustment of the
variable tap affects the bias voltage of both grids, making it very difficult
to prevent either one or the other of the two tube sections from cutting off.
A more practical form of the circuit is shown in Fig. 2.15. R% and R±
together form the effective plate load of the right half section and serve to
reduce the effective Ebb seen by this sec¬
tion. Thus ip2 is kept small and does not
produce a large bias through the common
cathode resistor. R2 is constant and small,
just enough to keep grid 2 slightly nega¬
tive with respect to its cathode. (A small
amount of grid current in the right-half
section is actually of little consequence.)
Rli should be large, both for maximum
gain and for wide operating range, since
the adjustment of Ri required for drift
Fig. 2.15. Practical form of Mil¬
compensation does move the operating-
ler circuit.
point of the left-half section over fairly wide
limits. (It has only a small effect on the right half.)
The condition for perfect balance now becomes
R$R 4
where
R% R4
The reader may convince himself, by using typical values in the circuit
of Fig. 2.15, that reasonable variations of a can be compensated for with¬
out causing tube cutoff.
In order to find the gain of the circuit or its sensitivity to changes in the
positive supply voltage, it is necessary first to compute the effective
cathode resistor seen by the amplifying section. This is best done by
writing an expression for the voltage change ek as a function of a current ik
flowing into the right-half section (see Fig. 2.136). Using Eqs. (2.13),
(2.14), and (2.39) and the fact that rp must be replaced by U>2 + R'l2j We
have
_ ikRk{rp2 T R'L2) , ii2Rk{R\/Rk)ek
Rk(v 2 + 1) + ^p2 + R'l-2 tf*(M2 + 1) + Tp 2 + R'l 2
_ 1_ikRk(rp2 T R'l2) ____
(2.46)
Rk(n2 + 1) + rP2 + R'l2 ~ M2-R1
Sec. 2.14] d-c amplifiers 77
*
Hence the effective cathode resistor of the left-half section (R0 of Fig. 2.14)
is
£k _ _Rk(rp2 T R'l2)_
(2.47)
ik Rk(n2 + 1) + 'f'p‘2 + R'l2 — M2^1
R„ = (2.48)
M2 + 1
where a is defined as in Eq. (2.42). Now the expression for gain found
previously for the triode amplifier with cathode bias [Eq. (2.28)] may be
used, with the result:
de7 M 1R LI
G =
der, R LI + Tpi + (mi + l)(^p2 T" RfL2)a/ (M2 + 1)
HiRli
if mi = M2 (2.49)
Rli + Tpi + a(r P2 + R'lz)
R2 T" Ri _ ijlRl
(2.57)
R2 — Ri Rl 4~ tp
_ jlR L@i
(2.58)
Rl 4~ 4“ (m 4“ l)Iik-(rP 4~ Rl)/[(h 4~ 1 )Rk 4- rp + Rl]
80 CONTROL SYSTEM COMPONENTS [CHAP. 2
In computing ep2, becomes the eg2 of Eq. (2.36) and egi is zero; hence
The simplest method for doing this is the bias battery (see Fig. 2.20).
If the quiescent voltage of the plate is eb volts and that of the grid ec volts,
the battery voltage must be eb — ec volts. Ideally the battery has zero
internal impedance and perfect voltage
stability; under these conditions the cou¬
pling is close to ideal and has the following
advantages:
1. No signal loss.
2. No power-supply drain.
3. Low-impedance coupling, hence good
frequency response and small noise pickup.
Fig. 2.20. Interstage coupling
4. Coupling introduces no time lag or with bias battery.
frequency distortion.
In practice this coupling has several serious disadvantages, and it is used
only very rarely. These disadvantages are:
1. Batteries have limited life.
2. Voltage does not remain constant, results in drift.
3. Practical batteries are physically large; hence the capacitance to
ground is large, and frequency response is adversely affected. Bulkiness
is also a very important disadvantage
when miniature size is a requirement.
Another coupling that can be used is
the gas-tube coupling (Fig. 2.21). In a
voltage amplifier only very little power
will be handled by the coupling, and for
this reason the smallest neon tubes, such
as the NE-2, are used. These tubes have
a voltage drop that is almost completely
independent of the current through the
Fig. 2.21. Gas-tube interstage tube, as long as the region of normal glow
coupling network.
is not exceeded.1 The voltage drop va¬
ries considerably from tube to tube, rang¬
ing from about 49 to 74 volts with an average value of 62 volts. The
current may vary from 0.03 to 0.3 ma, but the design value is 0.1
ma. In the design of this coupling, one must keep in mind that
glow tubes require a relatively high voltage to strike the arc. For the
NE-2 neon tube this is about 90 to 100 volts. Although the voltage is
nominally independent of current, there is a small increase as the current
is increased; the effect is as though there were an internal resistance of
about 10,000 ohms. The fact that ionization is involved in the conduc¬
tion processes through the tube gives rise to a very pronounced time lag
less rugged than vacuum tubes; i.e., they are easily damaged or destroyed
by relatively minor overloads. A further disadvantage is the large varia¬
tion of parameter values among units having nominally the same charac¬
teristics. This problem will undoubtedly become less severe as manufac¬
turing and selection processes are improved. Transistors cannot be used
at very high radio frequencies but this is of no consequence in a control
system. Finally, transistors are extremely sensitive to nuclear radiation.
Since transistor action takes place by means of minute atomic impurities
(the positive and negative charge carriers referred to below), any dis¬
turbance of this action by radiation particles has a significant effect on
the operation of the device.
A junction transistor may be thought of as a sandwich of three sections
of semiconductor crystal. The central section is referred to as the base
while the outside sections are the emitter and collector respectively. In
most transistors the junction between collector and base has a larger area
than that between emitter and base, but except for this the junctions are
identical. There are two types of transistors. In the p-n-p type the
collector and emitter consist of material in which the electric current is
carried by holes, or positive charge carriers, while in the base the cur¬
rent is carried by electrons or negative carriers. The other transistor type
is the n-p-n, where the base consists of positive material and the emitter
and collector of negative material.
The operation of a transistor may be explained by thinking of the two
junctions as rectifying junctions. The collector junction is biased in the
reverse direction, and in the absence of emitter current only a very small
inverse current flows in the collector circuit. The emitter junction is
biased in the forward direction, so that only a small voltage between
emitter and base is needed to cause relatively large emitter currents to
flow. The flow of emitter current through the base results in the produc¬
tion of a large number of charge carriers within the base, and these are
picked up by the collector, resulting in a flow of current in the collector.
Thus, we may define transistor action as the control of current in the
negatively biased collector junction by current in the positively biased
emitter junction. (The terms “positive” and “negative” are correct for
p-n-p transistors; they would be reversed for n-p-n transistors.) An
important transistor parameter is a, the ratio of collector current to
emitter current. In junction transistors a is slightly less than one, typical
values ranging from 0.95 to 0.98. In point-contact transistors a normally
exceeds one.
2.19. Analysis of Simple Transistor Circuits. In analyzing transistor
circuits it is convenient to assume small signal operation about a fixed
operating point just as in vacuum tubes. In the following discussion we
consider only junction transistors, since point-contact transistors are not
86 CONTROL SYSTEM COMPONENTS [Chap. 2
Wc
0 v
ocu Ie
rb =
A
die
dvi
ie = const.
(2.63)
= — rb (2.64)
di e ic = const.
ic)
Fig. 2.25. Transistor in grounded-base
and
connection: (a) actual circuit; (5) a-c dV‘
rc. — — rb (2.65)
equivalent circuit; (c) d-c equivalent dir ie — const.
circuit showing lco.
In the definition of rx we prefer to
introduce the current gain a already mentioned above. For the directions
assumed for the currents ic and ie, this parameter is defined by
dic
a = — (2.66)
dL V2 = const.
since the actual collector current in a p-n-p transistor is out of the collec¬
tor. From Eq. (2.62) we obtain that
die
(2.67)
die V2 — const. rb -F rc
Hence we have that
= a
rb + rc
or rx = a(rb + rc) (2.68)
Sec. 2.19] D-C AMPLIFIERS 87
A possible equivalent circuit from which Eqs. (2.61) and (2.62) could have
been obtained directly is shown in Fig. 2.256. This equivalent circuit is
not quite correct for d-c amplifiers since it indicates that ic = 0 when
ie = 0. In practice a small current Ico flows in the collector circuit even
when ie = 0. To account for this current the circuit of Fig. 2.256 may
be modified as shown in Fig. 2.25c.1 The magnitude of Ic0 depends on
the transistor and the temperature. In a typical transistor (General
Electric 2N43A) Ic0 ranges between 5 and 10 ^a at a temperature of
Fig. 2.26. The grounded-emitter circuit: (a) actual circuit showing load resistance and
battery; (6) equivalent circuit.
25°. The temperature dependence of Ico is one of the major factors caus¬
ing transistor d-c amplifiers to drift. The variation of Ic0 is given
approximately by
La = I cooek{T~ro) (2.69)
where Icoo = Ic0 at the temperature T0, and k varies from 0.06 to 0.09 with
an average value being 0.08.* This relation is valid for T0 of about 25°C
and T — T0 less than 50°C. It should be noted that for k = 0.07, Ico
doubles for every 10°C change in temperature.
Since the grounded-base connection discussed up to now has a current
gain of less than one, direct coupling of stages cannot produce current
gains greater than one, and the voltage or power gain for a multistage
amplifier would therefore be the same as for a single stage. For this
reason the most commonly used connection is the grounded-emitter con¬
nection shown in Fig. 2.26a. This connection is closely analogous to the
grounded-cathode connection for vacuum tubes. To obtain the equa¬
tions for this connection we may redraw the equivalent circuit of Fig.
2.256 in the form shown in Fig. 2.266. Note that the input current is
now ih = ii = —ie — ic, and the output current is ic = as before.
While it would be possible to write the equations for this circuit in the
same form as Eqs. (2.61) and (2.62), i.e., with the voltages as the depend-
1 Hunter, “ Handbook of Semiconductor Electronics,” McGraw-Hill Book Com¬
pany, Inc., New York, 1956, pp. 13.2-13.4.
* Ibid.
88 CONTROL SYSTEM COMPONENTS [Chap. 2
Solving for and i2 results in Eqs. (2.70) and (2.71), and by direct com¬
parison we find that
re[re + rb — a(rb + rc)] re
hie = re + rb — + rb
rc + rb + re — a{rb + re) 1 a
re re
hre
rc + rb + re - a(rb + rc) rc( 1 — a)
re + rb — a(rh + rc) a
hf e (2.72)
rc + rb + re — a(rb + rc) 1 — a
1 1
hoe
re + rb + rc — a(rb + rc) rc(l — a)
In making the approximations, advantage has been taken of the fact that
rc is usually very large, in the megohm range, while rb and re are typically
less than 1,000 and 50 ohms respectively.
The equations for the transistor in terms of the h parameters and the
subscript notation used are now being accepted as standard by industry.
The subscripts i, r, f, and o refer to input, reverse, forward, and output
respectively, while the subscript e refers to the fact that the parameters
are defined for the grounded-emitter connection. For the grounded-base
and grounded-collector connections the subscripts 6 and c are used. A
double subscript notation using the numbers 1 and 2, i.e., hu, hi2, h21, h22,
instead of h{, hr, hf, h0, is also in common use, and for some applications
z and y parameters are occasionally used.2
<3
M
Typical values:
hie ~ 2,000 ohms
V\ hre - 6x 10'4
hfe — 50
hoe- 2.5 x 10~5 mhos
o o
io)
Uf)
Fig. 2.27. Equivalent circuit for grounded-emitter circuit in terms of h parameters:
(a) exact circuit; (6) approximate circuit.
neglected. Similarly the voltage hreVz in the input circuit is usually much
less than the voltage drop through hie, so that the input impedance is
approximately hie. Thus the very simple approximate equivalent circuit
of Fig. 2.27b results. In this circuit the input circuit is completely
decoupled from the output, making the transistor a unilateral current
amplifier with the current gain hfe.
-15
-A. = PSD u nmn
^-
200
o -10
E 150
S" 100
-5
50
-0
0
0 -10 -20 -30
TC I
volts
that two families of curves are required: an input family and an output
family. From the input family one obtains hie, the slope of the curve at
the chosen operating point, and hre, the vertical distance between two
curves for which the difference in collector voltage is unity. Similarly,
from the output family one obtains hfe, the vertical distance between
curves for which the difference in base current is unity, and hoe, the slope
of the curve. From the shape of the curves it is apparent that the
parameters hie and hfe can be obtained with good accuracy, while hre and
hoe are difficult to get accurately. Fortunately, the important parameters
are the ones that can be obtained accurately. It should also be noted
that the output characteristics are almost parallel and equidistant for
constant increments of base current. This means that the transistor is
a good linear amplifier of the input current; i.e., a sinusoidal input current
will result in sinusoidal output current or voltage. On the other hand,
due to the curvature of the input characteristic, a sinusoidal voltage
applied to a transistor will in general yield a nonsinusoidal output. It is
therefore usually desirable to drive the base from a high-impedance, or
constant-current, source. It should be pointed out that the h parameters
are basically small-signal parameters just like the n, rp, and gm of a vacuum
tube, and that they vary with the operating point. They also vary with
the operating temperature. Thus the typical values given in Fig. 2.27
can vary by factors of two or three or more. Most manufacturers of
transistors do, however, supply charts showing what this variation is.
The equivalent circuit of Fig. 2.27a again does not take into account
the fact that with zero base current the collector current is not zero. In
order to obtain an equivalent circuit that properly takes into account all
of the d-c conditions of a transistor, we may go back to the circuit of
Fig. 2.25c, and convert it to the common-emitter form in terms of h
parameters. The details of this conversion are left to the reader, but it
results approximately in introducing an additional voltage source
— Icohfehre/hoe into the input circuit, and an additional current source
— hfeho into the output circuit (see Prob. 2.9).
2.20. Single-stage Circuits. The current gain, voltage gain, power
gain, input impedance, and output
impedance of a transistor may be
easily obtained from the four-ter¬
'R, minal representation shown in Fig.
2.29. In this representation it is as¬
sumed that the currents i\ and i2 and
Fig. 2.29. Four-terminal network. the voltages V\ and v2 are related by
Eqs. (2.70) and (2.71).
In order to find the current gain, we note that in Fig. 2.29, v2 = —i2Rl-
Substituting the value of v2 in Eq. (2.71), we immediately find that
Sec. 2.20] D-C AMPLIFIERS 91
hf e
(2.73)
1 "f" hoeR’L
hfe{R l/hie)
(2.74)
Vi h fe'vre
{eh
1 T" Rlhoe
hneh
oe,,jie
(hfehre) {hie)
y. = = h oe 1 - hoe (2.77)
Vi {hiehoe) {hie ~b Rg) _
The power gain is simply the product of current and voltage gain. It
should be noted that the expressions for current gain, input impedance,
and output impedance reduce approximately to one of the h parameters.
Also the voltage gain is approximately equal to the product of current
gain and ratio of load resistance to input resistance. These approximate
results could have been obtained directly by use of the approximate
equivalent circuit of Fig. 2.27h.
Although the grounded-emitter circuit is the one used most often, the
grounded-collector and grounded-base circuits are also used occasionally.
The form of these circuits for p-n-p transistors is shown in Fig. 2.30. For
n-p-n transistors, the battery voltage should be reversed. The simplest
way to obtain the various gains and impedances for these circuits is first
to compute new h parameters for them. These h parameters may then
be used instead of the common-emitter parameters in Eqs. (2.73), (2.74),
(2.76), and (2.77). This is possible since these equations describe the
characteristics of the generalized four-terminal network of Fig. 2.29. It
must be noted, however, that the approximations made in these four
equations apply in general only to the grounded-emitter connection.
92 CONTROL SYSTEM COMPONENTS [Chap. 2
^2e P 2c
%le he
he he he
^ lc hiehc “h (1 hre)V 2c
he (1 —b hfe)hc ~b hoeV 2c
h^c hiie
(2.78)
hfc — — (1 + hfe)
hoc
Table 2.1
Approximations
Zi Zo Gi Gv
•
made
Common 1 , Rl RLhoe « 1
hie hfe hfe h.
emitter hoe ttie Rg/hie » 1
° < h/ir’
flie'voe
<1
for details of the procedure. For the common emitter circuit, Shea’s
results can be summarized as follows:
1. The poorest situation results when the base is driven from a high-
resistance source, and the emitter is grounded.
2. Placing a resistor in the common emitter lead improves stability.
3. Reducing the output resistance of the driving source improves
stability.
4. A feedback resistor connected between the collector and base
improves stability.
It should also be noted that a large load resistor in the collector lead
tends to protect a transistor from thermal runaway, since an increase in
Ico causes a decrease of collector voltage. This is true even though the
voltage stability factor as defined by Shea increases with RL.
Another method of reducing drift is to use a difference amplifier1 as
shown in Fig. 2.31. This circuit is
completely analogous to the vac¬
uum-tube difference amplifier dis¬
cussed in Sec. 2.11, and its output
is effectively equal to the difference
in the two base signals multiplied
by a gain factor. Also, as far as
changes in Ico are concerned, the
output will be a function of the dif¬ Fig. 2.31. The transistor difference
ference of the currents in the two amplifier.
transistors. Thus the circuit will
discriminate against variations in Ico in the same way as the tube circuit
of Sec. 2.11 discriminates against changes in eh. The complete analysis
of the transistor version of the circuit is left as an exercise for the reader.
The difference amplifier is only one special case of using a transistor as
a compensating element to reduce drift in a d-c amplifier. Other methods
of using transistors, diodes, and temperature-sensitive elements in this
way are described in the literature.2 Most of these methods require that
the amplifier be adjusted for least drift after construction.
2.22. Multistage Amplifiers. When transistor amplifier stages are cas¬
caded to form a multistage amplifier it is necessary, just as in vacuum-
tube amplifiers, to design interstage coupling networks which will permit
proper voltage and current biases to be applied to the coupled transistors.
This problem is, however, often much simpler in transistor amplifiers than
in vacuum-tube amplifiers. This is due to three factors. First the col-
r _ _Aj_ Rg _ E\Rg
R\ + (RgRs)/(Rg + Rs) Rg + R% R\{Rg -f- Rz) + RgR%
Since Eh Ie, Rg, and R% are known, Ri may be found. The operating
point of the first stage, and therefore the d-c level of the amplifier output,
can be varied by adjusting Ri.
The second stage is an n-p-n stage; hence the base must be biased in the
positive direction. If the positive voltage A3 + E2 is very much greater
in size than the negative collector voltage of the first stage, the gain
loss in the coupling network is small. This fact would dictate a choice of
small collector operating voltage for the first stage. R± and A5 are
Sec. 2.23] D-C AMPLIFIERS 97
designed so that the base of the second stage operates at the correct value
of quiescent emitter current.
In the coupling between the second and third stages no resistance is
needed at all since the collector of the second stage and the base of the
third stage can both operate at the same positive voltage (approximately
E2). As a matter of fact, in low-level stages it is sometimes possible to
dispense also with resistor R$ and to let the input impedance of the third
stage act as the collector load resistance for the second stage. A very
simple circuit results.
The emitter of the third stage is connected to the positive voltage E\
so that the output voltage e0 can be zero for zero e*. Instead of using a
separate battery E2, a bleeder may be used between the positive supply
and ground, or more elegantly, a silicon diode (Zener diode) may be used
instead.1
2.23. The Chopper-stabilized D-C Amplifier. Although the difference
amplifier and the Miller circuit discussed in previous sections are capable
of reducing drift to a value low enough for d-c amplifiers of moderate gain,
these relatively simple circuits cannot meet the requirement for almost
zero drift that is made in the most critical situations. D-c amplifiers used
in analogue computers or amplifiers used to amplify the extremely minute
voltages generated by thermocouples are representative of this class of
problem.
For such applications it is necessary to “ detour ” around the drift prob¬
lem introduced by d-c amplifiers and to use a-c amplifiers instead. A
typical circuit to accomplish this is shown in block-diagram form in Fig.
2.33. Here the signal is passed first into a modulator, which converts the
d-c (or very low frequency) signal into a relatively high frequency signal
that may then be amplified in a standard a-c amplifier. Since a-c ampli¬
fiers with very high gain are relatively easy to construct and since they
do not drift, a very large amount of drift-free amplification may be
obtained, provided that the modulator itself is fj’ee of drift. After the
signal has been amplified to the desired level by the a-c amplifier, it is
passed through a demodulator to recover the original d-c, or low-fre¬
quency, information.
Although a number of electronic modulator and demodulator circuits
1 Hunter, op. cit., p. 13.10.
98 CONTROL SYSTEM COMPONENTS [Chap. 2
are available (see Chap. 6), most of them are subject to some drift.
Hence, for critical applications the electromechanical vibrator, or chop¬
per, is used (see Fig. 2.34). These devices have in recent years achieved
a very high degree of perfection; they are hermetically sealed and suffi¬
ciently rugged to withstand large amounts of acceleration, shock, and
(6)
Fig. 2.34. Electromechanical chopper: (a) disassembled; (b) schematic.
all three contacts are connected together. With this type of chopper
either the output or the input is always grounded. There is therefore no
possibility of a feedback connection being established between output and
input through the capacitance existing between the open vibrator con¬
tacts. Such feedback might result in undesirable oscillations.
The operation of the circuit is best understood by assuming the input
signal to be a low-frequency sine
wave and examining the wave¬
shapes at various points in the cir¬ Time
cuit, as shown in Fig. 2.36. Figure
2.36a shows the sinusoidal input
voltage. If it is assumed that the
input impedance of the a-c amplifier
is infinite compared to the resist¬
ance Ri in Fig. 2.35, the input volt¬
age (a) appears at the input to the
amplifier without attenuation when
the input contact of the chopper is
open; otherwise the voltage is zero.
This situation is illustrated in Fig.
2.365. Note that the amplifier
Fig. 2.36. Waveshapes of chopper-
input voltage still contains a com¬ stabilized amplifier.
ponent at the low signal frequency.
However, if we assume that the amplifier does not pass this low
frequency, then the amplifier output voltage will have the form shown in
Fig. 2.36c.1 This output will be passed to the low-pass filter only when
the output chopper contact is open; hence there results the voltage shown
1 It is shown in Chap. 6 that the amplitude of the signal component of the wave¬
shape of Fig. 2.366 is one-half of the input amplitude. The high-pass filters in the
a-c amplifier effectively subtract this component. The reader may convince himself
by performing this subtraction graphically that Fig. 2.36c is obtained.
100 CONTROL SYSTEM COMPONENTS [Chap. 2
in Fig. 2.36d. The low-pass filter serves to remove the chopper frequency
from the output, so that the final result is ideally an amplified reproduc¬
tion of the input (Fig. 2.36a). Note that, in the example carried through
here, the amplifier was assumed to have an even number of stages, so that
there is no sign reversal between the voltages of Fig. 2.366 and c. The
final reversal between input and output is due to the fact that the chopper
inherently applies a signal to the input at the same time that it shorts the
output, and vice versa. It should be clear that an odd number of ampli¬
fier stages would have resulted in a reversal between signals 6 and c and
no reversal between a and d.
Although the amplifier described above and others operating on the
modulator principle make it possible to achieve very high gain with
negligible drift, they have the very serious disadvantage that they can
amplify only very low frequency signals. The reader may easily prove
to himself by the method used in connection with Fig. 2.36 that, if the
input frequency had been equal to the chopping frequency, the output
would have been a d-c voltage. This indicates that amplification of sig¬
nals at chopper frequency is impossible (see also Chap. 6). Even for
signal frequencies of the order of one-tenth of the chopping frequency it
would be difficult to design a low-pass filter to follow the amplifier which
would remove the carrier and pass the signal without too much attenu¬
ation. The obvious means for alleviating this difficulty, i.e., using higher
chopper frequencies, is limited by the fact that most choppers are built
to operate at either 60 cps or 400 cps and that it is very difficult to build
an electromechanical device to operate reliably at frequencies in the
upper audio range. Hence, when it is desired to have high gain with
zero drift over a wide frequency band extending down to zero frequency,
it becomes necessary to employ a two-channel arrangement, like that
shown in Fig. 2.37. Here the signal is split into high- and low-frequency
components by means of the input filters shown preceding the amplifiers
in the two channels. Actually, the high-pass filters shown with the high-
frequency channel might be the usual RC coupling networks used in a-c
amplifiers. The crossover frequency at which the signal shifts from the
Sec. 2.23] D-C AMPLIFIERS 101
high-frequency channel to the low-frequency channel must be low enough
so that negligible energy at chopper frequency enters the modulator,
since this will give rise to spurious signals, as indicated previously. A
second low-pass filter, similar to the one shown in Fig. 2.35, is required
in the low-frequency channel to remove the chopper frequency from the
output. The summing circuit might take the form of the difference
amplifier described in Sec. 2.11. In this case it is, however, necessary
that there be a sign reversal in one channel relative to the other.
It should be noted that, unless special care is used in the design of the
transfer functions of the two channels, there will be a dip in the frequency
at which the signal shifts from one channel to the other. Although the
transfer functions of the two channels required to avoid this will not be
described here, an indication of how the problem can be solved approxi¬
mately is given below in connection with Eq. 2.80. Negative feedback is
Zf
also commonly used to stabilize the gain and has the additional advantage
of providing approximately constant gain for the composite amplifier even
if the two channels do not have exactly the same gain. Thus it is possible
to obtain an amplifier with substantially constant gain over a wide fre¬
quency range extending down to direct current, while at the same time
the drift is virtually zero.
In many applications, such as the operational amplifiers used in
analogue computers, it is not necessary to have constant gain at all fre¬
quencies. Large amounts of negative feedback are always employed in
these amplifiers, and the primary requirement is zero drift. In this case
the ideas concerning drift elimination by means of negative feedback dis¬
cussed in Sec. 2.3 may be employed. The chopper-stabilized d-c ampli¬
fier of Fig. 2.37 may be used as the high-gain, driftless preamplifier. Since
high gain is required only to combat drift, the high-frequency channel
shown in Fig. 2.37 can be omitted and replaced by a simple bypass con¬
nection of unity gain. The summing amplifier may then be followed
by several stages of d-c amplification to provide a sufficiently high loop
gain at higher frequencies to ensure that the over-all gain of the amplifier
102 CONTROL SYSTEM COMPONENTS [Chap. 2
Gi
%-Q 1 + (RiCis T 1)(J?2C2S T 1)_
G2[(RlCiS + 1) (R^C2.S + 1) -f~ Cj]
(2.80)
{RiCis T ^{RzCzs T 1)
When the numerator is solved for its roots, it is found that, unless the
ratio of R\C\ to R2C2 is greater than G\ (or less than 1/Gfi), the roots have
large imaginary components, indicating a pronounced dip in the frequency
response. Since this is undesirable, the ratio of the two time constants
should be kept of the same order as the gain. The adjustment is not,
however, very critical. The open-loop frequency response will then have
the appearance shown in Fig. 2.40. Note that the gain is very high for
0.0001 0.001 0.01 0.1 1 10 100 1000 104 105 106 107
co radians/sec
Fig. 2.40. Loop gain of Goldberg amplifier.
If the resistance of the coupling network becomes finite, both the effec¬
tive value of plate-load resistance seen by the tube and the effective posi¬
tive supply voltage will be reduced. The plate load becomes simply the
parallel combination of Rl and Ri + R2, or
/ _ Rl(Ri + R2)
(2.81)
Rl T R\ T R2
p _ (Ebb — Ekk)RL
Rl H- Ri H- R2
(Ri -f R2)Ehb T RlEicIc
(2.82)
Rl T R1 ~b R2
Both of these effects are indicated by line 2 in Fig. 2.416, which repre¬
sents the load line resulting from a finite R1 + R2. All the load lines
must meet at the point ep = Ekk, since there is then no voltage across R\
and R2. It is clear that, if the level of impedance of the coupling network
is not high enough, the tube will operate in an undesirable region with
relatively small gain and be able to handle only relatively small signal
swings. On the other hand, very high coupling impedance results in poor
high-frequency response and increased sensitivity to noise pickup. A
Sec. 2.25] D-C AMPLIFIERS 105
compromise is, therefore, usually necessary, and the levels of Ri and R2
are usually taken so as to keep line 2 in Fig. 2.416 fairly close to line 1.
The design procedure in detail is as follows:
1. Pick values of Rl and R\ + R2 giving sufficient gain and signal¬
handling capacity.
2. Decide on operating point (usually in the middle of the linear range).
3. Specify operating voltage of next grid (usually near zero volts).
4. Find Ri and R2 from relation
R1 _ Rkk
(2.83)
R1 + R2 £b — Ekk
(a) (b)
for this capacitor may be computed if the input capacitance of the follow¬
ing stage is known. This capacitance can be found from the published
values of interelectrode capacitances by methods given in standard texts.1
If this input capacitance is designated by C2, then the equivalent circuit
of the coupling network takes the form shown in Fig. 2.426, where Ra is
the output impedance of the previous stage and C1 the bridging capacitor.
The transfer function of the coupling network is found by the methods
discussed in Chap. 1. It is given by
c2 / R2
ex \R0 T Ri R2
R1C1S + 1
R0RiC\R2C2 9 R0RiCi -f- R0R2C2 T R2R\C\-b R\R2C2 . ^
R0 -b Ri ~b R2 Ro ~b Ri ~b R2
It may be shown that the best value of C\ is the one that makes
R1C1 = R2C2
1 See, for instance, Reich, “Theory and Applications of Electron Tubes,” McGraw-
Hill Book Company, Inc., New York, 1944, pp. 93-96.
10G CONTROL SYSTEM COMPONENTS [Chap. 2
e2 Ri 1
(2.85)
Ci \R0 Ri -\r R2/ \[RoR2C2/(Ro + R\ + R2)]s + 1
i.e., the zero in the transfer function exactly cancels one of the poles, and
the remaining pole is moved to a relatively high frequency. The transfer
function of the coupling network without the compensating capacitor C1
may be obtained by setting C1 = 0 in Eq. (2.84). Comparison of the
result of this with Eq. (2.85) indicates that the passband of the network is
increased by a factor of (R1 + R0)/R0 by use of the correct value of Ci.
2.26. Design of Triode Amplifiers with Neon-tube Type Interstage
Coupling Network. When the neon-tube coupling networks are used, the
design procedure is similar to that described in the previous section.
Since the operating current for neon tubes must fall within rather narrow
limits, one variable, namely, the impedance level of the coupling network,
is specified in advance. The effect of the coupling on the effective plate
voltage and plate-load resistance is similar to that found in the resistance-
coupled amplifier. The effective supply voltage is
R'l = * (2.87)
Kl T riff
where ine is the neon-tube current and Rg the grid-leak resistance of the
following stage (see Fig. 2.43a). The detailed design steps are then as
follows:
1. Specify the quiescent value of the voltage at the grid of the following
stage. If Ekk is given, this immediately specifies the value of Rg, since
Sec. 2.26] D-C AMPLIFIERS 107
2. Select a value of Rl, which, together with the value of Rg decided on
in (1) above, will result in an effective load line giving sufficient gain and
signal range.
3. The proper number of neon tubes to use between stages is now
determined by noting that the sum of the quiescent grid voltage deter¬
mined in (1) plus the neon-tube drop must result in an operating point for
the first stage that will permit the required signal swing in the first tube.
This requirement is complicated by the fact that the voltage drop of neon
tubes picked at random may vary from about 49 to 74 volts. Hence, in
order to avoid selection of neon tubes, it is desirable to pick Rl in such a
way that the full range of neon-tube voltages will not force the operating
point of the first stage into undesirable regions.
This procedure is best illustrated by an example. Assume that ec2, the
quiescent second-grid voltage, is —15 volts and that a plate-signal swing
of 10 volts peak is present. Let Ebb = 300 volts. The effective plate
supply voltage E'bh is less than this,
the amount depending on the value
of R'l. Suppose two neon tubes are
used. The operating-point voltage
for the plate of the first stage may
then vary from 83 to 133 volts with
variations in neon-tube voltages
(see points A and B, Fig. 2.44).
The signal swing will cause the plate
Fig. 2.44. Effect of neon-tube tolerances
voltage to vary from 73 to 93 volts
on operating point.
or from 123 to 143 volts around
these two operating points. It is clear from the diagram that the
lowest value of plate voltage (73 volts) wou'd require a positive grid
voltage if the lower plate-load resistance R'Ll were used. Hence the
larger value shown in the figure, RfL2, would be indicated. Another
possibility is to use three neon tubes. Considerations similar to those
discussed would then indicate a possible plate-voltage variation between
122 and 217 volts. This would probably be satisfactory if the plate-load
resistance were R'L1, but the tube would be close to cutoff with the larger
plate-load resistance.
It should be noted, incidentally, that the variation of plate operating
point required to accommodate random values of neon-tube voltages is
accomplished by the zero balance adjustment required in every d-c ampli¬
fier. This adjustment must be designed with the proper range to handle
the variations discussed here, and other possible tolerance effects. The
adjustment should usually be placed in the first stage of the amplifier, so
that all stages may be adjusted together.
The question of neon-tube tolerance is one of the simpler examples of
108 CONTROL SYSTEM COMPONENTS [Chap. 2
For a particular value of Rk, a “static” load line (i.e., the line for R = *>)
can be drawn for the cathode follower from considerations similar to those
employed in amplifier stages. When the plate current is zero, the plate-
to-cathode voltage is Ebb — Ekk. This locates one point on the static load
line. When the plate-to-cathode voltage is zero, the plate current is
(Ebb — Ekk)/Rk• This locates another point and results in a line such as
the one shown in Fig. 2.455. This load line may be used to analyze the
no-load behavior of the cathode follower. When the output voltage
ek = 0, the plate-to-cathode voltage is equal to Ebb. This point may
therefore be considered.to be the quiescent operating point.
Sec. 2.28] D-C AMPLIFIERS 109
Fig. 2.46. Design of difference amplifier: (a) schematic; (6) load lines.
PROBLEMS
2.1. A cathode follower using a 12AX7 vacuum tube has a cathode resistor of
200 kilohms; Ebb = 300 volts; Ekk = —300 volts. Vibrations of the grid cause n to
change by ±5 per cent; the change of plate resistance is negligible. With grid
grounded, find the change of output voltage caused by the vibrations.
2.2. Compute the drift due to variations of Ebb and Ekk for the difference amplifier
(Fig. 2.11). After obtaining the exact expression, simplify the result by assuming
that the fi’s and rp’s of the two tube sections are identical and that Rk » rp.
2.3. Show that the drift in a properly adjusted Miller circuit (Fig. 2.13) due to
variations of Ebb is given by Eq. (2.50) if the assumption is made that the yu’s of the
two tube sections are identical.
2.4. Figure 2.47 shows a circuit that may be used to add three signals. Find the
expression for e0 as a function of ei, e2, ez. Assume that all three tubes have the same
n and rp and that Rk rp.
2.5. (a) Determine the drift caused by a variation of Ebb on the output of the cir¬
cuit in Prob. 2.4. (b) Repeat for a variation of Ekk- (c) Repeat for a change in
heater voltage.
2.6. (a) Find the gain and output impedance of the circuit of Fig. 2.48. Assume
the two tubes are identical. (6) Find the drift due to change of Ebb. (c) Find the
drift due to change in heater voltage.
2.7. The circuit (Fig. 2.49) shows a series-type amplifier which is supposed to have
the feature that drift due to heater-voltage variation is canceled in it if R is correctly
adjusted, (a) Find the value of R that provides perfect cancellation of filament-
Problems] D-C AMPLIFIERS 113
voltage variation effect. (6) Determine the drift due to changes in Ebb- (c) Deter¬
mine the gain e0/ei. Assume that the tubes are identical.
Fig. 2.49
2.8. When transistor circuits are to be analyzed by means of the node method the
transistor equations are most conveniently expressed in terms of y parameters as
follows:
ii = y litfi + 2/12^2
ii = 2/21^1 + 2/22^2
2.10. Obtain the h parameters for the grounded-base transistor amplifier in terms
of the common-emitter parameters [see Eq. (2.79)]. Use the approximations that
hre <K 1 and that hiehoe <£ h/e.
2.11. Show that the approximate expressions for the voltage gain, input impedance,
and output impedance of the grounded-collector transistor amplifier as given in
Table 2.1 are correct.
2.12. Compute the modified h parameters for the common-emitter transistor ampli¬
fier with a resistance Re connected between emitter and ground. Find the effect
of Re on the input impedance voltage gain and current gain. Assume that Rehoe <$C 1.
2.13. Repeat Prob. 2.12, but consider a resistance Rf connected between collector
and base. Assume that Rf » Ke.
2.14. Find the gain v2/(vn — v;2) for the difference amplifier of Fig. 2.31. Also
find the input impedance seen by vn.
2.16. Find the drift in v2 due to change in Ico for the difference amplifier of Fig. 2.31.
114 CONTROL SYSTEM COMPONENTS [CHAP. 2
2.16. The circuit of Fig. 2.51 shows a resistance-coupled d-c amplifier with the
following characteristics:
It is desired to optimize the gain and the gain-bandwidth product as a function of the
resistance level of the coupling network. For this purpose the gain and gain-band¬
width product are to be computed for R\ + R2 = 2 megohms, Ri + R2 = 1 megohm,
and Ri + R2 = 0.5 megohm. For the purpose of this problem the bandwidth is
defined as the frequency in radians per second for which the gain of the amplifier
is 0.707 times the value at direct current. The input impedance of the second stage
of the amplifier may be assumed purely capacitive with a value of 150 nnf. Assume
further that the quiescent value of the grid voltage for both tubes is —1 volt, that
the output impedance of e; is zero, and that the load impedance seen by the second
stage is infinite.
Sketch a curve of the behavior of the gain and the gain-bandwidth product as a
function of R\ + R2. Comment on the result.
2.17. The d-c amplifier shown in Fig. 2.52 is to be designed to permit a maximum
load-voltage swing of ±50 volts into the 10-kilohm load. Quiescent value of input
and output is 0 volt. Find values for Ri, R2, Rz, and fib; determine the number of
neon tubes required between the stages and the required adjustment range for en2.
Fig. 2.52
CHAPTER 3
POWER AMPLIFIERS
1 See, for example, Reich, “Theory and Applications of Electron Tubes,” McGraw-
Hill Book Company, Inc., New York, 1944, chaps. 7, 8.
2 See, for example, Members of the staff of the Department of Electrical Engineer¬
ing, Massachusetts Institute of Technology, “Applied Electronics,” John Wiley &
Sons, Inc., New York, 1943, chap. 10, art. 4.
3 Valley and Wallman, “Vacuum Tube Amplifiers,” Radiation Laboratory Series,
vol. 18, McGraw-Hill Book Company, Inc., New York, 1947, chap. 9.
115
116 CONTROL SYSTEM CONPONENTS [Chap. 3
ation of vacuum-tube amplifiers are defined by the grid bias. For class A
operation the grid is biased so that plate current flows throughout the
input-signal voltage cycle. In class B operation the grid is biased to cut¬
off so that plate current flows only when the grid is driven in the positive
direction by the grid signal. Thus if the input signal is a sine wave, the
plate current flows for one half the cycle. Class AB operation is inter¬
mediate between these first two. In class C operation the grid is biased
beyond cutoff, usually to at least twice the cutoff value. As is readily
seen, the transfer from signal voltage at the grid to plate current is grossly
nonlinear except in class A operation. If desired, the subscript 1 or 2
may be added to the designation of the class of operation to denote the
magnitude of grid signal anticipated. If the grid is driven positive with
respect to the cathode, thus drawing grid current, the subscript 2 is
added; otherwise the subscript 1 is used.
Fig. 3.1. Class B operation as shown on a typical transfer characteristic, which shows
plate current plotted against grid voltage, with the voltage from plate to cathode of the
tube held constant. The grid-signal amplitude shown is causing class Bi operation.
In 'push-pull operation, two tubes are required. The two tubes are
driven by signal voltages that are 180° out of phase. Thus when one
tube is nearing cutoff, the other tube is entering its operating range. The
resultant composite transfer is linearized, and distortion is reduced.
Although the theoretical values are somewhat higher, under actual oper¬
ating conditions the plate efficiency of class A amplifiers is about 30 per
cent; of class B amplifiers, about 60 per cent; and of class C amplifiers, 70
to 80 per cent. Thus it may be seen that it is usually worth while to
employ class B operation, even though the design procedure is somewhat
more involved than that for class A amplifiers.
The definition of class B operation requires that the tube be biased to
cutoff, as shown on the transfer characteristic in Fig. 3.1. Since plate
current flows for only one half of the grid-voltage cycle, the resultant
plate-current wave is seriously distorted. A second tube is added to the
circuit that will operate on the other half of the signal-voltage wave, and
the resultant plate currents are added in the load. In a-c applications
Sec. 3.2] POWER AMPLIFIERS 117
Ebb
■/Vr- ■ Level
0 -
0-
Input
from
preceeding
stage
A signal on the input transformer drives one grid positive and the other
grid negative. The plate currents flow in opposite directions in the out¬
put transformer, thus canceling any d-c component of flux in the trans¬
former core. In Fig. 3.3 the composite plate characteristics of the tubes
are shown. Composite plate characteristics are commercially available
for tubes such as the 6L6 and 807 that are commonly used in push-pull
circuits. The proper plate-battery voltage and grid-bias voltage must
be carefully chosen to establish the proper operating point. In Fig. 3.3
the curvature of the characteristics at low plate current has been exagger¬
ated. We note that, strictly speaking, this is not class B operation, since
both tubes conduct for small signal voltages. If the characteristics are
adjusted so that the tubes are completely cut off, the composite character¬
istics are no longer straight lines and a certain amount of distortion is
introduced.1 If we assume that the characteristics are essentially
straight lines, it is possible to derive an equivalent circuit for the oper-
1 Reich, op. cit.} art. 8.1.
118 CONTROL SYSTEM COMPONENTS [Chap. 3
Fig. 3.3. Composite plate characteristics for push-pull operation. Strictly speaking,
this is class AB operation.
fl6g
Pi oad = PR = a2Rh (3-3)
jv + a2RL,
Sec. 3.2] POWER AMPLIFIERS 119
For eg, g, and Rl constant we may determine the optimum turns ratio by
differentiating load power with respect to a and setting t he result to zero.
The result is
a2 (3.4)
which is the familiar statement that the load impedance must be matched
to the impedance of the generator for maximum power transfer. Occa¬
sionally one is led by this familiar result into assuming that the converse is
true. The reader may show that, if the internal impedance rp of the
generator can be adjusted by choosing the
proper tube, it should not be made equal to -ww—
the load impedance but rather should be made lp\
■o2Rl
as small as possible, for maximum power
transfer. O
~ Peg
\02ja>L
In general, the load is not purely resistive.
If the load consists of a servomotor, for
Fig. 3.5. Equivalent circuit
instance, it will have a large inductive react¬ with reactive load.
ance. Under these circumstances the equiva¬
lent circuit will be as shown in Fig. 3.5. The load impedance is referred
to the primary of the output transformer. For this type of load the plate
current is
dip
— ixeg = rpip + oPRiiv + cl2L (3.5)
dt
The plate current is no longer in time phase with the generator voltage,
and no longer does the simple load
line describe the operation of the cir¬
cuit. From Eq. (3.5) we see that
only when the rate of change of plate
current is zero will the operating
point lie on the load line. When the
plate current is increasing, the volt¬
age across the inductance opposes
the increase, and when the current is
decreasing, the inductance tends to
lpZ
maintain the flow. It can be shown1
Fig. 3.6 Operating locus for inductive that, for a sinusoidal input signal,
load. the operating point follows an ellipse
whose major axis is inclined some¬
what to the load line, as shown in Fig. 3.6. The ratio of the distance
along the major axis to the minor axis is equal to the ratio of the load
1 Dow, “Fundamentals of Engineering Electronics,” John Wiley & Sons, Inc.,
New York, 1945, p. 274.
120 CONTROL SYSTEM COMPONENTS [Chap. 3
Fig. 3.7. Basic circuit of the grounded-emitter push-pull amplifier with p-n-p junction
transistors.
1 4 Rl
(3.6)
1 + 3B rs
Vl = T\\i\ + T \2X2 /o
v2 = r21ii + r22i2
1 Ibid.
2 Shea, “Principles of Transistor Circuits,” John Wiley & Sons, Inc., New York,
1953, p. 34.
vT*.
122
o.A.
T12^21
Ho — f 22 (3.8)
r n + rs
-O- •-O-
rm^b <
cp <
Fig. 3.10. Push-pull composite characteristics for class B operation. The two curves
are matched at the operating point Vc = —60 volts.
drift introduced in this stage do not have so serious an effect as noise and
drift introduced at lower signal levels. The d-c power amplifier differs
from the a-c power amplifier in several respects. First, transformers
cannot be used for phase inversion at the input or for load matching at
124 CONTROL SYSTEM COMPONENTS [Chap. 3
not limited by the grid potential, and the conducting thyratron is effec¬
tively a short circuit from anode to cathode. If the anode current is not
limited by the load, the tube will be damaged by the excessive current
flow. There is a small voltage of 10 to 20 volts, depending on the gas,
across the tube that is approximately independent of current, owing to
the mechanism of gas conduction.1 In order to stop conduction, the
Fig. 3.13. Typical thyratron firing characteristics. Grid potential for initiation of
conduction is plotted against positive anode potential for a typical thyratron.
