Vous êtes sur la page 1sur 136

e x a m p l e 2.

20 a n N - r e s i s t o r c u r r e n t d i v i d e r Now
consider the more general current divider having N resistors, as shown in Figure 2.38.
It can be analyzed in the same manner as the two-resistor current divider. To begin, the
element laws are

i0 = −I (2.101)

vn = Rn in , 1 ≤ n ≤ N. (2.102)

Next, the application of KCL to either node yields

i0 + i1 + · · · iN = 0 (2.103)

and the application of KVL to the N − 1 internal loops yields

vn = vn−1 , 1 ≤ n ≤ N. (2.104)

Finally, Equations 2.101 through 2.104 can be solved to yield

i0 = −I (2.105)
Gn
in = I, 1≤n≤N (2.106)
G1 + G2 + · · · GN
1
vn = I, 0≤n≤N (2.107)
G1 + G2 + · · · GN

where Gn ≡ 1/Rn . This completes the analysis.

As was the case for the two-resistor current divider, the preceding analysis shows that
parallel resistors divide current in proportion to their conductances. This follows from
the Gn in the numerator of the right-hand side of Equation 2.106. Additionally, the
analysis again shows that parallel conductances add. To see this, let GP be the equivalent
conductance of the N parallel resistors. Then, from Equation 2.107 we see that

I
GP = = G1 + G2 + · · · GN (2.108)
vn

from which it also follows that

1 1 1 1
= + + ··· (2.109)
RP R1 R2 RN

where RP ≡ 1/GP is the equivalent resistance of the N parallel resistors. The latter result
is summarized in Figure 2.40.

83a
.. .

F I G U R E 2.40 The equivalence 1- + ----- 1- –1


1- + ... + ------
of parallel resistors; for N = 2, R1 R2 RN Rp = -----
R 1 R2 RN
RP = R1 R2 /(R1 + R2 ).
.. .

Finally, the two current-divider examples illustrate an important point, namely that
parallel elements all have the same voltage across their terminals because their terminals
are connected directly across one another. This results in the KVL seen in Equations 2.80,
2.81, and 2.104, which state the equivalence of the terminal voltages.

83b
e x a m p l e 2. 28 b a s i c c i r c u i t a n a l y s i s m e t h o d Solve
the circuit in Figure 2.58 using the basic method.
Step 1 is to assign the branch variables. Figure 2.59 shows the circuit with the variables
properly assigned.

In Step 2, we write the constituent relations:

vS = −V (2.153)

v1 = i1 R1 (2.154)

v2 = i2 R2 (2.155)

v3 = i3 R3 (2.156)
v4 = i4 R4 (2.157)
v5 = i5 R5 . (2.158)

In Step 3, we write the KVL and KCL equations. The KVL equations with respect to
the loop choice shown in Figure 2.60, are

vS + v1 + v2 + v4 = 0 (2.159)
−v2 + v3 = 0 (2.160)
−v4 + v5 = 0 (2.161)

R1
R2

R3 F I G U R E 2.58 Circuit example.


+
V
-
R4

R5

97a
v
i1

+ i3
R1
- v2
v1
+ R2 -
+
- v3
i2 -
F I G U R E 2.59 Circuit with R3
vS +
properly assigned variables.
V
+
-
i4
iS R4 +
- v4
R5 i5

- v5 +

(a)

L1 L2

F I G U R E 2.60 Loop and node


choice.
+
(b)
V
-

L3
(d) (c)

97b
At node (a), the KCL equation is

i1 − i2 − i3 = 0. (2.162)

Notice that nodes (b) and (c) are connected by a wire, so they yield only one KCL
equation

i2 + i3 − i4 − i5 = 0. (2.163)

Lastly, at node (d), we have


i4 + i5 − iS = 0. (2.164)

Combining the constituent relations with KVL equations, we obtain

−V + i1 R1 + i2 R2 + i4 R4 = 0 (2.165)
−i2 R2 + i3 R3 = 0 (2.166)
−i4 R4 + i5 R5 = 0. (2.167)

By adding Equations 2.162 2.164, we have

iS = i1 . (2.168)

Eliminating i2 and i4 and substituting back into Equations 2.166 2.167 gives us

R2
i3 = iS (2.169)
R2 + R3
R4
i5 = iS (2.170)
R4 + R5
 
R22 R24
V = iS R1 + R2 + R4 − − (2.171)
R2 + R3 R4 + R5
 
R2 R3 R4 R5
= iS R1 + + . (2.172)
R2 + R3 R4 + R5

As a quick sanity check of the solution, one might notice that the equivalent resistance
of the network around the voltage source is R1 + R2 R3 + R4 R5 , which is correctly
shown by Equation 2.172.

97c
e x a m p l e 2.33 v o l t a g e - c o n t r o l l e d r e s i s t o r Thus far
we have dealt with resistors that have a fixed resistance. However, like dependent
sources, we can also have resistors whose values depend on other parameters. As an
example, Figure 2.71 depicts a voltage-controlled resistor whose resistance RX is a
function of vI .

Let us suppose we are interested in determining vO as a function of vI for

RX = f (vI ) = Ro vI

where Ro is some known constant. Let Ro = 5 k/V.


First, R1 and R2 form a simple voltage divider, and since R1 = R2 , we have vI = V/2.
Second, RL and RX also form a voltage divider. Therefore,

RX
vO = V
RL + RX
Ro vI
=V
RL + Ro vI
5 k/vI
=V
10 k + 5 k/vI
vI
=V
2 + vI
V
2
=V V
2+ 2

V2
= .
4+V

Substituting V = 5 V, we find that vO = 25/9 V.

R1 = 100 k Ω RL = 10 kΩ

+
F I G U R E 2.71 Circuit with V 5 V vI vO
voltage-dependent resistor. -
R2 = 100 k Ω
RX = f(vI)

107a
2.7 A F O R M U L A T I O N S U I T A B L E F O R A
COMPUTER SOLUTION *
Thus far we have seen several circuit examples that we solved by writing a
set of equations based on the constituent relations for the elements, KVL,
and KCL. There were as many independent equations as unknown variables,
which allowed us to solve for any variable by simple algebra. The same set
of equations can be written in matrix form so that they are amenable to a
computer solution. For example, the circuit in Figure 2.1 analyzed using the
basic method in Section 2.3.5 resulted in ten equations and ten unknowns.
These ten equations are summarized as follows:

v1 = i1 R1 (2.202)
v2 = i2 R2 (2.203)
v3 = i3 R3 (2.204)
v4 = i4 R4 (2.205)
v5 = V (2.206)
−v5 + v1 − v2 = 0 (2.207)
+v2 + v3 + v4 = 0 (2.208)
−i5 − i1 = 0 (2.209)
+i1 + i2 − i3 = 0 (2.210)
i3 − i4 = 0. (2.211)

The ten unknowns are v1 , v2 , v3 , v4 , v5 , i1 , i2 , i3 , i4 , and i5 . The equations can be


rewritten so that constant voltages and currents appear on the left-hand side of
the equation.

0 = v1 − i1 R1 (2.212)
0 = v2 − i2 R2 (2.213)
0 = v3 − i3 R3 (2.214)
0 = v4 − i4 R4 (2.215)
V = v5 (2.216)
0 = v1 − v2 − v5 (2.217)
0 = v2 + v3 + v4 (2.218)
0 = −i5 − i1 (2.219)
0 = i1 + i2 − i3 (2.220)
0 = i3 − i4 . (2.221)

107b
This set of equations can be written in matrix form as follows:
    
0 1 0 0 0 0 −R1 0 0 0 0 v1
 0  0 1 0 0 0 0 −R2 0 0 0 v2 
    
 0  0 0 1 0 0 0 0 −R3 0 0  
    v3 
 0  0 0 0 1 0 0 0 0 −R4 0 
 
   v4 
V 0 0 0 0 1 0 0 0 0 0  v5 
 
 = 
 0  1 −1 0 0 −1 0 0 0 0 0  
     i1 
 0  0 1 1 1 0 0 0 0 0 0  
     i2 
 0  0 0 0 0 0 −1 0 0 0 −1 
 
    i3 
 0  0 0 0 0 0 1 1 −1 0 0   i4 
0 0 0 0 0 0 0 0 1 −1 0 i5
(2.222)

This matrix equation is in the form

b = Ax

where x is a column vector of unknowns and b is the column vector of drive


voltages and currents. This vector of unknowns can be solved by a computer
using standard linear algebraic techniques such as Cramer’s rule. In fact, the well
known SPICE software package uses methods such as these to solve circuits.5

5. The examples in this chapter focused on linear circuits, which result in a set of linear simulta-
neous equations. However, the fundamental method of solving circuits based on KVL, KCL, and
constituent relations applies equally well to nonlinear circuits. A nonlinear circuit might contain
nonlinear circuit elements with constituent relations such as v = i3 R, or, i = K(ev/VT − 1). The
resulting set of equations that arise will be nonlinear. Computer solution of such circuits makes use
of another technique called linearization, which is discussed in Chapter 4. Further discussions of
linearity are in Chapter 3, and a further treatment of nonlinear circuits is in Chapter 4.

107c
e x a m p l e 3.9 e v e n m o r e o n t h e n o d e m e t h o d As
we discussed earlier, it is often inefficient to use only Kirchhoff’s laws to analyze com-
plicated circuits. For example, if we simply modify the example we saw on page 192
to the circuit shown in Figure 3.18, Kirchhoff’s laws alone will not be able to solve it
easily. We will use the node method to solve the problem and the node assignment in
Figure 3.19.

With respect to the node assignment in Figure 3.19, we have the following equations:

V − e1 e2 − e1 e3 − e1
+ + =0 (3.29)
R1 R2 R3
0 − e2 e1 − e2 e3 − e2
+ + =0 (3.30)
R4 R2 R6
e1 − e3 e2 − e3 0 − e3
+ + = 0. (3.31)
R3 R6 R5

We can rearrange the terms and express the equations in matrix form:

 1 1 1
   V 
+ + − R1 − R1 e1
 R1 R2 R3

2

3     R1 
    
 − R1 1
+ 1
+ 1
− R1  e2  =  0 
 R2 R4 R6    
 2

6
    
− R1 − R1 1
R3
1
+R +R 1
e3 0
3 6 5 6

i1

+ i3
R1
- v2
v1 R2
+ -
+
v3
i2 -
R3 F I G U R E 3.18 Circuit with
+
appropriate branch variables.
V i6
- R4 + v
i4 6
iS + -
R6
- v4 i5
R5

- v5 +

135a
e1

R1 +
- R v2
2
v1
+ -
+
v3
F I G U R E 3.19 Node R3 -
assignments of the circuit. +
e2
V
-
+ v
6
R4 + -
R6
- v4
R5

- v5 + e3

Standard matrix techniques can be used to solve for the unknowns. Let us assign the
following values to the resistors and voltage source:

V=5V

R1 = 50 

R2 = 100 
R3 = 100 
R4 = 75 

R5 = 75 

R6 = 150 .

Then we have:
 1 −1 −1
   1 
e1 10
 25 100 100
   
 −1 3 −1     
 100  e 
150   2 
=0
 100  
−1 −1 3
100 150 100 e3 0

Solving for e1 , e2 , and e3 , we have e1 = 35/11 V, e2 = 15/11 V, and e3 = 15/11 V.


Notice that e2 = e3 , that is, there is no current going through resistor R6 . Since R2 = R3
and R4 = R5 , the symmetry of the network between node e1 and ground splits the
current going into node e1 evenly, thus causing the same voltage drop.
135b
e x a m p l e 3 . 12 a m o r e c o m p l e x d e p e n d e n t - c u r r e n t
s o u r c e p r o b l e m As a more complex example of the node analysis of a
circuit containing dependent sources, consider the analysis of the circuit shown in
Figure 3.28. This circuit has two dependent sources: one VCCS and one CCVS.
In addition, its resistors are labeled with their conductances for convenience.