Fig. 3.14. Relation between anode and bias voltages for combination of a-c and d-c
bias.
| + sin- (i _ g) (3.9)
where E{ is the signal voltage and Edc is the d-c bias, which is assumed to
be approximately equal to the amplitude of the a-c bias. For full control
of the firing angle from 180 to 0°, E{ varies from zero to 2Edc. Thus the
required signal range for full control is determined by the amplitude of
the a-c bias. This can also be shown by computing the gain, or the
change in firing angle degrees per volt of input signal, by differentiating
Eq. (3.9) with respect to E{. This gives
d(f>a 1
(3.10)
dEi ■\Z2EiEdc E,2
If this is evaluated for E{ = Edc, that is, for <f>a = 90°, the gain becomes
d<t>a
(3.11)
dEi 4>a = 90° Edc
and is thus shown to be inversely proportional to the d-c bias. Thus the
gain can be increased by decreasing the d-c bias and therefore the peak
value of the a-c bias. Beyond a certain point, however, this reduction
leads to difficulties. One of the important advantages of the control
method using a-c bias is that the intersection of the grid voltage with the
critical control line is sharp and not greatly affected by small variations
of the critical voltage due to temperature or random effects. This advan¬
tage is largely lost if the amplitude of the a-c bias is reduced too greatly,
and erratic firing may result.
The firing angle can also be controlled by shifting the phase of the a-c
bias relative to the phase of the anode. In this method a fairly large a-c
grid voltage is used, usually without d-c bias. The major disadvantage
of this method is the difficulty of shifting the phase of the grid voltage.
Although a number of circuits are available for this purpose,1 they usually
1 Reich, “Theory and Applications of Electron Tubes,” McGraw-Hill Book Com¬
pany, Inc., New York, 1944, pp. 510-515.
Sec. 3.6] POWER AMPLIFIERS 129
eL = es — eT (3.12)
Fig. 3.16. Voltage and current wave¬
where er is a constant, usually taken
shapes for circuit in Fig. 3.15.
as 15 volts or perhaps neglected,
Having found cl at point (a), we write
di _ eL
(3.13)
dt L
thus establishing the slope of the current wave. The initial slope of er is
found by multiplying the initial value of di/dt by R. The increment over
which this initial slope is extended depends on the accuracy required for
the results. If it is sufficient to find the initial slope, the maximum value,
and the point at which er once more becomes zero, the calculation is fairly
simple. We know that the maximum value (or zero slope) of the iR volt¬
age wave falls at point b, since at that point eL and therefore di/dt are
zero, and
er — es — er (3.14)
As yet, we have not determined point b. We pick a point b such that the
integral of cl from a to 6 times R/L is equal to the value of er at point b.
This may be seen by noting that
r di
eL = (3.15)
LJt
Therefore Jb eL dt = fib Ldi
=0Jla
(3.16)
@rb
Now ib = (3.18)
R
1 P. T. Chin and E. E. Moyer, Controlled Rectifier Circuits, Trans. AIEE, vol. 63,
p. 501, 1944.
Sec. 3.6] power amplifiers 131
(3.19)
where Es is the amplitude of the supply voltage, ET the tube drop, co the
line frequency in radians per second, and 0a the firing angle. The time
origin (t = 0) is the time at which the tube fires.
The instantaneous load current is best found as a function of time by
use of Laplace transform methods. The Laplace transform of the voltage
of Eq. (3.20) is
Es(co cos 4>a + s sin 0a)
(3.21)
s2 + CO2
Z R + Ls — L (3.22)
Es co cos 0a + s sin 0a _ ET 1
(3.23)
L (s2 T w2) (s T R/L) L s(s T R/L)
Y/\ _
sin (cot + 0a — 0) — sin (0a — 6) exp ( — ^ t
Et R ,
1 — exp (| — j-1
i (3.24)
~R
expression can be used to find the instantaneous value of the load current
at any time t, it cannot be solved directly for the time at which the current
goes through zero, i.e., the extinction point, because of the simultaneous
presence of exponential and trigonometric functions of time. It is, how¬
ever, possible to obtain the same information indirectly by assuming an
arbitrary value of cct and finding the value of firing angle <f>a required to
make the current go through zero at this point. If i(t) in Eq. (3.24) is
set equal to zero, the equation becomes
ErZ
sin (cd + (f>a — 6) — sin (0„ — S) exp ( — 1
) ESR
(3.25)
This equation may be rearranged to give
<t>a — 0 + 7T
where both 0a and 0C are measured from point 0 in Fig. 3.16. Sub¬
stituting for cot in Eq. (3.26), we obtain
0a = 6 + 7T — sin-1
1_1
1
^ S3
EtZ
■e-
1 — exp
1
1
1
1_1
1-1
1
o
coL
3
term (R/coL)((f)c — <f>a) is large (greater than about 3), all the exponential
terms become negligibly small. Then, approximately,
0a 6 + 7r — sin-1 — 0c + 0a
EtZ
or 0c 6 + 7r — sin-1 (3.29)
ESR
This indicates that, if the inductance is sufficiently small and if the differ¬
ence between extinction and firing angles is sufficiently large, the extinc¬
tion point becomes independent of the firing point. A typical set of
Fig. 3.17. Relation between firing angle and extinction angle for half-wave circuit with
inductive load. Tube drop is neglected.
If the inductance is sufficiently small and if the firing point comes suffi¬
ciently early so that (i2/o>L)(0c — 0O) > 3, the extinction angle may be
found from Eq. (3.29). Under these conditions the average current
becomes
where X = coL, Z = (jR2 + X2)1^, and where we have used the fact
that 6 = tan-1 uL/R. If the tube drop ET is negligible, this becomes
+ S)
0C ~ 7T + 26 — 0O
But if the inductance is very large, d « tt/2. Hence approximately
0c ~ 2tT 0a (3.34)
This result implies that for very large inductance the conduction period
for negative values of instantaneous supply voltage is approximately
equal to the conduction period for positive voltage; i.e., the shaded area
in Fig. 3.16 is equal to the crosshatched area. The average current is
therefore approximately equal to zero. This is borne out analytically
by substituting Eq. (3.34) into Eq. (3.31). If the tube drop is neglected,
the result will be seen to be zero. In practice the load current is, of
course, not zero, since Eq. (3.34) is only approximate; however the effect
of large inductance in severely reducing the average load current should
be evident.
Sec. 3.6] POWER AMPLIFIERS 135
Curves of average current as a function of firing angle for several values
of ooL/R are given in Fig. 3.18. These curves are obtained as follows:
First, a number of values of </>c — </>a are chosen and used in Eq. (3.28) to
determine <f>a. Then (f>c can also be found, and the current is then
obtained from Eq. (3.31). The tube drop has been neglected, and the
Fig. 3.18. Normalized average load current versus firing angle for half-wave thyratron
circuit. Tube drop is neglected.
A-c
conducts during the negative half cycle of a-c supply voltage and provides
a low-resistance discharge path for the load current. Thus the load cur¬
rent is not forced down by the negative supply voltage, as is the case in
the half-wave thyratron circuit without a diode, but instead it follows a
more slowly decreasing, exponential discharge curve. This is shown in
136 CONTROL SYSTEM COMPONENTS [Chap. 3
Fig. 3.196, which shows both the voltage and current waveshapes. If the
tube drop of the diode is negligible, the load voltage during negative half
cycles of a-c supply voltage is zero, and the waveshape is identical with
that observed on a half-wave circuit with a pure resistance load. Hence
the average load current is equal to the average current found for pure
resistance, (E8/2ttR){\ + cos 0O).
3.7. The Full-wave Thyratron Amplifier with Inductive Load. The
full-wave circuit shown in Fig. 3.20 is more commonly used than the half¬
wave circuit considered in Sec. 3.6. Since there are two voltage pulses
per cycle of supply voltage, a smoother current flow to the load results.
The analysis of the full-wave cir¬
cuit is complicated somewhat by
the fact that under certain condi¬
tions the load current may be con¬
tinuous. This happens, in fact,
A-c
<? supply <?
Fig. 3.20. Full-wave thyratron circuit. Fig. 3.21. Voltage and current wave-
Rg limits the current drawn by the con- shapes in full-wave thyratron circuit,
trol grid.
whenever the extinction angle of one tube as given by Eq. (3.28) occurs
later than the firing point of the other tube. Typical waveshapes of
voltage and current, for both the continuous and discontinuous case, are
shown in Fig. 3.21. Since in normal operation the firing points of the two
tubes are 180° apart, the condition for continuous current flow is
0c 0a ^ TT
where both 0C and 0a are measured with respect to the same origin and on
the same tube. Substituting this relation into Eq. (3.28) gives
EtZ 1 — exp ( — Rtt/uL)
0a, < 0. — sin-1 (3.35)
ESR 1 + exp ( — Rtt/uL)
Sec. 3.7] POWER AMPLIFIERS 137
for the condition for continuous current flow. In making this substitu¬
tion, it should be noted that the last term of Eq. (3.28) must have the
value 7r rather than zero.
The determination of average voltage and current is now reasonably
straightforward. If the circuit operates in the discontinuous mode, the
voltage and current are found exactly as for the half-wave circuit except
that the averages must be taken over one half cycle. Thus
1 Pc
E&v = - / (Es sin $ — Et) d<f)
^ J <t>a
If the inductance is small and if the firing point comes early enough,
expressions similar to Eqs. (3.32) and (3.33) can be derived for the full-
wave circuit; in fact the current is simply twice the value given in these
equations.
When the circuit operates in the continuous mode, the average voltage
and current are obtained more easily, since it is only necessary to integrate
from one firing point to the next. This is made clear by inspection of
Fig. 3.216. Thus, for the continuous mode the average voltage is given
by
-j r <£a+ir
Eav = - / (Es sin </> - Et) d(f>
. 4>a
A/ = _
(3.40)
A$a (L/R)s + 1
In this equation k is the slope of the curve of current versus firing angle
and is a function of the operating point around which the small changes
AI and A<£a are taken. •
Sec. 3.9] POWER AMPLIFIERS 139
3.9. The Thyratron Amplifier with a D-C-motor Load. Thyratron
amplifiers are quite often used to control low- and medium-power d-c
motors. A simple half-wave circuit for unidirectional speed control is
shown in Fig. 3.23 with all voltages and currents labeled. The motor
field is assumed to be excited by a separate d-c supply and is assumed to
be fixed.
A d-c-motor load differs from loads
considered up to now in that it gen¬
tmJ
erates a counter emf Eg = kv f2, where
12 is the speed and kv the velocity volt¬
age coefficient (see Sec. 4.15). When
the speed is constant, this counter
emf is a fixed voltage and may be rep¬
resented by a battery in the load cir¬ Fig. 3.23. Half-wave thyratron circuit
with motor load.
cuit. As a result of the presence of the
counter emf, the thyratron may not fire when the grid potential passes the
critical value, and it is convenient, therefore, to define three modes of
operation for the circuit. These modes are defined by reference to Fig.
3.24.
When the grid passes the critical value in region 1 of Fig. 3.24, the tube
cannot conduct, since the voltage on its plate is negative. However,
unless the grid signal consists of short pulses, the tube will conduct as
soon as the instantaneous a-c supply
voltage exceeds the counter emf Eg (point
a in Fig. 3.24). Thus in mode 1 changes
in firing angle that do not go beyond
point a have no effect on the current
flowing in the motor. A firing angle
occurring anywhere between points a and
b in Fig. 3.24 puts the circuit into oper¬
Fig. 3.24. Operating modes of ating mode 2. In this mode the circuit
thyratron amplifier with motor
operates normally, with the tube firing
load.
as soon as the critical grid potential
is exceeded. Mode 3 is defined by a firing point occurring in region 3.
The armature current is always zero in this mode. The three modes are
defined mathematically by the following three equations:
where </>« is the firing angle. Note that the sin-1 functions used in these
equations are the so-called principal values. Note also that, for a given
value of (f)a, the circuit can be in only one of two modes. Thus for
4>a < tt/2 the circuit goes from mode 2 to mode 1 as Eg is increased. For
<t>a > 7r/2 the circuit goes from mode 2 to mode 3 as Eg is increased.
The analysis of the operation of the motor is simplified considerably if
it is assumed that the inductance of the armature circuit is negligible.
This is true for most d-c motors, particularly those designed for use with
thyratrons, because, as will be shown later, the presence of inductance
reduces the maximum current and developed torque. Thus, in the fol¬
lowing analysis we assume that the inductance is zero. Also, it is con¬
venient, although not necessary, to neglect the tube drop Et-
If a firing angle and the speed of the motor are given, the average arma¬
ture current can be found as follows: If the circuit is in mode 1, the tube
conducts during the time interval bounded by points a and b in Fig. 3.24.
Hence the average current is
r 4>b
I = (Es sin </> — Eg) d<t>
2ir Ra <t>a
1
[Es (cos 4>a — cos (t>b) — Eg(4>h — </>«)] (3.42)
2 7rRa
(3.43)
Hence (3.44)
If the circuit is in mode 2, the thyratron fires at the firing point <f>a,
which is now no longer determined by Eg but may have any arbitrary
value. Hence
V —sin"1 (Eg/Ea)
For mode 1:
s -J
(3.48)
These two equations can be used to plot curves of developed torque versus
speed for different firing angles. Such a set of curves is shown in Fig.
3.25. These curves have been normalized with respect to maximum
torque and speed of the motor. The points of transition between modes
are also indicated in this figure. Note that the transition from mode 1 to
mode 2 is not abrupt; in fact it can be shown by differentiating Eqs. (3.47)
and (3.48) that the slopes of the two curves are the same at the transition
point. The same is true for the transition from mode 2 to mode 3.
A number of interesting points can be deduced from Fig. 3.25. Note
that, for firing angles less than 90° and for small developed torque, the
motor operates almost entirely in mode 1. Its speed is therefore not con-
142 CONTROL SYSTEM COMPONENTS [CHAP. 3
trolled by the firing angle; in fact it appears from the curves that the
no-load speed for all firing angles less than 90° is equal to the maximum
speed. If the no-load speed of a motor is to be controlled, firing angles
of more than 90° must be employed. At these firing angles the circuit
operates, however, in a highly nonlinear on-off region, i.e., the boundary
between mode 2 and mode 3. Under certain conditions this may give
rise to a nonlinear type of oscillation. Thus, consider the circuit shown
in Fig. 3.26. Suppose that for some reason the motor is running too fast.
The feedback circuit acts to retard the firing angle. If the motor is
running at no load, the circuit goes into mode 3, and the armature current
is cut off. The motor now coasts until its speed is less than the reference
speed. If there is a relatively large time lag in the controller, the motor
may slow down too much before the thyratron fires again and may then
again overspeed and coast. The action resembles the bouncing of a ball
being pushed along a horizontal plane by a block moving at uniform
speed.
By differentiation of Eqs. (3.47) and (3.48) it can be shown that the
slope of the curves for zero developed torque is zero. This means that
the speed regulation near zero torque is very great; i.e., the motor must
slow down considerably to support a relatively small torque change. On
the other hand, near full torque the curves are all approximately linear.
They may, therefore, be used in determining a quasi-linear transfer func¬
tion, as described in Sec. 4.16. In general, the effect of the thyratron
amplifier can be approximated by a series resistance which tends to
increase the motor time constant.
3.10. The Effect of Armature Inductance. In the preceding section
we have considered the armature inductance of the motor to be negligible.
However, since the ratio of inductive reactance at line frequency to the
resistance is the significant parameter determining the effect of the
inductance, this assumption is not always completely justified. In
Sec. 3.10] POWER AMPLIFIERS 143
general, however, it is true that the inductance, even if not completely
negligible, is small.
An accurate analysis of the effect of inductance was shown in Sec. 3.6
to be a rather involved task. However, if the inductance is small, its
effect may be taken into account approximately without too much diffi¬
culty. We assume, therefore, that the inductance is sufficiently small
so that 2ir,}La <<C Ra, where / is the line frequency.
During the period of tube conduction, the armature current may be
thought of as consisting of two components, an a-c component due to the
a-c supply voltage and a d-c component. If the inductance is small, the
a-c component will be approximately sinusoidal in shape, particularly
toward the end of the conducting period when the initial switching
transient can be assumed to have decayed. Also, if the inductive reac¬
tance is small compared to the resistance, the amplitude of the a-c com¬
ponent is approximately Es/Ra. Thus, the only effect of the inductance
on the a-c component of the armature current is to delay it, so that its
cycle is behind the a-c supply voltage by the angle tan-1 (2TrfLa/Ra)- The
d-c component of the current cannot be affected by the inductance at all.
Hence, we find that the instantaneous armature current during the period
of tube conduction is approximately
_ tan-1 J (3.49)
‘ti'a / -tv a
Once the thyratron has been fired, it continues to conduct until the plate
current goes through zero. By setting the instantaneous current equal
to zero in Eq. (3.49), we find that
the extinction point is given by
(3.50)
d_
(Es sin </>) = Es cos
d($> <t> = 4>b
1 /2tr/LaV
AA = - ? 2 (3.53)
2 V Ra ) Ve;- - e
A/
AA J_ /2t/L0V (3.54)
Ve.2 E 2
2ttR 47T^a \ Ra /
ferred over the half-wave circuit discussed in Sec. 3.10. The operation of
the circuit is quite similar to the half-wave circuit as long as the current
is not continuous. In fact, all that has been said concerning operating
modes, speed-torque curves, effect of inductance, etc., applies to the full-
wave circuit, except that, for a given firing angle and speed, the armature
Sec. 3.12] POWER AMPLIFIERS 145
current and therefore the torque are twice as great as in the half-wave
circuit. However, as was pointed out in Sec. 3.7, under certain condi¬
tions the current in the full-wave circuit may be continuous. If the
inductance is negligible, this can happen only for negative speeds and is
therefore rather unlikely during normal operation. However, if the
armature inductance is sufficiently large, the current may become con¬
tinuous if the speed is low enough and the firing angle small enough. The
extension of the theory to this case is straightforward and is left to the
reader.
3.12. Thyratron Circuits for Reversing Control. Two commonly used
circuits permitting bidirectional motor-speed control are shown in Fig.
Control
signal 9 9
Fig. 3.29. Thyratron circuits for bidirectional motor-speed cbntrol. Both are half¬
wave circuits.
3.29. Both of these circuits are half-wave circuits in which one tube con¬
trols the motor when it is running in one direction while the other tube
controls it in the opposite direction. The transformer connections in the
circuit of Fig. 3.296 are such that the two tubes conduct on alternate half
cycles. This is necessary, since there would otherwise be a short circuit
through the tubes and transformer windings whenever both tubes con¬
ducted simultaneously. This problem does not exist in the simpler cir¬
cuit of Fig. 3.29a; however, in it also the two tubes can conduct only on
alternate half cycles. For smooth transfer of control from one tube to
the other near zero motor speed, the grid bias is usually so adjusted that
for zero signal input both tubes fire for a short part of their respective
half cycle. An increase of signal in the positive direction advances the
firing angle of one tube, say, tube 1, and retards the angle of the other
tube (tube 2), until eventually only tube 1 conducts. For negative
signal the action is reversed.
When only one tube conducts, the analysis of the circuit is exactly the
same as that carried through in Sec. 3.9. However, in the interval near
zero speed the effect of the two tubes operating together must be con-
146 CONTROL SYSTEM COMPONENTS [Chap. 3
sidered. For the sake of simplicity we assume that the armature induct¬
ance and the tube drop are negligible.
The waveshape of the armature current is shown in Fig. 3.30 for oper¬
ation in mode 2. The average armature current is found by integrating
over the two pulses of current occurring per cycle. We have, therefore*
Suppose now that 0ai = </>«o — A0a and 0a2 = tt + 0aO + A</>a. This
implies that for zero signal the firing angle of tube 1 is </>ao and that the
firing angle of tube 2 comes 180° later. Also the signal A</>a advances the
firing angle of one tube by the same amount that it retards the firing
angle of the other. Substituting into Eq. (3.55), we obtain after some
reduction
Since the average developed torque Qd = ktI*v and since the counter emf
Eg = kvtt, as in Sec. 3.9, we convert Eq. (3.56) into a relation between
speed and torque:
This relation holds only when both tubes are conducting and are in
mode 2. If it is assumed that the grid-control circuit limits variations
of firing angles to the range between 0 and 180° on each tube, then
Eq. (3.57) applies'only if A0« meets all of the following conditions:
Sec. 3.12] POWER AMPLIFIERS 147
fc„12 (3.58)
sin-1 < 0oO — A <t>a <7r — sin-1
X
j Jcv 12
— sin < 0ao + A<t>a <7r + sin-
X
A composite set of speed-torque curves for the motor can now be plotted
by use of Eqs. (3.47), (3.48), and (3.57), keeping in mind the limitations
Fig. 3.31. Speed-torque curves for bidirectional thyratron control circuit. Armature
inductance and tube drop are neglected; <f)ao = 135°.
dJQa (3.59)
d!2
148 CONTROL SYSTEM COMPONENTS [Chap. 3
JRg
(3.60)
kvkt(l 0ao/7r)
Low-pass
fitter ' \
er Output
VWV-1- Retoy \ \ er
Input
lut V-/
R C.1 -m
I o
o h-oOut
but if the load acts as a low-pass filter and is essentially sensitive only to
the average value of the pulse chain, it is shown below that the amplifier is
quite satisfactory and in fact “linear” over most of the operating range.
If the input-signal amplitude is large, the pulses are long and the average
value of the output is high. The limit on input-signal amplitude is
reached at the point at which the relay remains closed for the entire
period.
The relay used in the amplifier must be a polarized relay, i.e., one that
will close to one contact for a positive voltage applied to the coil and to
the other contact for a negative voltage applied. The same effect may,
however, be obtained by using two unpolarized relays which are in con¬
junction with diodes. In any physical relay there is a deadband, or
152 CONTROL SYSTEM COMPONENTS
range of coil voltage that is too small to activate the relay. All practical
relays also display hysteresis; i.e., a larger value of coil voltage is required
to close the relay than that required as a minimum to hold in the
relay.
The major advantages of relay amplifiers are simplicity, ruggedness,
and economy. The major disadvantage is the fact that the load must
supply a good deal of smoothing if the operation is to be acceptable.
However, it is shown below that, for a typical motor with an integration
and a single time constant, a good replica of the input to the relay ampli¬
fier is possible at the shaft output.
3.15. Relay-amplifier Stability. Before discussing the input-output
relations of a relay amplifier with feedback, we consider the stability of
such a device. The stability of feed¬
Cm back systems containing relays is perhaps
h <-
most easily investigated by use of the
describing-function method of Kochen-
>f j burger.1 In this method the highly non¬
linear input-output relation of the relay
ir is replaced by a quasi-linear describing
A >
>f i\ ■ function. This describing function is
c. the relation between the magnitude of
i
r
i
-> h <-
a sinusoidal input signal and the magni¬
Fig. 3.37. Relay characteristic. tude and phase angle of the fundamental
component of the output. Kochen-
burger has shown that a relay with the input-output characteristics shown
in Fig. 3.37 has the describing function
4 sin
h2 /—a (3.61)
7T Ir
A — h — . A T h
COS —l _ COS" 1_
2L 2Ir
(3.62)
A — h _ j A + h\
cos —i
P — /4
~2i^ + cos -2ir)
In Eqs. (3.62) A and h are the deadband and hysteresis, respectively, of
the relay, as indicated in Fig. 3.37.
The remainder of the feedback loop making up the system is assumed
to be linear and has the transfer function Hlm The system is stable if
Fig. 3.38. Typical inverse Nyquist diagrams for a feedback system containing a relay.
has at least two lags. Note that H2 is a function of Ir, while Hi is a func¬
tion of frequency. Intersection of the two curves indicates instability.
For the relay amplifier with feedback shown in Fig. 3.39, Hi is given by
kRi k'
Hi = (3.63)
RiR2Cs 1 + Ts
(R1 + R2) 1 + Ri T R2
where k is the amplifier gain. Therefore the Nyquist plot will have the
form shown in Fig. 3.40. Since the curve of l/Hi does not intersect that
Fig. 3.39. Typical relay-amplifier circuit. Fig. 3.40. Inverse Nyquist diagram for
relay amplifier of Fig. 3.39.
for — H2, the implication is that the system is stable no matter how large
the gain k is. However, at high gains, previously neglected lags, e.g., the
time required for the relay to close after the coil is energized, may cause
the frequency locus to bend around as shown in Fig. 3.38, thus possibly
causing instability. Similarly, if the filter break frequency is raised to a
154 CONTROL SYSTEM COMPONENTS [Chap. 3
value such that other lags in the loop, such as the mechanical and/or elec¬
trical lags of the relay itself, become important, the frequency locus bends
to the left around the origin, as in the case of increased gain. If it
intersects the amplitude locus, instability is indicated.
3.16. Relay-amplifier Static Characteristic. The output of a relay
amplifier with feedback is a series of pulses, positive or negative, of ampli¬
tude Eb and with length and repetition rate determined by the input.
Usually the relay amplifier is em¬
'm ployed in an application where only
Eh —- the average value of this train of
" A pulses is important.
~e2~e\ In order to determine the static
*1 *2
relation between the input voltage
V A
ein and the average output voltage
-Ef
Em, we shall consider the circuit in
Fig. 3.39, with the static character¬
Fig. 3.41. Pull-in and drop-out voltages
istic between e and em as shown in
defined on the relay characteristic.
Fig. 3.41. Let us suppose that, at
zero time, to, e = 0. Thus the relay is open, and em = 0. Let a step of
voltage, Exn in magnitude, be applied, as shown in Fig. 3.36. Then
Exa - e2 = (3.65)
or t\ = —Ti In (Ein — e2) (3.66)
R2 Ri
e [1 Eh [1 e-(t-tl)/T21
Ri + R2 Ri T R2
+ e2e-(t-t')/T> (3.67)
It will be noted that Eq. (3.70) is the same as Eq. (3.64) plus the term
due to the initial voltage on the capacitor. All the segments of the volt¬
age curve will be exactly the same as those described by Eqs. (3.67) and
(3.70). Equation (3.64) is the special case when no charge is on the
capacitor. The relay closes again at £3, when e = e2; this time is given by
During the period of time U — U the output voltage is — Eb, and during
the period t2 — h the output is zero. These two lengths of time represent
the cycle of the output-voltage wave. The average output is therefore
— Eb(t 2 — t\)
Em = (3.72)
h - tl
ei - Eiu „■ + Eb Rl
Ri T R2 R i + R2
Eb In
777
e2 ~ Ain
R2 I
r>—;—T-,—h Eb
77TR 1
R1 T R2 R1 T R2
or Em =
p _ rp -^2 I jp
Ti ln e2 - Ein + ln 1 iD^i + R2 ^ b Ri + R2
T2 e\ — Ex R2 , T7, R1
62 — Ex + Ei
R1 T i?2 R1 T R2,
(3.73)
1 + 2ei/(Eb — E in)
1 + 2e2/ (Eb — Ein)
(3.74)
1 - Eb/(Ehl -2d)
1 - Eb/(Ein - 2e2)
By examination of this relation we can see that it gives the required nega¬
tive real result only for the region
e2 < Ein < Eb + gi. For Ein < e2
the relay does not close; the output
is zero. For Ein > Eb + ei, the relay
remains closed and the output is Eb.
It may be shown that, if e\ and e2
are small with respect to the battery
voltage, the slope of Em versus Ein is
essentially constant over the operat¬
ing range. The exact relation be¬
tween input-voltage magnitude and
the average of the output voltage can
Fig. 3.42. Static characteristic of the
be determined by the use of Eq.
relay amplifier, normalized to the relay¬
(3.75). In Fig. 3.42 is shown Em
closing voltage Eb. The contact volt¬
age was ten times the closing voltage in versus Ein normalized to Eb.
this normalization. 3.17. The Frequency Response of
a Relay Amplifier. The limitation
on the frequency response for a sinusoidal input of a relay amplifier is
determined to a first approximation by the time constant T of the
filter or the chatter rate of the output pulses. Actually, however, this
chatter rate or output-pulse repetition rate depends on the input-signal
magnitude, and in the limit, where Ein is either greater than Eb or zero,
the relay remains always closed or always open. Since the relay amplifier
is a nonlinear device, the entire concept of frequency response is con¬
trived, and a measure for one type of input signal will not hold for
another. For a step-function input, the relay amplifier is particularly
fast and essentially establishes the required average value of em in one
opening and closing.
The response of a relay amplifier to sinusoidal signals depends on the
amplitude and the frequency of the input signal. For signals whose peak
amplitude at the input of the relay is less than the value required to close
the relay, there is no output. For instance, if the closing voltage of the
relay is 10 volts, a signal whose peak amplitude is less than 10 volts will
not close the relay. Since the low-pass filter attenuates the amplitude
of the input signal, the peak amplitude must be computed at the input
to the relay. At-the break frequency of the filter, for instance, the filter
Sec. 3.17] POWER AMPLIFIERS 157
reduces the input amplitude by a factor of 0.707; thus the minimum input
signal amplitude must be 14.14 volts peak. Essentially a relay amplifier
will operate properly for normal-amplitude signals up to the break fre¬
quency of the low-pass filter. In Fig. 3.43 are shown the amplitude
reduction and phase shift for a typical relay amplifier whose relay-closing
voltage is 10 per cent of the battery voltage applied to the relay terminals.
The frequency response shown in Fig. 3.43 was obtained by smoothing
the pulse-train output in order to recover the input-signal frequency com¬
ponent. When the input-signal frequency is equal to or greater than the
Load
m-
Gate B
^ A-c supply
winding
Control Q
voltage \~)
Gate
Control winding
winding
(a) U>)
Fig. 3.45. Series-connected magnetic amplifier, and hysteresis loop of core material.
tional magnetomotive force is added to each core and causes the cores to
become saturated for some part of the a-c cycle. Suppose that, at a cer¬
tain point in the cycle, reactor 1 of Fig. 3.48 saturates. When this hap¬
pens, the flux can no longer change, or
t =0 (3-77)
Hence the reactor can no longer generate any counter emf and, by
assumptions 1 and 3, now represents a short circuit, not only on its gate
winding, but also on its control winding. As a result the control winding
of reactor 2 is now connected only across the resistance of the control cir¬
cuit, which, by assumption 2, is negligibly small. The entire control
circuit seen by reactor 2 therefore represents a short circuit, and even
though this reactor is not saturated, the voltage across both of its wind¬
ings must also be zero. The flux in the core must therefore be constant.
Hence all the supply voltage appears across the load. During the next
half cycle the process repeats, with the roles of the two reactors reversed.
In Fig. 3.50 are shown the resulting waveforms of voltage, flux, and cur¬
rent. In drawing these curves, we assume that the circuit is operating
in the steady state and that the reactors are normally excited. At time
t — 0, reactor 2 is just coming out of saturation, and the flux in reactor 1
has the value (/>a. This value is chosen to meet the requirement that the
circuit is in steady-state conditions, i.e., that 4>i = <t>a for cot = 2ir,
47r, ....
As time increases from zero, the flux in both reactors increases, with
d(f>/dt exactly equal to the value required to make the reactor counter
voltage equal and opposite to the applied voltage. This state of affairs
continues until cot — a, at which point reactor 1 saturates. As indicated
in the preceding paragraph, the flux in both reactors now remains con¬
stant, and the voltage drop across the reactors is zero. The supply volt¬
age therefore appears across the load, and load current flows. As soon as
the supply voltage passes through zero, d<t>/dt can again be supplied by
both reactors. The reactors therefore are again able to generate counter
emf, and the load voltage and current go to zero, until at cot = t + a, reac¬
tor 2 saturates again and the cycle repeats.
The waveshape of the load-current pulses is the same as that of the
load-voltage pulses, since the load is assumed to be resistive. During the
first pulse, i.e., for a < cot < tt, reactor 1 is saturated, and its load and
control circuits are therefore completely decoupled. The fact that the
load current flowing in its gate winding produces mmf has no effect on its
control winding. In reactor 2, however, the flux is constant at a value
that is less than the saturation value. According to the ideal form of the
hysteresis loop assumed for the core material (Fig. 3.47), there can there-
Sec. 3.19] POWER AMPLIFIERS 163
fore be no net magnetomotive force in this reactor. Hence the mmf of
the gate winding is opposed by an equal and opposite mmf produced by
the control winding, or
During the next half cycle the load current flows in the opposite direction,
but since reactor 1 is connected in the opposite direction to the control
where II is an average over one half cycle. In Fig. 3.48 is shown one
164 CONTROL SYSTEM COMPONENTS [Chap. 3
Ic = Is (3.80)
■Kc
Thus we find that the current gain of the series magnetic reactor is given
by
II = Nc
Kt (3.81)
Ic Ng
Here again, all quantities are average values. The d-c power gain is
ElIl Nc2Rl
(3.83)
ECIC Ng2RC
Since EL and II are average quantities, the power output given by this
expression is that obtained after rectifying and filtering the current out¬
put and then passing it through the load. If the power content of the
a-c current pulses’coming from the magnetic amplifier is of interest (as it
Sec. 3.20] power amplifiers 165
K- - W.F' 13 84)
where F is the form factor of the current wave shown in Fig. 3.50d.
In the preceding discussion, use has been made of the approximation
that Rc, the control-circuit resistance, is negligibly small. Since in prac¬
tical circuits there is always some resistance in the control circuit, the
question arises of how large Rc may become, while still being considered
negligibly small. As has already been demonstrated, only one reactor
saturates at any one time, and the other one is unsaturated. The unsatu¬
rated reactor acts exactly like a transformer. The small control-circuit
resistance is therefore reflected to the gate winding by means of the
standard transformer relation:
As long as this reflected resistance is very much less than the load resist¬
ance Rl, the voltage drop across the reactors will be negligibly small com¬
pared to the load voltage during periods of conduction. A ratio of Rl to
R'c of 10 or more is usually considered adequate to ensure that the assump¬
tion of negligible Rc does not result in excessive errors. If R'c is greater
than Rl, the reactors operate in a different mode, which is not considered
here.1
3.20. Time Constant of the Series-connected Amplifier. The rela¬
tions derived in the previous section were obtained on the basis of steady-
state operation. When the control voltage is varied, however, the
dynamic performance of the amplifier must also be considered. This
performance is affected chiefly by the time constant of the control wind¬
ing. There is a slight residual time lag in the load winding which is not
considered here.
In order to find the time constant of the control winding, we must
determine its effective inductance Lc. We may write
Ee = Rcic + Ne - Nc
d
= Rcic + Nc (</>i — ^2) (3.86)
d<f>d
(3.88)
die
El = Es - Eg (3.89)
where co = 2x/. The average supply voltage Es is equal to Eg for the case
where the reactor does not fire, i.e., where a = ir. Hence
Es = 8Ng(t>sf (3.93)
From Eqs. (3.89), (3.92), and (3.93) we have now
El = SNg(f)sf — 4:Ng<f)sf(l — cos a)
= 4W„/0S(1 + cos a) (3.94)
But 0i( 1 + COS a) = 0S — ( — 0S cos a) = <j>d (3.95)
Sec. 3.21] power amplifiers 167
from Fig. 3.52, so that, finally, the desired result is
El — 4 Ngf<t>d (3.96)
All that is required now to find Lc is to relate (f)d to ic. By use of Eqs.
(3.85) and (3.86), we have
El _ IlRl _ NcRl
(3.97)
Ic Ic Ng
Hence, from Eqs. (3.86) and (3.96)
Lc _ Nc2Rl
Rc MN^Rc
Note that the time constant is inversely proportional to the supply fre¬
quency. This indicates that, in order to obtain high response speeds,
magnetic amplifiers should be operated from high-frequency supplies.
A commonly used figure of merit applied to power-amplifying equip¬
ment is the product of power gain and bandwidth. This figure of merit
reduces to a very simple result for magnetic amplifiers if the bandwidth is
defined as the reciprocal of the time constant. Using the relation for d-c
power gain developed in the previous section [Eq. (3.83)], we find that
K„ ~ = 4/ (3.100)
rent pulses do not induce pulses in the control circuit, as m the series-
connected amplifier, and the control current consists of a steady d-c cur¬
rent. However, by assumption 1 of Sec. 3.19, the net mmf on any core
must be zero during periods when it
is not saturated. Hence during
these periods there must be a current
in the gate winding such that
Ngig = Ncic (3.101)
When both reactors are unsaturated,
this current flows in opposite direc¬
tions through the two gate windings.
The load current is therefore zero
during this time, and only ig circu¬
lates between the two gate windings.
The waveshape of the two gate¬
winding currents is shown in Fig.
3.54c. The shaded portion repre¬
sents the load current. The average
or d-c value of gate-winding currents
must be zero, since there are no rec¬
tifiers in the circuit. Hence the
average value of a single load-cur¬
rent pulse must be equal to the value
of gate-winding current during the
period when the reactor is not satu¬
rated. For each reactor, therefore,
Fig. 3.54. Waveshapes pertaining to N I
parallel-connected magnetic amplifier. II = (3.102)
where II is the average value of the load current over half a cycle.
Since each reacfor furnishes only half the load-current pulses, the total
Sec. 3.21] POWER AMPLIFIERS 169
j, _ II _ 2Ne
(3.104)
Kl L Nn
The voltage gain is
El _ IlRl _ 2NcRl
Ke = (3.105)
Ec IcRc
The average or d-c power gain is
-tt' _ 4Nc2Rl (3.106)
p ~ ~n7rc
and the rms or a-c power gain is
tv- 4lNc2Rl p<>
_
(3.107)
^ ~ Ng2RC ^
where F is the form factor of the load-current pulses.
The time constant of the parallel-connected amplifier can be obtained
by methods similar to those used in connection with the series amplifier.
An important difference is that in the parallel circuit the two gate wind¬
ings permit the flow of a circulating current which increases the total time
constant. The path of the circulating current is shown in Fig. 3.55a and
may be seen to be equivalent to a short circuit coupled to the control
winding. This equivalence is made somewhat clearer in Fig. 3.555, in
which the circuit of Fig. 3.55a has been redrawn with the lower reactors
turned upside down and the connections to the load removed.
By means of straightforward circuit analysis1 it can be shown that, if
the coefficient of coupling between the two circuits is unity, the effective
time constant is given by
r=b ■tvc
+ wQ (-3-108)
1 See Sec. 4.3 and Eq. (4.18), where this same problem is discussed in a different
connection.
170 CONTROL SYSTEM COMPONENTS [Chap. 3
where Lc and Lg are the effective inductances of the control and gate wind¬
ings, respectively, while Rc and Ra are the corresponding resistances.
The inductance of the control circuit may again be defined as in Eq.
(3.88)
d<f>d
Lc = N (3.88)
c dic
El 2N (if (t>d
El
or 4>d (3.109)
2NJ
Hence Eq. (3.88) becomes
The voltage gain is given by Eq. (3.105) and may be substituted. Hence
T — ^L
(3.113)
R N°* (3.115)
c Nc2
and the time constant becomes
2Nc2Rl
T = (3.116)
NgtRJ
In order to compare the performance of the parallel- and series-connected
Sec. 3.22] POWER AMPLIFIERS 171
amplifiers, it is again instructive to determine the ratio of d-c power gain
to time constant. From Eqs. (3.106) and (3.116) this ratio is found to be
Kv 4:Nc2RL/Ng2Rc
= 2/ (3.117)
T 2N c2Rl/N g2Rcf
Thus it appears that the parallel-connected amplifier has a somewhat
poorer performance than the series circuit. It is clear from the preceding
analysis that this is due entirely to the short-circuit path existing in the
gate-winding circuit.
3.22. Magnetic Amplifiers with Feedback. The performance of mag¬
netic amplifiers can be improved greatly by the use of positive feedback.
The feedback can be applied in a
number of ways. By way of illus¬
tration we consider the series-con¬
nected amplifier with feedback ap¬
plied through a special feedback
winding. In a typical circuit such
as the one shown in Fig. 3.56, the
alternating load current is rectified
and applied to feedback windings
N/ on each of the cores. The feed¬
back windings are connected in op¬
position just as the control windings
are. and the effect of the current Fig. 3.56. Magnetic amplifier with
feedback.
flowing in these windings is the
same as that of the current flowing in the control windings. The equal
ampere-turn relationship shown to exist in the simple reactor must hold
here as well, and we have, therefore,
B ± Nj (3.120)
Nf
Nc 1
then IL = It (3.121)
Na 1 - B
We may also write that
NcRl 1 Ec Nt
El = IlRl = /< Ri (3.122)
Na 1 - B Re Ng 1 - B
The current gain is available from Eq. (3.121) and the voltage gain from
172 CONTROL SYSTEM COMPONENTS [Chap. 3
Le ~ Nc w (3-88)
The relation between <f>d and EL [Eq. (3.96)] is unchanged by the feedback
winding; hence, by the same reasoning as that employed in Sec. 3.20,
N2Rl
(3.124)
4iV/(l - B)
Lc _ Nc2Rl
(3.125)
Wc ~ ±fNg2Rc( 1 - B)
Since inductance varies as the square of the number of turns, we have
Nl Lc
(3.126)
N2
so that for falling transients the time constant is given by
The performance of the circuit may again be judged by the figure of merit
defined above. From Eqs. (3.123) and (3.125) we have that the ratio of
power gain to time constant for rising transients is given by
Kv 4/
(3.128)
T (1 - B)
Sec. 3.23] POWER AMPLIFIERS 173
Thus it is clear that, although the gain and time constant are increased by
positive feedback, there is a net advantage in the use of feedback.
The expressions obtained indicate that the gains and time constant
become infinitely large as B —> 1. This is not found to be true in practice,
primarily because the accuracy of the assumptions made in the derivations
begins to diminish. In particular, the assumption of the square hysteresis
loop (assumption 1 of Sec. 3.19) becomes less and less tenable, and the
actual shape of the hysteresis loop must be considered as the feedback
factor approaches unity.
It should also be noted that a tacit assumption made in the above
analysis was that the rectifiers are perfect, i.e., that they do not permit
any reverse current to flow. In practice, some reverse current always
does flow and has the effect of slightly reducing the feedback factor B.
3.23. The Self-saturated Amplifier. The most commonly used method
of providing feedback in magnetic amplifiers is the method of self-satura¬
tion. A number of circuits employing this principle are shown in Fig.
ftr
/wv*h
§ Load
o—A/W—
o
°
[W-
3.57. In all of these circuits, rectifiers in series with the gate windings
produce a d-c component in the gate-winding current which is equal to the
average (over one half cycle) of the load current. It can be shown1 that,
if the rectifiers are ideal, i.e., if they do not permit any reverse current
flow, the self-saturated amplifier is equivalent to a feedback amplifier in
which the feedback factor B is equal to unity. Hence, as indicated above,
the simplified analysis of the previous paragraphs no longer yields reason¬
able results, and a different approach must be used. We present here
only an empirical point of view, which takes the experimental control
characteristic of the amplifier as its starting point and deduces all further
results from it.
A typical self-saturated-amplifier control characteristic is shown in
Fig. 3.58a. Note that the load current is almost maximum when the
control current is zero and that a “bias’’ value of negative control current
is therefore required to produce minimum output. This is due to the fact
that the rectifiers in the circuit cause the amplifier to saturate itself when
1 W. J. Dornhoefer and V. H. Krummenacher, Applying Magnetic Amplifiers,
Elec. Mfg., March, 1951, vol. 47, part I, pp. 94 ff., April, 1951, part II, pp. 112 ff.
174 CONTROL SYSTEM COMPONENTS [Chap. 3
there is no control current and that the action of the control current is
actually to “desaturate” the cores. It can be shown1 that the sloping
portion of the characteristic from A to B in Fig. 3.58a is identical to the
back, or demagnetizing, portion of the hysteresis loop of the core material.
In order to obtain approximate equations for the gain and time constant
of a self-saturated magnetic amplifier, it is convenient to approximate the
empirically determined curve of Fig. 3.58a by three straight lines, as
K- - w. - mm NJ- (3130)
In these equations Icb is the “bias” value of control current which results
in the minimum output II min, Es is the control current resulting in full out¬
put, and NI = Nc(ICs — Eb) is the number of control ampere-turns required
to vary the output from maximum to minimum. Note that NI is a con¬
stant that depends only on the back slope of the hysteresis loop of the core
material and becomes smaller as the quality of the core material is
improved. The voltage gain is given by
Lc = E'NC2
(3.133)
Rc 2 Rc(NI)N0f
This represents the over-all time constant for rising transients when the
rectifiers are blocked. For falling transients, where circulating paths may
exist through the rectifiers (depending on the exact circuit used), the time
constant will be increased as discussed previously [cf. Eq. (3.108)] and
may be written in general:
(Nc> N/\
(3.134)
2NgfRc(NI) \RC ^ Rj
where Rg is the resistance of the circulating path in the gate winding.
3.24. The Ramey Magnetic Amplifier. A magnetic amplifier operating
on a somewhat different principle
from the ones described above has Dy D2 R,
"L
«- H—ww
been described by Ramey.1 A cir¬
cuit diagram for a single-ended am¬
plifier of this type is shown in Fig. © o<
N,c o! No
0,
3.59. Note that rectifiers are em¬ L0- Control 3
Gate
ployed both in the control and load winding winding
winding, and that on the control side Fig. 3.59. The Ramey magnetic ampli¬
an a-c bias voltage is inserted in ad¬ fier-basic circuit.
dition to the control signal ec.
The operation of the amplifier may be explained most easily by assum¬
ing that ec is a d-c voltage. The direction of this voltage is such as to
apply reverse bias to the rectifier D i. The a-c bias voltage is in phase
1 R. A. Ramey, On the Mechanics of Magnetic Amplifier Operation, Trans. AIEE,
vol. 70, part II, pp. 1214-1223, 1951.
176 CONTROL SYSTEM COMPONENTS [Chap. 3
with the supply voltage es. Thus, if eb = Eb sin cot, es = Es sin cot, where
Eb and Es are maximum values. Suppose now that at time t = 0 the core
is saturated so that the flux 0 = 0S. During the first half cycle the bias
voltage eb is in a direction opposing ec, and the direction of es is such that
the rectifier Z>2 is biased in the reverse direction. No current flows in
either winding until the instantaneous amplitude of eb exceeds ec. At this
time the voltage difference eb — ec is applied to the control winding, and
the directions of the voltages are such that the core is demagnetized.