To analyze the circuit in Figure 3.28, we redraw it as shown in Figure 3.29. Here, the
VCCS is replaced by an independent current source having value Ĩ, and the CCVS is
replaced by an independent voltage source having value Ṽ. Note that the new indepen-
dent voltage source is not a floating voltage source because it is connected to ground
through the known voltage V.
The circuit in Figure 3.29 can be analyzed by the node method presented earlier. Since
ground is already defined in the figure at Node 5, Step 1 is already complete. To
complete Step 2, the node voltages are labeled as shown. The voltages at Nodes 1 and
2 are the unknown node voltages e1 and e2 . The voltage at Node 3 is set by the original
independent voltage source, and is labeled accordingly. The voltage at Node 4 is also
known since the new voltage source is an independent source, and it is labeled as such.

Next, we perform Step 3, writing KCL for Nodes 1 and 2 in the process. This yields

G1 (e1 − V − Ṽ) + G2 (e1 − e2 ) − I = 0 (3.64)

Node 4
Node 1
- v + e1

G1

+ G2 I
ri
-
Node 2
V F I G U R E 3.28 A circuit with two
e2 dependent sources.
G3
Node 3 i

+ G4 gv
V
-

Node 5

145a
Node 4
Node 1
- v + e1
V + V˜
G1

+ G2
V˜ I
-
F I G U R E 3.29 The circuit from
Node 2
V
Figure 3.28 redrawn with
e2
independent sources. G3
Node 3 i

+ G4 I˜
V
-

Node 5

for Node 1, and


G2 (e2 − e1 ) + G3 (e2 − V ) + G4 e2 + I − Ĩ = 0 (3.65)
for Node 2. Equations 3.64 and 3.65 can be restated as
 
   I
G1 + G2 −G2 1 G1 0 G1 V
e1  
=  . (3.66)
−G2 G2 + G3 + G4 e2 −1 G3 1 0  Ĩ 

Following Step 4, Equation 3.66 is solved for e1 and e2 . This yields


 
   I
1 G2 +G3 +G4 G2 1 G1 0 G1 V
e1  
=  
e2  G2 G1 +G2 −1 G3 1 0  Ĩ 


1 (G3 +G4 )I+(G1 (G2 +G3 +G4 )+G2 G3 )V+G2 Ĩ+G1 (G2 +G3 +G4 )Ṽ
=
 −G1 I+(G1 G2 +G1 G3 +G2 G3 )V+(G1 +G2 )Ĩ+G1 G2 Ṽ

(3.67)

145b
where

 = (G1 + G2 )(G2 + G3 + G4 ) − G22 . (3.68)

Finally, we use Equations 3.67 and 3.68 to solve for i and v, the branch variables that
control the CCVS and the VCCS, respectively. This yields

i = Ĩ − G4 e2
1  
= G1 G4 I − (G1 G2 + G1 G3 + G2 G3 )(G4 V − Ĩ ) − G1 G2 G4 Ṽ (3.69)

v = e1 − V − Ṽ
1  
= (G3 + G4 )I − G2 G4 V + G2 Ĩ − G2 (G3 + G4 )Ṽ , (3.70)


which completes the node analysis of the circuit in Figure 3.29. Note that KCL was used
at Node 5 to derive the first equality in Equation 3.69.
To find the actual values for Ĩ and Ṽ, we now substitute Equations 3.69 and 3.70 into
the element laws for the CCVS and the VCCS, respectively. This yields

r  
Ṽ = ri = G1 G4 I − (G1 G2 + G1 G3 + G2 G3 )(G4 V − Ĩ ) − G1 G2 G4 Ṽ (3.71)


for the CCVS, and

g  
Ĩ = gv = (G3 + G4 )I − G2 G4 V + G2 Ĩ − G2 (G3 + G4 )Ṽ (3.72)


for the VCCS. Finally, Equations 3.71 and 3.72 are jointly written as
 
 − gG2 gG2 (G3 + G4 ) Ĩ
−r(G1 G2 + G1 G3 + G2 G3 )  + rG1 G2 G4 Ṽ
 
g(G3 + G4 ) −gG2 G4 I
= (3.73)
rG1 G4 −rG4 (G1 G2 + G1 G3 + G2 G3 ) V

and then solved simultaneously to yield


 
g(G3 + G4 ) −gG2 G4 (1 − rG3 ) I

Ĩ r(G1 G4 + gG3 ) rG4 (G1 G2 + G1 G3 + G2 G3 ) V
= . (3.74)
Ṽ  + rG1 G2 G4 − gG2 (1 − rG3 )

The actual values of the dependent sources are now known. Finally, to complete the
node analysis, at least to the point of determining e1 and e2 , Equation 3.74 is substituted

145c
into Equation 3.67 to yield
 
G3 (1 + rg) + G4 (1 + rG1 )  − G2 G4 − rG1 G3 G4 − gG2 (1 − rG3 ) I

e1 g − G1 G1 G2 + G1 G3 + G2 G3 − gG2 (1 − rG3 ) V
= .
e2  + rG1 G2 G4 − gG2 (1 − rG3 )
(3.75)

Now, with Equations 3.74 and 3.75, all node voltages are known and so all
branch variables may be computed explicitly.
As was the case for the circuit in Figure 3.26, it is also possible to apply the simple node
analysis described in Subsection 3.3 to the circuit in Figure 3.28. However, for the latter
circuit, the savings in time is not as great because some effort and thought is needed
to express i and v explicitly in terms of e1 and e2 . Furthermore, since these expressions
can be obtained in several different ways, the simple analysis becomes somewhat ad hoc
when applied to the circuit in Figure 3.28.

To begin the simple node analysis of the circuit in Figure 3.28, we express i and v
explicitly in terms of e1 and e2 . The ability to do so will be needed to carry out the spirit
of Step 3. From the definition of v in Figure 3.28, it is apparent that

v = e1 − V − ri. (3.76)

Thus, v can easily be expressed explicitly in terms of e1 and e2 once i is so expressed.