The flux therefore decreases from the saturated value 0S until the bias
voltage eb is again less than ec. At the end of the first half cycle the flux is
given by the expression
1 r ait = ir — sin-1
where a is the firing angle and where A2 is the area under the supply volt¬
age as shown in Fig. 3.60. The output voltage consists of pulses, and its
average value is proportional to the area shown in Fig. 3.60 as A3. It
Sec. 3.24] POWER AMPLIFIERS 177
should be clear that although a d-c control signal has been assumed for
this analysis, an a-c signal at supply frequency will work just as well.
It will be noted that the amplifier operates on a shared-time principle.
During the first half cycle the control voltage charges the core, and during
the second half cycle the supply voltage discharges it. For this reason
the response time of this amplifier is less than one cycle, and it is therefore
much faster than conventional amplifiers.
A disadvantage of the amplifier is that the control signal must be able
to furnish sufficient power to magnetize the core completely in one half of
a cycle of the supply voltage. Thus the input power required tends to be
larger than for the self-saturated amplifier described in Sec. 3.23, in which
charging of the core takes several cycles. With modern high-grade cores
the required magnetizing current is, however, quite small, and typical
Reactor
amplifiers of this type may have an input impedance greater that 10,000
ohms.
The basic circuit shown in Fig. 3.59 can produce only a half-wave,
unidirectional load current. A somewhat more complex circuit using two
reactor cores is shown in Fig. 3.61. As shown, this circuit produces a
full-wave, unidirectional current in the load, but by reversing the rectifiers
in the lower section the circuit is converted to furnish a half-wave, reversi¬
ble current to the load.
It has been the purpose of these sections to introduce the reader to the
analytic assumptions usually made in the treatment of magnetic ampli¬
fiers and to derive by their use the gains and time constants of several sim¬
ple representative types. This will enable the reader to assess the oper¬
ation of these devices in a control system. For design procedures and
information on problems of cascading, inductive loads, and the like, the
reader is referred to the texts by Storm1 and Geyger2 and to the current
periodical literature.
1 Storm, “Magnetic Amplifiers,” John Wiley & Sons, Inc., New York, 1955.
2 Geyger, “Magnetic Amplifier Circuits,” McGraw-Hill Book Company, Inc., New
York, 1954.
178 CONTROL SYSTEM COMPONENTS [Chap. 3
PROBLEMS
3.1. A half-wave thyratron amplifier is used to control the current in a pure-resist-
ance load of 25 ohms. The supply voltage is 220 volts, and the tube drop is negligible,
(a) Obtain the “small-signal” gain dl/d(j>a for <f>a = 90°. (b) The firing angle of the
circuit is controlled by use of combined a-c and d-c bias with the a-c bias shifted 90°
from the a-c plate supply. If the a-c bias is 55 volts rms and if the d-c signal voltage
is Ein, find the over-all “small-signal” gain dl/dElu for an operating point defined by
<f>a = 90°. Assume that the critical firing potential is very small compared to 50 volts.
3.2. A half-wave thyratron amplifier of the form shown in Fig. 3.15 is used to con¬
trol the current in a generator field. The field has a resistance of 50 ohms and an
inductance of 0.5 henry, and the thyratron amplifier is connected to a 220-volt,
60-cps single-phase supply. It is desired to vary the field current from 0.5 to 1 amp.
(a) Find the range of firing angles required. (6) Find the approximate gain A//A<£a
at the mid-point of the required current range, i.e., for a current of 0.75 amp. Assume
that the tube drop is negligible.
. .
3 3 The firing angle for the thyratron of Prob. 3.2 is controlled by use of combined
a-c and d-c bias, with the a-c bias shifted 90° from the a-c plate supply. If the a-c
bias is 110 volts rms, find the over-all gain of the circuit, dI/dEs, where Es is the d-c
signal voltage, for the case I = 0.75 amp. Assume very small critical grid potential.
3.4. Find the average load current in the half-wave thyratron circuit of Prob. 3.2
if the load resistance is 100 ohms, if the inductance is 0.1 henry, and if the firing angle
is 45°. Assume that Eq. (3.29) applies. The tube drop is 10 volts.
3.5. The load of the full-wave thyratron circuit of Fig. 3.20 consists of a 10-ohm
resistance and an inductance of 0.1 henry. Plate voltage is 110 volts rms, 60 cps,
and tube drop is 15 volts, (a) Find the firing angle 4>a for which the load current is
just continuous. (6) Find the load current for firing angles <£a = 30° and <£a = 150°.
3.6. In connection with the thyratron-controlled motor, show that the change-over
from mode 1 to mode 2 operation and from mode 2 to mode 3 operation is not
abrupt.
3.7. A half-wave thyratron circuit used to drive a d-c motor has a supply voltage
of 220 volts, 60 cps. The thyratron tube drop is negligible. One of two motors is
used with this circuit. One of them has an armature resistance of 0.6 ohm, an induct¬
ance of 0.001, and a Kt of 1 ft-lb/amp, while the other has a resistance of 0.7 ohm,
negligible inductance, and a Kt of 0.9 ft-lb/amp. In all other respects the motors
are identical. Find the ratio of maximum stall torques available from the two motors.
3.8. A relay amplifier with feedback of the sort shown schematically in Fig. 3.35
uses relays which close when the voltage across their coils exceeds 40 volts; they open
when the voltage drops to 25 volts. The amplifier used between the RC filter and
the relay coils has a gain of 20, and the time constant of the filter is 0.1 sec. For a
10-volt peak input signal, find the maximum frequency to which the amplifier can
respond.
3.9. A series-connected magnetic amplifier like that shown in Fig. 3.45a has 5,000
turns on its control winding and 2,000 turns on its gate winding. The control resist¬
ance is 10 ohms, and the frequency of the supply is 60 cps. A load resistance of 100
ohms is connected to the amplifier, and the control voltage is adjusted to produce an
average load current of 0.5 amp. At time t = 0, a 400-ohm resistance is shunted
across the 100-ohm load, (a) Find the control current for t < 0. (b) Find the load
and control currents at the instant after the 400-ohm resistance is added, (c) Sketch
the behavior of the average load current as a function of time after t = 0. Neglect
the pulsed nature of the output, (d) Draw an equivalent circuit that, on the average,
behaves like this magnetic-amplifier circuit.
CHAPTER 4
D-C MACHINES
push-pull amplifier. The currents flowing in the two halves of the field
winding are therefore in opposition, and their effects tend to cancel; hence
the mmf produced in the air gap of the generator is proportional to the
difference of the two currents. Although this method of exciting the
generator field is by no means the only one used, it has the advantage of
permitting reversal of the mmf, and therefore reversal of the d-c motor,
without the need of actually reversing the field currents. It is therefore
almost universally used when the generator field is driven by an electronic
control amplifier.
The generator used in the, Ward-Leonard system may be any one of
several types. In low-power applications a simple d-c generator may
be used, but for power levels above a few hundred watts the power ampli-
179
180 CONTROL SYSTEM COMPONENTS [Chap. 4
NgP(t> 12
E0 = (4.1)
V
where Eg = generated voltage, volts
Na = total number of armature conductors
P = number of field poles
0 = total magnetic flux per pole, webers
12 = speed of rotation, radians/sec
p = number of parallel paths in armature winding
For a particular machine, Na, P, and p are constants, and therefore Eq.
(4.1) may be written in the simpler form
Eg = /ci0£2 (4.2)
1 See, for instance, Bull, ‘‘DC Machinery,” John Wiley & Sons, Inc., New York,
1947, p. 46.
2 Ibid,., chap. 11. Also Dawes, “A Course in Electrical Engineering,” vol. 1,
“Direct Currents,” 4th ed., McGraw-Hill Book Company, Inc., New York, 1952,
pp. 415ff.
Sec. 4.2] D-C MACHINES 181
When a generator is used for power amplification, as in the Ward-
Leonard system, it is driven at constant speed. Hence SI in Eq. (4.2) is a
constant that may be combined with ki, so that only cf> and Eg are variable,
0 being the independent variable. When the generator is open-circuited,
</> depends only on the field current.
The relationship between the flux and the field current is ordinarily
expressed graphically in the form of a hysteresis loop or saturation
Fig. 4.2. Stator, armature, and brush rigging of a typical d-c tachometer.
where Nf is the number of turns of the field and (R is the reluctance of the
magnetic path expressed in suitable units. A combination of Eqs. (4.2)
and (4.3) results then in a linear relation between field current and gener¬
ated voltage:
Eg = kgif (4.4)
The field circuits of almost all d-c generators are highly inductive, since
a large number of turns are used to obtain the required magnetomotive
force with a minimum of field current.
Hence, the inductance must be con¬
sidered, as in the circuit shown in Fig.
4.4. We have
E, = R/if + L, (4.5)
Fig. 4.4. The equivalent circuit of
a d-c generator.
if the armature current is zero or con¬
stant. By use of Laplace transforms we can solve for the field current
Sf WRf)ti, (4.6)
Rf -T LfS 1 -j- TfS
where Tf = Lf/Rf is the field time constant. In this equation the resist¬
ance Rf is the total resistance of the field circuit together with the output
resistance of the driving source considered as a Thevenin generator;1 Ef is
the Thevenin voltage of the source.
1 Everitt and Anner, “Communication Engineering,” 3d ed., McGraw-Hill Book
Company, Inc., 1956, p. 119.
Sec. 4.2] D-C MACHINES 183
When a center-tapped field is used with a push-pull circuit, as shown in
Fig. 4.1, an equation very similar to Eq. (4.6) can he derived. If the
push-pull amplifier stage consists of two vacuum tubes with an amplifica¬
tion factor fj, and a plate resistance rp and if the voltages applied to the
grids of these tubes are eg and — eg, respectively, then an equivalent circuit
for the push-pull arrangement takes
rP
the form of Fig. 4.5. Here the
arrows indicate the assumed posi¬
tive direction of the voltages and
currents. With the current as¬
sumed as shown, the effective mmf
is Nf(ifi — if 2), if both halves of the
field winding have the same number
of turns, N/. R'f represents the
Fig. 4.5. Equivalent circuit of a d-c gen¬
resistance of each field winding erator with push-pull field.
alone; M is the mutual inductance
between the two halves; Ebb and Zb are the voltage and output impedance,
respectively, of the amplifier power supply.
The mmf produced by the two fields may be found as a function of the
control voltage eg by use of Kirchhoff’s laws. In writing the mesh equa¬
tions for the circuit, it is convenient to let
Z = rp + Rrf + LfS + Zb . -v
and ZM = Zh-Ms 1 ]
k\£lk(f)Nf/Rf a (kg/Rf)Ef
(4.11)
Tfs+l Ef = Tfs + 1
184 CONTROL SYSTEM COMPONENTS [Chap. 4
Fig. 4.6. Phase lag produced by hysteresis. Fig. 4.7. Gain change produced by
hysteresis.
The signs of M in Eqs. (4.12) and (4.13) must be the same but can be
either positive or negative, depending on the sense assumed for the wind¬
ing of the coil representing the eddy-current circuit. The sign of Ne in
Eq. (4.14) must, however, correspond to the signs chosen for M in such a
way that the effect of the eddy current ie is to oppose the field current if.
This is required by Lenz’s law. Simultaneous solution of the three equa¬
tions then vields
7 T i Af
77 A fLe NJI
kEf l 1lf + RjRe RfRe ■) (4.15)
LfLe - M*
5+1
RfRp.
This equation can be simplified if we assume that the magnetic paths for
both the eddy-current circuit and the field winding are identical. With
186 CONTROL SYSTEM COMPONENTS [CHAP. 4
Le = ^L, (4.10)
This assumption also implies perfect coupling of the two circuits, so that
M = VlZ} = L, |r (4.17)
paths of the armature and field circuits are essentially at right angles to
each other and the coupling is very small. Compound generators are not
ordinarily used in servomechanisms, but we mention mutual inductance
here in anticipation of the discussion of the more elaborate generators
such as Rototrols and Amplidynes, where the effect is not always
negligible.
In order to assess the general importance of mutual inductance in d-c
generators, we compute the relative magnitudes of the voltages induced
in the armature by speed action and
those induced by mutual induct¬
ance. For this purpose, consider
an armature winding rotating in an
alternating field, as shown in Fig.
4.9. The armature voltage gener¬
ated in each turn of the armature
winding is given by
Fig. 4.9. Simplified schematic of arma¬
ture winding. (4.19)
where </> is the flux in webers linking the particular turn in question. The
flux threading a particular turn is a periodic function of the space position
of that turn. Thus the turn made up of the conductors A and B is linked
by the maximum amount of flux for the armature position shown in the
figure. This position is defined as 6 = 0. As the armature rotates, the
flux decreases to zero as 6 becomes 90°, reverses sign (relative to the turn
under discussion), reaches a maximum for 6 = 180°, goes to zero again,
and again reaches a maximum when 6 = 360°. With the usual shape of
the field poles the function of flux versus 6 is a flat-topped wave, but for
purposes of analysis it is convenient to assume a sinusoidal wave. If the
flux at any particular position of the armature also varies as a sinusoidal
function of time, we may write
4> = 4>m cos 6 cos tot = k+If max cos tct cos 6 (4.20)
where to and If max are the frequency and amplitude, respectively, of the
field current and k$ is a constant, linearity of the magnetic path being
assumed here. Hence, the differentiation indicated in (4.19) gives
dd
= —k^If max( to cos d sin ut + cos ut sin d (4.21)
Eg = kgtf (4.4)
convenient to rewrite Eqs. (4.24) and (4.25) in such a form that Ef and ia
are the independent variables and V0 and if the dependent variables:
The first term in this equation represents the open-circuit voltage, and
except for the very small mutual-inductance term in the numerator, this
term is identical with the expression for the open-circuit voltage already
derived [Eq. (4.11)]. The second term shows the effect of the load cur¬
rent; hence the quantity in brackets multiplying the load current repre¬
sents the output impedance. It is seen that the output impedance con¬
sists primarily of Ra + Las, the resistance and self-inductance of the
armature. The effect of mutual inductance can be estimated approxi¬
mately if we assume that kg y> Mafs and Rfy>Lfs. This can, of course,
be true only at low frequencies; however, if these relations do hold, the
mutual-inductance term reduces approximately to Mafskg/Rf and has the
form of an additional inductance in the armature circuit, which either
subtracts from or adds to the self-inductance La (depending on the sign of
Maf relative to the sign of kg). In a noncompound generator the entire
effect is usually extremely small.
4.5. Figure of Merit. A figure of merit used very commonly for all
types of power-amplifying equipment is the product of power gain and
bandwidth. In order to compute this figure for the generator as a func¬
tion of some of the design features, it is first necessary to give precise
definitions for speed of response and power gain. Both of these quantities
will be defined for a generator connected to a pure resistance load; and
since the power gain is a function of the value of the load resistance, the
definitions are based on the use of a standard load resistance equal to the
generator output resistance. The response speed is then defined as the
frequency in radians per second of the sinusoidally varying control voltage
Ef which results in a reduction of the amplitude of the load voltage by a
factor of 1/\/2 from the very low frequency value. This is the commonly
used half-power-point definition. The power gain is defined as the ratio
of the power dissipated in the standard load resistance to the power
supplied to the field by the control voltage, both powers to be measured
at direct current or at very low frequency.
For a generator with.an output impedance that is purely resistive (i.e.,
Sec. 4.5] d-c machines 191
for which the effects of the mutual and self-inductances in the armature
circuit are negligible), the speed of response is given by
1 _ Rf
(4.29)
Tf + Te L/(l + TV TV)
where Rf and Lf are the resistance and inductance of the field circuit,
respectively, and Te is the eddy-current time constant (see Sec. 4.3).
The power input is
El
(4.30)
Rf
and the power output is
Et
Po = Ri (4.31)
.Ra T Rl,
(4.33)
4:RaLf(l + T e/ Tf)
The quantities entering into this expression are still related to some
extent. Thus, kg is proportional to the number of turns on the field-coil
winding [see Eqs. (4.1), (4.3), and (4.4)], while Lf is proportional to the
square of the number of field turns. Hence, increasing the number of
turns will not affect the figure of merit. By similar reasoning it may be
shown that the number of turns of the armature, the number of poles, the
type of armature winding, etc., all have a negligible effect on the figure of
merit. On the other hand, since both kg and Lf are inversely proportional
to the magnetic reluctance, any decrease of reluctance will result in an
over-all improvement. Such a decrease might be produced by a reduction
of the air gap or the use of a higher grade of iron. The figure of merit is
proportional to the square of the speed, since kg is proportional to speed.
This indicates that higher performance can be obtained from a given
machine by running it at higher speed. This fact is used in some aircraft
generators.
The particular figure of merit computed here gives an indication of
which of two generators is “ better,” if the only criteria are power gain and
bandwidth. However there are other criteria, and therefore other figures
of merit, that may be more pertinent in a given situation, and hence no
single figure of this sort is ever adequate to give a complete evaluation of
192 CONTROL SYSTEM COMPONENTS [Chap. 4
Both the pilot generator and the main generator are driven at constant
speed by a prime mover not shown in the figure. The pilot generator has
at least two fields, one of which is a series field L i and the other the control
field. In the figure a pilot generator with a single push-pull control field
is shown; however many pilot generators, particularly those designed for
heavy-duty industrial service, may have four or five separate control
fields in addition to the series field. This multiplicity of control fields
makes it possible to build a control system for fairly high performance
without the use of vacuum tubes. Such a system is shown in Fig. 4.13.
Since the basic operation of a Rototrol is not affected by the number of
control fields used, we shall concentrate on the single-control-field type in
the following paragraphs and show later (Sec. 4.8) what changes are
necessary in the theory when additional control fields are used.
4.7. Analysis of the Rototrol Generator. The mmf responsible for the
air-gap flux in the pilot generator is the sum of the ampere-turns of the
control field and the series field. For a given value of control-field current
and series-field current the generated voltage of the generator may be
determined from an average magnetization curve of the sort described
1 W. H. Formhals, Rototrol, a Versatile Electric Regulator, Westinghouse Engr.,
vol. 2, pp. 51-54, Mdy, 1942.
Sec. 4.7] D-C MACHINES 193
(4.35)
Eg kg
* ,N i,
(4.36)
tf + m11
Ec
^l (4.37)
R 4“ Ls
where R is the total resistance and L the total inductance of the armature
circuit. Note that R may include armature reaction and L mutual-
inductance effects, as explained in Sec. 4.4. Combining Eqs. (4.36) and
(4.37) gives then
kglf
ii = (4.38)
R - (kgN i/N/) + Ls
The current i\ is the field current for the main generator. If this gener¬
ator is linear, then its open-circuit output voltage is
E0 = k2ii (4.39)
Eo (kg/Rf) k2
(4.40)
Ef (Tfs + 1 )[R- (kgNi/Nf) + Ls]
kgN!
R = Rr = (4.41)
N,
Hence the transfer function for a critically tuned Rototrol becomes
E0 kgk2 1
(4.42)
Sf RfL S(Tfs + 1)
The free $ in the denominator substantiates the statement made above
that the critically tuned Rototrol is an integrating device.
The effect of eddy currents in the iron of either the pilot generator or
the main generator is fundamentally the same as for the simple generator.
196 CONTROL SYSTEM COMPONENTS [Chap. 4
For the pilot generator, Eq. (4.18) applies without change, and eddy cur¬
rents simply increase the effective field time constant. For the main
generator, it can be shown by arguments similar to those employed in
connection with Eq. (4.18) that eddy currents simply decrease the total
inductance L of the armature circuit. Demonstration is left to the reader.
Mutual inductance between the control circuit and the armature cir¬
cuit also has essentially the same effect here as in the simple generator,
and an analysis similar to that carried out in connection with Eq. (4.28)
leads to the result that Eq. (4.40) must be changed to read
Eo (k2/Rf)(kg - Mafs)
(4.43)
Ef kgNi T (kg — Mafs)Mafs
(T fS + 1 )R —
~N7 + Ls+ R, + Lfs
Eo k\t i if i2 = 0
Eo k2t2 if ii = 0
and R\ and R2 represent the total resistance, including that of the voltage
source in each of the two control circuits. This result may be generalized
to more than two control circuits.
4.9. The Regulex Generator.
Instead of using a series field for
the self excitation of the pilot
generator, as in the Rototrol, it is
also possible to use a shunt field.
The resulting two-stage generator
is called the Regulex and has been
Fig. 4.15. A schematic of the Regulex gen¬
used in systems built by the Allis-
erator.
Chalmers Company. A schema¬
tic diagram of this machine is shown in Fig. 4.15. Here again the
pilot generator is shown with a push-pull field designed for excitation
by an electronic amplifier; however this generator, like the Rototrol,
is often built with several control fields. The theory of operation is
198 CONTROL SYSTEM COMPONENTS [Chap. 4
very similar to that of the Rototrol, and by proper adjustment of
the shunt-field resistance Ri, it is possible to obtain integrating action from
this system. The complete transfer function has a slightly different form
from that of the Rototrol; its computation is left to the reader.
4.10. The Amplidyne Generator. A generator often used in control
systems is the Amplidyne, first described by Alexanderson, Edwards, and
Bowman.1 The Amplidyne is essentially a two-stage generator, like the
Fig. 4.16. Amplidyne principles: (a) <£/ is the flux set up by the control- field current I/;
4>Q is caused by current Iq. (b) Quadrature brushes short-circuited; </>d opposes <£/,
creating a constant-current generator, (c) Addition of compensating winding con¬
verts machine to Amplidyne.
Rototrol described in the previous sections; however the two stages are
combined in a single machine with a single armature winding. The
Amplidyne therefore tends to be somewhat smaller in size than a Rototrol
of the same rating.
The principle of operation is as follows: Consider the simple d-c gener¬
ator of Fig. 4.16a with a control-field current I/. This current produces a
flux <f)f, and if the machine rotates in the direction shown, the voltage eq
induced in the conductors will be in the direction indicated by the con-
1 E. F. W. Alexanderson, M. A. Edwards, and K. K. Bowman, Dynamoelectric
Amplifier for Power Control, Trans. AIEE, vol. 59, pp. 937-939, 1940.
Sec. 4.10] D-C MACHINES 199
ventional system of dots and crosses. If a current is taken from the
brushes, an armature-reaction flux <j>q due to this current appears at right
angles to <£/. In conventional generators every attempt is made to sup¬
press the armature reaction, but in the Amplidyne a magnetic path is pro¬
vided to encourage it. Furthermore the brushes are short-circuited so
that the maximum amount of 4>q is produced for a given control-field cur¬
rent. The armature-reaction flux will then also induce voltage in the
armature conductors. This voltage is in a direction indicated by the dots
and crosses of Fig. 4.166. The second set of brushes shown in that figure
is used to connect this voltage to the load. We refer to this voltage as the
direct-axis voltage ea. Since the quadrature flux (f>q is very much larger
than the control flux <f>f, the direct-axis voltage is very much larger than
the quadrature-axis voltage eq. The additional amplification takes place
in the magnetic circuit of the machine. The armature-reaction flux is
somewhat analogous to the pilot-generator armature current of the
Rototrol system.
When the output brushes of the machine shown in Fig. 4.166 are con¬
nected to a load, the resulting current flowing in the armature will again
result in an armature-reaction effect, and the reader may convince himself
that this effect opposes the control-field flux. The load current is there¬
fore limited to a value such that the mmf of this armature reaction is
equal to the magnetomotive force produced by the control field If. In
other words the machine acts as a constant-current generator. Its load
current is proportional to the control-field current and essentially inde¬
pendent of speed and direction of rotation of the armature. Referred to
as a Rosenberg generator, this type of machine has been used for railroad-
train lighting service since 1905, and as the Metadyne it has been used in
diesel-electric locomotive drives.1
In servo systems the constant-current feature is usually undesirable
and is eliminated by passing the load current through a compensating
field so located that the armature reaction of the load current is canceled
by the mmf of the compensating field. This is shown in Fig. 4.16c. The
addition of the compensating field converts the machine to an Amplidyne,
a machine delivering a voltage that is only moderately affected by the load
current. Amplidynes are occasionally built with a number of different
control fields for purposes similar to those already discussed in connection
with the Rototrol. Sometimes the quadrature brushes are not short-
circuited directly; instead they are connected through field windings that
may be located in either the quadrature or direct axis to obtain certain
special operating characteristics such as, for instance, integrating action.
A major problem in Amplidynes is sparking at the brushes, due to the
armature reaction on which the operation of the machine depends.
1 Pestarini, “ Metadyne Statics,” John Wiley & Sons, Inc., New York, 1952.
200 CONTROL SYSTEM COMPONENTS [Chap. 4
Special brushes and construction of the field poles are used to alleviate
this difficulty. Nevertheless Amplidynes often spark very severely at
the quadrature brushes when delivering large output voltages.
4.11. Analysis of the Amplidyne. The following analysis of the Ampli-
dyne is based on a paper by Bower1 and assumes complete linearity of
the electric and magnetic circuits, neglects effects of eddy currents in the
iron, and assumes the machine to be running at constant speed. The cir¬
cuit under discussion is shown in Fig. 4.17, and the resistances and
inductances indicated in that figure in all cases represent the entire resist¬
ance and inductance of each of the three circuits. We shall also be con¬
cerned with mutual effects between the circuits and shall distinguish
Rf
-vWf
T Control
Ef /) Lf field
the quadrature circuit. This parameter could be lumped with the resist¬
ance Rq to form an “effective” resistance R'q, as discussed in connection
with the simple generator (Sec. 4.4). Ndq represents the speed voltage
in the quadrature circuit resulting from load current in the direct circuit.
It is not much affected by brush shift but is a function of the amount of
compensation supplied by the direct-circuit compensating coil. It is
assumed to be positive for an undercompensated machine. Therefore, if
Ndq is positive an increase of load current causes a decrease of the quadra¬
ture current and hence the output voltage.
For the direct circuit we have
This equation has the form of a generated voltage, Nqdiq, set equal to the
sum of the impedance drop and the terminal voltage. All the remarks
made previously concerning the signs of the mutual terms apply here
without change. Ndd is the same type of parameter as Nqq in Eq. (4.46).
Nfdif is the speed voltage generated in the direct circuit by direct cou¬
pling from the control field. It is zero if the brushes are on the neutral
axis. Nfd is positive when the brushes are shifted in the direction of
rotation.
The primary purpose of this analysis is the determination of the
Thevenin equivalent of an Amplidyne; hence we wish to solve Eqs. (4.45)
to (4.47) in such a way that the open-circuit voltage and output imped¬
ance become evident. This is most easily done by rearranging these
equations in the form
(U-v ,/>
Ef — idZdf = Zffif + Zqfiq
— $dZdq — Zfqif -fi Z„„% qqvq (4.48)
— idZdd — Zfdtf + Zqdlq + Vo
Zff Rf L/S
Zfq — Nfq — MfqS
Zd q Ndq — MdqS
Z dd Rd + Ndd + LdS
Sec. 4.11] d-c machines 203
made exactly equal and opposite to Rq, and the open-circuit output volt¬
age becomes approximately
p __ p N f qN gd/RfL q
(4.52)
tj° s(Tfs + )
1
indicating that the output voltage (at low frequencies) is the integral of
the input voltage. If an Amplidyne is to be used in this way, a series coil
with an adjustable shunting resistor may be used to adjust Nqq.
The second term of Eq. (4.51) represents the component of the output
voltage resulting from direct coupling between the control field and the
output circuit. If the brushes are not shifted from their neutral position,
N/d is theoretically zero; hence the only coupling remaining is that caused
by mutual inductance. This will also be quite small in an Amplidyne
that is properly compensated. Thus, the entire term is usually neglected.
It should be pointed out, however, that while the resulting low-frequency
approximation to the Amplidyne transfer function is usually satisfac¬
tory, one occasionally finds an unusual system in which there is a minor
loop or a parasitic loop with a wide passband. In such a system the
presence of a term like the second term of Eq. (4.51) that does not vanish
at high frequencies may make the difference between a stable system and
one that oscillates at high frequencies.
The output impedance is given by a much more complicated expression
than the open-circuit voltage, and our simplifications must therefore
become more extensive. We assume that the brushes are located exactly
on the neutral axis, so that
NdqNqd NdfNfqNqd
Rq T~ Nqq R,(Bq + N„) s
Z0 — Rd + Ndd T* RdS +
TqS + 1 (TfS+ + )
1 )(Tqs 1
_ R, (4.54)
Tfs + 1
In this expression the first three terms represent simply the self-imped¬
ance of the load circuit, and it is almost obvious that it would be a compo¬
nent of the output impedance. The first fractional term represents the
effect of imperfect compensation on the output impedance. It is positive
in an undercompensated Amplidyne, and in many commercial machines it
Sec. 4.12] D-d Machines 205
is several times larger than the first term at low frequencies. This term
may be represented as the impedance of a resistance in parallel with a
capacitance, as shown in Fig. 4.19, and it may be so represented in an
NdaN
dq+V qd R
R VAAAAA-
Rq + N qq
RC = T, C
C=
Ndq Nqd
Rd+ Ndd Ld
■K
-wwvw-
-\MAA-
Nfq Nqd
_ Ndq Nqd
Co ~ Rf^Rq+Nqq^ Ef
Rq + Nqq Vo
(Tf s + 7)(Tq s + /)
The values of gain and output resistance depend, of course, on the size
and other construction features.
The reader should note that one effect of the parallel RC circuit in the
equivalent circuit is that the quadrature-circuit time constant Tq does not
necessarily appear in the transfer function of a loaded machine. The
demonstration of this fact is left as an exercise (see Probs. 4.3 and 4.4).
206 CONTROL SYSTEM COMPONENTS [Chap. 4
however, the brush-contact drop is a very large part of the total voltage
drop, since the brushes are short-circuited. The exact nature of the
brush-contact voltage depends on the brush material and contact pres¬
sure. The general shape of the voltage as a function of current is similar,
however, to that found in electric arcs: a high voltage at small currents,
decreasing as the current increases (see Fig. 4.21a). The result is that a
small quadrature voltage can produce only a very small quadrature cur¬
rent, but as the quadrature voltage increases, the current becomes larger
much more rapidly, sometimes showing a sort of trigger action due to the
brush-voltage function. As a result of this action, the over-all hysteresis
curve (output voltage versus control-field current) characteristically has
a shape similar to that shown in Fig. 4.216, with a deadband effect at the
center. When an Amplidyne is used in a control system, this deadband
tends to reduce the loop gain and the accuracy as the input and output
approach correspondence. There may also be a tendency to hunt, if the
deadband is very severe.
Sec. 4.15] D-C MACHINES 207
4.14. Comparison of the Various Generator Types. A question of
considerable practical interest is which of the three generators, the Ampli-
dyne, the Rototrol, or the Regulex, is “best” for the application at hand.
While we can give no complete answer to this question, a short discussion
relating the various features of the machines may be helpful.
The Rototrol and Regulex are very similar. They are essentially two
simple generators operated in a cascade connection, equipped with a
special feedback connection that makes the output proportional to the
integral of the input. In other respects their characteristics do not differ
from those of two simple generators designed to be used together. They
would have the same maximum power output, the same overload capac¬
ity, the same size and weight, and about the same speed of response. The
integrating feature is of interest primarily in control systems that do not
use electronic amplifiers to drive the generator fields. If electronic con¬
trol amplifiers are used, the integration can be performed more easily and
probably more accurately within the control amplifier. Thus, if a given
power output is required, there seems little reason to choose either the
Rototrol or the Regulex in preference to two simple generators.
The Amplidyne is also essentially a two-stage generator; however, the
two stages are combined on a single armature. In performance the
Amplidyne is probably somewhat poorer than two separate generators.
This is due to the various interactions taking place inside the machine.
The problem of rather heavy sparking at the brushes has already been
mentioned. Furthermore, although a number of different characteristics
may be obtained from the Amplidyne by brush shifting or different wind¬
ing arrangements, it seems clear that two separate generators driven by
an electronic amplifier should permit much greater flexibility in the adjust¬
ment of the characteristics to meet a particular requirement. The chief
advantage of the Amplidyne is undoubtedly its compactness. Since it
uses only a single yoke and armature, it occupies less space and weighs less
than two equivalent cascaded generators would. Hence Amplidynes
are often preferred in aircraft applications.
4.15. D-C Motors with Separate Excitation. When one of the d-c
generators discussed in the previous section is used to control the speed of
a d-c motor, the arrangement used is usually the familiar Ward-Leonard
system shown in Fig. 4.1, where the motor field is separately excited
by a fixed supply and the two armatures are connected. The transfer
function relating the speed or position of the motor shaft to the control
signal applied to the generator will depend on the type of generator
used. Since all the generators discussed thus far may be replaced by
equivalent circuits of the Thevenin type, it is only necessary to find the
general transfer function of a motor connected to such an equivalent
generator.
208 CONTROL SYSTEM COMPONENTS [Chap. 4
The linear analysis of motors requires (1) the assumption of viscous
friction at the output shaft, (2) linearity of the electric and magnetic
circuits, (3) negligibly small armature reaction. For the moment, no
attempt will be made to justify these assumptions, but the effects of the
nonlinearities and armature reaction will be considered later.
The system shown in Fig. 4.22 will be considered. The motor is con¬
nected to a generator with open-circuit voltage E0 and output impedance
Z0. It has an armature resistance Rm and an inductance Lm. The arma¬
ture has a moment of inertia J, and the coefficient of viscous friction
between the armature shaft and a stationary reference system is B. The
armature current is ia, and the motor shaft position is 6. The motor is
connected to a load requiring a load torque Ql.
A E0 - Ec
^a (4.59)
Zo + Rm + LmS
Finally the sum of all torques applied to the motor shaft must be zero, or
The three equations, (4.55), (4.59), and (4.60), contain five unknowns;
hence we can solve for three as a function of the other two. We consider
E0 and Ql the independent variables and solve for 0. The solution is
(Eo/kv) - (QLZ/kvkt)
(4.61)
s[(JZ/kvkt)s + (BZ/kvkt) d- 1]
a_ K/kv QhR/kykt
(4.64)
s[(JR/kvkt)s + 1] s[(JR/kvkt)s + 1]
where /, _ 1 (4.66)
1 fc„( 1 + BRa/kA)
7 _ Ra (4.67)
and
BRa + kvkt
Since all the parameters defining k\ and k2 are constant in a linear motor,
the speed-torque curves are parallel straight lines, and they are equidis¬
tant for equal increments of E0. A typical set of such speed-torque
curves is shown in Fig. 4.23.
Of greater practical interest than the determination of the speed-torque
curves from the transfer function is the reverse process, namely, obtaining
the transfer function from the speed-torque curves. This is often a con¬
venient procedure in practice when the speed-torque curves have been
measured or are otherwise available. The moment of inertia of the motor
armature must be known. Suppose that it has the value J and suppose
an external torque Q0 to be applied to the motor. Setting QL in Eq. (4.65)
equal to Q0 + J dO/dt and using Laplace transform notation, we obtain
sd = kiE0 — k2Js2d — k2Q„
A
z _ kiE0 k2Q0
or (4.68)
s(k2Js -j- 1)
If the values of Aq and k2 from Eqs. (4.66) and (4.67) are substituted into
Eq. (4.68), it will be found that an expression almost identical in form
with Eq. (4.61) is obtained. The only difference is that the armature
impedance Z in Eq. (4.61) has been replaced by the resistance Ra. Thus
the transfer function that has been obtained from the speed-torque curves
is correct if the armature inductance is negligible.
Sec. 4.16] D-C MACHINES 211
by a nonlinear source of some sort (see Fig. 4.24). The curves are
assumed to be plotted for integral increments of some input variable x
corresponding to the generated voltage E0 in Eq. (4.65). At the oper¬
ating point P the average vertical distance between the curves is equal to
the parameter kh while the slope of the particular curve passing through
P is — fc2. By direct analogy with Eq. (4.68) the transfer function at the
operating point P is then given by
kix — k2Q0
(4.69)
s(/c2*/s -f 1)
Since the speed-torque curves are not linear and not equidistant, the
parameters ki and k2 will vary with the operating point.
It should be clear from the above discussion that the transfer function
obtained from the speed-torque curves can show only a single time con-
212 CONTROL SYSTEM COMPONENTS [Chap. 4
stant, namely, the inertia time constant. Its accuracy depends, there¬
fore, on the assumption that other time constants are negligible.
4.17. An Equivalent Circuit for the D-C Motor. It is sometimes con¬
venient to replace the motor by an equivalent electric circuit element.
Such an element can be found if Eqs. (4.55), (4.59), and (4.60) are solved
for the armature current in terms of E0 and Ql. The result may be put
in the form
Ql/kt
E0 +
(J/kvkt)s + B/kvkt
(4.70)
1
Z0+ Rm + LmS + (.J/kvkt)s + (B/kvkt)
If for the moment we ignore the load torque, it is seen that the denomina¬
tor represents an impedance consisting of the elements ZQ, Rm, Lms, all
connected in series with a capacitor
Rm <-m
J/kvkt in parallel with a resistance
-VWNA-
B/kvkt. Furthermore, it may be
Motor
termino/s
J B__ demonstrated that a current source
kvkT kykj
Qh/kt connected in parallel with
the combination J/kvkt and B/kvkc
Fig. 4.25. An equivalent circuit of a d-c
motor.
properly represents the load. The
circuit therefore takes the form of
Fig. 4.25, where we have omitted the output impedance Z0 of the gener¬
ator. It can be shown that the voltage across the capacitor is the counter
emf of the motor, and the speed can therefore be obtained by dividing the
capacitor voltage by lcv.
Sec. 4.19] D-C MACHINES 213
If friction is neglected and if there is no shaft load, the equivalent cir¬
cuit is a simple RLC series circuit, and all equations pertaining to the
motor will have the form of the equations of this circuit.
4.18. Inaccuracies in the Motor Analysis. The assumptions of linear¬
ity, etc., that were made to facilitate the analysis will now be examined
briefly. Nonlinearity of the electric circuit is confined primarily to the
brush-contact resistance and will cause a slight deadband effect in the
motor’s curve of speed versus input voltage. This effect was discussed
in connection with the Amplidyne (Sec. 4.13) but is so small in a motor
that it is usually neglected. A much more pronounced deadband effect is
produced by the static friction between the rotating and stationary parts
of the motor; it is sometimes serious enough, particularly in small motors,
to cause instability of the servo of which the motor is a part. Aside from
this effect, however, the fact that the friction drag is rarely proportional
to speed is of no great consequence, since it was shown that the friction
term can usually be neglected in the motor transfer function.
Armature reaction in motors results in a reduction of flux, just as it
does in generators; hence both kt and kv are reduced slightly when large
armature currents flow. It is seen therefore that armature reaction in a
motor cannot simply be absorbed in an equivalent resistance as it is in
the generator. However, since the field is usually operated in saturation
and since compensating windings and commutating poles are often used
on motors to improve commutation, armature reaction is not usually a
major factor in motor performance.
Nonlinearity of the magnetic circuit can have no effect on the perform¬
ance, since the motor field is maintained (in the absence of armature reac¬
tion) at a constant strength.
4.19. Other D-C Motor Types. A motor that is fairly often used in
low-power servo applications is the split-field series motor, a circuit of
\vhich is shown in Fig. 4.26. These
motors are usually driven directly from
a push-pull electronic amplifier (see Sec.
3.13), or by a polarized relay, as shown
in Fig. 4.28, rather than from a gener¬
ator. For this reason their power out¬
put is limited to about 50 watts. The
Fig. 4.26. The split-field series
two series field coils are wound in such motor.
a way that, for the current directions
shown in Fig. 4.26, the mmf’s oppose; hence the flux is approximately
proportional to the current difference, if the magnetic circuit is assumed
to be linear. Since the armature current is the sum of the two field
currents, the developed torque is
Qd = kt(Ji + /2)(/1 - U) (4.71)
214 CONTROL SYSTEM COMPONENTS [Chap. 4
dE i dE 2
(4.73)
dlo Ii, 0=0 dh 12, 0 = 0
Note that these equations are not written in Laplace transform notation,
since they contain products of the variables. Solving (4.74) for 7i and
72, we get
E\ZS — E%Zm — kv0(E\ — E2)
Z/- Z.iR
(4.75)
E<iZs — E\Zm — kvd(E\ — E2)
Zs2 - Zm2
Let us suppose now that E1 and 772 vary an equal and opposite amount
from a quiescent value E0; i.e.,
Ei = E0 4- AE (4.76)
E2 = Eq — A E
Then the substitution of (4.76) and (4.75) in (4.71) gives for the developed
torque
q _ 4.kfEo AE $ktkv0(AE)2 ,
^ “ (Zs2 - Zm2) (Zs2 - Zm2)(Z8 - ZM) 1 j
We see that, under the assumptions made, the motor does not supply
electrical damping and that the only damping is due to mechanical fric¬
tion. As a matter of fact, in the absence of friction the motor shaft posi¬
tion is the double integral of AE, the control voltage. This behavior
differs quite markedly from that observed in the separately excited motor
discussed in the previous sections.
It should be noted, however, that in general the damping supplied is
proportional to the negative of the slope of the speed-torque curve (see
Sec. 4.16). Hence, at large speeds and large control voltages, the split-
field series motor is very heavily damped. The implications of this state¬
ment and the effect of this nonlinear damping on a servo loop in which a
split-field series motor is used will not, however, be pursued here. As
216 CONTROL SYSTEM COMPONENTS [Chap. 4
has already been mentioned, the motor is often used with a polarized
relay in a circuit such as is shown in Fig. 4.28, where either one field or
the other is excited. The combined nonlinearities of the relay and the
motor are then usually such that only
the most approximate analysis is pos¬
sible,1 and the servo must be designed
more by experience and trial and error
than by any sort of exact analysis.
4.20. The Field-controlled Motor.
Still another motor-control method
Fig. 4.28. Relay-operated split-field
used in servo systems employs the
series motor.
field-controlled motor, in which the
armature is supplied with a constant current. Since the control amplifier
needs to supply only the relatively small field current, this drive may be
used in moderately large power applications without including a generator
in the loop to provide additional power amplification. In a sense the
motor acts as a power amplifier itself. A generator is, however, needed to
supply the required constant armature current; hence the amount of
equipment needed is still about the same as for a Ward-Leonard system
of similar power rating. The only advantage from the point of view of
installed power capacity is that the constant current might possibly be
furnished directly from the a-c line by a judicious arrangement of trans¬
formers and rectifiers. This equip¬
ment might be lighter, less expen¬
sive, and easier to maintain than a
rotating generator.
The motor is commonly built with
a center-tapped field coil to permit
Fig. 4.29. A schematic diagram for the
direct operation by a push-pull am¬ field-controlled d-c motor.
plifier. Since it has already been
shown in Sec. 4.2 that a push-pull field can always be replaced by an
equivalent single-ended field, we consider this type here, as shown in Fig.
4.29. If linearity of the magnetic path is assumed, the developed torque
is given by
Qd = kJatf (4.80)
The field current is given by
Et
V (4.81)
Rf(l + Tfs)
where Tf = Lf/Rf.
If the armature has a moment of inertia J and a coefficient of viscous
ft — (kjg/Rf)Ef _Ql
s(Tfs + 1 )(Js + B) s(Js + B) V y
Here again we note that the motor does not provide any electrical damp¬
ing; the only damping is due to mechanical friction. As in the split-field
series motor, if friction is negligible, the motor acts as a double integrator.
Very often the armature current for this type of motor is not obtained
from a true constant-current source, but rather an approximately constant
current is obtained by connecting the armature to a constant-voltage sup¬
ply through a high resistance. If the motor is permitted to run only at
speeds low enough that the counter emf remains negligibly small com¬
pared to the supply voltage, the armature current will remain almost
constant. This method of supplying the armature has the advantage of
doing away with an expensive and complicated constant-current supply.
A large amount of power is, of course, wasted in the resistance, but since
the drive can be used only for power outputs of about 500 watts or less,
this is a relatively minor disadvantage.
If the voltage supplying the armature is V and if the total impedance in
the armature circuit (including series resistance, armature resistance, and
inductance) is R + jwL, the armature current is given by
. _ V - Ec
(4.83)
u R + jo>L
where Ec, the counter emf, is given by
Ec = kvifd (4.84)
The developed torque is still given by
Qd ktialf (4.85)
Note that Eqs. (4.83) through (4.85) are not written in Laplace transform
notation, since they involve, on the right side, a product of two variables.
In order to get an approximate result, we make a linear approximation of
the two nonlinear equations as follows:
For small variations of if and d around some operating value, Eq. (4.84)
may be approximated by the linear term of its Taylor expansion
dEc 6E,
A Er. = Aif + Ad
dif Ec = Ec 66 Ec = Ec
\
= kv60 Aif + kvif0 A6 (4.86)
where Ec0, 6a, and i/a are the operating-point values of the quantities in
question. Similarly, Eq. (4.85) becomes
A Qd kitao Aif —h kiZfo Ata (4.87)
218 CONTROL SYSTEM COMPONENTS [CHAP. 4
— AEC
At a (4.88)
R d- Ls
since V is a constant. If now Eqs. (4.86), (4.87), and (4.88) are combined
with the torque balance [Eq. (4.78)] we obtain
PROBLEMS
4.1. Obtain the transfer function of the Regulex generator system (Fig. 4.15) and
determine the value of Ri required for critical tuning. The slope of the magnetiza¬
tion curve of the pilot generator is ki volts/amp of field current, while the slope of the
magnetization curve of the main generator is k2 volts/amp.
4.2. Show in what way the transfer functions of the Amplidyne are affected by
the removal of the compensating winding in the direct-axis circuit. Indicate in
what way the machine becomes a constant-current Rosenberg generator, and find
the transfer function relating control-field current to load current.
4.3. The Amplidyne shown in the equivalent circuit of Fig. 4.20 is loaded by a pure
resistance R0. Assuming that the inductance Ld is negligible, find the transfer func¬
tion V0/fif. Note that the time constant Tq does not appear in this transfer function.
4.4. The Amplidyne shown in the equivalent circuit of Fig. 4.20 drives a d-c motor
having a moment of inertia ./, armature resistance Rm, a counter-emf constant kv,
and a torque constant kt. The inductance of the armature and the mechanical fric-
Problems] D-C MACHINES 219
lion of the armature are negligible. Also neglect the inductance La of the Amplidyne.
Find the transfer function 6/Ef, where 6 is the motor shaft position. Note that the
time constant Tq does not appear in this transfer function.
4.5. The following two tests are performed on an Amplidyne to determine its
characteristics:
a. Open-circuit step-function test: For a step function of field current of 1 ma the
open-circuit output voltage is E0 = 10(1 — e~50t) volts.
b. Short-circuit step-function test: For a step function of field current of 1 ma the
short-circuit output current is 70 = 1(1 — e_500<) amp.
This Amplidyne is used to drive a d-c shunt motor which has an armature resist¬
ance of 1 ohm, kv = 0.1, kt = 0.1, J = 0.0002 in a consistent set of units; friction,
armature inductance, etc., are negligible.
Assume the simplest possible equivalent circuit for the Amplidyne; i.e., all mutual
inductances are zero, and the inductance of the direct axis is negligible. Find a rela¬
tion for the transfer function from Amplidyne field current to motor speed.
4.6. A d-c motor has the following ratings:
a. Find the ratio BR/kvkt in terms of these ratings, where B = coefficient of viscous
friction, R = armature resistance, kv = voltage constant, kt = torque constant.
b. Determine this ratio for a motor for which
4.7. Show that the characteristic speed-torque curves of the field-controlled d-c
motor shown in Fig. 4.30 are similar to the speed-torque curves of the split-field series
motor (Fig. 4.27). Show that the quasi-linear transfer function of the motor is,
therefore, similar to that of the split-field series motor.
Fig. 4.30
CHAPTER 5
able component of the error as the average value Em, and although the
probability curve indicates that the random error extends over a very
large range (theoretically infinite), the randomness is usually measured by
a, the standard deviation defined in Eq. (5.1). It is convenient to normal¬
ize these errors with respect to the range of the measured variable and to
agree on the following definitions:
. . range of variable
Accuracy = ---=-- (5.2)
m
t. . . A range of variable
.Precision = --- (5.3)
<7
The accuracy and precision as defined in these equations increase for the
better instrument, which is in accordance with normal usage of these
words. Note that the error contributing to poor accuracy is in the nature
of an error in calibration. It is normally desirable that transducers be
linear, i.e., that the output variable be directly proportional to the input
variable. Any departure from linearity would therefore give rise to a
repeatable error in a nominally linear transducer and thus reduce the
accuracy as here defined. Also, in many servomechanism applications,
two transducers with closely matched characteristics are used, one at the
input and the other at the output. Any mismatch between these trans¬
ducers would result in a repeatable error and adversely affect the accuracy
of the system. In the following discussion of errors in transducers we
shall be concerned primarily with repeatable errors, since the random
errors contributing to lack of precision are primarily a function of the
workmanship and care expended in the construction of the transducer.
In addition to accuracy and precision, another desirable characteristic
of transducers is high input impedance; i.e., the transducer should require
a minimum amount of power for its operation. This is particularly
important in transducers operating at the input to a servomechanism,
since one of the primary functions of an automatic control system is
accurate power amplification between input and output. In some cases,
it is required that only an infinitesimally small amount of power be
extracted from the input; photoelectric devices are often used in such
applications. Other desirable characteristics are rapidity of response
and low noise output.
5.2. Elementary Operation of Synchros. The most common form of
electromagnetic transducer used to convert an angular position of a
rotating shaft into an electric signal is the synchro (also called selsyn or
autosyn). It is typical of the class of induction transducers commonly
employed in control and measurement applications, and the results of the
discussion on synchros are applicable with some modifications to these
other transducers also.
222 CONTROL SYSTEM COMPONENTS [Chap. 5
where 6 and a are the angles, measured from the same reference, of the
pole axes of the generator and control-transformer rotors, respectively.
Eom is the maximum value of the output voltage. It normally is equal to
\/2 X 57 volts in 115-volt synchros. The phase shift (3 between input
and output voltages is the sort of phase shift found between primary and
secondary voltages in a transformer and is due to the resistance of the
windings. Since the ratio of resistance to reactance is usually quite
small, /3 also is usually of the order of only a few degrees and is often
neglected.
In operation in a servo system the rotor of one of the synchros, usually
the control transformer, is connected to the output shaft, and the input
is applied to the other rotor. The voltage e0 is applied to the amplifier,
which feeds the motor and turns the output shaft. Hence the servo
operates to make the output voltage from the control transformer zero, so
that, under quiescent conditions, 6 — a = 90°. It is convenient, there¬
fore, to define an angle 5 = a + 90°, in order that the condition for zero
output voltage may be h = 6. In this case
e0 = Eom sin (5 — 0) sin (cct — (3) ^ Eom(& — 6) sin (cct — (3) (5.5)
5 — 6 = a m sin cost
This output voltage is shown in Fig. 5.5 together with the voltage applied
to the generator, the so-called reference voltage. The phase angle (3
Control-
tronsformer
output
voltage
Reterence
voltage
wvwww
Fig. 5.5. Control-transformer output-voltage waveforms.
is taken to be zero. The figure shows the 180° phase reversal of the out¬
put voltage relative to the reference when the difference 5 — 0 becomes
negative. The output voltage is an example of a suppressed-carrier
amplitude modulation. This may be shown by writing equation (5.6) in
the equivalent form:
two currents ia and % are alternating at the carrier frequency to. Their
magnitude depends on 6, the rotor angle of the generator, and although
ideally they are in time phase, in practice a small difference in the phase
angles of the impedances of the stator results in a small phase angle 7
between the currents. Thus, let
ia = Ia sin Oit
ib = h sin (a)t — 7) (5.8)
e0 = coMc[Ia cos (a + 30°) cos cot + Ib cos (« — 30°) cos (cot — 7)]
(5.10)
By expanding the cos (cot — 7) term, this becomes
Ia and Ib, the amplitudes of the two stator currents ia and % shown in Fig.
5.6, are periodic functions of 0 (a “negative amplitude” refers to a voltage
180° out of phase with one having a positive amplitude). The salient
rotor construction commonly used in synchro generators may be expected
to make Ia and Ib nonsinusoidal functions of 6, but for the present it is
convenient to ignore this fact and to assume a sinusoidal variation.
Inspection of Fig. 5.6 indicates that, with 6 = 30°, Ia = 0 and, with
6 = —30°, Ib = 0. Furthermore both Ia and Ib are positive as shown,
when 6 = 0°. Hence, if we let Iam and Ibm be the maximum values of Ia
228 CONTROL SYSTEM COMPONENTS [Chap. 5
bm
cos 7 — 1 cos 26
/ am 2 \/S y/Z
cos \6 — a — tarn (5.14)
1 + (ir~ cos
/
7—I i+ sin 26
am
V3
In arriving at this result, we have used the fact that sin (6 + a) may be
written as sin [26 — (6 — a)]. Note that, if the stator circuits are
balanced, Iam = hm and 7 = 0; hence the relation reduces to the ideal
COS (6 — a) =0
(5.15)
or 6 - a = 90°
However, if for any reason the stator circuits are not exactly balanced,
either Iam differs from Ibm or 7 has a nonzero value, or both, so that an
additional angle, which by Eq. (5.14) is a function of 26, must be added in
Eq. (5.15) to reduce the inphase voltage output of the control transformer
to zero. This additional angle represents an error in the synchro system.
The sixth-harmonic error component indicated in Fig. 5.7 can be shown
by a similar argument to be caused by the salient-pole construction of the
generator rotor. This results in a nonsinusoidal variation of the currents
Ia and h with rotor angle that may be expressed in a Fourier series.
Owing to symmetry, only the odd harmonics will be present, and it can
be shown1 that the third, ninth, etc., harmonics cannot exist because of the
balanced three-phase arrangement of the stator. Hence, if we consider
harmonics up to and including the ninth, there will be present, in addition
to the fundamental, only the fifth and seventh harmonics of 6. Thus, if
it is assumed that the stator impedances are all balanced so that Iam = Ibm,
we have
I a — 11 cos (6 -f- 60°) + h cos (56 — 60°) T Ii cos (/ 6 T 60°) ,_
h = I\ cos {6 — 60°) + h cos (56 + 60°) + 17 cos (76 — 60°)
Here Ii, Ib, and 17 are the maximum amplitudes of the fundamental, fifth,
and seventh harmonics, respectively, and are assumed to be the same for
1 See, for instance, Kerchner and Corcoran, “Alternating Current Circuits,” John
Wiley & Sons, Inc., New York, 1938, pp. 268ff. The discussion given there is for a
polyphase circuit, but a similar argument can be used in the single-phase synchro
system to show that the third space harmonic of stator current cannot exist in the
lines. •
Sec. 5.3] SYNCHROS AND RELATED DEVICES 229
Ia and lb- The phase angles for the fifth- and seventh-harmonic terms
can be deduced from inspection of Fig. 5.6 and the facts that, for 6 = 30°,
Ia — 0 and, for 6 = —30°, h = 0. Both I5 and may be negative. If
it is assumed that y in Eq. (5.12) is zero and if Eq. (5.16) is then sub¬
stituted into (5.12), there results, after some reduction, the new relation
for zero inphase output:
1 + Q2
2 {[y cos a — (1 — v2)Q sin a]2 + sin2o-j
[1 - (1 - v2)Q2]2 + 4Q
(5.19)
and the phase angle of the output voltage relative to /30, the phase angle
at standstill, is given by
1 + (1 + v2)Q2\ 1
(3 — /30 = tan-1 + tan-
1 + (1 - v2)Q2J KV cot a — (1 — v2)Q,
(5.20)
Comparison of Figs. 5.8 and 5.9 indicates that the point of minimum
output voltage coincides quite closely with the point at which the phase
shift is 90°, at least for relatively low velocity. Hence the error intro¬
duced into the servo system by the velocity effect is approximately equal
to the value of a for which the output is minimum. This value, desig¬
nated as crm, may be obtained by differentiating Eq. (5.19) with respect to
(t and setting the derivative equal to zero. The result is given by
2 vQ
tan 2dm (5.21)
1 + (1 - v2)Q2
and is plotted in Fig. 5.10 for various values of Q. For the relatively low
velocities usually encountered, 1 — v2 « 1, and the error is small enough
so that tan 2am « 2crTO. Hence for low velocities Eq. (5.21) may be
reduced to
_ 57Q
degrees (5.22)
v 1 + Q2
0 0.1 0.2 0.3 0.4 0.5. 0.6 0.7 0.8 0.9 1.0
v — @/(jj
Fig. 5.10. Variation of <rm as a function of velocity and Q [see Eq. (5.21)].
will be given here, it is possible to show1 that under these conditions the
output voltage is given by
En =
UQ - j) - [cos <7 + jQ(l — v2) cos <r + jv sin a]
1 - (1 - V2)Q2 + 2jQ CO
where co is the line frequency in radians per second, the other symbols are
as defined in connection with Eqs. (5.19) and (5.20), and a is the difference
between the velocities of the generator and control transformer. If a is
1 Ibid. Equation (5.23) is not given in Weiss’s work but may be derived from the
results given there.
234 CONTROL SYSTEM COMPONENTS [Chap. 5
zero, Eq. (5.23) can be reduced to a form equivalent to Eqs. (5.19) and
(5.20). If p = 0, i.e., if the generator rotor does not rotate, Eq. (5.23)
reduces to the simple expression
(5.25)
It is evident from Fig. 5.8 that the separation between maximum and
minimum values of output voltage decreases with velocity. This may
be interpreted as a reduction in gain dEa/da of the synchro system with
velocity. Since we are again primarily concerned with the inphase com¬
ponent of the output, it is instructive to compute the ratio of a small
change of inphas^ output voltage as a function of a small change in a.
Sec. 5.5] synchros and related devices 235
This result can be derived by use of Eqs. (5.19) and (5.20), and a simple
expression results if this is evaluated for a = 0:
This expression is plotted in Fig. 5.12 for Q = 3; note that the change in
gain is not very great until the
velocity becomes greater than one-
third of synchronous speed.
5.5. Two-speed Systems. The
static accuracy of servomechanisms
using synchros can be made much
higher than the static accuracy of
the synchros themselves by the
simple expedient of gearing up the Fig. 5.12. Variation of gain as a function
rotor of the control transformer of speed.
relative to the actual output. The
generator must, of course, be geared up in the same amount relative to the
true input. Then if the static synchro error is e, and if both synchros
make N revolutions for each revolution of the output or input, the error
at the output will be reduced to e/N. This assumes, of course, that the
gears do not introduce an additional error; but if precision gears are used,
a considerable improvement in accuracy is possible. Commonly used
values of N are 36, 31, and 25.
Unfortunately, a system in which the synchros are geared up by a factor
of N will have N — 1 false points of synchronization. Thus, for instance,
assume N = 36. Then for every 10° rotation of the output the control
transformer rotates through 360°. Hence, if the system is in equilibrium
for a given value of the input, it will also be in equilibrium when the out¬
put is displaced from its correct value by 10°, 20°, etc. This difficulty
may be overcome by using a second set of synchros arranged to make one
revolution for each revolution of the output or input. These synchros
are referred to as the one-speed synchros and serve to bring the servo
into approximate alignment. When the error in the one-speed system is
sufficiently small, the one-speed system may be switched off and the
N-speed system allowed to take over to complete the final accurate align¬
ment of the over-all system. A system of this sort is shown schematically
in Fig. 5.13. For the system to work properly it is, of course, necessary
for the stators of the two pairs of synchros to be clamped into their sup¬
porting frames in such a way that the two synchros tend to line up the
output to exactly the same point.
It should be noted that, in all cases where a two-speed synchro trans¬
mission is employed, the one-speed transmission is used only for approxi-
236 CONTROL SYSTEM COMPONENTS [Chap. 5
mate synchronization and is not operative during normal operation of the
servo. Hence the performance characteristics of a servo considered in the
design are always those observed with the high-speed system in control.
A number of circuits are available to perform the required switching
between the high-speed and one-speed systems. One simple circuit
utilizing neon tubes as switches is shown in Fig. 5.14. Its operation
N-speed synchros
depends on the fact that the amplifier has an infinite input impedance and
may be explained as follows: The signal coming from the one-speed con¬
trol transformer is amplified in Vi and V2 and is applied through the trans¬
former T2 to the neon tubes. If the signal is large enough, the neon tubes
fire and provide a low-impedance path for the one-speed signal to the
amplifier input. At the same time the signal from the high-speed control
smaller than the input impedance of the amplifier and the resistance of
the neon-tube circuit. The turns ratio of the high-speed input trans¬
former must be so chosen that the high-speed signal is never sufficiently
large to fire the neon tubes.
For a given small rotation of the output away from synchronism, the
output voltage of the iV-speed synchro is N times as high as that of the
one-speed synchro. Therefore, the loop gain with the high-speed channel
operating is inherently N times as large as with the one-speed system
switched on. An advantage of the circuit described here is that the extra
amplification included in the one-speed channel permits the loop gain to
remain approximately constant as the system is switched from one chan¬
nel to the other. This contributes to smooth operation.
R
Amplitude of
Error
(o) Switching circuit using clipper (b) Wave shape of amplifier input signal
occurs outside this zone, the A-speed synchro will cause the system to
move away from the correct correspondence point at 0 to a false point
such as A. The switching circuit normally has hysteresis; i.e., it takes a
larger voltage to switch the system to the one-speed channel than it does
to switch to the A-speed channel. If the hysteresis is so great that it
takes a voltage greater than e2 to reconnect the one-speed channel, then
the system will simply lock in at the false point A. More probably, how¬
ever, the system switches back to the one-speed channel at a voltage less
than e2, so that the servo oscillates between the voltages e\ and e2.
Neither mode of operation is, of course, desirable.
If es is the voltage output of the one-speed control transformer at the
point at which the switch connects the system to the A-speed channel,
we see from Fig. 5.16 that for proper operation es < ke sin (tt/N) , or approx¬
imately es < keTr/A, where ke is the synchro constant. If this inequality
1 For instance, Ahrendt, “ Servomechanism Practice,” McGraw-Hill Book Com¬
pany, Inc., New York, 1954, pp. 59-64; James, Nichols, and Phillips, “Theory of
Servomechanisms,” Radiation Laboratory Series, vol. 25, McGraw-Hill Book Com¬
pany, Inc., New York, 1947, p. 84.
Sec. 5.5] SYNCHROS AND RELATED DEVICES 239
is only barely satisfied, then, although the system will lock in at the right
place, it may hesitate slightly immediately after the switch has operated,
since the voltage from the V-speed synchro is very small. Hence in prac¬
tice es should be about )^/ce7r /N.
Actually, errors exist in both synchros; in fact it is the existence of
these errors that makes the extra complication of the two-speed system
necessary. A possible range of one-speed and V-speed error is shown as
the dotted lines in Fig. 5.16. Suppose we let ei and be the maximum
error of the one-speed and V-speed synchro systems, respectively. For
identical synchros, ei will be approximately equal to en- Then, by inspec¬
tion of Fig. 5.16 and by use of the sort of reasoning used above, we find
that for proper operation
es < ke (T LI
\N ~ N “
6N
V
(5.27)
or, if ei = €n,
The absolute-value bars are used because ei and may be either positive
or negative. As for the case of zero synchro error, it is desirable to do
more than to barely satisfy this inequality.
Since only the magnitude of es, not its phase, is usually used to actuate
the switching circuit, es cannot be negative. Hence inequality (5.28)
gives us an upper limit on the gear ratio V that may be employed between
the true output and the V-speed selsyn. Specifically
N + 1 < A (5.29)
Nil
Thus for an error of, say, 0.5° (this is rather larger than usually observed,
but serves to illustrate the principle), V would have to be less than 360.
Actually V seldom exceeds 100 in practice, since the switch cannot be
expected to operate always at exactly the same voltage and a safety factor
is required. Furthermore es cannot actually be zero but must have a
finite value. Very large values of V are not justified if the gearing error
becomes comparable to the synchro error or if other components of the
servo limit the maximum accuracy obtainable. Another factor limiting
maximum gear ratios is the fact that the moment of inertia of the control-
transformer rotor is reflected by the square of the gear ratio to the motor
shaft (see Chap. 8), and with small motors and large ratios this may result
in a considerable increase of the effective moment of inertia seen by the
motor. Finally, the top design speed of the synchros may limit the maxi¬
mum gear ratio that should be employed.
One further problem of two-speed synchro operation must be discussed
240 CONTROL SYSTEM COMPONENTS [Chap. 5
here: the problem of the 180° ambiguity found in two-speed systems hav¬
ing an even gear ratio. In Fig. 5.17 is shown the variation of output
voltage of the two synchros near 0° and 180° for systems having odd and
even gear ratios. Considering the odd gear ratio first, we see that the
slope of the curves at 180° for both synchros is negative; hence 180° is not
a point of stable equilibrium for the servo system. If the servo happened
to be energized, with the error exactly 180°, any small disturbance would
cause the A-speed voltage to increase and make the system drive away
from the 180° position. As the voltage increases, the one-speed system
takes over and brings the servo to the proper position at 0° error. For
the even ratio, however, the slope of the A-speed curve is positive both at
0° and at 180°. Therefore, since at 180° the one-speed system is switched
off, the 180° point is a point of stable equilibrium at which the servo may
come to rest.
It is often desirable to use even gear ratios, particularly the ratio 36,’
since this permits dials calibrated for 360° per revolution and 10° per
revolution to be attached directly to the one-speed and 36-speed synchro
shafts, respectively. To overcome the difficulty discussed above, a
Total one-speed
One-speed voltoge with
stick-off voltage signal after
stator shift
voltage goes to zero at the same point as the iV-speed voltage. The effect
is shown in Fig. 5.186. Note that with the proper amount of stick-off
voltage the signal of both synchros goes to zero near 180° with a negative
slope.
It should be noted that the voltage waves of Fig. 5.18 represent the
amplitude of a-c signals, with negative amplitude referring to a signal
180° out of phase with one having positive amplitude. Hence the con¬
stant stick-off voltage must be a constant a-c voltage, in phase with the
synchro output. This voltage is easily added to the one-speed synchro
output by means of a small transformer in a circuit, such as the one shown
in Fig. 5.19.
A-c
supply
Ia — Im cos (6 + 60°)
(5.30)
h = Im cos (6 — 60°)
9d9 CONTROL SYSTEM COMPONENTS [Chap. 5
where Im is the maximum value of Ia and Ib. The discussion is also sim¬
plified by assuming that the differential generator has unity turns ratio
(this is standard practice) and that the coefficient of coupling between
stator and rotor circuits is unity. Under these conditions, standard
transformer theory shows1 that the ampere-turns acting along any axis
in the differential-generator stator must be equal and opposite to the
ampere-turns acting along the same axis in the rotor. It is convenient
here to choose the axis passing through the leg 0'2' of the rotor (see Fig.
5.20). The ampere-turns contributed by the rotor along this axis are
Control
transformer
angle between the axis under consideration and the axis of the winding
contributing the mmf. Thus, the angle between the reference axis
and the leg l'O', carrying current I'a + I'b, is 60°, and the contribution
to the mmf along the reference axis must be as given by the second term
in (5.31). The angles between the three stator windings and the axis
0'2' are (60° — </>), ( — 0), and (120° — 0) for the legs 10, 02, and 03,
respectively. The rotor angle 0 is measured as shown in Fig. 5.20. Note
that the direction in which the current is assumed to flow in the winding
is important. The stator ampere-turns contributed to axis 0'2' are
N(Ia + Ib) cos (60° — 0) + NIa cos 0 + NIb cos (120° — 0) (5.32)
In a similar fashion, and by using leg 0'3' as the reference axis, we find
We see that under these simplifying assumptions the form of /' and I'b is
the same as that of Ia and Ib, except that the angle 6 — <j> has been sub¬
stituted for the angle 6. Hence, the effect of the differential generator is
to change the control-transformer angle for zero output voltage to
8 = 6 — <f> rather than 8 = 6, the value
obtained when the differential generator
is not present. Thus the shaft angle of 9
the differential generator is added (or
<p 8 Servo
subtracted) from the angle of the input.
This may be represented schematically ^IG' °’21' An eci^RaJfnt^ block
. diagram showing the effect ot the
as m Fig. 5.21. It is, of course, possible dilierential generator.
to use more than one differential genera¬
tor when more than one additional input must be inserted into the system.
In practice, differential generators are subject to the same sort of errors
as other synchros.1 Also, the coefficient of coupling between stator and
rotor circuits is less than unity, and therefore the differential generator
requires a fairly large magnetizing current, which must be supplied by
the synchro generator. This current produces heating in the generator,
and it places a limit on the number of differential generators that can be
excited by a synchro generator of a given size. The magnetizing current
may be reduced quite considerably, at least as far as the synchro generator
is concerned, by the use of a synchro capacitor. This is a set of three
capacitors, usually connected in delta, and “potted” in a single can.
They are ordinarily connected across the three lines connecting the
synchro generator and differential generator. Synchro capacitors are also
sometimes used in simple generator-control-transformer systems, particu¬
larly when one generator supplies several control transformers. Since
the impedance of the control transformer is usually quite high, a smaller
capacitor is required in this application.
5.7. Synchro Repeaters. A synchro repeater (also called synchro
motor) is externally and in electrical characteristics identical to a syn¬
chro generator. The chief constructional difference is that the rotor of a
synchro repeater usually has a vibration damper (see Chap. 8) on it. In
typical systems the repeater is connected to the generator as shown in
Fig. 5.22. The operation may be explained qualitatively by thinking of
the generator as transmitting a magnetic-field pattern to the stator of the
■Primory
■■
winding
Output
■III
-Secondary-
windings
Fig. 5.25. A schematic of an E pick- Fig. 5.26. Schematic of a variable-
off designed for rotary motion. reluctance transformer.
PROBLEMS
5.1. The inductive reactance of each leg of the stator winding of a type 1 CT con¬
trol transformer is 1,000 ohms and that of the type 1 G generator used to excite it is
100 ohms. Nominally the ratio of resistance to reactance in both machines is 1:4,
but as a result of a bad connection in one stator lead, the resistance ratio in this one
leg is increased to 1:3.9. Find the maximum quadrature voltage from the control
transformer if the maximum inphase output voltage is 57 volts rms.
6.2. Determine the maximum second-harmonic error in a synchro transmission in
which the ratio of Iam/hm = 0.99. Assume that all other errors, including the quad¬
rature error, are zero.
6.3. Find the maximum value of sixth-harmonic error in a synchro transmission in
which the fifth space harmonic is 20 per cent of the fundamental and the seventh
10 per cent of the fundamental. Assume all other errors negligible.
5.4. A servo system using a two-speed synchro transmission is supposed to have
an effective velocity-error constant of at least 1,000. The ratio of inductive react¬
ance to resistance in the control transformers is 4. Assuming that the servo gain
constant cannot exceed 5,000, find the maximum gear ratio permissible between the
output shaft and the high-speed synchro.
5.5. A servo system uses a two-speed synchro transmission with a 36:1 ratio
between high- and low-speed synchros. The error in both synchros may be as large
as 0.3°; the synchro constant for both is 1 volt /deg. The switching circuit used to
switch the system between the high- and low-speed synchros has the form shown in
Fig. 5.14. The neon tubes fire when an rms voltage of 60 volts is applied to them;
they go out when the voltage is 40 volts. Assume that the transformers T\ and T%
have unity turns ratio; find the extreme values of the combined amplifier gain and
transformer T2 turns ratio that will permit reliable switching.
5.6. What is the value of the stick-off voltage required to prevent ambiguity in a
system using a two-speed synchro system with 36:1 gear ratio? The synchro con¬
stant for both synchros is 1 volt/deg. Neglect synchro error.
design with plate load Rl and cathode-biasing resistor RK• The feature
that makes the circuit act as a phase-sensitive demodulator is the a-c
plate-supply voltage. This supply is referred to as the reference volt¬
age, and it must be of the same frequency as the input signal. As a
result of the a-c supply, the amplifier is cut off every other half cycle,
operating only during the half cycle in which the plate voltage is positive.
If the phase of the input signal is such that during the on period of the
amplifier the upper grid in Fig. 6.1 is positive and the lower one negative,
more current flows in the upper triode than in the lower, and hence the
249
250 CONTROL SYSTEM COMPONENTS [Chap. 6
plate of the upper triode is more negative than that of the lower. If the
phase of the input is reversed, the polarity of the output reverses. Typi¬
cal output waveshapes obtained from the circuit for various phase rela¬
tions between input and reference are shown in Fig. 6.2. In drawing
these figures, it is assumed that the gain of the tubes is independent of the
magnitude of the plate voltage; this is a good assumption during most of
the time that the voltage is positive. Figure 6.2c shows the effect of an
input signal 90° out of phase with the reference; note that the average, or
d-c, output is zero. The discriminator is therefore seen to reject quad¬
rature signals. In general, if the input is less than 90° out of phase with
the reference, the d-c component of the output is reduced from the value
it would have if the signal and reference were exactly in phase. A typical
output waveform is shown in Fig. 6.2d.
In all cases shown, the output contains a large a-c ripple component.
This ripple is normally highly objectionable because it results in satura¬
tion, heating, and general malfunctioning of amplifiers and power devices
following the discriminator. Hence low-pass filters are usually used in
the discriminator output. For most applications these are simple RC
networks of the type described in Chap. 1.
The output waveforms shown in Fig. 6.2a and 6.2b are those obtained
from a standard half-wave rectifier. Hence the discriminator circuit
shown in Fig. 6.1 is referred to as a half-wave discriminator. Some of
its other features are that it provides gain between input and output and
that it delivers a push-pull output signal. If the two halves of the circuit
are exactly identical, zero input results in both zero d-c and zero ripple
output. In practice, however, perfect balance of the circuit is difficult to
achieve, and there is usually some output for zero input. Since the cir¬
cuit unbalance may vary as a result of temperature and supply-voltage
changes, the output produced by the unbalance may also vary; i.e., the
circuit may drifts This is not unreasonable since the circuit is in part a
Sec. 6.2] DEMODULATORS AND MODULATORS 251
where co is the carrier frequency and (3 is the phase shift that might exist
between the carrier and the reference. The output of the discriminator is
the product of the input and the demodulating gain function of Eq. (6.1):
v0(t) = GF(t) sin (cot — (3) ^ + - ( sin cot + i sin Scot + \ sin Scot +
2 7T \ o 5
1 . 1 . , 2/1 0 , . 1 . ,
v0{t) = GF(t) cos /3 —\- - sm cot-l ^ cos 2cot + «—= cos 4oot
7T 2 7T \o 0-5
The useful part of the output is provided by the first term in the first
bracket.
All other terms represent ripple. We note that the original modulation
F(t) has emerged from the discriminator multiplied only by the constant
factor (G/tt) cos (3, without any phase lags or amplitude changes depending
on frequency. We conclude from this that the processes of modulation
and demodulation do not inherently introduce any phase lag (or lead) into
the signal channel and that the phase lag normally found is entirely due
to the ripple filters. As mentioned previously, these are almost always
required to eliminate the objectionable effects of the ripple on the circuits
following the discriminator.
Equation (6.5) shows that the gain between input and useful output is
proportional to the cosine of the phase shift between the carrier and the
reference. The desirability of keeping this phase angle small if maximum
gain is to be secured is therefore evident. However it is also clear that
phase shifts of the order of less than 10° are not of any great consequence.
Phase shift also has some effect on the magnitude of the ripple. This can
be appreciated qualitatively by inspection of Eq. (6.4), since the second
bracket, which contributes a large ripple component, is proportional to
the sine of the phase angle and therefore vanishes if the carrier and refer¬
ence are in phase. A somewhat more quantitative idea of the effect of
phase angle on output ripple is obtained by computing the magnitude of
the harmonics from Eq. (6.4) as a function of |8. The results of such a
computation are given in Table 6.1 for the case of GF(t) — 1. The table
Harmonic Magnitude
0 0.318 cos j8
1 0.5
2 0.21 Vl + 3 sin2 /3
4 0.043 Vl +15 sin2 0
6 0.018 Vl + 35 sin2 /3
8 0.01 Vl +63 sin2 /3
Sec. 6.2] DEMODULATORS AND MODULATORS 253
shows clearly that the ripple voltage increases with /?, but since the largest
ripple-frequency component, the fundamental, is independent of /3, one
would not expect the rms value of the ripple to increase very much as /3
changes from 0 to 90°. It is instructive in this connection to compute the
rms value of the ripple for these two extreme values of (3. Using only the
harmonics listed in Table 6.1, we find approximately that for (3 = 0° the
rms value of the ripple is 0.544, while for (3 = 90° the ripple increases to
0.70. Thus it is clear that, although phase shift does increase the ripple
somewhat, the effect is not of any great importance in a half-wave
discriminator.
In order to find the maximum frequency of the input modulation F(t)
that can be handled without ambiguity by a half-wave discriminator, we
let
F(t) = V sin ojst (6.6)
v0{t) = GV cos jS - sin cost + x/\ cos (co — oos)t — Iq cos (co + cos)t —
7T
The term yielding the useful part of the output is still the first term of the
first bracket, GV (cos (3)(1/t) sin cost, with all else constituting ripple.
Thus the lowest harmonic of the ripple is shifted downward by the signal
frequency and becomes co — cos. It is clear that, when
CO — cos = cos
or cos = )q^co (6.8)
the signal frequency and the frequency of the lowest harmonic of the rip¬
ple are the same, and it is no longer possible to distinguish between signal
and ripple. Hence we have the important and quite universal result that
the absolute upper frequency limit of a half-wave discriminator is equal to
one-half the carrier frequency. Practical reasons such as the difficulty of
constructing a ripple filter with a very sharp cutoff characteristic usually
restrict the maximum permissible signal frequencies to values considerably
below this absolute limit. Thus a commonly observed rule of thumb
applied to the design of feedback systems is that the loop-gain crossover
frequency should be no more than one-tenth of the carrier frequency if a
half-wave discriminator is used.
Equation (6.7) contains additional information that is useful in the
design of ripple filters. Usually these are simple, low-pass RC networks
254 CONTROL SYSTEM COMPONENTS [Chap. 6
sider first a typical circuit, operating on the same principle as the one
shown in Fig. 6.1, and discuss its operation qualitatively.
The circuit of a triode full-wave discriminator is shown in Fig. 6.4. The
dots shown on the transformers are the conventional polarity markings;
hence, when the upper two triodes conduct, the lower two are disabled by
negative plate supply, and vice versa. Suppose now that the phase of the
input is such that the grid of the uppermost tube is positive when the
upper two triodes conduct. Then the output signal e0 is positive in the
direction shown by the arrow. During the next half cycle the lower two
triodes conduct, and the connection of the input transformer is such that
the lowermost grid is now negative. Hence the output signal is again
positive in the direction of the arrow. Thus for each cycle of the refer¬
ence we get two pulses of the output, or a full-wave signal. Had the
phase of the input relative to the reference been inverted, so that the
uppermost grid was negative during the conduction period of the upper
two tubes, we would have found the output to consist of negative pulses.
If the phase of the input were such that during the conducting period of
the upper triodes the uppermost grid were first positive and then negative,
the output would consist of a pulse first positive and then negative. By
this sort of reasoning one can obtain the output waveshapes shown in
Fig. 6.5. We note again that a signal 90° out of phase with the reference
will result in zero output, and if the phase shift between signal and refer¬
ence is somewhere between 0 and 90°, a reduced d-c output results.
In analyzing this circuit we make the same assumption we made in the
analysis of the half-wave discriminator, namely, that the gain of the cir¬
cuit is a constant during the conduction period of either set of triodes.
Also, as has already been noted, the operation of the circuit is such that,
256 CONTROL SYSTEM COMPONENTS [Chap. 6
if during one half cycle of the reference the output is positive for positive
input, then for the next half cycle the output is positive for negative input.
Hence, the circuit can be assumed to
multiply the input by the square
wave shown in Fig. 6.6; this differs
from the square wave considered in
connection with the half-wave dis¬
criminator primarily in that its d-c
level is zero. The magnitude is arbi¬
Fig. 6.6. Approximate demodulating
trarily taken as G, the gain of the am¬
function of full-wave discriminator.
plifier in one of the switching modes.
The discriminator gain function may be expanded in the Fourier series:
then the output given by the product of (6.2) and (6.9) may be expanded
into the form
2 4/1 0 . , 1
v0(t) = GF(t) cos (3 -I U COS 2bit + -—- COS 4co£
7T 7T \6 6 ‘ 5
1
+ -—= cos 6cot + • •
5 • 7
— (sin (3) - sin 2cd + sin 4cot + -=^-= sin 6cot + (6.10)
7r \3 3-5 5-7
The useful part *of the output is F(t) multiplied by the time-invariant part
Sec. 6.3] demodulators and modulators 257
of the series,
with all other terms contributing only to the ripple. Here again it may be
seen that the input modulation F(t) emerges unchanged by the processes
of modulation and demodulation. Also the effect of (3 on the gain is
identical with that found already for the half-wave discriminator. In
fact, comparison of Eqs. (6.4) and (6.10) shows that, except for a factor
of 2, which can be absorbed into the gain factor G, the only difference
between the outputs of the full-wave and half-wave circuits is that in the
former, the fundamental frequency component of the ripple is absent.
Hence the table of harmonic magnitudes constructed for the half-wave
discriminator (Table 6.1) applies also to the full-wave discriminator if we
set GF(t) = Yl and let the first harmonic be zero.
The table may again be used to find the approximate rms value of the
ripple voltage and to estimate the effect of d on it. Using only the har¬
monics up to and including the eighth, we find that for (3 = 0° the rms
value of the ripple is 0.216, while for /? = 90° it goes up to 0.473. (Note
that these values are computed for GF{t) = y in order to effect a direct
comparison with the half-wave circuit.) It is clear that for d = 0 the
ratio of rms ripple voltage to useful output is less than half as large in the
full-wave as it is in the half-wave circuit. This is due to the absence of
the fundamental, which is responsible for the major share of ripple voltage
in the half-wave discriminator. It is also apparent that the effect of
phase shift between signal carrier and reference is more serious in increas¬
ing the ripple, and it is therefore somewhat more important here to keep
this phase angle low.
We note that the amount of filtering required to reduce the ripple out¬
put following a full-wave discriminator to a permissible amount is much
less than that needed to produce an equal attenuation of the ripple pro¬
duced by a half-wave circuit. This is due both to the smaller rms value
of the ripple and to the fact that the lowest ripple harmonic is the second
harmonic rather than the fundamental of the carrier frequency. Specifi¬
cally, let us assume that a double-section RC filter, producing an attenu¬
ation proportional to the square of the frequency, is employed as a filter.
Then, for the same reduction in ripple relative to d-c output, the filter
used with the full-wave circuit may have about three times the passband
(as defined by the reciprocal of the RC time constant of the two filter sec¬
tions) of the filter used with the half-wave circuit. The superiority of the
full-wave discriminator in causing reduced attenuation and phase lag of
desired signal information is, therefore, evident.
In order to find the maximum frequency of the input modulation F(t)
258 CONTROL SYSTEM COMPONENTS [CHAP. 6
2.2 2
v0(t) = GV cos (3 - sin oost + — sin (2co — aos)t — — sin (2oo + a>s)£ +
7T 07T 07T
4
— (sin (3) - cos (2co — C0S)£ — - COS (2oj + Ws)£ + • • • (6.13)
7T o o
As before, the first term of the series, GV (cos (3)(2/ir) sin oost, represents
the useful signal, with all other factors being ripple, and it can be seen
that the lowest harmonic of the ripple is 2co — oos. The upper limit on
cos is reached when this lowest ripple frequency coincides with the signal
frequency, i.e., when
oos = 2co — 00s
or oo s — oo (6.14)
Thus the absolute upper limit is equal to the carrier frequency, a limit
twice that of the half-wave discriminator. As was mentioned in connec¬
tion with the half-wave circuit, the actual useful upper limit is normally
considerably less than the theoretical maximum because of practical
difficulties of filter design. However, if these difficulties are assumed to
be identical in the two cases, the fact that the practical upper frequency
limit of the full-wave discriminator is twice as high as that of the half¬
wave discriminator still holds true.
It should be pointed out that the result concerning the maximum theo¬
retical frequency usable with a full-wave discriminator is not universally
true but depends on the fact that the input signal was assumed to have a
sinusoidal carrier. If the carrier were an accurate square wave, identical
with the one assumed for the discriminator demodulating function (Fig.
6.6), the processes of modulation and demodulation would have subjected
the signal to two successive multiplications by a square wave. This, at
least theoretically, would be equivalent to multiplication by a constant,
so that there would be no upper frequency limit and, incidentally, no
ripple in the output. While this result is primarily of theoretical interest,
owing to the difficulty of generating the required accurate square waves,
it does point out the desirability of square-wave modulation and demodu¬
lation in general.
As discussed in connection with the half-wave discriminator, it is possi¬
ble to take advantage of the fact that the ripple is concentrated in rela¬
tively narrow bands around the harmonics of the carrier to design an
improved type of ripple filter incorporating band-rejection filter sections.
The use of this type of filter is of considerably greater practical importance
Sec. 6.4] DEMODULATORS AND MODULATORS 259
with the full-wave than with the half-wave discriminator. For the half¬
wave discriminator the extra complication involved hardly seems justi¬
fied, since the discriminator itself is not optimum and there is little point
in trying to improve an inferior circuit when a better circuit is available.
This argument does not apply, however, to the full-wave discriminator,
for it represents the best that can be achieved without increasing the cir¬
cuit complexity manyfold. Since the lowest harmonic of the ripple is the
second and since the next one (the fourth) is almost five times smaller (see
Table 6.1), a relatively large improvement in filtering efficiency is obtained
by use of a single rejection filter to remove the band of frequencies around
the second harmonic. All other harmonics can be removed easily by a
standard RC low-pass filter, and, as mentioned in connection with the
half-wave discriminator, this part of the filter attenuates only the higher
frequencies and has therefore a minimal effect on the signal band. Here
again it must be kept in mind that the rejection filter must attenuate
effectively all the frequencies in a band that is centered around the second
harmonic of the carrier and has a width equal to twice the maximum fre¬
quency expected in the signal envelope.
The improved performance possible with the full-wave discriminator is
obtained at the expense of considerable additional circuit complexity.
This is easily appreciated by comparing Figs. 6.4 and 6.1. In order to
get accurate full-wave output, the two halves of the circuit of Fig. 6.4
must be carefully balanced; i.e., the gain must be the same no matter
which triode pair is conducting. If this is not the case, then in Fig. 6.5a,
for instance, every other pulse will be larger than the two adjacent pulses
(see Fig. 6.10a), and the output will then contain a component at the
fundamental of the carrier frequency. <? Ref 9
shown in Fig. 6.1. As in the other two circuits discussed, the important
feature making this circuit a discriminator is the a-c plate supply, which
disables the circuit every other half cycle. The output is a half-wave
signal similar to that shown in Fig. 6.2, but since this is a single-ended
circuit, the output voltage cannot reverse sign and is always negative
during periods of tube conduction. The magnitude of the d-c component
of the output varies continuously from a large negative value for a large,
inphase signal input to smaller negative values as the input decreases to
zero, reverses in phase, and increases again in the opposite direction.
Hence, if it is necessary that the d-c output be zero for zero input, some
sort of d-c potential-shifting device of the type described in Sec. 2.17 must
be employed in the output. The output contains a large ripple compo¬
nent when the input signal is zero, and on the average the ripple is con¬
siderably larger in this circuit than in the push-pull circuit considered in
Sec. 6.1. Thus a much larger amount of filtering is required to reduce the
a-c component of the output to a permissible value. A further disadvan¬
tage of a single-ended circuit is that it has no drift compensation.
The analysis of this circuit proceeds in most essentials like that of the
push-pull half-wave discriminator given in Sec. 6.2. The lowest ripple
harmonic is the fundamental of the carrier, and therefore the upper limit
on the envelope frequency is one-half the carrier frequency. The circuit
operation cannot, however, be explained in terms of multiplication of the
input signal by a square wave, since this would not account for the half¬
wave output obtained for zero input. A more reasonable assumption
here is that the sum of the input and the sinusoidal reference is multiplied
by the square wave. This assumption is justified by noting that the
plate current of a triode may be expressed by
Zp —l-
Tp
When the analysis is carried out on this basis, it will be found that both
the quiescent d-c value and the magnitude of the ripple are determined
primarily by the magnitude of the added reference voltage. This voltage
should, therefore, be kept as small as possible, but on the other hand it
must be large enough so that the sum of input and reference is always
positive during the on period of the circuit. It is clear in summary that,
although the circuit appears to be simpler than the one shown in Fig. 6.1,
the extra filtering and d-c potential-shifting circuits that must be used
with it largely nullify this apparent simplicity.
Both the circuits shown in Figs. 6.1 and 6.4 can be converted to cathode-
follower circuits if it is not required that the signal be amplified and if the
low output impedance and low drift afforded by cathode followers is
desirable. In Fig. 6.8 is shown a half-wave cathode-follower circuit cor-
Sec. 6.5] DEMODULATORS AND MODULATORS 261
responding to the circuit of Fig. 6.1. Again its operation depends on the
fact that the tubes are disabled by the reference signal every other half
cycle. It can, therefore, be analyzed exactly as the circuit of Fig. 6.1
was, and its performance is identical except that the gain is slightly less
than unity. The reason for grounding the reference transformer at the
midtap of the winding is to provide a negative voltage for the cathode
return during the operating half cycle. The placement of the tap need
not be accurate. The full-wave counterpart of this circuit can be built
up just as the full-wave circuit of Fig. 6.4 was built up from the half-wave
circuit of Fig. 6.1. Construction of a detailed circuit is left as an exercise
for the reader.
6.5. Diode Discriminators. All the triode discriminator circuits dis¬
cussed thus far drift to a greater or lesser degree because they combine the
functions of discriminator and d-c amplifier. For this reason diode dis¬
criminators are often preferred, particularly in systems having sufficient
Input
On the other hand, if the signal transformer is not at fault, then it should
be possible by the adjustment of any one of the four resistors to produce a
true full-wave-rectified output. However, it may then be found that there
is some d-c output for zero input. If this is undesirable, then another
adjustment must be provided. Assuming that the transformer taps
cannot be moved, one of the resistors in circuit bad and one in circuit bed
of Fig. 6.9 must be made adjustable to permit both zero d-c output for no
input and a balanced full-wave output to be achieved.
The waveform shown in Fig. 6.105 results from the addition of a full-
wave-rectified voltage and a sine wave at fundamental frequency but 90°
out of phase with the reference. This effect cannot be produced by
resistance unbalance but may be caused by unequal capacitances between
the two halves of the reference winding and ground. This in turn may be
caused by a reference transformer that does not have an electrostatic
AYV
(a) id) (c)
Fig. 6.10. Output waveforms resulting from circuit unbalance.
shield between its primary and secondary windings. The effect may be
simulated by a small capacitor connected from either point 5 or point d
to ground (Fig. 6.9). The remedy is to use a transformer with a Faraday
shield or to add compensating capacitors between points 5 or d and ground.
In Fig. 6.10c is shown the result when both the effects described occur
simultaneously. Both resistive and capacitive adjustments must then
be made to convert the output to the balanced full-wave form.
In addition to the balancing problem, another point of interest in
this circuit is the maximum input signal for which the circuit will still
function properly. To determine this, assume that the circuit is properly
balanced and that the lower two diodes (Fig. 6.9) are conducting. The
voltage across either one of the upper diodes can then be found by
KirchhofFs laws. Let e0 be the output voltage and ea, eb, etc., the voltage
of points a, 5, etc., all with respect to ground, and let eR be the reference
voltage across one half of the reference-transformer secondary [that is,
eR — A(eb — ed)]', then the voltage across the rectifier in arm ab is given
by ea — (e0 + eR). However, during the period when the lower set of
rectifiers conducts, eQ = ec = — ea; hence the rectifier voltage is 2ea — eR.