One relatively convenient way to express i explicitly in terms of e1 and e2 is to combine
KCL applied at Nodes 1, 3, and 4. This results in

i = I + G2 (e2 − e1 ) + G3 (e2 − V). (3.77)

The first term on the right-hand side of Equation 3.77 is the current through the
independent current source, and the second term on the right-hand side is the cur-
rent through the resistor labeled G2 . These two currents combine at Node 1, and their
sum exits Node 1 through the resistor labeled G1 . Finally, the combined current passes
through Node 4 and the CCVS, before entering Node 3. At Node 3, the combined
current also combines with the current through the resistor labeled G3 , and together
they exit Node 3 as i. The last term on the right-hand side of Equation 3.77 is the
current through the resistor labeled G3 . Thus, Equation 3.77 does express KCL applied
to Nodes 1, 3, and 4. Finally, the substitution of Equation 3.77 into Equation 3.76 yields

v = e1 − V − r (I + G2 (e2 − e1 ) + G3 (e2 − V)), (3.78)

which expresses v explicitly in terms of e1 and e2 .

Next, we apply the simple node method, beginning with Step 3, yielding

0 = G1 v + G2 (e1 − e2 ) − I, (3.79)

145d
for Node 1 and

0 = I + G2 (e2 − e1 ) + G3 (e2 − V ) + G4 e2 − g v (3.80)

for Node 2. At this point Equations 3.79 and 3.80 still contain v. However, upon
substitution of Equation 3.78, they can be rewritten as
 
G1 + G2 + rG1 G2 −G2 − rG1 (G2 + G3 ) e1
−G2 − g − rgG2 G4 + (1 + rg)(G2 + G3 ) e2
 
1 + rG1 G1 (1 − rG3 ) I
= . (3.81)
−1 − rg G3 (1 + rg) − g V

Finally, following Step 4, Equation 3.81 can be solved to yield


 
G3 (1 + rg) + G4 (1 + rG1 )  − G2 G4 − rG1 G3 G4 − gG2 (1 − rG3 ) I

e1 g − G1 G1 G2 + G1 G3 + G2 G3 − gG2 (1 − rG3 ) V
= ,
e2  + rG1 G2 G4 − gG2 (1 − rG3 )

(3.82)

which is identical to Equation 3.75, as it should be. The main point here is that while
the application of the simple node analysis described in Section 3.3 to circuits containing
dependent sources can result in less work, it also generally becomes less structured. This
is because, as part of the analysis, it is necessary to determine the variables that control
the dependent sources explicitly in terms of the unknown node voltages before the node
analysis is actually completed. It may not always be obvious how to do this in a simple
way. For this reason, when it is necessary to carry out a well-structured node analysis,
such as when the analysis is to be computerized, then the node analysis presented in this
subsection is preferred.

145e
3.3.4 T H E C O N D U C T A N C E A N D S O U R C E M A T R I C E S *
As we saw earlier in Equation 3.27, when a resistive circuit is linear (that is,
when its resistors and dependent sources are all linear), the equations resulting
from Step 3 of a node analysis can be formulated as a matrix equation, which
takes the form
Ḡ ē = S̄ s̄. (3.83)
Here, ē is a vector of the unknown node voltages, s̄ is a vector of the known
independent source amplitudes, and Ḡ and S̄ are known matrices, referred to
here as the conductance and source matrices, respectively. Examples of such
equations can be seen in Equations 3.27 and 3.66.
As previewed in the discussion following Equation 3.27, the matrices Ḡ
and S̄ have a very special structure. This structure allows us to skip the details of
Step 3 of a node analysis, and derive the two matrices directly from the topol-
ogy of the circuit. This also facilitates the computerization of a node analysis.
Alternatively, the special structure of the two matrices can be used to check
our work during Step 3. For simplicity, in this subsection we will examine
the structure of Ḡ and S̄ that arises from circuits that contain neither floating
voltage sources nor dependent sources. However, it is possible to extend our
observations to accommodate these sources as well.
The special structure of Ḡ and S̄ can be exposed by studying the partial
circuit shown in Figure 3.30. By the end of Step 3 of a node analysis, one
expression of KCL has been derived in terms of the unknown node voltages for
each node having an unknown node voltage. In the case of the partial circuit in

Node 1
e1

G1

I
e2 e3

F I G U R E 3.30 A partial circuit. G2


Node 3
Node 2 G3

V
+
V
-

145f
Figure 3.30, the corresponding expression of KCL for Node 3 is

G1 (e3 − e1 ) + G2 (e3 − e2 ) + G3 (e3 − V ) − I = 0. (3.84)

In writing Equation 3.84, KCL has been taken to state that the sum of the
currents exiting a node must vanish. Next, we rearrange Equation 3.84 as

−G1 e1 − G2 e2 + (G1 + G2 + G3 )e3 = G3 V + I. (3.85)

By writing KCL as in Equation 3.85, the special structure of the expression


becomes apparent. For example, the conductance of each resistor connected
to Node 3 contributes positively to the coefficient of e3 , and negatively to the
coefficient of the node voltage at the other end of the resistor. This is because
e3 acts to drive currents out from Node 3, while the other node voltages act
to drive currents in to Node 3. The same observation holds for the coefficient
of the grounded independent voltage source, except for a change in sign due to
the fact that the corresponding term is moved to the opposite side of the equal
sign. We also see that the current source enters positively into Equation 3.85,
once its term is moved to the opposite side of the equal sign, since it sources
current into Node 3.
Now consider assembling Equation 3.85, and its counterparts from the
other nodes in the circuit, in the form of Equation 3.83. Each expression of
KCL becomes a row within Equation 3.83. For the sake of discussion, let us
assume that these rows are ordered according to the number of the node for
which they are written, and further that the node voltages in ē are listed in order
of their corresponding node numbers. In this case, Equation 3.85 enters into
Equation 3.83 as
     
. e1 .
.  e2  . I
     
.  e3  .  V
     
−G1 −G2 G1 + G 2 + G 3 · · ·    · · ·  
 0  e4  = 1 G3  . .
     
.   .  .  . 
  .    . 
. .
. . .

(3.86)

Thus, we see that Ḡ is a matrix of conductances. A diagonal element at the


position [m, m] in Ḡ is the sum of the conductances connected Node m. An
off-diagonal element at the position [m, n] in Ḡ, m  = n, is the negative of
the conductance connecting Nodes m and n. This is true even for the zero
elements within Ḡ since a zero conductance indicates the absence of a resistor,

145g
or no connection. As a consequence of this structure, Ḡ is symmetric about its
main diagonal, at least in the absence of dependent sources.
Similarly, the matrix S̄ contains the coefficients of the sources. For each
independent current source, there will be a +1 in its column in S̄ at Row m if
the source enters Node m, a −1 if the source exits Node m, and a 0 otherwise.
For each grounded independent voltage source, the conductance connecting it
to Node m will appear in Row m of its column in S̄, including zeros to indicate
the absence of a connecting resistor.
Again, the structure of Ḡ and S̄ can be seen in Equations 3.27 and 3.66.
Consider, for example, the matrices in Equation 3.27. The [1,1] element of Ḡ is
G1 + G2 + G3 because the resistors labeled G1 , G2 , and G3 are all connected to
Node 1. Similarly, the [2,2] element in Ḡ is G3 +G4 because the resistors labeled
G3 and G4 are both connected to Node 2. The [1,2] and [2,1] elements in Ḡ are
both −G3 since G3 connects Nodes 1 and 2. Since the voltage source connects
to Node 1 through the resistor labeled G1 , but does not connect to Node 2, the
[1,1] element of S̄ is G1 and the [2,1] element is zero. Similarly, since the current
source enters Node 2, but does not connect to Node 1, the [2,2] element of S̄
is +1 and the [1,2] element is zero. Thus, the matrices in Equation 3.27 could
have been derived by inspection of the circuit topology only.

145h
3.4 L O O P M E T H O D *
We have already seen several examples of a complementary relationship
between voltage and current, so it should come as no surprise that there is
a simplified analysis method based on an astute choice of current variables that
closely parallels the method in the preceding section. Here we choose current
variables that flow in loops, that is, in closed paths. By this definition, the current
flowing into any node will always be identically equal to the current flowing
out, so KCL is identically satisfied. As in Chapter 2, we continue to define loop
currents until every element is traversed by at least one loop current. To illus-
trate, let us define a set of current loops for the circuit we previously analyzed,
as in Figure 3.31. KCL at Node 1 gives

(i1 + i2 ) − i1 − i2 = 0 (3.87)

which is identically zero for all values of i. Thus because KCL is automatically
satisfied for this choice of current variables, we have to write only KVL and the
constituent relations. Combining these in one step, we obtain

−V + (i1 + i2 )R1 + i1 R2 = 0 (3.88)


−i1 R2 + i2 R3 + (i2 + I )R4 = 0. (3.89)

Now rewrite to place the source terms on the left:

V = i1 (R1 + R2 ) + i2 R1 (3.90)
IR4 = i1 R2 − i2 (R3 + R4 ). (3.91)

i2

R1 Node 1 R3 Node 2

+ F I G U R E 3.31 Loop currents.


V R2 R4 I I
-
i1

145i
By Cramer’s Rule,

V(R3 + R4 ) + IR4 R1
i1 = . (3.92)
(R1 + R2 )(R3 + R4 ) + R1 R2

The voltage across R2 can now be found from Equation 3.92 and

e1 = i1 R2 . (3.93)

Equations 3.92 and 3.93 can be reduced to Equation 3.8 by simple algebra.

145j
e x a m p l e 3 . 13 l o o p m e t h o d Let us use the loop method to analyze
the circuit depicted in Figure 3.18 in our previous example. Figure 3.32 shows our choice
of the loops for this circuit.

The corresponding loop equations are

−V + i1 R1 + (i1 − i2 )R2 + (i1 − i3 )R4 = 0 (3.94)

(i2 − i1 )R2 + i2 R3 + (i2 − i3 )R6 = 0 (3.95)

(i3 − i1 )R4 + (i3 − i2 )R6 + i3 R5 = 0. (3.96)

By rearranging the terms into matrix form, we obtain


    
R1 + R2 + R4 −R2 −R4 i1 V
    
 −R2 R2 + R3 + R6 −R6  i2  =  0  .
−R4 −R6 R4 + R5 + R6 i3 0

Assigning the same values to the voltage source and resistors,

V=5V
R1 = 50 

(a)

R1
R2

i1 i2 R3
F I G U R E 3.32 Circuit with
+ properly assigned current loops.
(b)
V
-
R4
R6
i3

(d) (c)
R5

145k
R2 = 100 

R3 = 100 
R4 = 75 
R5 = 75 

R6 = 150 

we obtain
    
225 −100 −75 i1 5
    
−100 350 −150 i2  = 0 .
−75 −150 300 i3 0

Solving, we have i1 = 2/55 A, i2 = 1/55 A, and i3 = 1/55 A. As a sanity check, the


current flowing through R6 is i2 − i3 = 0, as desired.

145l
e x a m p l e 3 . 17 s u p e r p o s i t i o n a p p l i e d t o a b e e h i v e
n e t w o r k Superposition and a bit of creativity can also be used to solve more
complicated resistive networks. Figure 3.45 shows a resistive network containing an
infinite plane of resistors in a beehive shape. Each of the resistors have a resistance
value R. What is the equivalent resistance Reqv when looking into port A-B?

One of the key ideas of this problem is to properly choose a reference node or a
ground node for measuring voltages of the internal nodes in the network. Referring
to Figure 3.46, we take ground at infinity. Then, we introduce a current IP into node A
using a current source, and draw IP out of node B using another current source. If we
can compute the resulting voltage VP between nodes A and B, then we can obtain the
effective resistance between A and B as

VP
Reqv = .
IP

Our circuit has two sources, one injecting a current IP into the network, and the other
drawing a current IP out of the network. We will determine the voltage VP using
superposition by adding the voltages across A and B resulting from each of the current
sources acting alone. Figure 3.47 shows the circuit with the current source at B

A
F I G U R E 3.45 An infinite plane
resistive network. Each resistor has
B
a resistance value R.

153a
IP

A
+
F I G U R E 3.46 Introducing VP
-
ground into the network. B

IP

i3 IP

A
i2
i1 +
F I G U R E 3.47 The circuit with VP1
only the current source at A being -
B
applied.