Similarly the rectifier voltage in leg ad is given by 2ea + eR. The refer¬
ence voltage eR is positive, since the lower rectifiers conduct. Hence, if
ea is positive, the voltage across the rectifier in arm ab becomes positive
when ea exceeds eR/2, and the rectifier conducts. For negative ea the
264 CONTROL SYSTEM COMPONENTS [CHAP. 6
9 Ref
UmJ
and functions the same way. During the half cycle when the diodes con¬
duct, the output and input are connected together, and during the next
half cycle, the output is zero. A slightly different version of this circuit
is shown in Fig. 6.115. Here R2 Ri, so that, during the periods when
the diodes conduct, the signal is essentially shorted to ground. Both of
these circuits when properly adjusted will have zero d-c and ripple output
when the input is zero. This is not true in the circuit shown in Fig. 6.11c,
whose output is always a half-wave-rectified wave. The amplitude of the
output signal is increased by an input that is in phase with the reference
and decreased by a signal that is out of phase with the reference. Thus,
although this circuit is exceedingly simple, it requires a d-c potential-
shifting circuit to make the average value of the output wave zero when
the input is zero, and it usually requires more extensive filtering. Placing
two circuits of the form of Fig. 6.11c back to back results in the push-pull
circuit shown in Fig. 6.lid. The reader will recognize that the circuits of
Fig. 6.11c and d are the diode counterparts of the triode circuits of Fig.
6.7 and Fig. 6.1, -respectively.
Sec. 6.6] demodulators and modulators 265
its resonant frequency near the carrier frequency for which the chopper
is designed. However, the resonant frequency of the reed should not be
exactly equal to the carrier frequency since very large changes in phase of
the oscillation of the reed would result from relatively small frequency
changes. A compromise is therefore necessary; in a typical design the
resonant frequency of the reed is 80 cps for a chopper designed to operate
at 60 cps.1 There is, of course, still some phase shift between the reed
and the reference signal, and in order to obtain maximum output, it is
necessary to shift the phase of either the voltage applied to the vibrator
coil or the signal voltage. A disadvantage of a vibrator constructed in
this way is that it can be used for only one carrier frequency, and, in
general, vibrators are not very practical for frequencies in the upper audio
range. On the other hand, vibrators possess a number of important
advantages over electronic demodulators: they are practically free of drift,
their noise level is very low, and the balancing problem discussed in con¬
nection with the ring demodulator is absent. They are, furthermore,
very compact, simple to use in a circuit, and, despite the fact that there
is a moving part and electric contacts, their life is quite acceptable.
Hence they are replacing electronic demodulators in many applications.
6.7. The Clamping Discriminator. A discriminator operating on a
principle not found in the discriminators described thus far is shown in the
? Ref ?
circuit of Fig. 6.14. In this arrangement the RC circuit in series with the
grid of each of the triodes is a grid-leak biasing circuit. It keeps the
triodes cut off except during a short time when the reference goes through
its positive peak. At this time the grids are driven positive, the capaci¬
tors Ci are recharged, and the output capacitor C2 is connected to the
input and charges to the value of the input signal. Two triodes are
1 Ibid., p. 108. The chopper referred to in this reference is the Synchronous Con¬
verter No. 75829-1 made by Brown Instrument Company, Philadelphia, Pa.
Sec. 6.7] DEMODULATORS AND MODULATORS 267
needed in order that the charging current may flow in either direction. If
the amplitude of the reference signal is very large compared to the cutoff
bias voltage of the triodes, the period of tube conduction is very short,
and it may be assumed that the charge placed on the capacitor C2 repre¬
sents the value of the input signal at the instant that the reference passes
through its peak; i.e., the circuit samples the voltage existing at the input
at this instant. The load impedance connected to the circuit is assumed
to be infinite. Hence the capacitor holds, or ^clamps,” the signal at this
level until the application of the next positive reference-voltage peak. A
typical form of the output for a signal in which the carrier is in phase with
the reference is shown in Fig. 6.15. Note that a constant-amplitude
input ideally results in a d-c output, while for a varying input the output
has a steplike appearance. The ripple content of the output is much less
than for more conventional discriminator circuits: To appreciate this
fully, the reader should keep in mind that the circuit described here is a
half-wave circuit; i.e., the signal is sampled only once per cycle.
In the circuit design, the RC time constant of the grid-leak circuits
should be of an order of magnitude larger than the period of the reference
signal in order that the capacitor shall not discharge significantly during
one period. The grid-leak resistor must be large enough to prevent
excessive grid current from being drawn at any time. The output
capacitor C2 should have such a value that the RC time constant of this
capacitor in combination with the resistance of the conducting tubes will
be very short compared to one period of the reference signal so that the
capacitor will charge up to the full signal level during the short time that
the tubes conduct. On the other hand the time constant with the tube
not conducting should be very much greater than a period of the reference
to ensure that the signal is properly held during the time between sam¬
pling instants. Hence the load impedance should be several orders of
magnitude greater than the plate resistance of the triodes, and a cathode
follower is preferably used here. It is clear that the improvement in
performance is obtained at the expense of a somewhat more complicated
circuit. This is particularly true if, as is commonly done, the reference is
268 CONTROL SYSTEM COMPONENTS [CHAP. 6
second.1 Hence an increase in the sampling rate makes the circuit useful
for higher-frequency signals, and it also reduces the amount of ripple at a
given input frequency. One way to double the sampling rate is to use
the full-wave equivalent of the clamping discriminator, shown in Fig.
6.16. The connection of the reference transformer is such that when the
upper pair of triodes conduct, the lower pair is cut off, and vice versa.
The operation of the triodes is the same as already described. If we sup¬
pose the upper end of the input transformer to be positive when the upper
pair of triodes conduct, a positive charge is placed on the output capacitor;
during the next half cycle the lower half of the transformer is positive and
the lower pair of tubes conduct, placing another positive charge on the
capacitor. Thus two samples per reference cycle are obtained.
By modifying the reference signal, the circuit of Fig. 6.16 can be
1 Truxal, “Automatic Feedback Control System Synthesis,” McGraw-Hill Book
Company, Inc., Netv York, 1955, pp. 500ff.
Sec. 6.7] DEMODULATORS AND MODULATORS 269
arranged to yield four samples per reference cycle. This is done by apply¬
ing a reference pulse to the tubes at the instants that the carrier, assumed
to be in phase with the reference, passes through the 45° and 135° points
(see Fig. 6.17). Obviously a fairly elaborate shaping circuit would be
required to obtain the required reference signal, and the additional com¬
plication would probably be worth while only if several discriminators
could all be operated from one reference source in a particular system.
A very simple method for deter¬
mining the frequency response of a
clamping discriminator is given by
Truxal,1 who shows that the transfer
function may be expressed as
1 — e -Ts
H(s) = (6.16)
Ts
The absolute values of the gain and the phase shift are sketched in
Fig. 6.18. In practice, only the portion below co/2 is of interest, since co/2
is the limit of unambiguous transmission. The gain at this limit is 2/tt,
and the phase lag ir/2 radians. Note that the sampling frequency co is
equal to the reference frequency for a half-wave clamping discriminator
and to twice the reference frequency for a full-wave discriminator. Thus
the results for the upper limit of transmission obtained here are identical
with those obtained with the simpler discriminator circuits discussed at
the beginning of this chapter.
In order to find the magnitude and frequency of the ripple harmonics
in the output, a Fourier analysis must be made on a typical output, such
as that obtained, for instance, by an input with a sinusoidal modulation.
1 Ibid., p. 507. Note that Eq. (6.16) is Truxal’s expression multiplied by 1/T.
Thisois done to make the gain unity for zero frequency.
270 CONTROL SYSTEM COMPONENTS [Chap. 6
Such a Fourier analysis can be made most simply by first finding the
Laplace transform of the output and determining the harmonics from this
by partial-fraction expansion.
The Laplace transform of the output signal obtained for a sinusoidal
input modulation of frequency cos is derived by Gardner and Barnes1 in
connection with their discussion of the solution of difference equations.
(sin cosT)(esT — 1)
F(s) (6.19)
s(e2sT — 2esT cos wsT + 1)
The poles of this function are found by solving for the zeros of the denomi¬
nator; they occur for s = 0 and s = j(27rn/T ± cos), n = 0, ± 1, ±2, . . . .
Hence F(s) may be written in the equivalent form
CO
F{s) = - +
s X
71 = — oo
s
kn
j[(2Tvn/T) + C0S]
k'n
+ s + j[(2Trn/T) — cos]
(6.20)
Since 2t/T = co, the sampling frequency, it may be seen from Eq.
(6.20) that the harmonic frequencies are ruo ± ccs, a result completely
analogous to that obtained previously for the simple half-wave discrimi¬
nator [Eq. (6.7)]. The kn may be evaluated by standard techniques
which, although fairly cumbersome, offer no difficulties in principle. The
result of Eq. (6.20) may then be put into the form
/COS <J0ST — l\ / s
\ 2Ten + o)sT ) \s2 + [(27m/T) + cos]2
1Gardner and Barnes, “ Transients in Linear Systems,” John Wiley & Sons, Inc.,
New York, 1947, p. 299.
2 Truxal, op. cit., pp. 507-517.
Sec. 6.8] DEMODULATORS AND MODULATORS 271
The second fractions of the two products are seen to be the transforms,
respectively, of sin [(27m/T) + ws]t and cos [(27m/T + us]t. Hence the
amplitude of the ripple component at (27m/T) + cos is obtained by taking
the square root of the sum of the squares of the coefficients of these two
factors, with the final result for the amplitude of the nth harmonic:
a _ sin (cosT/2)
(6.22)
n irn + (cosT/2)
, sin (to.r/2)
(6.23)
n irn — (a.T/ 2)
/ 1 Is /
A h
y X
bl
nI V
V
Output
Fig. 6.19. Output of full-wave modulator.
rates as high as 10 kc and (2) excellent linearity for signals down to 0.1 milli¬
volts in amplitude. There is little temperature drift so long as the input
impedance is kept low. Figure 6.20 shows a simple transistor chopper.1
The square-wave drive applied to the bases of the transistors alternately
switches one transistor on and the other off. For the half cycle in which
A conducts, B is cut off; thus the input is connected to the output through
the output transformer. For the other half of the cycle, A is cut off and
the path from the input is open-circuited. At the same time B conducts
and shorts the output through resistor R. While a common emitter cir¬
cuit could be used here, the common collector was chosen because the
static collector voltage for zero collector current is only about 1 mv. This
voltage appears across the output transformer in the same polarity for
both transistors, and if the transistors are matched, this is a constant d-c
value and thus does not appear at the output terminals.
where G is the gain during the conducting period. Clearly if, because of
variation of tube parameters, 0 changes, the output for a given input
changes. In particular, if the input is zero and the tube characteristics
are not exactly identical, some square-wave output will result. Thus if
the characteristics of one tube change relative to those of the other, the
amplitude of this square-wave output can change; this constitutes drift.
For the diode circuit the cause of drift is somewhat different and
depends on the fact that a diode will conduct some current when reverse
voltage is applied. Hence, if we consider the half-wave circuit of Fig.
6.11a, we find that during the period of nominal nonconduction some signal
may get to the output, which instead of oscillating between the signal
amplitude and zero will instead oscillate between full-input amplitude
and reduced-input amplitude. It is clear that, if temperature changes or
other factors cause the back current of the diodes to change, then the
magnitude of the output square wave changes for a given input, and we
again have drift.
Since a major reason for the use of a modulator in a servo is to prevent
the drift that is caused by d-c amplifiers, a modulator that drifts is of
relatively little usefulness. For this reason the most commonly used
modulator, particularly for systems using a 60- or 400-cps carrier, is the
electromechanical chopper, already discussed previously (Secs. 2.23 and
6.6). It may be connected as shown in Fig. 6.13, but with the input and
output terminals reversed, or it may be connected as a half-wave modu¬
lator, as shown in Fig. 2.35. In either case its good square-wave output
and lack of drift are the primary advantages.
PROBLEMS
6.3. A two-section low-pass filter used to smooth out the ripple generated by a half¬
wave discriminator operating on a 60-cps carrier has the transfer function
It is desired to reduce the rms value of ripple output to less than 10 per cent of the
useful d-c output. Assume the input signal is in phase with the reference and find the
time constant T required.
6.4. Repeat Prob. 6.3 for a full-wave discriminator. Find the ratio of time con¬
stants required in the two discriminator types.
6.6. Design a ring demodulator like that in Fig. 6.9. Assume both transformers
have unity ratio. The maximum expected value of the input signal is 20 volts rms.
The diodes are 1N34, having a maximum inverse voltage rating of 60 volts and a
maximum current rating of 50 ma. The circuit operates into a load resistance of
10,000 ohms, and it is desirable to keep the output impedance of the discriminator to a
minimum. Find the values of the resistors R required.
6.6. Find the current rating of the secondaries of the reference transformer and
of the rectifiers required in the power discriminator of Fig. 6.12. Express the result
as a ratio of maximum load current.
6.7. In the clamping discriminator shown in Fig. 6.14 the triodes have a plate resist¬
ance of 1,000 ohms when they conduct. The pulses applied to the grids cause the
tubes to conduct for one-twentieth of a cycle. During this time the capacitor should
charge to 95 per cent of the correct input value. The capacitor should hold the cor¬
rect value with a loss of less than 5 per cent during the period that the tubes do not
conduct. The pulse-repetition frequency is 400 cps. Find the value of C2 required,
and also the minimum value of load resistance.
CHAPTER 7
A-C MOTORS
--o A-c
Fixed
phase supply
lM)J
A-c
omplifier
-Sj
Control
phase
C1
1
) Motor
Gears
-1- Synchro
|
generator
A-c
£- KfJ-0 supply
Control
transformer
Input
Fig. 7.1. All a-c servo using a two-phase motor.
error between input and output. The error voltage from the control
transformer is amplified in a standard a-c amplifier and applied to the
control winding of the motor. The 90° phase shift between the control
and reference voltages may be obtained either by connecting a capacitor
in series with the reference winding or by incorporating a simple phase-
shift network in the amplifier. Neither method offers any particular
difficulty, but the second has the advantage that the phase shift is not
affected by the motor operating conditions. Another small departure
Sec. 7.2] A-C MOTORS 277
from conventional amplifier design results from the highly inductive input
impedance of a-c motors. To improve the effective load power factor, a
resonating capacitor is usually connected across the amplifier output
terminals.
It should be noted that it is not necessary to pass the signal from the
control transformer through a phase discriminator; the motor acts as its
own discriminator. Hence many simple control systems require no d-c
amplification at any point in the loop, and therefore the problem of drift
associated with d-c amplifiers is eliminated.
Another desirable feature is that the motor, in common with all induc¬
tion motors, requires no electric connections, such as brushes or slip
rings, between its stator and rotor, and friction is thus reduced to an
absolute minimum. This results in very smooth operation and is one
of the reasons why the induction motor is used even in systems where the
error signal is d-c and must be passed through a chopper or other modu¬
lator to actuate the motor.
A-c servomotors are built for power outputs ranging from x/i to 1,000
watts; however they are most commonly found in applications requiring
less than 10 watts. The efficiency is usually quite low, 5 to 20 per cent
being typical. This, in addition to the fact that the motors often must
operate near zero speed for extended periods of time, necessitates the use
of an external blower on units of more than 10-watt output rating to pro¬
vide proper cooling. Such a blower is usually driven by a separate
single-phase induction motor.
7.2. Construction Features. The stator frames of a-c servomotors are
all constructed in essentially the same way although motors of different
manufacturers have external differences, and various mounting styles are
available. Several types of commercially available servomotors are
shown in Fig. 7.2. The stators have a standard distributed winding to
improve the space distribution of flux density around the air gap. It
is desirable to have a flux distribution as close to sinusoidal as possible,
because space harmonics in the flux tend to produce notches and other
irregularities in the speed-torque curve of the motor. Most 60-cps
motors are wound for two or four poles, but 400-cps motors may be
wound for as many as 10 or more poles, to reduce the operating speed.
The two phases of the winding may be identical; however, in some
motors the control winding is designed for a much higher voltage than
the reference winding, and in some cases the control winding is center-
tapped. These special features make it possible to connect the motor
directly to the output stage of the amplifier without the need of an
output transformer. It is also possible, particularly in the smaller
motors, to design the stator winding in such a way that only a small frac¬
tion of the power input to the motor is supplied by the control amplifier,
278 CONTROL SYSTEM COMPONENTS [CHAP. 7
the major share of the power coming from the reference supply. All of
these special design features serve to reduce the size and weight of the
control amplifier.
There are three basically different rotor designs: the squirrel-cage rotor,
the solid-iron rotor, and the drag-cup rotor. Squirrel-cage rotors are
quite similar to those used in standard induction motors except that, in
order to reduce the moment of inertia, the diameter is usually much
smaller than the length (see Fig. 7.3). Squirrel-cage windings are subject
rotor
S • (b)
Fig. 7.3. Cutaway view of a-c servomotor with squirrel-cage rotor. Note long slim
design and skewed rotor slots.
Fig. 7.4. (a) Exploded view of a drag-cup a-c servomotor; (6) internal construction of
a drag-cup motor.
ho =
E 2o N, Ei
(7.1)
R2 + j'^fE2 Ni R2 + flirfLi
In these equations Ei and E2 are induced voltages per phase, and Ni and
N2 are the effective turns of the stator and rotor, respectively. R2 and
L2 are the resistance and inductance per phase, respectively, of the rotor,
and / is the line frequency. The additional subscript o indicates that the
quantities so marked are measured at standstill. If the rotor is permitted
to rotate, the relative velocity between the rotating magnetic field and the
rotor conductors decreases. Hence both the voltage induced in the rotor
and the rotor frequency decrease, becoming zero when the rotor speed
becomes synchronous, i.e., equal to the speed of the magnetic field. It is
easily demonstrated that the induced voltage and frequency are propor¬
tional to the slip S, defined as
In the second form of Eq. (7.4) the effective rotor resistance R2/S has
been broken up into two components, R2 and i?2[(l — S)/S\. This is
done to indicate two components of rotor power. The component due to
the currents flowing through R2, that is, \I2\2R2, is the rotor copper loss;
hence the remaining power, \I2\2R2[(1 — S)/S]f is proportional to the
mechanical power developed by the rotor.
In order to determine the stator current, we note that, whether or not
the rotor turns, the points of both maximum induced voltage and maxi¬
mum rotor current must rotate around the rotor in synchronism with
the stator magnetic field. The rotor current therefore causes a demag¬
netizing mmf N2I2 rotating at synchronous speed (note that a squirrel-
cage winding is effectively a single turn; hence the mmf is numerically
X* R'z 4
■ww^ ww—ORHP—
-►/t
fi
R,m \X,m
equal to I2). Since the magnetic flux in the air gap is fixed if the applied
voltage is fixed, a current must flow in the stator to produce a magnetizing
force exactly equal to the demagnetizing N2I2\ i.e.,
We use the symbol /{, since the current required to cancel the mmf N2I2
is only the power-producing component of the stator current. There is
also a magnetizing component IM, which is required to set up the rotating
magnetic field. From Eqs. (7.5) and (7.4), we get
_ n ft)
1 R'2 + jX'2o + R'2[( 1 -S)/S\ 1 ;
QQ = 2|/{|
The factor 2 arises from the fact that the analysis has been made on a
per phase basis. Since there are two phases, the actual power is therefore
twice the power per phase. From Eq. (7.2) we see that 1 — S = 12/12s,
where 12s is the synchronous speed. Hence
q = -||/:p§ (7.8)
When the unbalanced set of voltages is applied to the motor, the theory
of symmetrical components indicates that the motor acts as though both
symmetrical sets of voltage were applied simultaneously. Hence, if we
assume that the electric and magnetic circuits making up the motor are
linear, we may say that the torque developed by the motor is equal to the
difference between the torque produced by the positive-sequence voltage
and that produced by the negative sequence.
Equation (7.9) gives an expression for the torque produced by a bal¬
anced set of voltages; suppose this set to be the positive sequence. The
expression for torque produced by the negative sequence must have the
same form as (7.9) except that, if the slip is to retain its zero reference at
positive synchronous speed, (2 — S) must be substituted for S. Since
Eq. (7.9) shows the torque to be proportional to the square of the applied
voltage we find that in general
or, if we let |Fm|2E(*S) = Qb(S)} the torque produced with balanced input
voltages, then is
VM i V M2
Q = Qb(S) Qb(2 - *8) (7.23)
VM
If we substitute for VMi and Vmi their values as obtained in Eqs. (7.21)
and (7.22), we find that
The presence of the sin 9 term indicates that the developed stall torque
is a function of the phase angle between the control and reference voltages,
so that the motor may be thought of as a phase discriminator. Normally,
of course, 6 is fixed at 90°, since this results in maximum torque. Equa¬
tion (7.25) also implies that the motor is approximately linear as long as
the speed is low, i.e., as long as Qb(S) ~ Qb(2 — S).
Equation (7.24) may be used to plot a complete set of speed-torque
curves for the motor, provided that Qb(S) is known for all S of interest.
This function is defined in terms of the impedances of the motor as in
Eq. (7.9) and can, therefore, be computed if these impedances are known.
The measurement of the impedances is, however, always rather difficult,
and in particular it is almost impossible to separate rotor and stator
impedances by means of simple measurements. In large motors the
approximation is therefore made that ZM is very large, so that its effect is
negligible; then we see from Eq. (7.9) that Qb(S) becomes approximately
q /m _ 2 |Em|2-K2$
M } ~ Tf. i^ + ^i+y^Xx + x^i2
If this approximation is valid, then all that needs to be measured is R\, R2,
and Xi + X2. The stator resistance Ri can be measured by a d-c test;
Ri + R2 and Xi + X2 are obtainable from a measurement of the motor
input impedance at standstill (S — 1); and if the further assumption is
then made that R2 is independent of speed, Qb(S) can be computed for
any desired value of S.
In small servomotors, particularly in those with a drag-cup rotor, the
assumption that ZM is very large is not particularly valid. Furthermore,
R2 depends to some extent on the rotor frequency because of skin effect
and eddy currents. Hence a computation of Qb(S), in addition to being
rather laborious, will not give very accurate results. The problem is
usually, therefore, completely bypassed by measuring the speed-torque
curve of the motor with balanced input, that is, Qb(S), rather than the
impedances. This is quite easy to do for speeds between zero and syn¬
chronous, and this type of curve is usually furnished by the motor manu¬
facturer as part of the motor characteristics. The only difficulty is that
Qb(S) must be known for values of S ranging between 0 and 2; i.e., a
i
ZM V M 1 2 R'2S
Qt(S) (7.26)
Z\ + Zm R2 + jSX2 + &[Z\Zm/{Z\ + Zm)\ ils
In these equations the symbol 12 stands for the transform of A12, etc. If
the small electrical time lag due to the inductance of the stator and rotor
circuits is not negligible, then, approximately,
AVC- BQa
(7.31)
(TlS + 1 )(JBs + 1)
. ai2
A ~ dVc
_ 21L_(sin 6 + k)Qb(S) + (sin 6 — k)Qb(2 — S)__
VM (1 + 2k sin 0 + fcj) _ dQt(2 - S) _ 2k gin e + /c2)
do do
(7.32)
and
B = -
<912 _4Q,_
dQ <9Qb($) , 07 • n I 7 o\ dQb(2 — *S) 07 • /) | 7
—(1 + 2k sin 8 + k2)-' -- (1 — 2k sin 0 + k2)
do do
(7.33)
As an example of the application of Eqs. (7.32) and (7.33), let us use the
simplified form of the expression for Qb(S) developed in the previous sec¬
tion [see Eq. (7.28)],
s
Qb(S) (7.34)
Cl + C2S2
and use this to find A and B at the operating point ordinarily of greatest
interest in servo applications, i.e., the point at which S = 1, k = 0. We
obtain
Us Cl "f ^2 •
A = 2
V~u Ci - C SU1
(Cx + C2)2
and B = 212s (7.35)
Cl - c2
In many servomotors, Qb(S) is so close to being a linear function of S that
C2 in Eq. (7.34) becomes negligible. Under these conditions, 1/Ci
becomes Qb(l) = Qbo, the blocked-rotor torque, and thus
A = 2 sin 6 (7.36)
VM
B = 2^ (7.37)
Qb 0
If, on the other hand, C2 is larger than Ci, both A and B become negative.
This represents a form of unstable behavior referred to as single-phas¬
ingwhich is discussed in more detail in a later section. By differ¬
entiating Eq. (7.34) it can be shown that, when C2 is greater than Ci,
Qb(S) peaks for 1 > S > 0, i.e., for positive speeds.
The process of partial differentiation described above is relatively
rigorous, but it is rather involved. Hence a simpler, although less accu¬
rate, method is often employed to obtain A and B. In this method the
speed-torque curves of the motor are replaced by parallel straight lines,
as shown in Fig. 7.10. The line for k = 1, 6 = 90° is drawn through the
points 17 = 17s, Q = 0 and 12 = 0, Q = Qbo, and the lines for various k’s
and d’s are drawn parallel to the first one and such that the blocked-rotor
torque is proportional to k sin 8. It is clear that, since this method con¬
siders only the end points of the actual speed-torque curve of the motor,
it cannot take into account any curvature in this characteristic. Also it
is apparent that-it does not yield the same answer as the partial-differ-
Sec. 7.6] A-C MOTORS 291
B = £ (7-39)
Q — QbO
Q& — Qbo
292 CONTROL SYSTEM COMPONENTS [Chap. 7
The input impedance per phase presented by the motor to the positive
sequence is essentially the impedance of the equivalent circuit of Fig. 7.6.
This impedance is
Zm[(R'2/S + jx']
Z = Zi + = Z{S) (7.41)
Zm + (R'2/S + jX'2)
jRMXM
where Z\ = Ri -\- jXi and Zm =
Rm + jX M
Vd = -jVM1
Vc2 = jV M2
yc = b cl + Vc2 — ~j(VMl — VM2)
Hence Eq. (7.44) becomes
F M2 V Ml
— kV M
m — Zc — y M2 — Vmi (7.45)
[Z(2 - S) Z(S)_
From Eq. (7.11)
VM2 — Vm — Vmi
so that finally
1 + k + [Ze/Z(2 - S)]
Vmi = V (7.46)
2 + [ZC/Z(S)] + [Zc/Z(2 - S)]
and similarly
1 - k + [Ze/Z(S)]
VM2 ~ Vm (7.47)
t
2 + [ZC/Z(S)] + [Zc/Z{2 - S)]
294 CONTROL SYSTEM COMPONENTS [CHAP. 7
The developed torque can now be found from Eq. (7.23) and is given by
the expression
1
Q = Qb(S) l + k +
Z(2 - S)
2 + Z(S) 1 Z{2 - S)
- Qb(2 - g) 1 - k + (7.48)
Z(S)
It is apparent that the presence of Zc results in a rather complicated
expression, especially since it is now necessary to know the input imped¬
ance of the motor in addition to the speed-torque curve for balanced volt¬
ages. The remarks made concerning the difficulty of computing Qb(S)
apply with even greater force to the computation of Z(S), since it is
necessary to find not only the magnitude but also the phase angle as a
function of S. This information is not ordinarily supplied by the motor
manufacturers. Thus, unless a number of rather drastic simplifying
assumptions are made or unless Z(S) is measured directly, Eq. (7.48) is
difficult to use directly. Figure 7.12 gives a typical set of speed-torque
curves for a motor assumed to have the very much simplified equivalent
circuit shown; note that the curves for Zc = 200 + jO are considerably
less regular than those for Zc — 0.*
It is interesting to note that the addition of Zc does not change the fact
established previously that the blocked-rotor torque is a linear function
of k. If in Eq. (7.48) we let S = 1, we obtain, after some simplification,
1 T R
Q — kQb o (7.49)
1 + (Zc/Z) I*
where R is the real part of Zc/Z. For speeds close to zero the approxima¬
tion holds that Z(S) ~ Z(2 — S) « Z( 1), so that Eq. (7.48) can be
simplified somewhat. The transfer function of the motor, although basi¬
cally unchanged, is, of course, also more difficult to find. In particular,
the determination of A and B by the “exact” method becomes consider¬
ably more cumbersome. It is, however, apparent from the appearance
of the curves shown in Fig. 7.12 that both the gain and damping for the
operating point k = 0, S = 1 will be reduced by the presence of Zc in the
control circuit.
7.8. Unbalanced Stator Windings. Quite commonly, motors are built
with more turns on the control winding than on the reference winding.
The advantage of this construction is that the voltage rating of the con¬
trol winding may be made high enough to make it possible to connect the
motor directly to the output tubes of an electronic power amplifier with-
* The equivalent circuit and the curves are adapted from Koopman’s paper referred
to earlier.
Sec. 7.8] A-C MOTORS 295
Fig. 7.12. Speed-torque curves for simplified equivalent circuit. (From Koopman)
turn out to be negative) in the control winding, and the motor may then
be analyzed as if it were a motor with equal windings but with an equiv¬
alent Zc in the control winding.
7.9. Single-phasing of an Induction Motor. Single-phasing refers to
operation of a polyphase induction motor on a single-phase supply. A
motor will not single-phase if, when driven by an external torque, it
always develops an opposing torque, i.e., if it has a positive damping
coefficient. Conversely, a motor that develops torque in the direction of
rotation may single-phase if the developed torque exceeds the friction
torque; such a motor exhibits negative damping. Although it is theoreti¬
cally possible to stabilize a servo with a motor having negative damping,
it is difficult and requires relatively complicated control circuits. Single¬
phasing is therefore considered undesirable in a servomotor. Since an
ordinary induction motor will run on single-phase, once brought up to
speed, it is necessary to incorporate special design features in motors
intended for servo use to prevent them from single-phasing. These
features are discussed in the following paragraphs. Two cases are con¬
sidered: (1) the control phase is short-circuited, and (2) the control phase
is open-circuited.
If the control phase is short-circuited, Vc = 0; hence k = 0, and by
Eq. (7.24) the torque developed by the motor becomes
S2 = (Rjy_ (7.52)
7tV + (X£ -T Xo)2
so that, if the peak is to occur for S > 1,
(R'2)2 > Ro2 + (X' + X0)2 (7.53)
Sec. 7.9] A-C MOTORS 297
Torque
_ Z(S)
V Ml = VM (7.54)
Z(S) + Z(2 - S)
Z(2 - S)
V M 2 = VM (7.55)
Z(S) + Z(2 - S)
Z(S) given in Eq. (7.41). The result, after some reduction, becomes
i - s
yc — -jvmKZm2 (7.59)
Ki + K2S(2 - S)
where
X' = 0
ZM jXm R2 (7.60)
Z\ -f- Zm Z\ = R\ + jX i
Then Eq. (7.59) becomes approximately
XM2 (fi/«.)
Vc - jV M (7.61)
(Ri+jXJRUl + [(XM/R'2)Wtt8)]*}
Note that the nonlinearity is of the “saturating” type; i.e., the voltage
at high speeds is less than the value that would be obtained if the tachome¬
ter were linear. Note also that the phase angle
between Vc and Vm depends on the power-factor
angle of the stator and becomes 90° as the power-
factor angle approaches zero.
The explanation of the operation of an a-c
tachometer given above is based essentially on
the so-called double-revolving-field theory of the
operation of single-phase induction machines.
Fig. 7.14. Flux patterns
While the results obtained are correct, a somewhat in the a-c tachometer.
clearer physical picture of the operation is
obtained by use of the cross-field theory. According to this theory, the
currents flowing in the main stator winding set up an alternating flux
<j>M in the air gap (see Fig. 7.14). When the rotor rotates, two types of
voltage are induced in it by the main flux, a transformer voltage Em and
a speed voltage Eq. This effect is explained in some detail in Sec. 4.4,
and it is shown there that the speed voltage Eq is in time phase with the
1 Koopman, op. cit., p. 322.
300 CONTROL SYSTEM COMPONENTS [Chap. 7
per ring. Neither their direction of rotation nor the developed torque
can, therefore, be changed easily. To make shaded-pole motors reversi¬
ble, they are built with two shading coils per pole. These coils are wound
with a relatively large number of turns, and their ends are brought out to
external terminals so that the current in the shading coils can be con¬
trolled. An exploded view of such a motor is shown in Fig. 7.15. Note
that a standard squirrel-cage rotor is used.
The primary advantage of this motor is its very much lower cost com¬
pared to standard servomotors. This makes the motor useful in many
noncritical commercial applications where low cost is a primary consider¬
ation. In such applications the motor is often used with relay control of
the shading coils, and in this way it is possible to design simple follow-up
systems employing no amplifiers and only the simplest and most rugged
components. Somewhat higher performance can be obtained by elec¬
tronic control of the shading-coil current. For best results the shading-
coil current should be 90° out of phase with the current in the main field
winding, and under these conditions the operation of the motor resembles
that of a two-phase motor.
302 CONTROL SYSTEM COMPONENTS [Chap. 7
such as is shown by the dotted line in Fig. 7.18a; then by analogy with
Eqs. (7.30), (7.38), and (7.39), we have
Torque Torque
the applied torque. Q0 and SlM are the stalled torque and maximum
speed, respectively, as determined by the straight-line approximation of
Fig. 7.18a.
PROBLEMS
7.1. Plot the speed-torque curve for a two-phase induction motor having the equiva¬
lent circuit shown in Fig. 7.19. The voltage per phase is 50 volts, and the two voltages
are 90° out of phase. The frequency is 60 cps, and the motor is wound for two poles.
Fig. 7.19
304 CONTROL SYSTEM COMPONENTS [Chap. 7
7.2. Assume that one phase of the motor of Prob. 7.1 is supplied with a fixed 60-cps
voltage of 50 volts and that the variable voltage applied to the other phase is 90°
out of phase with the fixed voltage. Plot speed-torque curves for variable voltages of
0, 10, 20, 30, and 40 volts.
7.3. Find the constants Ci and C2 [see Eq. (7.28)] for the motor of Prob. 7.1.
7.4. Find the approximate motor transfer function for the motor of Prob. 7.1 on
the assumption that the moment of inertia of the rotor is 0.1 oz-in.2 Consider two
operating points: (a) speed = zero; variable voltage = zero; (b) speed = zero,
variable voltage = fixed voltage = 50 volts. Variable voltage lags fixed voltage
by 90°.
. .
7 6 The following data are available on the Diehl FPE-25-11 motor:
Output, watts. 5
Poles. 2
Reference volts. 115
Control volts. 115
Frequency, cps. 60
Locked torque, oz-in. 5.5
Torque at 1,800 rpm, oz-in. 3.8
Moment of inertia, oz-in.2. 0.098
Assume that the torque equation for 115 volts on both phases is given by
T = Ci + C& s ~ slip
(a) Find the quasi-linear transfer function for the operating point of zero speed, zero
control voltage. (6) Find the quasi-linear transfer function for the operating point
of zero speed, full control voltage, (c) Find the approximate transfer function using
the straight-line approximation to the speed-torque curves. The control voltage is
90° out of phase with the reference voltage in all cases.
7.6. Assume that the motor of Prob. 7.1 is used as an a-c tachometer. If the
reference voltage is 50 volts, find the numerical relation between velocity and output
voltage.
CHAPTER 8
S = Ma = M~ (8.1)
■7777777/
system with the first of the electric networks. Mechanical force is then
analogous to voltage. The second, and more modern, analogue is the
mobility analogue, in which mechanical force is analogous to current.
As Firestone1 points out, the mobility analogue is the more convenient of
the two because the schematic diagrams look the same. Furthermore,
current is more closely allied to force than is voltage, since the methods of
measurement are similar. In Table 8.1 is shown a comparison of elec¬
trical quantities and analogous mechanical quantities.
Electric Electric Mechanical
11 Force
M
I tzJ
777777,
O
O
O
O
Electrical Electrical
Mechanical (mobility analogue) (series analogue)
Vjo-
4'
M
lT
o o
M y o o
O o is
£3
O O'* 77//////.
m
B V/////////7////////.
7Z//7.
(o) U)
Fig. 8.4. Series mass, spring, and dashpot.
The transfer function of the system shown in Fig. 8.4 is most easily
obtained by rearranging Eq. (8.5) as follows:
so that
y _ Bs _ (B/K)s
(8.7)
x Bs + K (B/K)s + 1
One more example may clarify a point concerning the handling of mass.
In Fig. 8.4a is shown a series-connected mass, spring, and dashpot. It
should be apparent that there can be no relative velocity across the mass.
One end of the mass is moving at exactly the same velocity as the other
end. The mass moves only relative to ground; this is the implication of
the schematic representation of mass shown in Fig. 8.46, which is identical
to Fig. 8.4a. Thus the mass in no way affects the transfer function of the
system for position or velocity.1 This may be more clearly seen by
redrawing the circuit as shown in Fig. 8.46. We may find the transfer
function in exactly the same manner as above.
I = (B/K)s + 1 (8'8)
(1) Y Ts
J “ 1 + Ts
(2) 1
1 + Ts
B
K
(3) T2 1 + TlS
X T i 1 + T#
A B\ A B\
1 Ki 2 Kx + A2
(4) Y Ki
X Ai + K 2 1 + Ts
rji _A_ B2
Ai + A2
Y 1 + T 2s
X 1 -T T\S
A B i -T B 2 Bi
T2 =
AT
(6) Y T is
J ~ 1 + 7Ts
Ai + A2
2 irT~
(7) F 1 + TV
1 ~ 1 + Tl8 log &<
A b2 b2 A B,
Tx = — + — t2
Ai A2 K2
(8) Y T is
~ 1 4 T2s -
A Bl
T2 = — + —
Kx K2
(9) Y T2 1 + TV
J ~ Til + T2s
qi
A b2
JL _“ qi
a
A_ _1
BxB2_
' K2 2 A2(£x + 5a)
(10) F Bx 1
X B i -f- B2 1 T 7’s
A
qi .A. BxBo
1
Ki(Bi + B2)
F Bx 1 + TV
X B\ + B2 1 + T2s
^ a #2 ni (Ax + K2)B\B2
l 1 — l2
k2 A\K2{B\ -\- B2)
See Note 1
(12) F AT 1 + TV
I
2 1 + TV
rji A Bx1 rp ± Bx + B2
'_ if, * K, + K,
See Note 1
i
Fig. 8.5 (Continued)
Sec. 8.3] MECHANICAL NETWORKS AND GEARS 311
3 xr (Bi + B2)K2
* In the steady state s = job.
Note 1. By choice of parameters T2 can be made larger or smaller than T\, and
thus either a —1 slope or a +1 slope can be synthesized.
. Note 2. The approximation is good for Tx » T2. To find the approximate rela¬
tion B2 is considered open at low frequencies, and Bx is considered shorted at high
frequencies. See Chap. 1 for discussion of approximate design methods.
Note 3. The approximate relation is obtained by considering X2 shorted at low
frequencies and K1 open at high frequencies. Multiplied out, the approximate rela¬
tion is
Y
X
Y
X
AiKo
-AW- -AW-
A3
~A/VV—^
-« ■ ^a- -
(a)
Hex nut
inertia weights
Ball bearing
Torque motor
rotor and
stator
Pickoff electric
spring stator
ond rotor
Magnetic flux
return
Damping rotor
Magnet
Magnet holder
(b)
high-frequency carrier (400 to 800 cps) is more damaging than the same
percentage shift would be in a 60-cycle carrier system.
McDonald1 suggests that a compact electromechanical unit may be
used to demodulate the a-c signal and apply it to a mechanical d-c lead
network and then convert the signal back into a-c form. The block
diagram of such a device is shown in Fig. 8.8a, and in Fig. 8.8b is
shown a cutaway drawing of the actual device. The torque motor acts
respect to the flux, eddy currents are set up in the copper, which in turn
produce a retarding torque.
The frequency response of a d-c motor with an inertia load and with no
damper has been discussed in Chap. 4 and its asymptotic a diagram is
shown in Fig. 8.11a. The break occurs at a frequency of kvkt/J\Ra
radians/sec, where kv is the generated voltage constant of the motor, kt is
Sec. 8.6] MECHANICAL NETWORKS AND GEARS 315
the torque constant of the motor, Ji is the moment of inertia of the rotor
and load, and Ra is the armature resistance. The reader may show that,
with the coefficients of the damper properly selected, the new a diagram
will be as shown in Fig. 8.116. The response is thus well damped. The
empirical design of these dampers long preceded the control-system
approach to the problem and has been used with equal success with a-c
motors. The damper is used in many simple a-c or d-c systems as the
only equalizer in the loop. In certain cases the use of this device makes
possible a satisfactory response in a very simple a-c control system, where
without the damper a more complex d-c system with an electronic
equalizer would be required in order to meet specifications.
The Lancaster damper has several disadvantages. Since it is at a high-
power-level point in the loop, it loads
the motor and may absorb an ap¬
preciable portion of its rated torque.
Another disadvantage is that, after
the mass J2 (see Fig. 8.9) has reached
an appreciable speed, it provides a
torque that tends to cause the motor
to overshoot the zero-error position;
this may cause poor synchronizing
performance. Also, in the hydraulic-
type damper the effect of temperature
on the coefficient of viscous damping
will change the frequency of the
breaks on the a diagram and thus can
decrease the effective damping.
A modification of the Lancaster,
or untuned, damper is the so-called
tuned damper.1 When, in addition
to the Viscous-fluid coupling, the free plate is coupled to the shaft by a
spring, the damper is said to be tuned.
8.6. Gears. Gears are used to reverse the direction of rotation of a
shaft, to provide self-locking action, to provide a right-angle drive, etc.
The most common use is to provide a change in shaft speed. For
instance, most electric motors are designed to operate at relatively high
speeds and low torque. The typical control application for motors is just
the reverse of this. A reduction in shaft speed and an increase in shaft
torque are the usual application for gearing between an electric motor and
its shaft load.
The spur gear and pinion is the most common type of gearing in control
1 Greenwood, Holdam, and MacRae, “Electronic Instruments,” Radiation Labora¬
tory Series, vol. 21, McGraw-Hill Book Company, Inc., New York, 1948, Sec. 11.6.
316 CONTROL SYSTEM COMPONENTS [Chap. 8
Fig. 8.12. (a) Spur gear and pinion; (6) helical gear; (c) bevel gear and pinion; (d) worm
and gear.
applications.1 The small gear is called the pinion (see Fig. 8.12). This
type of gearing is simple to manufacture, since the gear teeth are at right
1 For more complete information on the design of gears see Berard, Waters, and
Phelps, “ Principles of Machine Design,” The Ronald Press Company, New York,
1955.
Sec. 8.6] MECHANICAL NETWORKS AND GEARS 317
angles with the body, or gear disk. The spur gear is a high-efficiency gear
and is thus reversible. Generally the highest gear ratio that is practical
with spur gears is about 10:1 for one mesh. Figure 8.13 gives a compari¬
son of proportions for the American Standard 20°-involute fine-pitch
system and the 14^°-pressure-angle system as modified for fine-pitch
service.1
For many years gear design has been standardized on these two pressure
angles, 14^° and 20°. The 14^° design is the older of the two standards
and was the product of practical considerations rather than engineering
research.2 As modified for fine-pitch service, the 14^° design has an
undercut pinion, which decreases the gear strength. The smaller pres¬
sure-angle design theoretically provides smaller friction and less backlash,
but in practice the improvement is negligible. In addition, tests show
Fig. 8.13. Comparison of the standard pressure-angle designs for fine-pitch service.
{Martin)
that scoring of the gear teeth can be eliminated by increasing the pressure
angle to 25°. Especially for small pinions the higher pressure ratio pro¬
vides higher surface durability and greater beam strength. It is possible
to design high-pressure-ratio gears with stubbed teeth to increase beam
strength and to yield a more favorable contact ratio.
1 Contact ratio is defined as the length of the path of contact between
two teeth divided by the tooth pitch (circular pitch), along the path of
contact, and it gives the average number of teeth in contact at any time.
For pinions and small gears the contact ratio for 20°-pressure-angle gears
may be more than twice that of 14^°-pressure-angle gears.3 This allows
use of a thinner gear or one made of lighter material. Furthermore,
owing to the undercut pinions in 14^° gears, the gear action is not con¬
tinuous, and the composite error is greater.4 The 20°-involute fine-pitch
system is the present American Gear Manufacturers7 Association (AGMA)
1 L. D. Martin, Instrument Gears, Machine Design, vol. 26, no. 2, p. 129, February,
1954.
2 Ibid.
3 Davison, Practical Considerations in Instrument Gear Design, Product Eng.,
vol. 22, no. 9, pp. 183-187, 1951.
4 Ibid.
318 CONTROL SYSTEM COMPONENTS [Chap. 8
standard, and indications are that even higher pressure angles may be
adopted in the future.
Gears are usually manufactured by cutting or hobbing the teeth in the
blank with special machine tools. With small pinion gears there are two
other methods that may also be used. Small pinions may be extruded,
or they may be cold-drawn. Only the softer nonferrous metals such as
bronze and aluminum lend themselves to extrusion, but any metal with
good cold-working properties may be cold-drawn.
Cold-drawing has several advantages over the two other methods.
The working strength of the teeth is increased. The method produces a
smooth, hardened surface, and the dimensional accuracy of the gear is
greater than with the hot extrusion process. Surface hardness up to
50 per cent greater and durability of 200 per cent more than that obtained
in the other processes are possible.1
The helical gear (Fig. 8.125) and the herringbone gear are modifications
of the spur gear. In the helical gear the gear teeth are slanted or helical
with respect to the gear disk. The helical gear is a smoother-running
gear than the spur gear, since the entire tooth does not make contact at
the same instant with its mate. A disadvantage of the helical gear is the
side thrust due to pressure of the mating teeth. This may be overcome
by combining a right-handed and a left-handed helical gear on the same
shaft. If the two gears are joined as a unit, the result is called a herring¬
bone gear. The herringbone gear is used only in very high power applica¬
tions such as nautical propulsion units, and is of little interest, therefore,
for control systems.