153b
A
+
i4 VP2 F I G U R E 3.48 The circuit with
i5 - only the current source at B being
B
applied.
i6 IP

turned off, and Figure 3.48 shows the circuit with the current source at A turned off. Let
VP1 be the voltage across A and B when the current source at A acts alone, and let VP2
be the voltage across A and B when the current source at B acts alone. By superposition,
we know that

VP = VP1 + VP2 .

Referring to Figure 3.47, the current IP injected into node A will split evenly into three
currents, i1 , i2 , and i3 . We know that

i1 = i2 = i3

because the injected current faces a symmetric situation in each of the three directions.
Since, by KCL,

IP = i1 + i2 + i3 ,

we can write

IP
i1 = i2 = i3 = .
3

153c
Since the current through the resistor connecting nodes A and B is

i1 = IP /3

and since the resistance value of the resistor is R, we can write

IP
VP1 = Ri1 = R .
3

Similarly, referring to Figure 3.48, the current IP drawn out of node B comprises three
components i4 , i5 , and i6 , where

i4 = i5 = i6 .

Since, by KCL,

IP = i4 + i5 + i6 ,

we can write

IP
i4 = i5 = i6 = .
3

And in like manner, since the current through the resistor connecting nodes A and B is

i4 = IP /3,

and since the resistance value of the resistor is R, we can write


IP
VP2 = Ri4 = R .
3
Composing the expressions for VP1 and VP2 we get
2IP
VP = VP1 + VP2 = R.
3
Therefore,
VP 2
Reqv = = R.
IP 3

153d
e x a m p l e 4. 5 n o d e m e t h o d This example uses the device shown in iD
Figure 4.5. Recall that this device is characterized by the following device equation: +
+ v1 D1
V
iD = 0.1vD2 for vD ≥ 0, (4.17) -
-
+
iD is given to be 0 for vD < 0. v2 D2
Referring to the series connected nonlinear devices in Figure 4.14, determine iD , v1 , and -
v2 , given that V = 2 V.

We will use the node method to solve this problem. We first select a ground node and F I G U R E 4.14 Nonlinear devices
label node voltages as shown in Figure 4.15. We have one unknown node voltage v2 . connected in series.

Next, we write KCL for the node with the unknown node voltage. Recall that the
KCL equations in the node method are written directly in terms of the node voltages.
Accordingly,
0.1v22 = 0.1(V − v2 ) 2 . V
The term on the left-hand side is the current through device D2. Similarly, the term on
iD
the right-hand side is the current through device D1.
+
D1
Solving, we get V
-
V v2
v2 = .
2
D2
Given that V = 2V, we get v2 = 1V. We now obtain the remaining voltages and
currents by applying KVL and the relevant device laws. Thus,

v1 = V − v2 = 1 V
F I G U R E 4.15 Circuit with node
and voltages labeled.

iD = 0.1v22 = 0.1 A.

Notice that we could have also solved the circuit intuitively by realizing that the same
current flows through two identical nonlinear devices. Thus, the same voltage must
drop across both. In other words,

v1 = v2 .

Furthermore, by KVL
2 V = v1 + v2 .
Or, v1 = v2 = 1 V.

201a
e x a m p l e 4.8 making simplifying assumptions
Sometimes, there are a few special cases of interest that can be solved analytically by
making appropriate simplifying assumptions. The circuit in Figure 4.19 is one such
example. Here for variety, we will solve the circuit by a direct application of KVL and
KCL. KVL around the path containing the voltage sources and the diodes yields

−2E + vD1 − vD2 = 0 (4.28)

and KCL at the junction of the two diodes gives

iD1 + iD2 = IA . (4.29)

These two equations, together with the equations for the diodes of the form of
Equation 4.1, can be solved for the diode currents, assuming identical diodes.

Now, if we assume that the diode voltages are always positive enough to make
the −1 term in the diode equation negligible (for Equation 4.1, true within less than
one percent for all vD larger than 125 mV), then iD1 becomes

IA
iD1 = . (4.30)
1+e −2E/V TH

We can obtain this equation by following these steps. First, substitute in Equation 4.28
expressions for vD1 and vD2 in terms of iD1 and iD2 derived from the diode equations
(neglecting the −1 term). Second, obtain iD2 in terms of iD1 from this equation, substitute
in Equation 4.29, and simplify to get Equation 4.30.

The diode current is thus a hyperbolic tangent function of the voltage E, except for an
offset of IA /2.

iD1
+
+
E vD1
- IA
-
F I G U R E 4.19 Hyperbolic
tangent generator. -

+ vD2
E
- +
iD2

203a
e x a m p l e 4. 9 voltage-controlled nonlinear
r e s i s t o r Let us now determine vO as a function of vI for the circuit in
Figure 2.71 when
Ro
RX = f (vI ) = .
vI − 1 V
where Ro = 10 kV.

We have,

RX
vO = V
RL + RX
Ro
vI −1
v
=V Ro
RL + v −1 v
I

Ro
=V .
RL (vI − 1 V ) + Ro

Substituting, Ro = 10 kV and RL = 10 k,

10 kV
vO = V
10 k(vI − 1 V ) + 10 kV
V
=
vI
V
=
V
2

= 2 V.

203b
e x a m p l e 4.13 half-wave rectifier re-examined
As another example of piecewise linear analysis, we re-examine the half-wave rectifier
circuit for a sinusoidal input previously analyzed using graphical analysis in Section 4.10.
This time around, we will use a piecewise linear model for the diode, but use the same
graphical approach of Section 4.10.