Both the spur gear and the helical gear may be convoluted into various
shapes. If the large gear is laid out flat, the result is called a rack and
pinion. If the gear shafts are not held parallel but placed at an angle,
the gear teeth must be modified, and the result is a bevel gear and pinion
(see Fig. 8.12c).
The pitch of a gear is defined as the number of teeth on a disk with a
1-in. diameter and is thus a measure of the fineness of the gear teeth. For
fractional-horsepower applications the standard production gear pitches
are 32, 48, and 64. The 48 pitch is recommended for standard applica¬
tions. For applications in which extreme smoothness is desired, the 64
pitch should be used. Where smoothness is not important and long gear
life under heavy loads is the prime consideration, the 32 pitch may be used.
A gear that is somewhat different from those described above is the
worm and gear (see Fig. 8.12d). The worm is a form of screw thread
which bears on a large gear set with its axis at right angles to the axis of
the worm. The gear is turned by rotation of the worm. The main
1 E. H. Rathbone, Advantages of Cold-drawn Pinions, Product Eng., vol. 21, no. 12,
p. 114, 1950.
Sec. 8.6] MECHANICAL NETWORKS AND GEARS 319
advantage of the worm-gear drive is that ratios of 100:1 or more may be
obtained with one set of gears. Usually worm gears are self-locking; i.e.,
the gear cannot drive the worm. In some applications this is desirable,
but in a servo control system the nonlinear action may cause stability
problems. It can be shown, however, that worm gearing can be made
fully as efficient as spur gearing1 and thus may be made reversible.
Self-locking of a worm-gear drive occurs under conditions of heavy
friction or small lead angle. The efficiency of a worm and gear may be
derived by considering the wedge, as shown in Fig. 8.14. The forces act¬
ing on the wedge may be expressed as components.
Here $ is the total force between the bodies, X is the lead angle of the
worm, at the pitch diameter, 0 is the pressure angle, or the slope of the
tooth, and / is the coefficient of friction. There is also a component of
force in the z direction which tends to move the worm and gear apart, but
since this motion is restrained by the shafts, it does not enter into the
efficiency calculation. Solving Eqs. (8.9) and (8.10) simultaneously
gives
cos 0 sin X + / cos X
(8.11)
cos 0 cos X — / sin X
(8.11)
(8.12)
We shall define the efficiency of the gear train as the ratio of Eq. (8.12)
to Eq. (8.11). The result is
where T is torque and o> is speed. The second relation is that the speeds
are inversely proportional to the gear ratio,
COi N2
(8.18)
0J2 N1
The Newton’s law relation may be written for the output shaft load as
J = L2 = .. (8.19)
CL 2 S(ji2
Substituting for the torque and speed from Eqs. (8.17) and (8.18), we have
T \&\/&2 T1N2/N]
J = (8.20)
sojiN1/'N2 so^iN 1/N 2
S&i
Ti
J
m
Thus the load moment of inertia reflected to the motor shaft is reduced
(8.21)
by the gear ratio squared. Therefore with only a moderate gear reduc¬
tion the reflected load moment of inertia
becomes negligible. The moment of the Motor
gears themselves is quite often not negli¬
gible, however.
r2
In order to minimize the inertia of the
gear train, the designer is tempted to place
the maximum practical ratio in the first
set of gears, thus making the effect of the
Fig. 8.16. Motor and a two-mesh
following sets negligible. Let us examine gear train.
the situation to see if this procedure is
correct. Figure 8.16 shows the configuration under consideration. The
total inertia seen at the motor shaft is
Jm -Ji + {Ji + Js) (?) +Ji (s) (s) + Jl (?) (a) (8-22)
where the J’s are the inertias of the gear disks and the r’s are the radii of
the disks. We shall assume that the reflected load inertia is negligible;
the inertia of r4 would then also be negligible, but it must be included in
the computation in order to solve for the individual gear ratios. Nor¬
malizing (8.22) to the inertia of the pinion ./1 and assuming that the
322 CONTROL SYSTEM COMPONENTS [Chap. 8
pinion inertias Ji and J3 and the radii r4 and r3 are identical,1 we have
f (ulY (8.23)
7f = 1 + (£ + 1)fe) +7 1 v2r4/
+ :*! +
1 3
^
ri
ri1 \r2r4,
(8.24)
1 (8.25)
rp ?Y r2
For a given over-all ratio, rir3/r2r4 is constant. Thus, since r 1 and r3 are
constant, r2r4 is constant:
M* = = Kr
\7V over-all/
Letting
r i2 = r32 = K2
we have
Jr
J1
= 1 + K, + ^r2
r22
+ T27fl (8.26)
Differentiating Eq. (8.26) with respect to r2 and setting to zero will give
the r2 for minimum inertia:
d Jm 2r2 2 K2 4Ki
= 0 (8.27)
dr 2 \J 1 Kl rV r2i
r26 - K2W ~ 2KiK2 = 0 (8.28)
As an example, take the over-all ratio as 10 and let the pinions have a
pitch diameter of in. Then
K2 = (0.25)2 = 0.0625
(0.25) (0.25)12
and K1 = 0.39
Vio
As an approximation, neglect the r22 term, which will be about one-tenth
the constant term and will make only a 1 per cent change in r2. Then
1 The pinions are chosen as small as practical and in instrument gearing are usually
identical. The more general case of power gearing, in which load capacity must be
included, is considered below.
Sec. 8.7] mechanical networks and gears 323
Fig. 8.17. The reflected inertia as a function of over-all gear ratio with the number of
meshes as a parameter. The multimesh curves are computed for optimum apportion¬
ing of the ratio between the meshes. Note the relatively small improvement of four
and five meshes over three meshes. (Courtesy Reeves Instrument Cory.)
r-1,000
-800
-700
r 600
f-500
T400
r 300
:-200
x
4
r-100
-80 c
-70 £
“60 r
r~50
i S
cn
r-40 ^
o
=-30 'o
Z L_
r-20 >
x o
2 r n
-10
4-
5-
6—1
Fig. 8.18. A nomograph that solves the equations given in the text. Place a straight¬
edge on the right-hand scale at the over-all gear ratio. The straightedge must also lie
on the point giving the number of meshes. The straightedge then intersects the left-
hand scale at the optimum ratio for the first mesh. The process may be repeated for
the subsequent meshes. The example in the text may be checked using these curves.
('Courtesy Reeves Instrument Cory.)
reduction. Thus the pinion gears in the second and following meshes
must be made wider in order to support the increased forces. It will be
shown that the calculation for power gearing can be significantly changed
when this requirement is included.
There are two standard empirical relations that are used to calculate
Sec. 8.7] MECHANICAL NETWORKS AND GEARS 325
the required size of power gears, the Lewis beam-strength formula and
the Buckingham wear-load formula. In a first approximation1 both of
these relations can be reduced to
WDP2 = Km (8.29)
where W is the face width of the meshing gears, Dp is the pitch diameter of
the driving pinion, m is the gear ratio from motor to given pinion, and K
is a constant of proportionality. Thus it is seen that the face width
and/or the pitch diameter of the pinions must be increased as we progress
through the gear train. This increase must be made when the forces in
the gear train are an appreciable portion of the maximum allowable force
for the material and gear design used. When over-all gear ratios are
greater than 100 or when four or more pinions are used, this increase in
pinion size will result in a considerable change in the gear design.
From practical fabrication and utilization considerations, Peterson2
suggests that face width and pitch diameter of successive pinions be
chosen by the empiric relation
W_=Dp=Po (8.30)
Wo Dpo P
where P is the diametral pitch of the gear and the subscript 0 refers to the
motor pinion.
The solution for minimum inertia employing the criterion in Eq. (8.30)
is displayed graphically for various meshes in Fig. 8.19. It will be noted
that, for small over-all ratio and low number of total meshes, the results
from Fig. 8.19 are identical with those from Fig. 8.17, which was computed
for constant-sized pinions. For higher ratios, however, the use of Fig.
8.19 will result in a reduction in inertia of greater than 2 to 1.
The advantages of constant-sized pinions are, of course, all the advan¬
tages of standardization. It should be borne in mind also that, where the
force transmitted is negligible with respect to the strength of the gears,
all the gears may be reduced in thickness and Fig. 8.19 will still yield
optimum results. Figure 8.20 gives curves that allow calculation of the
individual gear meshes after the total number of meshes has been chosen
from Fig. 8.19.
A possible compromise between the two extremes of complete standard¬
ization and individual design exists for medium- and light-duty gear
trains. It is possible at the slow-speed end of the train to use gears of
lower pitch than those used at the high-speed end. The low-pitch gear
teeth are larger and stronger and are more suited for heavy duty. The
over-all cost of the gear train is kept low by employing standard gear
1 D. Peterson, Power Gear Trains, Machine Design, vol. 26, no. 6, p. 161, 1954.
2 Ibid.
326 CONTROL SYSTEM COMPONENTS [Chap. 8
f __1__1______i___L__J___III
1 2 3 4 6 8 10 20 30 40 60 80100 200 400 600 1,000
Total reduction ratio,m
Fig. 8.19. Reflected inertia as a function of gear meshes for power gears. {From
Peterson)
Fig. 8.20. Choice of individual meshes given over-all ratio and number of meshes.
Subsequent meshes are chosen by reducing ratio and number of meshes by the first
mesh choice. {From Peterson)
pitches; yet the strength requirements are met. This solution is widely
used in practice.
Quite often, the inertia of the gear train can be reduced by measures
other than optimizing the ratios. By using punched gear disks rather
than solid disks; the' inertia can be reduced to about 75 per cent of the
Sec. 8.8] MECHANICAL NETWORKS AND GEARS 327
solid-disk value. A punched gear disk is one in which much of the
material has been removed from the otherwise solid body of the disk.
Substitution of materials can work an even greater reduction. The
substitution of aluminum gears for steel reduces the inertia to 33 per cent
of its original value, and the substitution of aluminum for brass gears
reduces the inertia to 25 per cent of its original value. Of course, factors
such as tensile strength and durability must be considered when sub¬
stituting materials.
The use of nylon for medium- and light-duty gears is also a possibility.
The extreme lightness of nylon (specific gravity, 1.14) is its main attrac¬
tion, although it is also easy to fabricate, resilient, and corrosion-resistant.
Nylon has a low coefficient of dry friction, which in effect makes it self-
lubricating, and it resists wear and abrasion. Nylon’s ultimate tensile
strength is about 12,000 psi, and it should not be subjected to environ¬
mental temperatures of greater than 100 to 120°F because it rapidly
loses its tensile strength and form at high temperatures.1
8.8. Gear Ratio for Load Matching. In instrument gearing the load
is usually negligible, and the inertia of the gear train is the only factor
that must be considered. The gear ratio is chosen on the basis of accu¬
racy and/or loop gain. The motor is chosen for its time constant, includ¬
ing the inertia of the gear train.
In power applications, however,
the power requirements of the load
and the power capacity of the motor
become important. It is usually
necessary to choose the smallest
motor that is capable of supplying
the load in order to minimize size,
weight, and cost of the power pack¬
age. The over-all gear ratio will thus Fig. 8.21. (a) Torque versus speed and
(b) horsepower versus speed of typical
be chosen with this in mind. The
a-c servomotor.
number of meshes and ratios of each
mesh should be designed by the methods discussed in Sec. 8.7 for power
gear trains.
In Fig. 8.21 is given, as an example, the approximate torque and horse¬
power plotted against speed for a typical a-c servomotor. In this particu¬
lar case, the motor develops maximum power at Ns/2, but whatever the
shape of the motor speed-torque curve, the speed for maximum power
may be found. Assuming that the required load velocity is known, the
over-all gear ratio should be chosen so that the motor is operating at the
speed at which it develops maximum horsepower when the load is oper-
1 R. Zimmerli, Designing Fabricated Nylon Parts, Machine Design, vol. 26, no. 3,
March, 1954, pp. 153-159.
328 CONTROL SYSTEM COMPONENTS [Chap. 8
ating at its design velocity. The horsepower rating of the motor can be
determined by determining the speed-torque characteristic of the load
and reflecting it through the gear train to the motor shaft. For instance,
in Fig. 8.22 are shown the horsepower and torque curves for a load con¬
sisting of viscous damping and inertia.
The maximum velocity of the load and
the maximum acceleration are specified.
When the reflected load horsepower-speed
Design
horsepower curve and the torque-speed curve are
superimposed on the motor curves, the
motor output must exceed the load by a
given design factor, usually 1.5 or 2 to 1.
8.9. Backlash in Gears. Backlash in
gears is the looseness or play between the
input and the output of the train. Back¬
lash may be measured as the angular dis¬
placement through which the input can be
moved with the output fixed, or it may be
Design Load speed given as the linear distance along the pitch
velocity
circle through which a gear may move with
Fig. 8.22. Horsepower and
respect to its fixed mating gear. AGMA
versus N for typical load.
standards are given in inches and can be
converted to angles if the pitch diameter is known. The AGMA standard
backlash classifications are shown in Table 8.2, along with composite error
limits.
The composite error referred to in Table 8.2 is an effect which may be
considered separately from backlash. In the cutting of the gear teeth
and the locating of the center bore for the gear shaft, there will inevitably
develop a certain amount of eccentricity between the shaft and the pitch
circle as well as distortions of the pitch circle itself. The eccentricity
that these errors lend to the gear-tooth motion as the gear is turned is
called composite error. The distance between the extremes of this
eccentricity is the total composite error, while the variation over the
angle covered by one tooth is called the tooth-to-tooth composite error.
Backlash results in a nonlinear relation between input and output of
the gear train and has received considerable study in the past few years.1
For large-amplitude signals, backlash has little effect, but as the signal
amplitude is reduced, backlash becomes a more important factor.
If a frequency-response characteristic of a portion of a system is to be
taken, the standard procedure is to maintain a constant input magnitude
or even reduce the magnitude as the frequency is increased. Let us
1 Chestnut and Mayer, “Servomechanisms and Regulating System Design,” John
Wiley & Sons, Inc., New York, 1955, vol. II, sec. 8.1.
Sec. 8.9] MECHANICAL NETWORKS AND GEARS 329
assume that the system contains a gear train with some backlash. Since,
in most systems, the response drops off at high frequencies, the amplitude
of the motion of the input gear will be reduced as the frequency is
increased. On account of backlash, the output of the gear train will be
reduced over and above the reduction that linear theory would predict at
•
Class A
20-45. 0.004 -0.006
46-70. 0.003 -0.005
71-90. 0.002 -0.0035
Class B
20-60. 0.002 -0.004
61-120. 0.0015-0.003
121 and finer. 0.001 -0.002
Class C
20-60. 0.001 -0.002
61-120. 0.0007-0.0015
121 and finer. 0.0005-0.001
Class D
Any pitch. No measurable backlash
* Between two assembled gears at their tightest point of mesh. Backlash will be
increased when the low points of runout are in contact.
instability. Finally the double gear increases the train inertia and
requires a longer pinion.1
A second possibility is to reduce the number of gears in the train. If
the over-all ratio is fixed, this requires a larger ratio in the first stages than
would be dictated by minimum-inertia considerations. The final choice
of the number of gear meshes is thus a compromise between reduction of
the effect of backlash and reduction of inertia. The reduction of inertia
must usually, therefore, be achieved by other means than increasing the
number of gear meshes.
Another possibility for reducing backlash is to use a gear train with a
very fine pitch and to set the shafts so that the gears bear rather tightly.
This increases bearing and rubbing friction, however, and thus cannot be
carried too far. While it is theoretically possible to adjust the operating
1 G. W. Michalec, Precision Gearing, Machine Design, vol. 27, p. 202, February,
1955.
Sec. 8.9] MECHANICAL NETWORKS AND GEARS 331
Totally enclosed
C spring
Fig. 8.24. Spring-loaded spur gears. Two forms of tension springs are shown.
limits the allowable loads. When the shafts are parallel, however,
beveloid gear teeth have line contact and are thus equivalent in wear and
load-carrying ability to conventional gears.1 For small skew angles,
beveloid gears approach the loading potential of conventional gearing.
8.10. Ball -screw Actuator. Several possible ways of converting rotary
motion into linear motion exist. The rack and pinion and the worm and
pinion are two examples. Another example is the simple screw thread
and nut. If the nut is prevented from turning as the screw is turned, the
nut travels linearly along the screw. The ball-screw actuator differs from
1 Ibid,
Sec. 8.10] MECHANICAL NETWORKS AND GEARS 333
the nut and screw only in that the operating surfaces rest on ball bearings.
Figure 8.26 shows a ball-screw actuator. As the actuator moves along
the screw, the balls are left behind. The tube shown in the figure carries
the balls to the front of the actuator, where they reenter it. The advan¬
tage of the ball-screw actuator over conventional threads is the greatly
Fig. 8.26. The ball-screw actuator, (a) Typical ball-bearing screw assembly employ¬
ing threaded nut with single ball circuit. Projecting guide finger on return tube
deflects balls into tube for recirculation. (6) Heavy-duty ball-bearing screw assembly
which has three ball circuits and yoke-type ball deflectors. (Courtesy Saginaw
Steering Gear Division, General Motors Cory.)
MECHANICAL COMPONENTS
^ Keyed
Ball bearing =a Input No.l
Ball
bearing
T.M
2- 4
::c.- j^-Boll bearing
Output Z?
(B =mm i N\\\m=
t 5
Keyed =2 f Input No. 2
~Ball bearing
Fig. 9.1. Bevel-gear differential.
integral with gear 2, and gear 3 is integral with gear 4. Both of these two
pairs of gears are arranged to rotate freely about the output shaft. The
bevel gears 2 and 3 mesh with gears 5 and 6. Gears 5 and 6 also ride free
of their shaft on ball bearings. Consider first that input 1 is fixed; then
gears 3 and 4 cannot turn. If input 2 is turned in the direction of the
arrow, gears 1 and 2 will also turn, as will gears 5 and 6, all in the arrow
direction. Gears 5 and 6 will thus “walk” around on the stationary gear
3, and they will carry the output shaft with them, in the arrow direction.
Now if input 2 is fixed and input 1 is turned, the output shaft will still
turn in the direction of the arrow. When both input shafts are rotated
1 Soroka, “ Analog Methods in Computation and Simulation,” McGraw-Hill Book
Company, Inc., New York, 1954, Chap. I. Svoboda, “Computing Mechanisms and
Linkages,” McGraw-Hill Book Company, Inc., New York, 1948.
334
Sec. 9.2] MECHANICAL COMPONENTS 335
^B
Planet carrier
Plonet
gears
>B
Section A-A
More important, the transfers from the two inputs to the output are not,
in general, the same, and they are not independent of the gear ratios as in
the bevel-gear differential.1 Figure 9.36 shows a more symmetrical form of
spur-gear differential used in an airborne computer. In Fig. 9.3a gear 3 is
integral with input shaft 1. Gear pair 2 and 5 and gear pair 4 and 7 are
integral. Let us assume that input 2 is fixed and that input 1 is moved
1 Svoboda, ibid.
Sec. 9.3] MECHANICAL COMPONENTS 337
in the direction shown by the arrow. Gear 1 then cannot move, and gear
3 is moved by the input shaft in the direction shown. The two gear pairs
are thus driven, and they in turn drive gear 6, which is integral with the
output shaft. If input 1 is fixed, input 2 turns gear 1, which walks the
gear pairs around the stationary gear 3. We note that the output shaft
will only move if the ratio r3/r2 is not the same as r6/r6. We have the
further constraint that, if the input 1 shaft and the output shaft are to be
collinear, as shown in Fig. 9.3a, it must be true that
r3 + r2 = rb + r6 (9.1)
Output = r3 rb
Inputx r2 r6
Output _ r3 rb
(9.3)
Input2 r2 r6
where the positive sense for input 2 is the same as for input 1. The net
output is thus
Given either r3/r2 or rb/rb, we may then select the other ratio to give any
desired ratio of the two input-to-output ratios. While this flexibility is
useful in adjusting gains, the most common arrangement is an input-out¬
put ratio of 0.5 for both inputs. In this case, r3/r2 and rb/rb can equal 1.0
and 0.5, respectively.
9.3. The Universal Joint. The Hooke joint, or universal joint, is a
device used to couple two shafts that meet at an angle. In Fig. 9.4 is
shown a primitive form of the joint. The points Ai and A2 are the ends
of the yoke on shaft A, and the points Bi and B2 are the ends of the yoke
on shaft B. We are interested in the distortion of the angular motion
produced by this device. Let us erect a plane normal to the shaft A and
a plane normal to the shaft B, both passing through the point 0. Then
these two planes will meet at the angle a, the angle between the two
shafts. The two planes will contain, respectively, the locus of the point
A i and the locus of point B\. In Fig. 9 Ad are represented the two planes
and the loci of the two points Ai and Bx. If shaft A is moved through
an angle 6, Ai will move to A[ and Bi will move to B[. The angle AiOB1
and A[OB[ is always a right angle because it is held so by the physical
cruciform. Furthermore the projection of A[OB[ on the plane normal to
A is a right angle. This can be shown from the theorem of solid geometry
338 CONTROL SYSTEM COMPONENTS [Chap. 9
that the projected angle of two intersecting lines at right angles on a plane
parallel to one of the lines is always a right angle. The plane normal to A
is parallel to line 0A\, in fact it contains it. We have before us the prob¬
lem of finding the angle </>, or BiOB[, which the shaft B has actually
turned through. The point /3 on the line of intersection of the two planes
By
(£>) End
Fig. 9.4. The universal joint. View (c) shows the joint 90° later than (a); view (b) is
the end view of the joint with angular position as in (a).
may be defined as having the same vertical height as B[ and b[. Thus
b[(30 and B[(30 are right triangles, and we may write
b[(3
tan 6 = (9.5)
0(3
In Fig. 9.5 is shown the variation of </> about its proper value as a function
of 6 for a = 45°. The maximum error of about 10° occurs at 6 = 38°.
In Fig. 9.6 is shown the maximum error in 4> as a function of a and also
the value of 6 for which this error occurs. For small shaft angles this
error is negligible. For large shaft angles two universal joints, each pro¬
viding one half the total angle, should be
used. If the yokes connected to the cen¬
tral shaft between the two universals are
coplanar, the errors of the two joints can
be made to cancel.
In a velocity system the angular varia¬
tion in a universal joint produces spurious
frequency components in the output ve¬
locity which must be considered if a care¬
ful analysis is being conducted.
9.4. Strain Gauges. A strain gauge is
a device that measures minute motions of Fig. 9.6. Maximum of <j> — 6 as a
1 Hetenyi, “Handbook of Experimental Stress Analysis,” John Wiley & Sons, Inc.,
New York, 1950.
340 CONTROL SYSTEM COMPONENTS [Chap. 9
for this is that the signal derived from a strain gauge is usually at a very
low level and requires considerable amplification before it is useful. One
such strain gauge is the capacitor type. It consists essentially of two
plates so mounted on the object undergoing strain that they move relative
to each other as a result of the strain.
/////;
The relative motion of the two plates
Specimen causes a small change in the value of
capacitance between them. This is best
detected by the use of a Wheatstone
bridge, as shown in Fig. 9.7. A constant
Fig. 9.7. Typical capacitor strain value of a-c voltage is applied to the input
gauge showing bridge type of
terminals. A deflection of the observed
sensing circuit.
body will change the spacing of the
capacitor and unbalance the bridge, thus causing a voltage to appear at
the output terminals. The voltage is a function of C' and thus of strain;
the relationship is given by
bridge ease, the operation may be considered linear over the normal
operating range.
Symmetry can be improved, thus decreasing quadrature error, and
nonlinearity decreased, as shown in Fig. 9.10, by arranging resistor a to
Cross section A-A
(a)
Fig. 9.9. A nichrome-wire strain gauge and bridge circuit. (■Courtesy Baldwin-Lima)
since the centrifugal force of the rotating mass increases as the square of
the velocity. For the configuration shown in Fig. 9.11 the torques about
the pivot point may be summed as
M/co2(a + l sin a)l COS a = (Kx + Bx + Max)b cos a (9.10)
This equation is nonlinear, not only because of the co2 term but also
because of the presence of the sin a term on the left-hand side. This term
can cause nonsinusoidal oscillations under certain conditions. We shall
neglect this possibility and assume that the terms in the first parenthetical
expression may be represented by a “constant” K. Then taking the
transform of the simplified equation,
A
X MflK/b
A O
(9.11)
or M „s2 + Bs + K
(9.12)
J — const.
tangent, or
T dt
— ddy = tan ( — dOy) = -J -— (9.13)
J s^s
so that
ddy Tz
(9.14)
dt Wy Jsws
where is the velocity of precession.
By similar reasoning it can be shown that the gyro precesses about the
z axis when a torque is applied along the y axis, or
dez = Ty
(9.15)
dt J §co §
Again the direction of precession is such as to line up the spin axis with
the direction of the torque vector.
Equations (9.14) and (9.15) are only approximations, because they do
not consider the fact that a change in precession velocity, multiplied by
the rest inertia of the gyro, also represents a change in angular momen¬
tum, which requires torque. Also they neglect the effects of gimbal fric¬
tion, which result in braking torques when the precession velocities
become large. The equations are therefore reasonably accurate only as
long as the precession velocity is low and the torque variations small.
There are several interesting things about the results obtained thus far.
First, the precession is along an axis perpendicular to that of the applied
torque; yet had not the disk a velocity about the spin axis, rotation would
have taken place along the axis of the applied torque. Second, the
applied torque results in a velocity rather than an acceleration, and
finally the moment of inertia about the spin axis enters into the relation
rather than the moment of inertia about the axis of the applied torque.
All of these facts make the gyroscope somewhat puzzling on first
examination.
The phenomenon of precession may be examined in another manner
that is perhaps more closely related to familiar concepts. To show the
force required for precession, the velocity of precession will be assumed,
and the required force will be found. The coordinate system fixed in
space is as shown in Fig. 9.15. Assume that the disk is spinning with an
angular velocity cos and also that the disk is precessing with a velocity
cop. Consider a particle of the disk near the rim.1 This particle has a
linear velocity parallel to the x axis, which is due only to the precession.
This linear velocity of the particle varies as the particle is carried around
by the spin velocity. At point A in space the component of velocity in
the x direction has a maximum negative magnitude, and at point C the
1 S. T. Preston, The Mechanics of the Gyroscope, Sci. American Supplement, vol. 58,
no. 1501, pp. 24,057-24,058, Oct. 8, 1904.
346 CONTROL SYSTEM COMPONENTS [Chap. 9
Fig. 9.15. (a) The disk with respect to coordinates fixed in space. (6) The linear
velocity of a particle of the disk in the x direction, (c) The linear acceleration in the
x direction of a particle of the disk.
has shown that this process gives results identical to those already
obtained.
9.8. Gyroscope. Equations of Motion. When the gyroscope is used
as a control-system component, the elementary relations between torque
and precession velocity [Eqs. (9.14) and (9.15)] are not sufficiently accu¬
rate, and the effects of gyro rest inertia and gimbal friction must be con¬
sidered. This may be done rather easily if the arrangement of the gyro
disk, gimbals, and torque motors is such that during precession the spin
axis always remains in a plane perpendicular to the torque causing the
precession. A commonly used configuration meeting this requirement
is shown in Fig. 9.16. Note that the torque of one of the torque motors
is applied to the gyro through a bail-ring having a slot in which the gyro
spin axis is free to slide. This arrangement permits both torque-motor
stators to be fixed to the outer frame. In this type of system we may
consider the torques required to accelerate the gyro rest inertia and to
overcome gimbal friction as subtracting from the “applied” torque,
leaving an “effective” torque resulting in precession only. If the gimbal
friction is assumed to be viscous, then this reasoning modifies Eqs. (9.14)
and (9.15) as follows:
rp j d~dz ddz j dOy ^ i^
T‘ - W - B° Tt ^ Hi (9J6)
rp j d Qy dOy _ j ddz
y Jy dt2 v dt ~ sC°s dt K' }
348 CONTROL SYSTEM COMPONENTS [Chap. 9
where J z and Jy are the rest inertias of the gyro and gimbals measured in
the z and y axes, and Bz and By are the coefficients of friction in these
axes. The torques Tz and Ty are the components of torque perpendicular
to the spin axis, since components parallel to the spin axis do not result
in precession. Hence if the actual torques applied by the torque motors
are T'z and T'y and if dz and 6y are measured between the spin axis and a
line perpendicular to the two torque axes, then Eqs. (9.16) and (9.17)
become
d2dz dSz j ddy
T[ cos 6y — Jz - B, (9.18)
w dt ~Js0)S~dt
e2 J s^s
(9.22)
Ty s[J yj zs2 + (JyBz + JzBy)s + (JsCOs)2 + ByBz]
This transfer function is, of course, accurate only if the conditions under
which Eqs. (9.20) and (9.21) were derived are met. Normally, friction
is sufficiently small to make ByBz « Jsws; hence Eq. (9.22) is approxi¬
mately equivalent to
We note that the gyro response has the standard quadratic form, and if
friction is small, the response to step or other shock inputs is oscillatory.
The transient oscillations observed in a gyro are referred to as nutations,
since the gyro spin axis characteristically traces out a conical surface dur¬
ing these oscillations. The frequency of nutation, i.e., the undamped
natural frequency of Eq. (9.23), is seen to be
con (9.24)
Since Jy and Jz are usually larger than J8, the nutating frequency is
usually less than the spin velocity. In fact it can be demonstrated that
the nutation frequency of a thin disk spinning in space without gimbals is
equal to one-half the spin velocity.
The transfer function relating precession velocity to precessing torque
[Eq. (9.23)] is referred to as the direct transfer function of the gyro. How¬
ever, owing to the gyro rest inertia and gimbal friction, a torque applied to
one set of axes of a gyro also results in some gyro motion in the same axis.
The relation between torque and angular displacement along the same
axis is referred to as the cross-coupling transfer function and can be
By[(Jy/By)S + 1]
(9.25)
J zJ y j zBy —I- J yB z
J 2 C0S 5
T 2., 2 s2 + T 2., 2 S + 1
K'J s ds
A similar expression may be derived for the y axis. A large cross coupling
is usually undesirable in a gyro, and Eq. (9.25) indicates that it is mini¬
mized by reducing friction and by making the spin velocity as large as
possible.
The complete set of gyro transfer functions may be summarized in the
block diagram shown in Fig. 9.19. In this figure con is the nutation fre¬
quency given by Eq. (9.24), and t; is the damping coefficient, which, by
Eq. (9.23), is equal-to (l/2Jscos)(Bz \/Jy/Jz + By \0~JTy).
Sec. 9.9] MECHANICAL COMPONENTS 351
9.10. Gyroscope Applications. Two basic gyro types are used in prac¬
tice: the free gyro and the restrained gyro. A free gyro is mounted in a
set of gimbals so that angular motion of the supporting frame is not trans¬
mitted to it. Ideally the direction of the spin axis of a free gyro is there¬
fore fixed in space. Such gyros are used to provide a fixed reference
position, for instance, in the vertical and directional gyros found in auto¬
matic pilots and in inertial navigation equipment. Other uses of this
sort are in artificial horizons, in roll-stabilizing equipment, and so forth.
Since even the most carefully constructed gyros are subject to a certain
amount of random drift, the accuracy of the reference direction estab¬
lished by the gyro tends to deteriorate with time. In some cases this is
of no consequence. Thus, a gyro used to establish the vertical or direc¬
tional references in a guided missile having a life of only a few minutes
may drift a few degrees per hour without any adverse effect on the accu¬
racy of the missile. Drift in the gyro used as the artificial horizon in an
airplane may not be important if the gyro can be reset occasionally.
However, in more critical applications the gyro is not completely free but
is equipped with small torque motors so that it can be precessed. In this
way the gyro can be automatically slaved to a more accurate primary
reference. Thus, a gyro used to establish a vertical may be slaved to a
pendulum of some sort. The position of the gyro is continuously com¬
pared to that of the pendulum, and any difference in the positions is used
to precess the gyro to make its position correspond to that of the pendu¬
lum. The precessing torque applied to the gyro is kept so small that the
maximum precession rate is only a little larger than the maximum drift
expected. In this way the gyro cannot follow rapid fluctuations of the
pendulum position, and the system acts merely to keep the gyro spin axis
aligned with the average pendulum position. Thus the gyro acts not so
much as a primary reference but as an integrating or smoothing device
applied to the actual reference. In a similar way directional gyros are
often slaved to a magnetic compass which provides the primary north-
south reference.
A somewhat different principle is used in the gyrocompass used on
ships. This is essentially a free gyro equipped with a weight that tends
to keep the spin axis in a horizontal plane. The weight hangs, supported
by two bearings on the spin axle, centered below the gyro disk. This
weight, together with the rotation of the earth, acts to precess the gyro
until the spin axis takes on a north-south position.
An application of the gyroscope that differs fundamentally from those
described above is the measurement of low angular velocity. In this
application, use is made of the fact that the precession velocity is directly
proportional to the precessing torque. Thus in one form of gyroscopic
velocimeter the gyroscope is equipped with synchro devices (see Chap. 5)
354 CONTROL SYSTEM COMPONENTS [CHAP. 9
that sense the relative displacement between the gimbals and the outer
frame. The output of the synchro is amplified and applied to the proper
torque motor in such a way that the resulting gyro precession returns the
output of the synchro to zero. This feedback system therefore acts to
keep the gyroscope aligned with the
outer frame. The torque devel¬
oped by the torque motors to ac¬
complish this result is, however, by
Eq. (9.14) directly proportional to
the precession velocity and there¬
fore to the velocity of the outer
frame. Commonly, the torque
motor is an electric motor, and the
torque is proportional to the voltage
Fig. 9.20. Gyroscopic velocimeter.
applied across the windings (see
Chap. 7). This voltage may there¬
fore be used as an indication of the velocity. A simplified schematic
diagram of such a system arranged to measure velocity in two axes is
shown in Fig. 9.20.
A similar, although somewhat simpler, device working on essentially
the same principle is the restrained, or rate gyro, shown in Fig. 9.21. In
Z axis
springs shown. Suppose now that the entire assembly is rotated about
the y axis. F-axis torque is applied to the gyro through the gimbal, and
the gyro begins to precess about the 2 axis. This precession is arrested by
the springs, which act to produce a torque on the gyro about the z axis,
opposing the z-axis precession. The torque generated by the springs
results in precession about the y axis, so that the gyro is able to follow
the rotation of the outer frame. The torque supplied by the springs
must, therefore, be precisely sufficient to precess the gyro at the rate at
which the outer frame is being rotated. Hence this torque is a direct
measure of the speed of the outer frame. The simplest way to measure
the torque is to measure the change in displacement of the gimbal relative
to the outer frame, since if the restraining springs are linear, this displace¬
ment is proportional to the torque. For maximum accuracy the spring
should not deflect appreciably, since the torque required to precess the
gyro about the y axis (fixed with respect to the outer frame) becomes
smaller as the spin axis of the gyro approaches the y axis.
If the angular momentum of the gyro is large," a large torque is gener¬
ated by relatively small velocities. Hence rate gyros can be made sensi¬
tive to very low angular velocities. With reasonable care a rate gyro can
be constructed that indicates the angular velocity of the earth, a velocity
of less than 0.0007 rpm. It should be noted, incidentally, that the
velocity measured by a rate gyro is always with respect to an inertial
reference fixed in space; furthermore since velocities about the x and z
axes of Fig. 9.21 do not have the action described, the rate gyro is in
general sensitive only to the y component of the angular velocity applied
to it.
The transfer function of the rate gyro can be obtained from Eq. (9.20).
The spring acts to make Tz proportional to dz; hence Eq. (9.20) may be
rewritten as follows:
J, o),
6, Jzs2 -f- Bzs -f- Kz
sdr (9.27)
Q j s&s/Bz *
z “ (J'/B')8 + 1 Uy
developers of the device state that for the usual accelerations of the z
axis encountered in practice, this output is negligible.
In most applications to inertial navigation three HIG gyros are
mounted on a stable platform to detect rotation about the x, y, and z
axes. The output signal from each gyro is amplified and applied in the
proper fashion to servomotors which control the rotation of the stable
platform about the three axes. Thus, the platform is maintained at a
fixed attitude with respect to inertial space.
The torque motors mounted on the shaft extension of the inner gimbal
of the HIG gyro may be used to apply an additional precessing torque to
the gyro. This torque will produce an output signal even when the input
angle 6y is zero. This output will be amplified and applied to the servo¬
motor with the result that the stable platform tilts, or turns, at a rate pro¬
portional to the torque input. Thus the platform position can be con¬
trolled by currents applied to the torque motor. It is necessary to do
this since the gyros will tend to maintain the platform fixed with respect
to an inertial reference in space. It is, however, normally desirable that
the platform be maintained at a fixed attitude (usually horizontal) with
respect to the earth. Since the earth rotates in space a platform mounted
on a fixed point on the earth’s surface would therefore have to be rotated
at the same rate as that of the earth. In a moving vehicle the rotation of
the platform would have to be at a greater or lesser rate depending on the
direction of motion of the vehicle.
In addition to the stable platform, a complete inertial navigation
system makes use of accelerometers. In its simplest form an accelerom¬
eter consists of a spring-mass system as shown in Fig. 9.23. By Newton’s
law a force is needed to accelerate the mass. Whenever the frame of the
358 CONTROL SYSTEM COMPONENTS [Chap. 9
line drawn perpendicular to the table, makes an angle 6t with the reference
line. Suppose first that there is initially no error, i.e., 0V = dt, and let
the vehicle accelerate with a linear acceleration a. If the accelerometer has
a gain constant Ka, then it will deliver an output voltage Kaa. As a
result of the acceleration the vehicle moves along the surface, and there¬
fore the angle of the true vertical with respect to the reference changes.
Pickup
Fig. 9.24. The inertial navigation problem: (a) coordinate system; (6) single-axis
block diagram.
If R is the radius of the earth, the acceleration of the true vertical with
respect to the inertial reference is
6 = a/R (9.28)
A torque must therefore be applied to the gyro mounted on the table such
that the table angle 6t also accelerates this amount. If HIG gyros are
used, a fixed voltage applied to the torque motor results in a fixed velocity
of rotation of the table. Therefore, we must integrate the accelerometer
output once to obtain the angular velocity of the true vertical and apply
the output from the integrator to the torque motor. If the gain of the
integrator is Ki and if the ratio of table angular velocity to gyro input
360 CONTKOL SYSTEM COMPONENTS [Chap. 9
voltage is Kt,
0t = KtKifKa a dt
or
dt = RtKiKa a (9.29)
Using Eq. (9.28) gives
dt = KtKiKaRdv (9.30)
Since the object of the torquing is to keep the table horizontal we desire
that d t = dv and therefore
KtKiKaR = 1 (9.31)
of accelerations along the surface. The stable platform acts very much
like such a pendulum.
The discussion of the inertial navigation problem given here has per¬
force been quite brief. Nothing has been said about the three-axis prob¬
lem, and about the necessity of feeding corrections for earth rotation and
other factors into the system. For further information on this subj ect the
reader is referred to the literature.1
PROBLEMS
9.1. A worm-gear differential is shown in Fig. 9.25. One input is the rotation of the
shaft Xi. The second input is the linear displacement of the worm by moving the
9.3. Demonstrate that the nutating frequency of a thin disk spinning in free space
is equal to one-half the spin velocity of the disk.
9.4. Write the transfer function of a free gyro from Tz to 6y.
9.6. Write the transfer function of a rate gyro from Ty to 0Z and construct a block
diagram. *
9.6. In Fig. 9.27 is shown a simplified diagram of a possible gyroscopic accelerom¬
eter. Derive the transfer function.
9.7. In Fig. 9.28 is shown a simplified diagram of a monorail car. Derive the equa¬
tions of motion and investigate stability.
oc axis
Prime
mover
Control
input
Fig. 10.2. A valve-controlled constant-pressure hydraulic system.
valve is used which permits the hydraulic fluid to pass directly to the
sump when the valve is in the off position. Operation of the valve diverts
part of the fluid away from the sump to the actuator, and in the fully on
position all the fluid goes to the actuator. In this type of system the
supply pressure is low whenever the valve is in the off position but rises
as oil is diverted to the actuator. No pressure regulator is required
except possibly as an emergency relief valve, and since fixed-stroke
hydraulic pumps are inherently constant-flow devices, neither is any flow
regulator required.
If a fixed-stroke pump is used with a pressure regulator in the constant-
pressure system, the power delivered by the pump must be essentially
constant, since both the flow and pressure are constant. Hence this sys¬
tem is very inefficient, particularly if the duty cycle is such that relatively
large peak loads occur only occasionally. Except for the power that is
delivered by the hydraulic motor or actuator to its load, all the power
delivered by the pump is eventually converted into heat, which raises the
oil temperature. Hence fairly elaborate cooling is usually required in
this type of system to maintain the oil temperature below its permissible
maximum.
Although the constant-flow system does not suffer from this particular
disadvantage, it is not used as commonly as the constant-pressure system.
This is due to the fact that the control of the actuator is not so positive
with this type of system as with the constant-pressure type. It might
be noted incidentally that in the constant-flow system the load is not
locked rigidly when the valve is in the centered position. Depending on
the application, this may or may not be an advantage.
In this chapter we shall consider the pump-controlled hydraulic control
system and the various components employed. In the following chapter
the valve-controlled system and its components will be discussed.
10.3. The Pump-controlled Hydraulic System. The basic diagram of
the pump-controlled system is shown in Fig. 10.1. The most common
type of variable-stroke pump used in a system of this type is the piston
pump, shown in Fig. 10.3. The prime mover turns the shaft to which the
cylinder block is attached. The piston-return plate, or wobble plate,
is fixed. As the cylinder block rotates, it carries with it the pistons which
bear on the wobble plate. The fluid-valving plate is positioned in such a
manner that the cylinder volume is increasing as the piston passes the
inlet port and decreasing as the outlet port is passed. Thus the fluid that
was taken in through the inlet port is expelled through the outlet port.
The flow rate of the piston pump can be changed by changing the angle
of the wobble plate with respect to the cylinder block. This adjustment
can be made mechanically, hydraulically or electrically. The ease of this
adjustment and the fact that maximum flow rate is not affected by pro-
3C6 CONTROL SYSTEM COMPONENTS [Chap. 10
viding for this adjustment make the piston pump the most widely used
high-performance variable-stroke pump. In standard piston pumps,
considerable reaction force is experienced at the lever that controls the
wobble-plate angle. This reaction force is usually in such a direction as
to reduce the flow to zero, although it is possible by offsetting the pivot to
neutralize this force for a particular load and indeed to reverse the net
force. Owing to this reaction force, it is necessary to provide a stage of
power amplification to operate the pump stroke control if the stroke is to
be controlled electrically. This stroke control could be of the type dis¬
cussed in Sec. 11.1, for instance.
The fluid is passed from the outlet port of the variable-stroke pump
through the high-pressure line to the motor. Any of the various types of
hydraulic pumps theoretically work equally as well when operated as
motors; however, motors are almost always built with a fixed stroke.
Thus, a piston type of motor has a wobble plate fixed at a particular angle
(usually 30 to 40°). The high-pressure fluid is admitted by the valving
plate to a cylinder and forces the piston back. This can be accomplished
only by the cylinder’s turning with respect to the stationary wobble plate,
thus turning the output shaft. The cylinder which has been filled by the
fluid is then carried around to the exhaust port and the fluid expelled to
the low-pressure line by the returning piston. Other types of pumps and
motors that are discussed in a later section can also be employed in such a
system with varied degrees of success.
All hydraulic transmissions must have some sort of make-up system,
to replenish the fluid lost by leakage, and must also have an overload
relief valve, for protection in case of load seizures, etc. A typical replen¬
ishing system and relief protection are shown in Fig. 10.4. In the usual
case of bidirectional operation, either line may be the high-pressure line;
thus a relief valve should be placed in both lines. Similarly, replenishing
is usually provided in both lines to provide for the possibility of operation
in either direction for long periods of time. Care must be taken that the
Sec. 10.4] pump-controlled hydraulic systems 367
fluid picked up from the sump is not aerated. Defoaming agents are
sometimes added to the hydraulic fluid in order to reduce its compressi¬
bility and to improve hydraulic efficiency. The replenishing feature adds
a nonlinearity to the operation of the system since it places a lower limit
below which the return-line pressure may not drop.
The relief valve also contributes a nonlinearity to system operation.
The pressure in the high-pressure line is proportional to the developed
torque on the motor. Thus a relief valve sets an upper limit on the
torque that can be produced by the motor. In the case of a pure-inertia
load this sets an upper limit on the motor acceleration and, in the case of a
viscous-damping load, an upper limit on the motor velocity.
Fig. 10.4. Hydraulic transmission with replenishing valve and overload relief valve.
Fl = KLTd (10.2)
Thus the pump flow does not all go into useful work in the motor, and as
we shall see below, an additional portion of flow must be accounted for.
To show that the pressure across the motor is proportional to developed
torque, we recognize that the angular displacement of the motor shaft is
directly related to the amount of oil (qm) forced through the proper chan¬
nels in the motor. Thus
qm = Kj (10.3)
lb in 3 in.-lb
V in. 21 X F sec = TJ (10.5)
sec
Fm = KJ (10.6)
therefore Td = (10.7)
The back pressure built up by the load not only causes leakage flow in
the motor and in the pump as well, which is accounted for by the term Fl,
but also causes the fluid in the lines to compress. The compressibility
of the hydraulic fluid is partially due to the minute quantities of air
entrained in it and to the expansion of the tubing with pressure, but the
major factor is the oil itself, which is compressible, as is any solid or fluid.
Usually the compressibility is a second-order effect except when the sys¬
tem operates at very high pressures and when there is a fairly large
amount of oil in the system. Volume compressibility is essentially pro¬
portional to pressure. Thus the compressibility flow is proportional to
the rate of change of pressure
Fc = Kcsp (10.8)
and forces it into the high-pressure line. The compression volume in the
high-pressure line is
Vi Api
Agi = (10.9)
7
where q = oil volume, in.3
p = pressure, psi
V = volume of pipe, in.3
7 = bulk modulus of fluid, psi
Ap = change in pressure in line, psi
and where the subscript 1 refers to the high-pressure line.
Similarly the compression volume in the low-pressure line is
V2 Ap2
Aq2 = (10.10)
7
Aqi = —A q2 = A qc (10.11)
Then if
V
V,= V2 = ^ (10.12)
Lj
where V is the total oil volume under consideration, we may rewrite Eq.
(10.9) as
Since the two volumes are the same, and if the replenishing valve does not
open, we have also that
^ = A- = Ke in.5/lb (10.15)
Fp — Fm + Fc + F l (10.16)
Using the flow relations above and given a knowledge of the load, it is
possible to derive a transfer function from pump stroke x to output shaft
position 6.
10.5. Transfer Function of the Pump-controlled System. In order to
derive the transfer function of the pump-controlled system, it is necessary
to determine the relation for the shaft load. Let us assume for generality
that the load consists of inertia, viscous damping, and an arbitrary load
torque. The relation in terms of developed torque and shaft position
will then be
Td = (,Js2 + Bs)6 + TL (10.17)
Geor pump
OI 70
L_ Pump speed, 1,200 rpm
CD
>
Fig. 10.6. Efficiency versus operating speed for the three most common types of
pumps. {From Kaay et al.)
Fig. 10.7. Fixed-stroke vane pump with two inlet ports and two outlet ports.
vanes are the weak point in vane pumps, and the problem of chipping and
breaking has not been completely solved. However, it will be noted here
that the gradual wear of the vane edges during operation does not increase
the leakage flow as wear on the pistons would.
The reciprocating mass in the vane pump is quite small as compared
with the piston pump, and the flow of fluid is smoother. Along with
these factors the simplicity of construction contributes to the quiet,
smooth operation and the long life of the vane pump.
Vane pumps can be constructed with a variable stroke. A simplified
sketch of a variable-stroke vane pump is given in Fig. 10.8. When a
variable stroke is provided, only one inlet and outlet port may be used.
Because of the cam ring and end clearances required for the variable-
stroke pump, its volumetric efficiency will be lower than a fixed-displace¬
ment pump of the same size. In addition, the delivery volume of a
variable-displacement pump is one-half that of the fixed-stroke type.
1 Ibid.
374 CONTROL SYSTEM COMPONENTS [CHAP. 10
to cancel the hydraulic force oil the gears. The method is successful in
reducing the breakaway pressure required by the motor at starting to 10
to 20 psi, a negligible value.
10.10. The Ball Pump. The ball pump, shown in Fig. 10.12, is basi¬
cally a modification of the piston design. The balls are forced in and out
radially by the cam plate and hydraulic pressure, thus displacing fluid, in
the case of a pump, or forcing rotation, in the case of a motor. The inlet
and outlet ports are inside the ball race. The stroke may be made vari¬
able by allowing the cam-plate eccen¬
Eccentric
Block cam plate tricity to be adjusted. In motor oper¬
(.stationary)
[rotates) ation, high-pressure fluid is admitted at
the port marked in and forces the balls
radially outward. This may be accom¬
plished only by rotary motion of the
ball race. The ball pump provides a
line contact between ball and race, while
the conventional piston pump provides
'Stationary
a surface contact between piston and
Fig. 10.12. Ball pump.
cylinder. Thus the problem of leakage
due to ball wear or eccentricity is severe in the ball pump. The radial
design has the advantage over the conventional axial piston pump of
simplicity and ease of manufacture.
10.11. Hydraulic Transmission Lines. While the effects of very short
transmission lines, if they provide adequate cross-sectional area, may be
neglected, it is often necessary to consider the pressure drop in a line of
moderate length; and in long lines the effect of the dynamic flow charac-
Fig. 10.13. (a) Laminar flow with a parabolic velocity distribution; (b) turbulent flow
with approximately ellipsoidal velocity distribution.
teristics upon the transfer function of the control system must also be
considered.
Hydraulic flow may be divided into two types: first, viscous flow, or
laminar flow, and, second, turbulent flow. In laminar flow the particles
of fluid flow smoothly without jumbling. The flow may be thought of as
concentric tubes of fluid sliding on one another, after the manner of an
expanding telescope. The innermost tube has the highest velocity, while
the tube nearest the boundary is stationary. The tubes of flow do not
Sec. 10.11] PUMP-CONTROLLED HYDRAULIC SYSTEMS 377
exist in turbulent flow. The individual oil particles are constantly inter¬
mixing. The two types of flow are shown diagrammatically in Fig. 10.13.
Under low pressures and for fluids of moderate to high viscosity the flow
may be laminar, but as the pressure is increased, the flow becomes turbu¬
lent. Flow may be characterized by a dimensionless constant Nr, the
so-called Reynolds number, after Osborne Reynolds, whose classic experi¬
ments in 1883 defined the types of flow. The Reynolds number is defined
as
DV
Nr — — (10.19)
Pi - P°- = gj (10.20)
1 The concept of the Reynolds number has been found useful throughout the field
of hydrodynamics, and the parameter D is in general some characteristic dimension
of the system.
* N. M. Sverdrup, Theory of Hydraulic Flow Control, Product Eng., vol. 26, no. 4,
p. 165, 1955.
2 See any text on fluid mechanics, e.g., Dodge and Thompson, “Fluid Mechanics,’’
McGraw-Hill Book Company, Inc., New York, 1937.
378 CONTROL SYSTEM COMPONENTS [CHAP. 10
Reynolds number, R
Fig. 10.14. Friction coefficient versus Nr for pipes.
Tr(pi — Pz)DA
(10.22)
128juL
where p is the absolute viscosity of the fluid. The relationship for flow
shapes other than circular pipes may also be derived. The absolute
viscosity p in Eq. (10.22) is related to the kinematic viscosity v in the
definition of the Reynolds number by
(10.23)
Fig. 10.15. Density (p) and kinematic viscosity (v) of Univis J43 versus temperature.
where y is the bulk modulus of the oil, psi; p is the density in lbs/in.3, and g
is the acceleration of gravity. For most hydraulic fluids the velocity is in
the neighborhood of 4,200 ft/sec. Hence a 100-ft line is equivalent to one
wavelength for a signal frequency of 42 cps and would be considered long
at this frequency. It could be considered short at frequencies of less than
4 cps.
The parameters determining the behavior of a transmission line are
the surge, or characteristic impedance ZQ) and the propagation constant
a + j(3. The characteristic impedance is the complex ratio of the sinus¬
oidal voltage to current in a wave passing down the line without reflection
from the end. The propagation constant is a complex number a
where a is the attenuation constant expressed in nepers per unit length,
and (3 is the phase shift constant expressed in radians per unit length. It
is shown in standard texts on transmission line theory that the velocity of
propagation is co//5, where oo is the signal frequency in radians per second.2
For electrical lines it has been shown that the characteristic impedance
is given by
w + W? / _ I tan-i A
(uC)2 / 2 a>L
co 2LC V2
a
Rh tr£>4
The hydraulic inductance per unit length is, by direct analogy with
electric circuits, the ratio of pressure drop per unit length to the resulting
rate of change of flow. If the pipe has a cross section of A in.2 and the
fluid has a density of p lbs/in.3, then the force exerted on a unit length of
fluid by the pressure drop pi along this unit length is pi A lbs. Also, the
unit mass of the fluid is p A lbs. Hence by Newton’s law
. pA 1 dF p dF
PlA = Y A~dt = ~gTt
T = Vl = JL
h dF/dt Ag
Finally the capacitance per unit length of line is the ratio of a change of
oil volume to a change of pressure in the unit length. For a fluid having
a bulk modulus of 7 psi the change in volume [see Eq. (10.9)] is
Aq = ^
T
Therefore Ch = ~ = —
Ap 7
radians/in.
382 CONTROL SYSTEM COMPONENTS [CHAP. 10
For the special case of the dissipationless line these expressions reduce to
In these equations:
p = pressure difference across the input terminals of the transmission
line, psi.
Pl = pressure difference across the load, psi.
F = flow at the input terminals of the line, in.3/sec.
Fl = flow at the load, in.3/sec.
Zl = Pl/ql = hydraulic impedance of the load, lbs-sec/in.5
I = length in line, in.
Z0 and a + j(3 are the characteristic impedance and propagation con¬
stant of one line.
The input impedance of the line is obtained from Eqs. (10.25) and (10.26)
from the fact that there are two lines, each with a characteristic imped¬
ance ZQ, and a propagation factor a + jfi-
In practice we are usually interested in obtaining the transfer function
of a system including a transmission line. Thus, suppose that the load is
a motor of the type discussed in Sec. 10.4, and the source is a variable
stroke pump. The first step is to find the hydraulic impedance ZL pre¬
sented by the motor. This is done by assuming the motor to have inertia,
viscous friction, and possibly leakage. The flow for a given pressure may
then be obtained by the methods outlined in Sec. 10.4. The parameters
Z0, a, and /6 of the hydraulic transmission line are assumed to be known,
and thus the input impedance Zt presented by the combination of motor
and transmission line may be computed by Eq. (10.27). This is the load
impedance that the pump sees, and therefore, given the characteristics of
the pump, F and p can be computed as a function of pump stroke.
Finally, Eqs. (10.25) and (10.26) may be used to determine pl and FL, and
thus the motor output.
When the product (a + j(3)l is small it is possible to simplify Eqs.
(10.25) and (10.26) by using only the first two terms of the Taylor expan¬
sion for the hyperbolic functions. The approximate expressions for the
equation are, then,
2/ 1
p = FlZl 1 + ~Z^S°L 2 (10.28)
If the constants are such that the second and third terms in the brackets
are small compared to unity, then a solution assuming compressionless,
inertialess flow is adequate, and the transmission line has negligible effect
on the system. Even if this is not possible, it is usually permissible to
neglect the power dissipated in the lines, and to use the expressions for
Z0, a, and d given by Eq. (10.24). This simplification gives
2p
p = flz i +y Zz^Ag ccl - 2gy CO n2 (10.30)
.zla
F = Fl i +j cd - co2Z2 (10.31)
2t 2gy
Equations (10.30) and (10.31) indicate that changes in pressure and flow
will be chiefly in phase angle if the load is essentially resistive, and will be
chiefly in magnitude if the load is reactive.
The expressions obtained here for hydraulic transmission lines may be
used also with pneumatic transmission lines, provided that the magnitude
of the pressure wave along the lines is small enough so that one can reason¬
ably define a constant bulk modulus. The reader is, however, referred
384 CONTROL SYSTEM COMPONENTS [CHAP. 10
also to Sec. 12.6 for an empiric relation of a somewhat different form that
has been obtained for pneumatic transmission lines.
PROBLEMS
10.1. In a particular pump-controlled hydraulic system like that shown in Fig.
10.1 the pump is rated at 0.738 in.3 per revolution, 3,600 rpm, and 1,000 psi at full
stroke. It may be assumed that the leakage flow is 5 per cent of the load flow at
full pressure. The fluid is Univis J43 at 70°F, and the pipe is %-in. copper with the
motor located 50 ft from the pump. Derive the transfer function from the percentage
stroke to motor-shaft-position output for a motor having a displacement of 0.378
in.3 revolution, a leakage of 5 per cent of full load flow at 1,000 psi, an effective
inertia of 50 ft-lb-sec2, and a coefficient of viscous damping of 10-ft-lb-sec. Neglect
all transmission-line effects except compressibility.
10.2. What is the Reynolds number at full load in Prob. 10.1?
10.3. For a coefficient of energy loss of 0.1 what is the pressure drop in the line in
Prob. 10.1?
10.4. Establish the input impedance at the pump of the transmission line and load
in Prob. 10.1 as a function of the driving frequency. Assume the line is symmetric.
10.5. Find the transfer function of the system of Prob. 10.1 (a) considering the
transmission line effects exactly, (b) using the approximate-transmission-line equations
for short lines [Eqs. (10.28) and 10.29)], and (c) using the approximate-transmission-
line equations for dissipationless lines [Eqs. (10.30) and (10.31)].
10.6. Figure 10.17 shows a simplified diagram of a common type of hydraulic trans¬
mission. The motion of the solenoid armature is amplified by the “stroke servo”
consisting of the pilot valve, power piston, and feedback link abc. The length of
ab of the feedback link is 0.5 cm, and the length be is 4.5 cm. The output of the
power piston is applied to the stroke arm of the hydraulic pump and varies the oil
delivery to the hydraulic motor. The motor is of the fixed-displacement type; hence
its speed is directly proportional to the amount of oil flowing through it.
In order to find the significant data on the stroke servo and solenoid, the linkage
is removed from the power piston at point c, and pivot c is then connected to
a fixed point so as to act as a fulcrum for the lever abc. Under these conditions it is
found that
1. The power piston does not move if the currents in the two halves of the solenoid
are equal.
2. The pilot valve moves down 0.002 in. and the piston moves down at the rate of
0.2 in./sec when the currents flowing into the solenoid differ by 1 ma.
3. When the valve is moved down 0.002 in. very suddenly, i.e., if a step function of
0.002 in. is applied to the valve, the initial acceleration of the power piston is 1,000
in. /sec2.
4. When an alternating current of constant amplitude but varying frequency is
forced into the solenoid, a strong resonance is observed in the solenoid response at
40 cps, and at higher frequencies the response drops off as the second power of fre¬
quency. At the resonant frequency the amplitude of the xs swing is three times that
observed at very low frequencies.
The following data are available on the pump-and-motor part of the transmission:
1. Pump speed is 4,000 rpm.
2. The pump delivery is proportional to y, being zero when y is neutral (as shown
in the figure) and 0.34 in.3 per revolution when y is 0.5 in. from the neutral.
3. The oil displacement of the motor is fixed and equal to 0.38 in.3 per revolution.
4. The motor has an effective moment of inertia of 0.015 lb-in.-sec2 and a coefficient
of viscous friction of 0.1 lb-in.-sec.
5. Owing to oil leakage from the high- to the low-pressure lines, the motor speed
will decrease by 3 rpm per inch-pound of torque applied to the output shaft.
6. Total oil volume in both oil lines is 12 in.3.
7. The effective compressibility of the oil used (this includes the stretch in the tube
walls) is 4.4 X 10~6 in.2/lb.
Assume linearity throughout the analysis of this system. Also it is permissible
to assume that the reaction force of the pump on the power piston and the hydraulic
reaction force of the pilot valve are negligible. Further, the volume of oil in the
stroke servo is so small that compressibility effects in that part of the transmission
may be neglected.
Problem: Find the complete expression for the transfer function 6/is, where 6 is
the position of the output shaft of the hydraulic motor, and is is the difference in
currents in the two halves of the solenoid. Sketch the asymptotic and exact fre¬
quency-response curve, and the phase-shift curve for the transmission. Neglect
effects above frequencies of 1,000 radians/sec.
CHAPTER 11
Z = ~~ X in. (11.1)
where Kv is the constant relating motion to now and has the dimensions
of (in.3/sec)/in. Flow may be con¬
verted to velocity of the actuator by
dividing by the area A of the piston.
Thus the transfer function from valve
position Z to Y is
The transfer function through the feedback link is the lever ratio
(11.4)
where the minus sign shows the change in direction. Figure 11.2 shows
a block diagram for the complete unit. The closed-loop transfer function
from X to Y is thus
Y _ a + b Kv/As _ (a + b)/a
J 1 + (aKv/bAs]) ~ JbAjaTQV+1 (1L5)
r
■/////.
1 /yAFT
'/////
Sump
7777
y
'/////
Sump
77777
I
Y/y/yy
LI
'7/7 7777, ^
Pump To actuator Pum/^ y. To actuator
'77777a. ryyyy.
7Q 7777
'/////.
Sump Sump
77777
u
[a) ib)
Fir;. 11.3. Two types of four-way spool valves.
\ f
Valve yy.
port
1 I could be manufactured with zero lap
and no bevel, the initial knife-edge
would rapidly round off in use.
Overlap Underlap
A valve with overlap has a dead
Fig. 11.5. Overlap and underlap on
lands of valve spool.
zone equal to the amount of overlap.
This nonlinearity results in loss in
sensitivity near the center and may cause instability in control systems.
Usually, therefore, valves with a small amount of underlap are preferred
for control-system applications. The disadvantage of the underlapped
valve is the leakage that exists even when the valve is centered.
Sec. 11.3] valve-controlled hydraulic systems 389
for maximum force per ampere. This is not so in the conventional valve,
since efficiency must be sacrificed in order to obtain linearity. Thus in
the PLM mode the actuator may be made smaller and lighter.
2. In conventional valves, the flow gain is usually reduced at small sig¬
nal amplitudes. This does not occur in PLM mode operation.
3. Small-signal sticking is eliminated since in a sense the PLM mode
provides the spool with a massive
dither (see Sec. 11.6).
4. Small-opening Bernoulli forces,
discussed in Sec. 11.5, become
unimportant.
5. Valve lock, discussed in Sec.
11.6, is a less serious problem because
(o) Conventional input signal for ideal the valve is kept in motion and be¬
proportional valve
cause it is not necessary to design for
Valve
exceedingly small tolerances, as is
opening required in the linear mode. Fur¬
thermore, the emphasis in the design
0
of the torque motor is on maximum
force rather than on linearity. The
increased force tends to override the
(b) Pulse train for no input signal; the normal small-opening effects.
overage flow is zero
6. Manufacturing tolerances may
be relaxed, since the flow need not be
linear with valve stroke. Round
valve ports, for instance, are per¬
fectly satisfactory.
7. Since the torque motor for the
PLM mode can be designed effi¬
(c) Pulse train for maximum input signal
ciently, a single-stage valve becomes
Fig. 11.6. Input waveshapes for pulse
operation of hydraulic valves.
a possibility in many applications
where two-stage valves were formerly
required. The single-stage PLM valve with its PLM modulator will
usually be more economical than the precision two-stage linear valve and
its driver.
Naturally the pulsed valve has several disadvantages compared to
linear valve operation:
1. The output flow is in the form of pulses of fluid; thus the load must
be such that its low-pass filtering effect is sufficient to smooth the flow
adequately. Fortunately, many loads do provide adequate smoothing.
One case where pulsed operation would not be desirable would be where
the load was driven, through a multimesh gear train, by a rather light
hydraulic motor. t The motor could follow the high-frequency pulsing of
Sec. 11.3] valve-controlled hydraulic systems 391
the fluid and destroy the gear train by vibrating within its initial backlash
limits.
2. An original disadvantage of the PLM mode was the rather complex
electronic pulse-length modulator required. When it is realized, however,
that the magnetic amplifier is almost ideally suited to the PLM mode, this
problem is eliminated.1 The normal output waveshapes of magnetic
amplifiers are portions of the sine wave of the a-c supply frequency.
Thus the magnetic amplifier may drive the valve directly by controlling
the instant of conduction.
Tests have shown that pulse operation will also improve the perform¬
ance of high-precision valves originally designed for linear operation.
The frequency response will usually be improved, and resonances and
dips due to deadband and friction will be eliminated.
The pulsed mode of operation is by no means limited to spool valves.
It may be applied to all the common designs. Indeed, the concept of
pulse operation opens up possibilities for entirely new types of valves and
hydraulic systems.
In a sense, the pulse mode of operation removes the burden of high
precision from the mechanical design of the valves and increases the
requirements on the electric drive for the valve. Fortunately, the prob¬
lems involved in designing electronic circuits to convert a continuous
signal into the form of a pulse-length-modulated wave are relatively easy
to solve. The circuit can be a modified form of the Eccles-Jordan trigger
circuit and can be built with six or seven transistors including the power
stage, or a compact magnetic amplifier can be designed to provide the
required drive.
The basic pulse repetition rate must be at least twice2 the highest fre¬
quency that the valve must follow, and for negligible phase shift due to the
pulsing, the pulse repetition rate should be about ten times the highest
signal frequency. In practice, a factor of three to five times the highest
signal frequency is satisfactory. If the repetition rate is too high, the
valve will be unable to follow it, and if the repetition rate is too low, the
frequency response of the valve is limited.
A single-stage servo valve designed for PLM operation with a 400-cps
repetition rate provided a frequency response that was flat to above 100
cps.3 The repetition rate of 400 cps was chosen so that the magnetic-
amplifier driver could operate from a 400-cps supply. The maximum
theoretical operating frequency would be one-half this value, or 200 cps.
Actual operation was successful up to about 150 cps.
1 Ibid.
2 B. M. Oliver, J. R. Pierce, and C. E. Shannon, The Philosophy of PCM, Proc.
IRE, vol. 36, no. 11, pp. 1324-1331, 1948.
3 Gibson, Tuteur, and Mapes, op. cit.
392 CONTROL SYSTEM COMPONENTS [Chap. 11
As was noted above, the output flow of the PLM-operated valve must
be smoothed by the load. In this sense, the PLM mode of operation may
be considered to be the hydraulic equivalent of the relay amplifier (see
Chap. 3) operated in the stable oscillation mode. In fact, the static
characteristic found experimentally by Chubbuck for the pulsed valve is
identical in form to the static characteristic of the relay amplifier derived
in Chap. 3 and shown in Fig. 3.42.
Indeed analogues of on-off electronic circuits and devices will provide
the hydraulic-servo designer with many interesting possibilities for
investigation.
11.4. Flow-Pressure Relations in Orifices and Valves. In Sec. 11.1
it was assumed that the flow through the pilot valve was directly propor¬
tional to valve opening. Actually this is true only if the pressure drop
across the valve is ignored or assumed constant. This assumption is not
supported by the actual physical facts. Let us consider the actual
relationships. After defining certain general flow relations in orifices we
shall discuss more specific relations in spool-type valves.
By definition, a valve or orifice is a portion of the flow path that has a
restricted flow area and across which a large pressure drop may be
developed. It may be shown by the law of conservation of energy1 that
theoretical volumetric flow through an
orifice is
F = A (11.6)
F = CvCaA = CA (11.7)
%&£/////// I
V77777777.
V
/
Sharp bend
in pipe
Fig. 11.8. Examples of flow separation in typical control-system applications.
downstream pressure drop; for example, the fuel flow must be limited in
case of combustion failure in a rocket.1 A smooth decrease in tube
diameter, called a venturi tube, is placed in the fuel line, as shown in
Fig. 11.9. The fluid velocity must increase through the constricted area.
From the law of conservation of energy the pressure head must decrease
as a result of the increase in velocity head. If the pressure drop is suffi¬
cient to result in a pressure below the vapor pressure of the fluid, cavita¬
tion will occur, as shown in the figure. Thus the equivalent cross section
of the tube is reduced, and the flow is limited at some upper limiting flow
rate. At normal flow rates the constriction of the venturi tube is so
slight as to cause only a negligible pressure drop.
P-0 Fc^
1 I
V77^\
'<£22}
P=Ps
Fa a I
Fb-A
7F77^\
Fz,Pz
///// W'
p — 0 Fd^d
V777}
0 20 40 60 80 100 120
Stroke x os a percent of diameter D
Equation (11.12) holds in the underlap region, where |rc| < x0. Outside
of the underlap region one term of the equation falls out, and only the
first or second term remains, depending on whether x > x0 or x < — xa.
Fig. 11.12. Flow plotted against pressure for an underlapped valve operating in the
underlapped region and with constant supply pressure.
The reason for manipulating the relations into the form of Eq. (11.12)
is that characteristic curves for the valve may be calculated and employed
in an analysis, as shown in Sec. 11.7. In Fig. 11.12 is shown the charac¬
teristic calculated from Eq. (11.12) of an underlapped valve operating in
the underlapped region with a constant supply pressure. It is assumed
here that area of the valve orifices is proportional to stroke. The orifice
coefficient used was 0.7, which is a reasonable assumption. It will be
noted that the characteristic is essentially linear in this region. This is
one of the reasons that the underlapped valve is preferred to the over¬
lapped valve. •
Sec. 11.5] valve-controlled hydraulic systems 397
valve configuration studied is shown in Fig. 11.14. The flow into the
chamber through de is essentially unrestricted if the opening x is small
compared to the circumferential length of the orifice; and if the flow is
assumed to be irrotational, nonviscous, and incompressible, the solution
of the flow pattern in the chamber is the solution of Laplace's equation
for the configuration. This solution1 yields a 6 angle of 69° when the
valve is square (</> = 90°) and when there is no radial clearance. With 6
known, the force on the valve spool can be found by conservation of
momentum of the flow in the valve chamber. Since the area de is large
with respect to the orifice area, the velocity of flow through de is negligible
as compared with the velocity of flow through the orifice. Axial pres¬
sures on the ends of the chamber are
In
/ equal. Thus the flow through the
Valve
body
777777a f
\
L // /
C\ Yi///f'/////7
F ,<t>
orifice produces a change in momentum
that is balanced only by an axial piston
force. This change of momentum may
7V h Yh be expressed in terms of the fluid flow by
YW7777777777777777777TYA7
Spool iI X
The minus sign arises from the fact that the expression is for the force by
the fluid on the container; this is the negative of the force required to
accelerate the fluid. By Bernoulli's equation the velocity v is given by
where p is the pressure drop across the orifice. Also, the flow rate F may
be expressed in terms of p by Eq. (11.7). Hence the axial force becomes
cose
\ P 9 \ P
= —2CAp cos 0 (11.14)
$T = 0A3pv
= - p- Fv cos e + p- L ^ (11.15)
g g ai
Fig. 11.15. Hydraulic-reaction force curves: (1) Ideal case, clearance = 0, radius = 0;
(2) clearance = 0.0001 in., radius =0; (3) clearance = 0.0005 in., radius = 0.0003
in.; (4) clearance = 0.0003 in., radius = 0. {From Lee, Blackburn)
component of the force. The second term is the transient force. For the
situation shown in Fig. 11.14 the transient force is negative, i.e., in a
direction to close the valve, when the valve is being opened and the flow
is increased. This is in accordance with the qualitative discussion of the
transient-force phenomena given above.
It is again convenient to convert Eq. (11.15) to a form such that the
force is a function of pressure, rather than flow. Using Eq. (11.7) we
obtain
The approximation is often made that the pressure drop across the valve
is constant. If this assumption is made, dp/dt = 0, and the last term in
(11.16) above vanishes. Equation (11.16) can then be written in the sim¬
ple form
Supply Supply
Fig. 11.16. Damping in hydraulic valves: (a) valve with positive damping; (6) valve
with negative damping.
Since the steady-state force acts to close the valve the equivalent Ber¬
noulli spring aids the centering springs usually employed with valves.
The damping term may, however, be positive or negative depending on
the sign of L. Usually a valve consists of several orifices in series, and the
Us for different orifices usually have opposite signs. Thus, consider the
valve shown in Fig. 11.16a. In this valve L2 is negative and Li is positive.
Orifice a produces negative damping proportional to Li and orifice d pro¬
duces positive damping proportional to L2. The net damping in the valve
is therefore positive, since L2 is greater than L\. In Fig. 11.165 is shown
a valve for which L2 is less than Lh and in this valve the damping from the
hydraulic reaction force subtracts from the viscous damping. Since the
hydraulic reaction force is a function of the pressure, it is therefore possi¬
ble for the valve of Fig. 11.166 to exhibit negative damping for sufficiently
high pressures. This may, of course, result in instability and violent
oscillation of the valve.
402 CONTROL SYSTEM COMPONENTS [Chap. 11
Valves have been designed that counteract the effect of the steady-state
axial force developed by the orifice.1 Figure 11.17 shows the configura¬
tion of a force-compensated valve and a detail of one of the ports. In this
device the inlet line to the chamber is constricted. The flow, entering at
an angle 6i, tends to close the valve as usual. The chamber of this valve
is shaped like a turbine bucket, and the flow leaving at an angle 02 is in
such a direction as to generate a reaction force that tends to open the
valve. In addition, the circulating flow entering the chamber at 03 also
aids in keeping the valve open. While the design of this valve is essen¬
tially empirical, Lee and Blackburn report a reduction in the steady-state
axial force by a factor of 50 compared with a conventional valve, and they
state that the redesigned valve is essentially perfectly compensated
throughout its operating range. The small notch shown dotted in Fig.
In Out
(A)
Fig. 11.17. Force-compensated valve. {From Lee and Blackburn)
11.17a simplifies the manufacture of the spool and does not appreciably
affect the compensation.
In addition to special shaping of the valve lands, there are several other
solutions to the problem of reducing axial forces on the valve spool. A
two-stage valve may be designed to solve this problem, or an entirely
different operating principle may be employed. Both of these possi¬
bilities are discussed below.
11.6. Radial Hydraulic Forces in Spool-type Valves. Radial or lateral
forces on valve spools, if not symmetrical around the periphery of the
spool, can cause the spool to move radially and lock or freeze against the
cylinder wall. This problem of hydraulic lock has become more acute
with the high pressures and low clearances of modern hydraulic systems.
Hydraulic lock may arise from several sources. One possible cause of
hydraulic lock is the radial component of the Bernoulli force discussed in
the previous section. In order to eliminate this possibility, the valve
ports must be made symmetric around the valve periphery. A second
possibility is dirt or metal chips that lodge in the radial clearance between
piston and cylinder, thus wedging the piston. This possibility can be
minimized by filtering the oil in the system.
1 Ibid.
Sec. 11.6] VALVE-CONTROLLED HYDRAULIC SYSTEMS 403
Fig. 11.18. Perfect right-circular cylindrical piston and cylinder, eccentric to one
another.
the leakage path if we assume that the flow is laminar and that the veloc¬
ity initially is zero. Thus even for an eccentric right-circular cylinder the
radial pressures on the piston balance around the periphery of the valve.
The question naturally arises as to the forces involved if this perfect
piston is somehow canted in the perfect cylinder, as shown in Fig. 11.19.
If we consider two leakage paths l\ and U on opposite sides of the piston,
the pressure distribution along the leakage paths will be as shown in Fig.
11.196 as a result of the interrelation of pressure head and velocity head.
Thus throughout the length of the leakage path there will be a net pressure
Pressure difference
tending to allign
piston axially
difference tending to drive the piston to the right in Fig. 11.19a or, in
general, toward the side with the constricted upstream flow. When the
piston arrives at the cylinder wall, it will align itself with the wall, and the
pressure drops along any leakage path will become equal. There will
thus be no net force holding the piston against the cylinder, and hydraulic
lock will not occur.
404 CONTROL SYSTEM COMPONENTS [Chap. 11
A somewhat more practical case than that of the perfect piston and
cylinder is the case of a tapered piston in a perfect cylinder. Figure 11.20
shows an eccentric tapered piston. If the clearance is small, the leakage
flow may be considered two-dimensional, and the relationship for laminar
flow between two flat plates may be employed with little error to find the
relationship between leakage flow and pressure.1
1 dp y2
(11.18)
3/x dx 4
where y is the distance between the plates. For an element of unit width
the volumetric flow will thus be
Fx =_1._dpy^
(11.19)
z 3^t dx 4
where 2 is the width.
For a linear taper the relationship between y and x along a leakage path is
y = c -f- kx (11.20)
where c is the clearance between piston and wall, as shown in Fig. 11.20,
and k is the taper of the piston. Solving for p by integration,
6 nFxk
V = z(c + kx)2 + Ci (11.21)
6 \xFxk 1
V = (11.22)
z (c + kx)2
Thus pressure is a parabolic function of x and varies with the taper k and
clearance c. A smaller c results in a lower pressure at any station along
1 I). C. Sweeney, Preliminary Investigation of Hydraulic Lock, Engineering, vol.
172, pp. 513-516, 580-582, 1951.
2 Dodge and Thompson, op. cit.
Sec. 11.7] VALVE-CONTROLLED HYDRAULIC SYSTEMS 405
the leakage path, as shown in Fig. 11.21. Thus there is a net radial pres¬
sure along the path in such a direction as to reduce the clearance. The
result is hydraulic lock.
It may be seen from Eq. (11.22) that, if the taper k is negative, the net
force will be in such a direction as to center the piston. For this reason,
since some tolerance must necessarily be accepted in any manufacturing
process, pistons are sometimes deliberately manufactured with a slight
negative taper in order to eliminate the possibility of a positive-tapered
piston being produced from a nominally zero-taper design. Quite often,
however, this solution to the problem of hydraulic lock results in excessive
leakage flow.
It has been found by Sweeny1 and others that, if several radial grooves
are cut into the surface of the piston land, to eliminate the pressure differ¬
ential around the spool, hydraulic lock can
be eliminated. The size of the grooving
should be large with respect to the clear¬
ances involved. Both the depth and
width of the grooves should be at least ten
times larger than the clearance if the clear¬
ance is of the order of a few ten-thou¬
Fig. 11.21. Pressure drops in leak¬
sandths of an inch, in order to permit free
age paths of tapered piston.
flow around the periphery. Sweeny found
that five or six of these grooves reduce the locking force to about 1 per
cent of its previous value in a typical spool.
Forces due to factors other than the pressure gradient caused by leak¬
age flow are also sometimes important contributors to hydraulic lock.
These factors include static friction, collection of dirt, and metal-to-metal
contact. Sweeny has found that these forces including hydrodynamic
lock itself build up rather slowly, taking 4 to 5 minutes to reach maximum
in some cases. A common solution to the static sticking problem is to
agitate the pilot valve at some high frequency either mechanically or by
superimposing an a-c component on the valve-driving signal. This
agitation is called dither and must be at a frequency high enough to
assure that the controlled elements are unable to follow the rapid motion
of the pilot valve.
11.7. Graphical Analysis of Control Valves. For large input signals
the nonlinearity of valve characteristics cannot be ignored. Analytical
methods of treating the problem are possible, but graphical techniques
appear more convenient. The similarity between the characteristic
curves has led to the extension of vacuum-tube analysis techniques to
hydraulic valves. In this section we shall consider several of these tech¬
niques in detail. It should be noted that these nonlinear analyses do not
1 Sweeny, op. cit.
406 CONTROL SYSTEM COMPONENTS [Chap. 11
permit the evaluation of equivalent gains or time constants for the ele¬
ments; thus the approximations considered in the following section, which
do permit this, are more often employed than these more exact methods.
It is possible to construct load lines on the valve characteristics that
relate flow and pressure, such as those shown in Sec. 11.4. A load line is
the pressure-flow relation for the load that is driven by the valve. Since
the valve and the load are in series, the same flow must exist in each, and
their two pressure drops must sum to the supply pressure. Since the load
line and the valve characteristic represent relations that must be simulta¬
neously satisfied, the point of operation must occur where the two curves
Fig. 11.22. A load line for a resistive load on an ideal valve characteristic. (Cunning¬
ham])
intersect. The load can be assumed to be so close to the valve that trans¬
mission-line effects can be ignored, or the effects of the line may be
included in the load. Two load lines of this type are illustrated in Fig.
11.22. The valve characteristic curves on which these load lines are
drawn are those of an ideal valve which has zero lap and square ports.
The relation between pressure across the load {p£) and flow through it is
assumed linear for the solid line. The load line is constructed relative
to the assumed supply pressure (ps). The actual operating point of the
system must lie on the load line at a point determined by the valve
opening (x). If x = 20, the operating point Q will be as shown in the fig¬
ure. Quite often the load presented to the valve has a nonlinear relation
between pressure and flow. We might refer to this as a nonlinear
hydraulic resistance. For instance, in Fig. 11.22 is shown dotted a
parabolic load1 line such as would be presented by an orifice. If the
Sec. 11.7] VALVE-CONTROLLED HYDRAULIC SYSTEMS 407
parabola passes through the operating point Q with the same slope as
the linear resistance, it will have an equation
(JlA F (o'!i
\in.2/ \min/
finA J_ (lit L hP
\gal / 12 \in./ 33,000 ft/min
(11.24)
where pressure is in pounds per square inch and flow is in gallons per min¬
ute. For a given supply pressure, curves of constant load horsepower
may be calculated from Eq. (11.25). These curves are hyperbolas and
are shown plotted for ps at 1,400 psi in Fig. 11.23. For a given value of
x we can find the maximum power that can be delivered to the load and
the pressure at which this occurs. From Eq. (11.12) we may retain the
first term for a zero-lap valve and write for a constant x that
pv = kQF2 (11.26)
2 ps
Vl = (11.28)
3
dF
ps = Pv + RlF + Jl (11.29)
ps — RlF, — pv = ^Rl + AF
or AF = V* ~ R'~ f v- (11.30)
1 Cunningham, op. cit., from which this example is adapted. Preisman, “ Graphical
Constructions for Vacuum Tube Circuits,” McGraw-Hill Book Company, Inc., New
York, 1943.
Sec. 11.7] valve-controlled hydraulic systems 409
Fig. 11.24. Construction for valve supplying a load with inertia. (From Cunningham)
curve, and a new line at the same slope locates the flow at the second
interval. The major approximation implied here is that the rate of
change of flow is constant throughout the time interval chosen. Accuracy
is improved by decreasing the time interval, but of course this requires
more computation. With our choice of 0.001 sec for the time interval,
about 12 construction steps are required to arrive at the operating point
at the intersection of the load line and the characteristic curve. In Fig.
11.25 is shown this response plotted against time. The response for a
linear approximation for the valve characteristic is shown dotted in the
same figure for comparison.
1 This corresponds to a hydraulic motor with a displacement of 0.38 in.3 per revolu¬
tion with a mechanical load consisting of a resistance of 0.0174 ft-lb-sec and an
inertia of 2.5 X 10-4 slug-in.3 per revolution.
410 CONTROL SYSTEM COMPONENTS [Chap. 11
dF
ps — (Rv + Rl)F + J (11.31)
dt
which has a solution
Ps f Rv + Rl A
F = - 1 — exp
xp (- -J t) (11.32)
Rv + Rl
and Pl = ps — FRV (11.33)
Fig. 11.25. Response of valve and load with inertia to a step input. (From Cunning¬
ham)
same for the linear approximation and the more accurate representation.
The value of Rv is 482 psi/gpm under this assumption. We see from the
figure that the flow increases more rapidly for the nonlinear construction,
thus implying that the effective resistance of the valve orifice is less at
small openings than would be assumed in the linear analysis.
It is possible to extend this construction technique to more involved
inputs. For example, consider a sinusoidal driving signal in x. Let
Since the values of x will be both positive and negative, we must employ
a composite characteristic curve. This second family of static curves is
plotted below the horizontal axis with the pv scale reversed and the ps
points coinciding.
Let us suppose that it is desired to calculate the values of flow at 10°
intervals in the sinusoidal variation; thus
Fig. 11.26. Construction for inertial load and sine-wave driving function. (From
Cunningham)
C7»
^ kT
g 600
E
h 2
40 400
1
20 200
0 0 0
-20 -200
-1
-40 -400
-2
-600
Fig. 11.27. Time response of a system with inertia for a sinusoidal driving function.
{From Cunningham)
conditions than a linear approximation, they are rather long and compli¬
cated, and they do not permit extension to general inputs. Under condi¬
tions of small driving signals it is possible to make good linear approxi¬
mations with resultant simplification of the analysis.
The first linear approximation is suggested by an examination of Fig.
11.12, the flow-pressure relations for an underlapped valve in the under¬
lapped region. The variation is almost linear in this region. The sys¬
tem is analogous to an electric bridge circuit, as shown in Fig. 11.28.
R = jU (11.36)
X leak
where ps is the supply pressure and Fleak is the total leakage flow with the
piston centered. The reader may show that
_xps
(11.37)
RRl + R2 — x1
Thus for small values of x we have the linear transfer function
FL _ Ps
(11.38)
x (RRl H- R2)
vacuum tubes are also basically nonlinear devices, since the plate current
as a function of plate or grid voltage is given by Child’s law:
Fi = F2 = Fl
(11.40)
P1 — P‘2 = PL
Pi + P2 = Ps (11.41)
Po = 0 (11.42)
where FL is the load flow and Pl is the pressure across the load, and where
other symbols are defined as in Fig. 11.11. Equation (11.41) shows that,
if pi goes up a certain amount, p2 goes down by the same amount.
The procedure is based on the assumption that, for small variations of
the variables about an operating point, the characteristics are essentially
linear. Thus, consider the nonlinear function between flow, pressure, and
valve displacement:
Fl = FL(pL,x) (11.43)
Differentiation gives
0F t dF t
dFL = —F dx + —F dpL (11.44)
dx dpL
Similarly the reader may compute Y and for given parameters establish
the transfer function for the valve. Equation (11.50) holds only in the
underlap region, and when the relation for the valve outside the underlap
region is found, it will be noted that the gain drops by a factor of 2. For
large-signal operation the high-gain region about center may sometimes be
ignored.
Once having accepted the restrictions implied in the small-signal
analysis, the operation of the valve with reactive loads of all sorts may be
considered. As a matter of fact, we may go further and consider the
effect of the mass and friction in the valve stem itself. To do this, we
assume that the input to the valve is a force source of finite stiffness. In
order to get a complete expression of the various reactions taking place
in the valve, the hydraulic reaction force must be considered. This force
has been discussed in Sec. 11.5 and was shown to be a function of pressure,
displacement, rate of change of pressure, and rate of change of displace¬
ment [see Eq. (11.16)]. Thus using the symbol for reaction force, we
can say
$h = $(pl,Pl,x,x) (11.51)
d5 . d& . . d& . d$ .
3* = Vl + ^Vl+^x + ~x
dpL^ ' dpL
— Ai Pl + X2 Pl + KhX + B (11.52)
The parameters Ai, X2, Kh, and Bh are considered constant for small varia¬
tions of the variables, and may be obtained by differentiation of Eq.
(11.16). If measured characteristics relating the reaction force to pres¬
sure and displacement are available, the parameters Ai and Kh may be
obtained from them. This procedure has the advantage that the actual
force characteristics rather than characteristics for an ideal valve are
used, but the parameters X2 and Bh cannot very well be obtained this way.
A compromise solution might therefore be used by which Ai and Kh are
determined from experimental curves and X2 and Bh from the ideal
relation.
416 CONTROL SYSTEM COMPONENTS [Chap. 11
PL = Gx - YfL (11.54)
Vl = ZlFl (11.55)
Displacement-
force feedback
1 l/r + zL
Ms2 + B's + K'
(11.56)
1 | G(\i + X2 s) Zl/T
“h Ms2 + B's + K' ZL + l/Y
SiG
(11.57)
(.Ms2 + B's + K')( 1 + YZl) + GZL(\, + X2s)
where B' = B + Bh, and K' = K + Kh. It was shown in Sec. 11.5 that
Bh is negative for valves for which the distance between the load outlets
and the sump orifices is less than the distance between load outlets and
pressure orifices (see Fig. 11.16). Therefore B' may be a negative quan¬
tity and the valve may be unstable and oscillate no matter what load is
connected to it. However, if the load outlets are midway between the
pressure and sump orifices, i.e., if Li = L2 in Fig. 11.16, Bh and X2 will be
zero. Under these conditions the valve may still be unstable if the load
contains an inertia component so that ZL — R + Ls. With this type of
load the denominator of Eq. (11.57) will be a cubic polynomial of s. In
order to determine whether or not the valve is stable the standard Routh
test1 may be applied to the denominator. For the case of ZL = R + Ls,
this test requires that for stability
The inequality shows that if Xi = 0 and B' > 0, the valve is stable. On
the other hand, if Xi <0, and B' = 0, the valve is always unstable.
Since the parameter Xi is negative in a valve in which the hydraulic reac¬
tion force is not compensated it is apparent that a certain amount of posi¬
tive damping (B' > 0) is required for stability.
We have seen here how a linear analysis can reveal valve instability.
By examining Eq. (11.58) the designer may adjust the various parameters
of the system to eliminate this possibility. The linear analysis has the
advantage not only of relative simplicity but also of revealing the approxi¬
mate transfer function of the unit. The accuracy of the analysis is
reduced, however, if large input signals destroy the assumed linear rela¬
tions about the quiescent operating point.
In this connection it should be noted that certain types of nonlinearities
will cause oscillations in spool-type pilot valves under conditions which
would allow the valve to be perfectly stable if the system were linear. To
take one example, there is evidence to indicate that static friction is
1 Chestnut and Mayer, “Servomechanisms and Regulating Systems Design,”
John Wiley & Sons, Inc., New York, 1951, Vol. I.
418 CONTROL SYSTEM COMPONENTS [Chap. 11
When the distance Y between the flapper and the orifice is very small the
flow at the flapper is controlled approximately by the cylindrical area
formed by extending the inner surface of the orifice pipe to the flapper.
This area is tD2Y, where D2 is the orifice diameter. Therefore, for very
small Y we have approximately
Fig. 11.31. Chamber pressure versus flapper travel, normalized. (From Nightingale)
Also, suppose that the piston is blocked so that the flow to the piston is
zero. Under these conditions F2 = F1 and therefore for F/7)2 < 0.25 we
have
(tt/4)C7)i2 \/pi — P2 = tCD2Y yjpi
P2 _ (Di/D2)4
or (11.62)
Pi 16(F/7)2)2 + (Di/7) 2)4
P2 = (7)i/7)2)4
(11.63)
Pi 1 H~ (7) 1/7)2)4
Force = ^ (11.64)
g
where F is the volumetric flow, v is the velocity of discharge, p is the den¬
sity of the oil, and g the acceleration of gravity. The flow and discharge
velocity can be found by the relations given above. This force can be
considerably reduced by arranging two nozzles, opposing one another on
either side of the flapper, as in the Moog valve discussed below.
The major limitation on the flapper valve is that the controlled-pressure
chamber must be kept rather small if the time constant of the valve,
which consists of the resistance of the upstream orifice and the capacity
of the controlled-pressure chamber, is to be kept small. Enlarging the
flapper orifice will improve the response of the valve, but at the expense
of increased flapper reaction force and leakage flow. These restrictions
place a limitation on the load that can be controlled by a flapper valve,
and as a result the flapper valve is most often employed as the first stage
of a two-stage valve.