Thus we start with the same circuit topology as in Figure 4.21a, except that the diode
is modeled using its piecewise linear approximation, the ideal diode, as shown in
Figure 4.32a. Assuming as before a ten-volt sinusoidal input voltage, as might be typical
in power supplies, we draw a succession of load lines on the piecewise linear characteris-
tics, Figure 4.32b, for representative values of the input wave, and plot the output voltage
point by point. The desired output voltage is vR , which in the graph is the horizontal

+
E = E o cos ( ωt ) +
- R vR
-
(a)
iD

Slope = -1/R

F I G U R E 4.32 Half-wave
rectifier: ideal diode piecewise
linear analysis.
vD
E
vR
(b)

vR

(c)
t

214a
distance from the operation point intersection to the input voltage. The resulting output
wave is shown in Figure 4.32c.
Comparison with the previous analysis, Section 4.10 and Figure 4.21, indicates that
at least for this problem, the simple ‘‘ideal diode’’ approximation yields a reasonably
accurate answer. As one would expect, the error mainly derives from the neglect of the
0.6-V drop across the diode. Clearly, the error would be more objectionable if the input
sinusoid had been of one volt peak rather than 10 volts.

214b
4.4.1 I M P R O V E D P I E C E W I S E L I N E A R M O D E L S
FOR NONLINEAR ELEMENTS *
The accuracy of the results of a piecewise liner analysis depends on the accuracy
of the model used. In this section, we will discuss the process of creating more
precise models of nonlinear elements when increased accuracy is desired.
To illustrate the process, let us use the diode as an example of a nonlinear
element. Thus far, we used the simple, ideal diode model. It is obvious from the
preceding example that the major effect in the model is when the voltage vD
across the diode is positive, and above about 0.6 V. Substantial improvement
can be made by adding a 0.6-V source in series with the ideal diode, as shown in
Figure 4.33a. The corresponding i v characteristic for the improved piecewise
linear model is shown in Figure 4.33b. Let us work a simple example using this
piecewise linear model.

+
iD
iD

vD

vD
+
0.6 V
0.6 V
-
-
(a) (b)

F I G U R E 4.33 Improved
piecewise linear diode models.
+
iD

iD
1
Slope = -----
Rd
vD +
0.6 V
-
vD
Rd 0.6 V

-
(c) (d)

214c
e x a m p l e 4. 14 a n o t h e r e x a m p l e u s i n g p i e c e w i s e
l i n e a r m o d e l i n g Let us rework the example containing a voltage source,
resistor, and diode in Figure 4.16 using the piecewise linear model for the diode from
Figure 4.33b. The behavior of this model, comprising an ideal diode in series with a
voltage source, can also be summarized in two statements:

Diode ON (vertical segment): vD = 0.6 V for iD > 0


Diode OFF (horizontal segment): iD = 0 for vD < 0.6 V (4.47)

Let us determine iD for E = 3 V and E = −5 V, given that R = 500 . According to


the piecewise linear method, we will focus on one straight-line segment at a time, using
linear analysis within each segment.

Vertical segment When iD and vD are in the vertical segment of their characteristic,
the circuit shown in Figure 4.34b results, and we can write

E − 0.6 V
iD = . (4.48)
R

Horizontal segment Figure 4.34c shows the corresponding circuit when the diode is
operating as an open circuit. In this segment,

iD = 0. (4.49)

Combining the results Intuition tells us that the vertical segment applies when
E > 0.6 V (the diode turns on) and the horizontal segment applies otherwise (diode
is off). Thus, when E = 3 V, Equation 4.48 applies, and

E − 0.6 V 3 − 0.6
iD = = = 4.8 mA.
R 500

Comparing to Equation 4.40, notice that the value of iD predicted by this improved
model is slightly lower than that predicted by the ideal diode model. The ideal diode
model did not account for the 0.6-V drop across the diode, and so overestimated
the current.

Equation 4.49 applies when E = −5 V, so

iD = 0.

214d
vR
+ -
iD
R +
+
E vD
-
0.6 V
-

(a) Complete model


vR
+ -

iD
R +
F I G U R E 4.34 Piecewise linear
+
analysis in the vertical and horizon-
E vD
tal straight-line segments using the - 0.6 V
diode model containing a voltage
source. -

(b) Vertical segment

vR
+ -

R iD
+
+
E vD
- 0.6 V
-

(c) Horizontal segment

Further improvement in accuracy can be realized by adding a series resistor Rd of


suitable value to the ideal diode and voltage source, as shown in Figure 4.33c.
The specific choice of resistor value depends on the application; one should
strive to make the characteristic match over the range of diode current expected
in the specific circuit (see Figure 4.35). We will illustrate the use of this model in
an example. More examples using these and other more complicated piecewise
models will appear throughout the book, and specifically in Chapters 7 and 16.

214e
e x a m p l e 4. 15 t h e d i o d e r e s i s t a n c e Choose values for Rd
for the piecewise linear diode model in Figure 4.33c assuming that the resistance must
provide a reasonable match for currents up to 0.4 A and 1 A. Assume VTH = 0.025 V
and Is = 10−12 A.

Figure 4.35 plots the v i characteristics for the diode. The figure shows that the resistance
value Rd1 = 0.1  provides a good match for the diode v i characteristics up to 1 A,
while the resistance value Rd2 = 0.2  provides a better match in the smaller current
range from zero to 0.4 A.

1.0
iD (A)

0.9
0.8
0.7 Slope = 10, Rd1 = 0.1 Ω
0.6
0.5
F I G U R E 4.35 Choosing a value
0.4 of the resistance in the piecewise
Slope = 5, Rd2 = 0.2 Ω
linear diode model.
0.3
0.2
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
vD (V)

214f
e x a m p l e 4.16 a more complicated piecewise
l i n e a r m o d e l Let us further rework the previous example using the piece-
wise linear model for the diode from Figure 4.33c. The behavior of this model,
comprising an ideal diode in series with a voltage source and a resistor, can be
summarized in two statements:

Diode ON (vertical segment): vD = 0.6 V + iD Rd for iD > 0.


Diode OFF (horizontal segment): iD = 0 for vD < 0.6 V.
(4.50)

Again, let us determine iD for E = 3 V and E = −5 V, given that R = 500  and


Rd = 10 .

Vertical segment When iD and vD are in the vertical segment of their characteristic,
the circuit shown in Figure 4.36b results, and we can write

E − 0.6 V
iD = . (4.51)
R + Rd

Horizontal segment Figure 4.24c shows the corresponding circuit when the diode is
operating as an open circuit. In this segment,

iD = 0. (4.52)

Combining the results For E = 3 V, Equation 4.51 applies, and so

E − 0.6 V 3 − 0.6 V
iD = = = 4.7 mA.
R + Rd 500 + 10

Equation 4.52 applies when E = −5 V, so

iD = 0.