11.10. Nozzle Valves. The Askania nozzle valve is a valve that has
a very low hydraulic reaction force.1 A schematic of the valve is shown
High in Fig. 11.32. The hydraulic flow is formed
pressure-^ into a jet by the movable nozzle, and the flow
Pivot is directed into one of two ports on either side
,—Jet pipe of the receiving block and conducted to the
'A load. The only force that must be supplied
'Input
r
f L
Centering \
<- by the input is that required to overcome fric¬
spring tion at the pivot and the inertia of the jet pipe.
Receiving
block
The major disadvantage of this device is its
VMsActuator
relatively high leakage. The frequency re¬
sponse of this device with the standard 8-in.-
Fig. 11.32. Askania jet-pipe
long, 1%2"in.-diameter jet pipe peaks at about
valve.
30 cps and has a damping ratio (f) of 0.2.
11.11. Slide Valves. Owing to the high cost of manufacturing spool-
type valves to the very close tolerances required by high-performance
systems and owing also to the high leakage of the flapper-type valves,
1 S. Z. Dushkes and S. L. Cahn, Analysis of Some Hydraulic Components Used in
Regulators and Servomechanisms, Trans. ASME, vol. 74, p. 595, 1952.
SEC. 11.11] VALVE-CONTROLLED HYDRAULIC SYSTEMS 421
efforts have been made to develop a valve with neither of these disadvan¬
tages. The slide valve1 shown in Fig. 11.33 is one possible answer. The
advantage of the slide valve is that the slide and block can be clamped
together, so that the two critical holes may be drilled through the block
and slide at the same time. The spacing between the holes is not critical.
Following this operation, the plugs and bushings are inserted to complete
the critical portion of the device. Exact alignment is thus assured for
all the orifices. The vertical spacing between slide and block may be
established by a milling operation. The slide is suspended by bars and
actuated by a torque motor of some sort, or the slide may be allowed to
rotate with respect to the block. In either case, the variation of area of
the orifices is essentially linear with respect to motion of the slide. The
slide valve can be force-compensated in the same manner as spool-type
Top
Side
valves, or the valve with rotary motion for the slide plate may be designed
so that the Bernoulli force of one nozzle cancels that of another.2 In fact,
in the configuration shown in Fig. 11.33, a component of hydraulic force
acts to open the slide if the slide is displaced from center. Unfortunately,
however, this opening force increases as the valve slide is opened, while
the Bernoulli force decreases. Thus these two components cannot be
relied upon to cancel each other. Figure 11.34 gives a plot of flow versus
displacement of the slider of a rotary-plate valve. Note the linear char¬
acteristic, and note that the leakage flow is only about 1.5 per cent of full
output flow, whereas the usual flapper-valve leakage flow is about 5 to
10 per cent of full-load flow. A practical disadvantage of the slide valve
is the scoring of the bushings and block by chips and dirt that have a
tendency to accumulate in the slide cavities.
Fig. 11.34. Flow versus displacement for a rotary slide valve. (From Lee)
2.0
1.0
0.8
■S 0.6
2 0.5
0.2
0.1
10 20 30 40 60 80100 200 300 400 600 1,000
Cycles per second
Fig. 11.36. Frequency response of actuator and slide valve. (Lee and Shearer)
In Fig. 11.386 the input varies the position of a movable sleeve in the
pilot valve. If the input is moved down, for instance, the high-pressure
oil is ported to the bottom of the actuator cylinder. The actuator moves
up, carrying the pilot-valve piston down by means of the feedback link,
thus stopping the flow of oil.
Figure 11.38c shows a Vickers two-land sleeve valve that operates in
a manner somewhat similar to the action of Fig. 11.386; however, the
principle of differential area is employed in the system. In Fig. 11.38c
the areas are in the ratio of 2:1, which is the most common arrangement.
If the input is moved down, the valve spool ports high pressure to the bot¬
tom of the actuator cylinder, and the top of the actuator cylinder is
opened to sump. The actuator will move up, and the sleeve will be
Input
pulled down by the feedback link, thus closing off the fluid flow. The
actuator follows the input with a reversal of direction. Now, if the input
is moved up, a rather different effect occurs. The high-pressure fluid is
ported to both the top and bottom of the actuator cylinder. The net
force is down.
It is also possible to construct very compact two-stage valves without
constructing the parts separately and connecting them by linkages.
Figure 11.39 shows a two-stage slide valve used in a hydraulic autopilot
control manufactured by Siemens (Germany) during World War II.
If the force developed in the left-hand chamber, p4A4, is equal to the force
developed in the right-hand chamber, p3A3, the spool is at rest. If the
input is moved to the right, for instance, the orifice is reduced in size
and p4 increases, thus increasing the force in that chamber and moving the
spool to the right. The new position of the spool will be such that the
chamber forces are once more balanced. The motion of the spool
ports oil to the main actuator. While this valve has been described1
1 R. Hadekel, Hydraulic and Pneumatic Servos, Automation, March, 1955.
Sec. 11.12] VALVE-CONTROLLED HYDRAULIC SYSTEMS 425
operate at 1,000 to 3,000 psi. The flapper-spool valve. More practical two-
stage valves of this type have been
arrangement of balanced nozzles at
developed.
the flapper reduces the force re¬
quired by the flapper, and the us of two chambers and two springs
eliminates the direct reliance on the value of the supply pressure
1 R. L. Scrafford, “ Hydraulic Servos Incorporating a High-speed Hydraulic-
amplifier Actuated Valve,” Trans. AIEE, part II, vol. 72, p. 175, 1953.
426 CONTROL SYSTEM COMPONENTS [Chap. 11
(a) Schematic
(b)
Fig. 11.41. The Moog valve.
larger outputs are also available. The leakage flow through the nozzles
at zero signal is typically about 2 per cent of full rated flow for the 8-gpm
model. This leakage flow cannot be reduced proportionally for smaller-
flow models. From test characteristics furnished by Moog, it would
appear that f, the damping ratio, is about 0.7, and the critical or
natural frequency of resonance is about 40 cps for the Model 500. Thus
Sec. 11.12] valve-controlled hydraulic systems 427
empirically
F 1 _ 1
ii - i2 _ (M/K)s2 + (B/K)s + 1 “ (s2/6.3 X 104) + (s/28.3) + 1
(11.65)
Fig. 11.42. a diagram for Moog Mode] 500 Fig. 11.43. Equivalent circuit of a
valve. (Courtesy Moog Valve Co.) portion of the Moog valve, where x is
flapper position.
resonant frequency of the second-stage spool is about 1,000 cps for the
typical Moog valve; thus it is well above the frequencies of interest. The
valve may be thought of as the equivalent of a bridge circuit, as shown in
Fig. 11.43. This circuit could not of itself cause the underdamped
response shown in Fig. 11.42. The flapper is flexible, and the interaction
between the nozzle flow and flapper position is perhaps responsible.
The Cadillac Model FC2 flow-control valve, as shown in Fig. 11.44, is
a two-stage valve similar to the Moog valve. Unlike the Moog valve,
however, the piston of the Cadillac valve is not directly restrained by
428 CONTROL SYSTEM COMPONENTS [Chap. 11
Fig. 11.45. Equivalent block diagram of the Cadillac Model FC2 valve.
negligible in this device. The time constant of the Cadillac FC2 valve is
determined by the gain of the feedback loop. Figure 11.45 gives a block-
diagram representation of the FC2 valve. The block diagram shown is
only one of several possible block diagrams that could be drawn. The
choice of force feedback is arbitrary. The flapper spring-mass system
has a natural or critical frequency that is
quite high and may be neglected. Like¬
wise the transfer function from piston
position to feedback spring force may be
considered a constant. Thus the trans¬
fer function of the closed loop from input
to output is
y _ m __fcife2fc8__
x 1 + fiB as2 + bs + c + kik2kzki
(11.66)
where n is the forward transfer function
and B is the transfer function through
the feedback path of the loop. The
Fig. 11.46. Frequency response of
Cadillac FC2 valve. (Courtesy
transfer function y/x thus has a simple
Cadillac Valve Co.) quadratic denominator, assuming that
the time constants of the several spring-
mass systems are very small. In the FC2 valve the loop-gain constant
/ci/c2/c3/c4 is about 314, from data furnished by the manufacturer, and
the response of the system is as shown in Fig. 11.46.
The frequency responses of the Moog and Cadillac valves are similar
with the exception that the FC2 valve amplitude response falls off more
gradually than the typical Moog characteristic. However, the damping
Sec. 11.13] VALVE-CONTROLLED HYDRAULIC SYSTEMS 429
factor of both valves may be adjusted within a limited range by changing
internal dimensions.
In all the hydraulic valves so far discussed, including the flapper valves,
the transfer is from the input variable to flow. Constant flow will cause
a constant velocity in the load element. If the over-all control system
controls the position of this element, the loop contains one integration.
Occasionally, an extremely high performance control system is required,
and the designer must provide two integrations within the loop to meet
dynamic accuracy specifications. Both the Moog and Cadillac valves
can be adapted to provide two integrations. In the Moog valve the cen¬
tering spring may be removed, while in the Cadillac valve the feedback
spring and lever may be removed. The force set up by the differential
pressure in the chamber will then move the piston at a constant velocity,
and the oil flow will constantly increase. Naturally, such a device must
be part of a feedback loop to operate properly, and the loop must be
properly equalized or compensated if it is to be stable.
In certain applications both the Moog Model 500 valve and the Cadillac
FC2 valve encounter problems with magnetic dirt building up around the
poles of the torque motor and hindering proper operation. Both con¬
cerns have developed dry torque-motor designs to overcome this difficulty.
11.13. Pressure-regulating Devices. Several different methods of
obtaining constant-pressure hydraulic supplies are used, depending on
the type of pump and the accuracy with which the pressure must be
controlled.
One type of pressure control is shown in simplified form in Fig. 11.47.
It consists of a spring-loaded valve connected from the high-pressure line
to sump. This type of pressure control is used with fixed-stroke pumps.
The flow from the pump is essentially constant at full flow, with the valve
porting to sump any oil in excess of that required by the load. There is
a considerable waste of flow and power in the regulating valve unless the
system is always operating at close to full load. The wasted power
results in a temperature rise in the hydraulic fluid and may require
430 CONTROL SYSTEM COMPONENTS [CHAP. 11
io) Primitive
Low
pressure
High
pressure
Pressure
adjust ■
lb) Practical
Fig. 11.48. Variable-stroke pump with pressure-controlled stroke for constant-pressure
service.
PROBLEMS
11.1. Show from Eq. (11.5) the result of making a in Fig. 11.1 equal zero.
11.2. In Fig. 11.49 is shown a typical spring-centered spool valve. Find the trans¬
fer function x0/$l for operation about the point xv = 0, pl = 0 for the hydraulic-
1 E. M. Greer, Hydraulic Accumulators, Machine Design, vol. 26, no. 1, p. 132, 1954.
* C. Cooke, Optimum Pressure for a Hydraulic System, Product Eng., vol. 27, no.
5, p. 162, 1956.
2 Ibid.
432 CONTROL SYSTEM COMPONENTS [Chap. 11
&
~\xv y
I
Y
Sump- D
High
pressure b
—b
Sump■
-p
Spring
7777,
Fig. 11.49
in. from the neutral position will just close one set of orifices. Assume that the ori¬
fices do not leak at all when they are closed and produce no steady-state hydraulic
reaction force. Also assume that transient hydraulic forces are negligible.
Each orifice is characterized by the two equations
where C, k = constants
g = acceleration of gravity
p = density of oil = 0.03 lb/in.3
Xi = amount of opening of fth orifice, in.
Fi = flow through ith orifice, in.3/sec
Pi = pressure drop across fth orifice, psi
$i = force of hydraulic reaction produced at ith orifice, lb
The following data are given:
Mass of valve piston =0.1 lb
Valve spring constant = 200 lb/in.
Valve coefficient of viscous friction = 0.04 lb-sec/in.
Mass of actuator piston = 1.5 lb
Area of actuator piston = 1 in.2
Friction of actuator piston = 1 lb-sec /in.
When xv = 0.005 in., when supply pressure = 2,000 psi, and when there is no load on
actuator, the piston moves at a rate of 5 in./sec and the total hydraulic reaction force
is 0.9 lb.
11.3. Assume that at each of the valving orifices shown in Fig. 11.49 the relation
between pressure flow and displacement is
Problems VALVE-CONTROLLED HYDRAULIC SYSTEMS 433
Ti = KFi Vpi
All the forces 5^ are in such a direction as to tend to close the valve.
The following data are given:
Mass of valve piston = 0.1 lb
Valve spring constant = 200 lb/in.
Valve coefficient of viscous friction = 0.04 lb-sec/in.
Mass of actuator piston = 1.5 lb
Area of actuator piston = 1 in.2
Friction of actuator piston = 1 lb-sec/in.
When xv = 0.005 in., when supply pressure = 2,000 psi, and when there is no load on
actuator, the piston moves at a rate of 5 in./sec and the total hydraulic reaction force
is 0.9 lb. When xv = 0, when supply pressure = 2,000 psi, and when a force of 1,000
lb is applied to the piston, the piston moves at a rate of 0.01 in./sec.
Find the transfer function x0/%i for operation about the point xv = 0, pl = 0.
11.4. Find the axial force developed at one port of a spool valve with a piston
diameter of 34 in- and a rectangular port of 0.1-radian width. Assume the stroke is
346 in- and the radial clearance is 0.002 in. Assume an orifice coefficient of 0.6 and an
operating pressure of 3,000 psi.
11.5. The flow relation of a hydraulic valve is given by
F = 25x s/pl
where F = flow, in.3/sec
x = valve-spool displacement, in.
pv = total pressure drop across valve, psi
The valve is connected to a constant-pressure supply of 1,500 psi. The sump pressure
is zero. A hydraulic motor with the following characteristics is connected to the
servo ports of the valve:
a. Find the response of motor speed versus time when x is given a step displacement
from zero to 0.01 in.
b. Find the response of motor speed versus time when x varies sinusoidally at a fre¬
quency of 10 cps with an amplitude of 0.01 in.
11.6. Calculate the orifice diameters for a flapper valve to operate at 4,000 psi.
Allow a maximum flow of 0.05 gpm through the nozzle. Calculate the reaction force
on the flapper when it is in the center of its operating range. The density of the oil is
0.03 lb/in.3.
11.7. Assume that the nozzle in Fig. 11.32 has been imperfectly made and that the
jet emerges at an angle of 10° from the center line of the jet pipe at 1,500-psi supply
pressure and an orifice diameter of 0.1 in. What is the effect?
434 CONTROL SYSTEM COMPONENTS [Chap. 11
11.8. A flapper valve operating at 1,500-psi supply pressure has an upstream orifice
with a diameter of 0.005 in. and a controlled-orifice diameter of 0.012 in. (see Fig.
11.50). The piston has a large area of 0.05 in.2 and a small area of 0.025 in.2 The
piston mass is 0.02 lb and the viscous friction between the piston and cylinder is such
that a force of 0.04 lb is required to move the piston at a rate of 1 in./sec. Obtain a
small-signal transfer function between flapper position and piston position for the oper¬
ating point defined by the fact that the piston is stationary in the middle of its allowed
travel, and there is no external force on it. Assume that the coefficient of discharge for
Upstream Q
orifice v Flapper-
Supply
pressure V7777A V/77777^7777^
^ ^ ^ V, Controlled
^ V/////////////M V,
I
V////////777/////////&.
Fig. 11.50
both orifices is 0.65 and that the compressibility of the fluid in the chamber is negli¬
gible. The density of the fluid is 0.03 lb/in.3
11.9. The flapper of the valve described in Prob. 11.8 is spring-restrained such that
the hydraulic reaction force at the quiescent operating point is balanced. The spring
constant of the spring used for this purpose is 100 lb/in. The flapper is controlled by
an electric torque motor which supplies a force of 0.05 lb/ma. The mass of the flapper
is negligible. Find the small-signal transfer function from input current to output
position of the piston for this system. Consider operation around the operating point
defined in Prob. 11.8.
11.10. A valve of the Moog type (see Fig. 11.41) has a second-stage valve piston
with a diameter of 0.1875 in. and a weight of 0.02 lb. The spring constant of each one
of the springs used to center the piston is 400 lb/in. The control orifices for the inlet
and sump are rectangular in shape and 0.03 in. wide. There are four of these rec¬
tangular holes spaced equally around the circumference of the sleeve at each land of
the valve piston. The first stage is a flapper valve having fixed orifices of 0.005-in.
diameter and variable orifices of 0.012-in. diameter. The flapper is controlled by a
torque motor supplying a force at the flapper of 10 lb/amp. The flapper is spring-
restrained, but has negligible mass or damping. The chamber volume is small enough
so that compressibility of the oil can be neglected.
The flow gain of this valve is measured by connecting the two outlet ports together
with a pipe large enough so as to produce negligible pressure drop. A flow meter is
then placed into the return line and the valve is connected to a 1,500-psi supply.
After subtracting the first-stage leakage flow, the flow gain is found to be 2-in.3/sec/ma
current difference into the torque motor.
Using small-signal methods about the operating point defined by load pressure =
load flow = 0, supply pressure = 1,500 psi, find the transfer function of the valve.
The density of the fluid is 0.03 lb/in.3
11.11. For the slide valve shown in Fig. 11.33, show that the reaction force tends to
open the valve when the slider is displaced from dead center.
CHAPTER 12
PNEUMATIC SYSTEMS
20(778) + 778
where q is the heat transferred to the fluid, p is the pressure in pounds per
square foot, work is the mechanical work done by the fluid, u is the inter¬
nal energy, V is the velocity in feet per second, z is the elevation, and v is
the specific volume, or volume per unit weight. Usually no heat is trans¬
ferred to the fluid, and thus q is zero. A process for which q is zero is
called an adiabatic process. Furthermore the difference in head or eleva¬
tion is negligible, and in a flow process no work is done by the fluid. Then
P1V1 P2V2 _ , V22 - F12
(12.2)
778 778 "2 "1 + 2g(778)
This equation applies whether the fluid is liquid, gas, or vapor. We shall
assume that our fluid follows the perfect-gas law pv = RT, where R is the
gas constant and T is the absolute temperature. For air, R is 53.3 ft/°R.
It can be shown2 that, for any perfect-gas process, the change in internal
1 S. L. Shearer and S. Y. Lee, Selecting Power Control Valves, Control Eng., vol. 3,
no. 4, pp. 73-76, 1956.
2 Binder, “Fluid Mechanics,” Prentice-Hall, Inc., Englewood Cliffs, N.J., 1949,
p. 197.
438 CONTROL SYSTEM COMPONENTS [Chap. 12
energy is
u2 — U\ = cv(T2 — T i) (12.3)
where cv is the specific heat of the fluid at constant volume. It can be
further shown1 that
1 R
cv = (12.4)
778 k - 1
For a general polytropic process
pvk = constant (12.5)
and for air, k = 1.4, if the process is adiabatic, and k = 1.0, if the process
is isothermal. Thus by combining Eqs. (12.3) to (12.5), Eq. (12.2) may
be written as
(*-!)/*'
V22 - Vi2 k Vi i - (2* (12.0)
2 k — 1 mi
wi 1
mi = (12.7)
9 Vi g
where w\ is the initial weight density and V\ the initial specific volume.
12.3. Pneumatic Flow through an Orifice. Let us consider the case
of adiabatic flow through a nozzle or orifice, as shown in Fig. 12.1. We
shall assume that V i, the velocity on the high-
pressure side, is negligible with respect to V2, the
velocity at the outlet. Thus Eq. (12.6) becomes
(k—l)/k
2k pi P2
P2 1 - (12.8)
k — 1 mi J> L
where v2 is the specific volume of the gas at the throat of the nozzle.
Substituting for V2 in Eq. (12.9) and employing Eqs. (12.5) and (12.7)
results in
(k+l)/k
Vi
W = A2 (12.10)
Vi
Figure 12.2 shows a plot of Eq. (12.10), the weight flow through an orifice
as a function of pi/pi. The maximum flow occurs at a ratio p^/px — 0.53.
1 Ibid.
Sec. 12.4] PNEUMATIC SYSTEMS 439
/F2 7tD dl =
7tD- dl y dV
(12.13)
8 gv 4 gv dl
VdV , fV'2dl
v dp + +J 2 g D 0 (12.14)
g
2 gv dV fdl
and dp -\- 2 -y- + 0 (12.15)
Vs D
Usually certain simplifying assumptions concerning the nature of the
flow may be made before integration is attempted. One type of flow is
isothermal, or constant-temperature, flow; the second type of flow is
adiabatic, or no-heat-loss, flow. Actual pneumatic processes lie between
these two extremes. Experimental work on many types of pneumatic
components such as flapper valves, capillary tubes, and tapered-pin
orifices have shown excellent agreement with theoretical isothermal flow.1
For isothermal flow pv = RT, and Eq. (12.15) can be integrated as
P i2 — P 22 =
gv i \ Vi Dt
(12.16)
Vi2Pi fl (12.18)
pl2 - P22 =
gv i D
For Reynolds numbers less than 2,000 the flow is usually laminar, and /
may be represented by the empirical relation found from the Staunton
1 C. R. Webb, Nonlinearities in a Pneumatic Controller, Trans. Soc. Instrument
Tech., vol. 7, no. 1‘, p. 9, 1955.
Sec. 12.6] PNEUMATIC SYSTEMS 441
diagram of
64
(12.19)
Nr
Substituting the relation for volumetric flow for V\ and rearranging, there
results
tt(pl2 - Pi2)DA
(12.20)
128/z/2pi
Equation (12.20) has been arranged for comparison with Eq. (10.22), the
Hagen-Poiseuille law for incompressible laminar flow in a pipe. While
there are some similarities, the relations are not the same. Figure 12.3
shows the error involved in using the incompressible relations instead of
the compressible relations for various Mach numbers.
12.5. Compressibility as an Equivalent Spring. Even if compressibil¬
ity is ignored in the flow through pipes, it should be considered for its
effect on the transfer function of the various system components. Just
as a viscous fluid adds viscous damping to various components, a com¬
pressible fluid acts as an equivalent spring. For instance, the resonant
frequency of a typical spool-type pilot valve was increased from 700
radians/sec to 950 radians/sec when 1,000-psi air was introduced into the
system; this result was due to the equivalent spring action of the air sup¬
ply. Usually the viscous-damping effect of compressed air is negligible
compared with friction and other sources of damping.
The spring rates of trapped compressed air can be important in
actuator cylinders. Figure 12.4* gives the equivalent spring constant of
compressed air initially at 100 psi trapped behind the piston in cylinders
of various sizes. The relation for the equivalent spring constant is
7TD ~p o
(12.21)
where k = spring rate, lb/in.
D = cylinder diameter, in.
L = cylinder length, in.
po = initial pressure behind piston, psi
12.6. Pneumatic Transmission Lag. Another effect that should be
considered in pneumatic systems is the transmission- or transport-lag
effect. Since air is much less dense than hydraulic fluid, the velocity of
propagation is lower. The approximate figure for the speed of sound in
air may be taken as 1,100 ft/sec, about one-fifth the speed of sound in
hydraulic oil. The actual signal-propagation velocity (group velocity)
will be less than this. If the pneumatic line were a perfect transmission
line, the output would reproduce perfectly any inputs, delayed in time,
however. The Laplace transfer function for a perfect transmission
line may be found by employing the real translation theorem.1 If
£[/(£)] = F(s), it can be shown that
1,000
800
700
600
500
400
300
250
200
150
C
£ 100
80
70
60
50
40
30
20
10
0.25 0.4 0.6 0.8 1.0 2 3 4 6 8 10
D cylinder diameter, in.
Fig. 12.4. Equivalent spring rates of compressed air. Trapped air was initially at
100 psi.
1 Gardner and Barnes, “Transients in Linear Systems,” John Wiley & Sons, Inc.,
New York, 1950, p. 236.
2 M. Bradner, Pneumatic Transmission Lag, Instruments, vol. 22, p. 618, 1949.
Sec. 12.7] PNEUMATIC SYSTEMS 443
Actual
'curve
'Delay time
3 sec plus
single time
constant of
14.5 sec
■} tas
y (12.23)
x 1 + Ts
where AP is the pressure drop and F is the volumetric flow. The pressure-
flow relations have been derived in Secs. 12.3 and 12.4. These relations
1 E. Stolarik, Notes on Pneumatics, Aero Dig., vol. 60, no. 1, p. 45, no. 2, p. 44,
1950.
Sec. 12.7] PNEUMATIC SYSTEMS 445
and pressure. For both Fig. 12.8 and Fig. 12.9 a correction factor for
other than standard conditions may be derived as
(12.25)
(12.26)
in.5/lb (12.27)
where F is the volumetric flow in cubic inches per second and P is the pres¬
sure in pounds per square inch. The value of capacitance is independent
of temperature but depends upon pressure and on the process of expan¬
sion. If the expansion is rapid, the process may be assumed to be
446 CONTROL SYSTEM COMPONENTS [Chap. 12
10’
10c
CD
CL
m.
c
\o
CD
CO
10 -1
CO
E
x:
o
CD
O
c
o
10 -2
co
CD
cr
10'3
-3
10 10-2 10'1
Diameter of tubes, in.
Independent of temperature
448
449
Fig. 12.14. Pneumatic-mechanical equalizers. A indicates supply pressure.
Plus Reset Plus Rate
450
Fig. 12.14. (Continued)
<v
03x
o
Fh
Pi
>>
E
PX
CQ
O03
c3
T3
in«o
T5a
ce
•p
o
CO
_
fc-
o
Plus Reset Plus Rate
■G
oa
o
r5
o
-A
>>
-aa>
w
ot-
<
z
'o
t-
o
HH
G
Oo
H
«
OPH
'O
o G
3
« .g
Ph "3
G
O
o
oo.
o
-G
pG
o
3
O£
-G
c«
G
-G
«
3
£
oc
2
o
-p>
G
451
452 CONTROL SYSTEM COMPONENTS [Chap. 12
Timing adjuster
Timing restrictor
Timing
capacity
Negative feedback
pressure
Sensitized diaphragms
providing mid point toad ■
Plug-in connectors
PVC tubes
Timing odjuster
Timing restrictor
Timing capacity
Sensitized diaphragms
b a
cd — cKp2 (12.28)
a + b a + b
where P = initial output pressure
c = gain constant of flapper, psi/in.
K = spring constant of feedback bellows, in./psi
Following the usual procedure of writing the linear equation about an
operating point, P may be struck out of Eq. (12.28). For the derivative
restrictor we may write
V = (1 + T1s)p1 (12.29)
where the constant T has the dimensions of a time constant. Its value
is determined by the area of the restriction and the volume of air at the
Air
Output supply
The form of Eq. (12.31) is rather different from that commonly used in
the process-control industries. There, the traditional method of analysis
would relate p and 6 by the differential equation containing the sum of the
derivative, proportional, and integral terms. An attempt is usually made
to cling to a physical interpretation of the separate terms by the introduc¬
tion of interaction terms, which turn out to be various ratios of the time
constants in Eq. (12.31). The interaction terms attempt to relate con¬
troller dial settings, which adjust the two restrictors in the example
above, to the time constants1 of the transfer function. As the advantages
of operational techniques become more widely known, the transfer func¬
tion and the frequency response of elements will themselves become a
part of physical reality for process-control engineers, and manipulation
of the time constants will be acceptable. This linearized analysis may
be used to develop transfer functions for the configurations shown in
Fig. 12.15 and the additional configurations given in the problems at the
end of this chapter.
12.8. A Pneumatic Control System. As an example of a pneumatic
control system that includes several of the more common types of compo¬
nents, we shall consider the pneumatic positioning system shown in Fig.
where M is the mass of the moving parts and B is the coefficient of viscous
damping. k\ is the coefficient of the spring 7; kb is the spring constant of
the bellows; and kc is the equivalent spring due to the Bernoulli force at
the valve orifice. Thus the transfer function from bellows force to valve
position is
x _ 1
(12.33)
ffbeiiows Ms2 + Bs + ki + kb + kc
The relationship between flapper position and bellows force may be
assumed linear in the operating range; however there is a pneumatic time
lag between flapper position and bellows pressure which is proportional to
bellows force
tlbellows _ '^flap
(12.34)
#flap 1 + TbS
-o e<
High U
f IB Cf
pressure
V
Fig. 12.19. The flapper valve and its electric analogue.
Figure 12.19 shows the pilot valve and its electric analogue. In the nor¬
mal operating range the ratio of the upstream pressure to the downstream
pressure across orifice B is greater than 2:1; thus the flow through the
orifice is constant. In the electric analogue this is represented by a con¬
stant current source I. The transfer function of interest is from a change
in the flapper orifice resistance Rc to the change in bellows pressure, repre-
1 W. Everitt and G. Anner, “Communication Engineering,” 3d. ed., McGraw-Hill
Book Company, Inc., New York, 1956, p. 231.
2 For a derivation that assumes linearity about an operating point, see H. A. Helm,
Frequency Response Approach to the Design of a Mechanical Servo, Trans. ASME,
vol. 76, p. 1209, 1954, and the appendix by R. Oldenburger.
Sec. 12.8] PNEUMATIC SYSTEMS 457
sented by the voltage eF. Unfortunately, as the bellows expands, its
equivalent capacitance changes; thus CF is a function of x. If the change
in bellows volume is only a small percentage of the total bellows volume,
this variation in C may be neglected. We may write by KirchhofPs law
that
I = iRc + icF (12.35)
or I = + cF (12.36)
rCc CU
where eF(0) is the initial value of the voltage eF. Suppose now that at
time t = 0 the resistance Rc is changed suddenly from the value RCi to a
new value Rc2. If we assume that a static condition exists prior to this
change then deF/dt in Eq. (12.36) is zero, and the initial value of the volt¬
age on the capacitance is
£f(0) = Rcil (12.38)
Substituting this value for e^(0) in Eq. (12.37) and letting Rc take on the
new value Rc2 we obtain
I _ eF
T CFscF — RciCfI (12.39)
s Rc 2
Thus the voltage eF approaches its final value in the familiar exponential
manner, and the time constant of the circuit is T2. When Rc is varied in
some other manner than stepwise, T2 varies during the process; however,
as an approximation for the time constant, some average value of resist¬
ance may be assumed.
PROBLEMS
12.1. Find the mass weight of flow for a smoothly rounded orifice 4i6 in- in diam¬
eter at 20°C. The upstream pressure is 1,000 psi.
458 CONTROL SYSTEM COMPONENTS [CHAP. 12
12.2. Find the equivalent time constant of a 5-in.-diameter cylinder 50 in. long with
a 1-lb piston.
12.3. Find the physical dimensions of a pneumatic network at 15°C and 1,000 psi
to synthesize the function 1 /(Ts + 1).
12.4. In Fig. 12.20 is shown a schematic of a so-called type II controller, which is
sometimes called a series controller because of the arrangement of the derivative and
integral restrictors.1 Find the transfer function from 6 to p.
Fixed Air
Output restrictor supply
Fig. 12.20
12.5. Find the transfer function of the parallel type III controller shown in Fig.
12.21, which is taken from Aikman and Rutherford.
Fixed Air
Output restrictor supply
<
P
Derivative xh
action Nozzle or
restrictor pilot valve
Integral
action —3 &
restrictor
Fig. 12.21
Fixed Air
Output restrictor supply
12.7. Find the transfer function of the type V controller shown in Fig. 12.23, taken
from Aikman and Rutherford.
Fixed Air
Output restrictor supply
460 CONTROL SYSTEM COMPONENTS [Chap. 12
12.8. Determine the transfer function of the lag4ead controller shown in Fig.
12.24, which is taken from Janssen.1 Assume that the gain of the flapper-nozzle
pilot valve is very high.
A
T
~m WiT^
K b
^ Pi P»
4 B -Pilot
valve
s\aaAaa.\v
Sn\\\W\W\^
V; Kf
r 1— Supply
J i—,pressure
vrvi
v;El vd:t
vN K
I Rri P+p
output
pressure
Fig. 12.24
12.9. Develop Eq. (12.16) from Eq. (12.15) by assuming constant weight rate of flow
and constant temperature.
PNEUMATIC COMPONENTS
Air Nozzle
Actuator
compressibility of the air in the actuator and the mass of the actuator
plus the mass of the load contributes an underdamped complex-conjugate
pair of roots. There is, in addition, the integration between valve posi¬
tion and actuator position. Reed finds the transfer from pressure ratio
to actuator position from experimental data to be approximately
x k
(13.1)
Ap s(0.01s + 1) (0.0004s2 + 0.008s + 1)
and a possible block diagram is shown in Fig. 13.2. It is seen from the
equation that the system becomes unstable if k is too large. It is possible
to equalize this system pneumatically by installing a compensator of the
form shown in Fig. 13.3. The nozzle-position actuator is connected to
the piston of fhe compensator, and the output of the compensator posi-
Sec. 13.2] PNEUMATIC COMPONENTS 463
Air volve
Pneumatic and Variable
control actuators nozzle Engine
Turbine discharge
pressure
R\
* 9
F_=_ksz
2 (/?,<;, s+/)(R2Czs+/)
100
80
60
40
20
10
T 8
^ 6
-O
i2
i
0.8
0.6
0.4
0.2
0.1
1 2 4 6 8 10 20 40 60 100 200 400 1,000
Time of operations, sec
Fig. 13.5. Weight comparison for storage units as a function of time of operation.
(Geyer and Treseder)
Fig. 13.7. Weight per unit of work versus rated work of actual production aircraft
actuators. (Geyer and Treseder)
Fig. 13.8. The intake stroke (1-2) is at inlet pressure. The compression
stroke will lie between the extremes of perfect isothermal compression
(2-3) and adiabatic compression (2-3'). The isothermal process requires
that the heat due to compression be extracted continuously to maintain
the gas at constant temperature, while the adiabatic process requires that
the cylinder walls be perfectly insu¬
lated so that no heat escapes. The
Discharge pressure
area enclosed by the cycle represents
:Isothermal compression
(pv-c)
the work of compression; thus the
isothermal cycle is to be preferred
Adiabatic compression
(pi/1A = c) since it represents the smaller value of
work. Even with isothermal com¬
pression, however, a good deal of
Inlet pressure
energy is required to pump a volume
Volume
of gas to an operating pressure of
Fig. 13.8. The pv diagram of an ideal
compression cycle. 800 or 1,000 psi. Let us say that an
ideal pump is to compress air, at the
rate of 0.001 lb/sec to 1,000 psig. If perfect isothermal compression is
assumed, we may use the familiar relation
WRT. p.
(13.2)
Fig. 13.9. The compression cycle of a Fig. 13.10. Efficiencies of multistage com¬
two-stage compressor with inter¬ pression cycles for pneumatic supplies.
stage cooling. {From Lucke, “ Engineering Thermodynamics,”
Columbia University Press, New York)
aspirator into the compression chamber may cause a diesel type of explo¬
sion at the high temperature and pressure in the chamber.
Much of the work actually done in compressing air cannot be recovered,
because it is impossible to design nozzles and loads to allow the air to
expand most efficiently, i.e., isothermally. Among other reasons this is
due to the time that would be required for such an expansion. Thus
over-all efficiencies of pneumatic systems are typically below 50 per cent,
and figures as low as 30 per cent are sometimes quoted.1 The efficiency
of the compressor alone will usually lie between 70 and 85 per cent.
An exception to the general rule of positive-displacement compressors
is found in jet aircraft, where the pneumatic supply may be the engine-
compressor discharge pressure, as in the example discussed in Sec. 13.1.
While this procedure is simple and convenient, it has at least two dis-
1 J. L. Shearer and S. Y. Lee, Selecting Power Control Valves, Control Eng., vol.
3, p. 72, March, 1956.
470 CONTROL SYSTEM COMPONENTS [Chap. 13
Pa
Fig. 13.11. Measured flow-pressure characteristic of a zero-lap valve as a function of
stroke x. (Shearer)
leakage flow in the valve is 0.005 in.3/sec, and the maximum no-load
velocity of the ram is 14 in./sec. In accordance with the principle dis¬
cussed above, the exhaust port of the valve has a maximum diameter
of % in., while the high-pressure port has a maximum width of only ^ in.
1 S. Y. Lee and J. L. Shearer, Proceedings of the National Conference on Industrial
Hydraulics, vol 8, p. 168, 1954.
472 CONTROL SYSTEM COMPONENTS [Chap. 13
Pneumatic
supply
Position feedback
Fig. 13.13. Diagram of a pneumatic positioning system.
1 10 100 1,000
cps
Fig. 13.14. The frequency response of the positioning system shown in Fig. 13.13.
(From Lee and Shearer)
High pressure
Fig. 13.16. Another type of two-stage, flapper-controlled pressure valve, called a non¬
bleed relay in the process-control industry.
controls the pressure pi. The second stage consists of the dual bellows
and the two nozzles n2 and n3 with their common flapper /23.
In equilibrium, piAi = p2A2, n2 is closed and n3 is almost closed by
flapper /23. If p 1 increases, the net bellows force carries n2 down, thus
forcing /23 away from n3 and allowing p2 to increase, returning n2 and w3
to equilibrium. If p 1 decreases, n2 moves up, leaving /23 to close n3 and
allowing p2 to vent to atmosphere through orifice 0 until the pressure
ratio is reestablished. The “nonbleed” appellation can be perhaps jus¬
tified since bleeding only takes place when either pi or p2 varies.
13.6. Pneumatic Motors. Rotary motors of any of the types discussed
under hydraulic components should theoretically operate properly under
pneumatic pressure, provided only that they receive proper lubrication.
At the present time., however, the only rotary-motor type commerically
Sec. 13.6] PNEUMATIC COMPONENTS 475
Suction
Pressure
1,500
1,000
E
Cl
L_
500
01_l_l_I_l
Suction 0 5 10 15 20
'Pressure Torque, ft-lb
pressed air not only acts as a lubricant but also allows an oil film to be
built up which seals the meshed gears and contributes to the volumetric
efficiency of the device by preventing leakage between the gears to sump.
Figure 13.19 shows speed-torque curves for a gear motor designed for
pneumatic service.
The piston-type hydraulic motor should be capable of operation as a
pneumatic motor, but a curious phenomenon is sometimes observed when
this application is attempted. After a short period of proper operation,
a sort of diesel action begins to take place. Apparently, the compressed
oil vapor suspended in the air begins to explode behind the cylinders. At
the same time, however, the motor is
refrigerated by adiabatic expansion
Pressure ■ rpm
730psi ■
Motor of the air in the chambers, and the
motor case remains cool. This diesel
valve
action probably reduces the efficiency
of the device and may damage the
mechanism if allowed to continue.
13.7. Actuators. Linear actuators
2,000 of the piston type used for hydraulic
s 1,800 systems are used also for compressed-
" 1,600 air service. Both spring-loaded and
3
Q.
=)
1,400 differential-area types are also in use.
o /_
1,200 The only construction difference ap¬
\ c>
1,000 On pears in the packing and seals. The
800
>
* “
spring-loaded actuator seems more
600
VC
\ b common in pneumatic systems, and
VP
\r5 it may take the form of a bellows
400
\c
200 VcT' \c or a Bourdon tube as well as the
\cp
0 piston type.
0.2 0.4 0.6 0.8 1.0
Output torque, ft-lb
Bellows and Bourdon tubes are
also useful as temperature-sensitive
Fig. 13.19. Speed-torque curves for a
pneumatic gear motor. and pressure-sensitive devices in
both hydraulic and pneumatic control
systems. When these elements serve as temperature-sensitive devices,
their principle of operation depends on the temperature coefficient of
expansion of the fluid enclosed. In the pressure-sensitive application,
the linear extension of the device depends on the pressure of the fluid
enclosed.
As a spring-loaded linear actuator, the Bourdon tube usually has much
the stiffer spring constant. While it is possible to design a bellows with a
very high spring rate for applications where only a small extension is
necessary, the Bourdon tube will usually exhibit better repeatability in
position versus pressure and will have the longer life. Bourdon tubes
Sec. 13.7] PNEUMATIC COMPONENTS 477
are constructed of steel or brass and typically exhibit excellent linearity
within their operating range. Typical characteristics are given in Fig.
13.20. Neither Bourdon tubes nor
bellows should be allowed to expand
beyond their free length if long life is
to be assured.
While bellows have been constructed
of many types of materials, the metal
bellows must be used in high-per¬
formance, high-pressure applications.
Metal bellows may be employed in
systems with pressures up to and be¬
yond 3,000 psi. Bellows are used in
many aircraft control systems as pres¬
sure-sensitive devices that correct for £=3
to
altitude, e.g., in bombsights and fuel to
03
l_
CL
controls to set the fuel-air ratio. O
Bellows can be constructed to sense o
E
3
differential pressure, as shown in Fig. 03
C
Cl
13.21. The pressures pi and p2 are
introduced to the inside and the out¬
side, respectively, of the large bellows.
The large expanding bellows is sealed
at both ends by the two small bellows
Fig. 13.20. Bourdon tubes: (a) sche¬
contained inside it. The motion of
matic; (b) typical characteristics.
the actuating rod is then a function of
the pressure difference. The unit is more compact and has a higher
sensitivity than two separate bellows connected by mechanical linkage.
a Position depends on
T P\~Pl
Fig. 13.21. Differential-pressure bellows. (Howard)
478 CONTROL SYSTEM COMPONENTS [Chap. 13
* J. H. Howard, Metal Bellows, Machine Design, vol. 26, no. 1, pp. 137-148, 1954.
f The silver-clad brass is not plated, but rather the solid silver is rolled on.
Wall thickness, in. 0.004 0.005 0.006 0.007 0.004 0.005 0.006
Approx, free length, || per convolu-
tion, in. 0.084 0.079 0.074 0.070 0.100 0.087 0.076
Maximum deflection,^ per convolu-
tion, in. 0.029 0.028 0.027 0.023 0.047 0.034 0.029
Spring rate,* * * § ** lb/in. per convolution
172 291 552 687 62 108 226
Flexibility, ff psi per convolution,
in/psi. 0.0040 0.0024 0.0012 0.0010 0.010 0.006 0.003
Maximum internal pressure, psi. 55 80 150 200 55 65 130
Maximum external pressure, psi.... 60 88 165 220 61 72 143
* J. H. Howard, Metal Bellows, Machine Design, vol. 26, no. 1, pp. 137-148, 1954.
f To obtain approximate inside diameter, subtract two times the wall thickness
from root diameter.
t May be decreased in every case, increased in some.
§ To obtain volume, multiply by bellows length.
|| To obtain total free length of bellows, multiply by number of convolutions.
H To obtain total maximum deflection of bellows, multiply by number of convolu¬
tions.
** To obtain spring rate of bellows, divide by number of convolutions,
ft To obtain flexibility of bellows, multiply by number of convolutions.
Sec. 13.7] pneumatic components 479
brass to 1.0 per cent for bronze. These figures are reduced by 2 by heat-
treating and by a factor of about 3 for extra-flexible convolutions.1
Design tables are available from manufacturers for the physical dimen¬
sions of all standard bellows. Table 13.2 is an example of one standard
table for lV^-in. convolutions.
E
3
E
X
O
E
c
CD
(_>
CD
QL
of
3
CO
CO
CD
c
o
CO
c
o
o
Fig. 13.22. Life expectancy of metal bellows as a function of pressure variation and
stroke. (Howard)
Bellows should never be expanded beyond their free length. The figure
for variable-pressure life must be used when pressure changes, even if the
bellows remains fixed. Internal flexing takes place as the pressure
changes, even if the ends are fixed, thus reducing the life. Several com¬
mon bellows configurations are shown in Fig. 13.23.
(b) Extension-compression
stop assembly
1 Ibid.
NAME INDEX
A-c servo (see Carrier-frequency servo; Amplifiers, a-c power, plate efficiency, 116
Servomotors) power transistors, 120-123
Accelerometer, 357-359 push-pull operation, 116
Accumulator, 430 transformer coupling, 117
pneumatic, 464, 467 vacuum-tube type, 124-126
Accuracy, 200, 221 d-c, 61-111
Actuator, hydraulic, 386 cathode-temperature variation, 65
pneumatic, 476-480 chopper-stabilized, 97-103
Air compressors, 467-470 definition, 61
Ambiguity in speed synchro systems, design considerations, cathode-fol¬
240-241 lower circuits, 108, 109
American Gear Manufacturers Associa¬ difference amplifier, dynamic load
tion, 317, 328 line, 110
Amplidyne, 180, 198-206 static load line, 110
analysis, 200 triode amplifiers, 103-108
inaccuracies in, 206 with neon-tube-type coupling
armature reaction in, 187, 199 network, 106-108
compensating winding, 187 with resistor coupling network,
deadband effect, 206 103-106
definition, 198 drift, 61-80, 94, 95
direct-axis time constant, 205 ideal coupling network, 84
direct-axis voltage, 199 interstage coupling network, 80-84
effect of nonlinear brush resistance, 206 linear analysis, 64
effectiveness of direct-axis compensat¬ noise, 61
ing winding, 201 power-type, 124-126
equivalent circuit, 205 with RC networks, 52-54
field time constant, 205 transistor, 84-97
hunting, 206 (See also specific types of amplifiers)
mutual inductance, 188, 200 Analogue, mechanical to electrical, 305-
open-circuit voltage expression, 203 307
output-impedance expression, 204 Armature reaction, in Amplidyne, 187
principle of operation, 198 in d-c motors, 213
quadrature-axis brushes, sparking at, definition, 187
200 effect of, in d-c generator, 180, 187
quadrature-axis time constant, 205 in d-c tachometer, 180
quadrature-axis voltage, 199 Artificial horizon, 351, 353
speed voltage parameter, 200 Askania nozzle valve, 420
transfer function, low-frequency ap¬ Aspirator, pneumatic, 475
proximation, 204 Asymptotic diagram, 14, 20
trigger action, 206 Autosyn, 221
Amplifiers, a-c power, 115 (See also Synchros)
classes of operation, 116
equivalent circuit, 118
inductive loads with, 119 Bail ring, 347
maximum power transfer, 119 Ball, bouncing, 142
483
484 CONTROL SYSTEM COMPONENTS