As illustrated using the diode, increasingly better fits to an actual nonlinear


device characteristic can be obtained by introducing more and more ideal
elements. For example, for the diode, increasingly better fits to the actual
diode characteristic can be obtained by introducing more and more ideal diodes,
batteries, and resistors. But again a price is paid; increased accuracy of the
model brings increased complexity. The proper compromise between simplic-
ity and accuracy is not always obvious. Start with the simplest model, then add
complexity to see if the solution changes in major ways.

214g
vR
+ -
iD
R +

+
E vD
- 0.6 V
Rd
-
(a) Complete model
vR
+ -

iD
R +
F I G U R E 4.36 Piecewise linear
+
analysis in the vertical and horizon-
E 0.6 V vD
- tal straight-line segments using the
diode model containing a voltage
Rd source and a resistor.
-

(b) Vertical segment


vR
+ -

iD
R +
+
E vD
- 0.6 V

Rd
-

(c) Horizontal segment

214h
e x a m p l e 4.21 d i o d e r e g u l a t o r To further illustrate the use of
incremental analysis, we examine the diode circuit shown in Figure 4.45, another crude
form of voltage regulator that is slightly better than our previous regulator. As before,
we assume that the supposedly DC source supplying the circuit in reality has 5 volts
of DC with 50 millivolts of AC superimposed. The regulator is designed to reduce this
unwanted AC component relative to the DC.

To understand how the circuit operates, first draw the DC subcircuit to determine
ID and VO , the operating point variables of the circuit. We will use the piecewise
linear analysis method (based on the piecewise linear model of the diode shown in
Figure 4.33c) to determine the operating point variables. Accordingly, Figure 4.45b
shows the DC subcircuit in which each diode has been replaced with its piecewise linear
model comprising an ideal diode, a 0.6-V voltage source and a resistor of value Rd . By
inspection From Figure 4.45b,
5V − 1.8V
ID = . (4.84)
R + 3Rd
For R = 1000  and Rd = 10 , a reasonable value for diode currents in the 1- to
10-mA range,
3.2
ID = = 3.1 mA. (4.85)
1030
Next, draw the incremental subcircuit, as shown in Figure 4.45c. Here we will use
the accurate v i relation for the diode from Equation 4.1 to compute the value of

R
+
+
50 mV AC
-
Total
vO
source
+
- 5 V DC
- R
(a) +
+
F I G U R E 4.45 Diode regulator. 50 mV AC
-
R
3 rd ∆vo
0 -
+
0
+ (c) Incremental AC subcircuit
+
5 V DC - 1.8 V VO
-
3 Rd -

(b) DC subcircuit

228a
the incremental diode resistance rd . This incremental resistance can be derived using
Equation 4.75, in which f is the diode v i relation. We have also seen that the applica-
tion of Equation 4.75 for the diode v i relation results in Equation 4.74, which directly
yields the value of rd as
25 mV
rd =  8.1 . (4.86)
3.1 mA
Now, from Figure 4.45c, we can write an expression for the small signal AC output
3rd 24.3
vo = 50 = 50 = 1.19 mV AC. (4.87)
3rd + R 24.3 + 1000
(From Equation 4.66, with vo equal to 1.19 mV we expect an error of about 2%
in the neglect of higher-order terms in the incremental analysis.)
The total DC output voltage of the regular can be found from the DC subcircuit,
Figure 4.45b,

VO = 1.8 + 3ID Rd (4.88)


−3
= 1.8 + 3 × 3.1 × 10 × 10 = 1.89 V. (4.89)

The fractional ripple at the input,

50 × 10−3
fractional ripple = = 10−2 (4.90)
5
and at the output,
1.19 × 10−3
fractional ripple =  10−3 (4.91)
1.89
so the ripple has been reduced relative to the DC by a factor of 10.

228b
e x a m p l e 4.22 s m a l l signal analysis using a
p i e c e w i s e l i n e a r d i o d e m o d e l In the diode regulator exam-
ple, we used the piecewise linear model for the diode when conducting the DC operating
point analysis, but reverted to the accurate diode equation when computing the small sig-
nal resistance. This example will illustrate that small signal analysis of nonlinear devices
can also be carried out by using their piecewise linear models for both the DC operating
point analysis and in computing the small signal device resistance. Of course, the accu-
racy of the results will depend on the fidelity of the piecewise linear model used for the
nonlinear device.

The example will be based on the simple diode-resistor circuit shown in Figure 4.46.
Let us suppose we are interested in the small signal values of the output voltage and
the diode current for a 50-mV incremental input. As promised, throughout this exam-
ple, we will use the piecewise linear model for the diode illustrated in Figures 4.33a
and 4.33b.

We start by drawing the DC subcircuit to determine the operating point variables


ID , VO as shown in Figure 4.46b. By inspection, we can write

5 − 0.6
ID = .
R
For R = 1000 , ID = 4.4 mA and VO = 0.6 V.

R
+
+ iD
50 mV
-
vO
+
- 5V
- R

(a) Total circuit id +


F I G U R E 4.46 A simple +
50 mV
diode-resistor circuit. -
R rd = 0 vO
+ 0 -
ID
0

+ VO (c) Incremental subcircuit


5V
- 0.6 V
-

(b) DC subcircuit

228c
Next, we draw the incremental subcircuit for the operating point given by ID = 4.4 mA
and VO = 0.6 V. Since we chose to use the piecewise linear model for the diode
throughout our analysis, we must derive rd based on this model. Since ID > 0, notice
that the diode is operating in the vertical segment of the piecewise linear v i curve shown
in Figure 4.33b. Since the reciprocal of the slope of this curve segment is zero, rd is also
zero. In other words, the ideal diode looks like a short circuit for incremental changes
in the current. Figure 4.46c shows the corresponding incremental subcircuit.

From Figure 4.46c, it is easy to see that the incremental change in the output voltage
for the 50-mV change in the input voltage is simply

vo = 0.

Similarly, the incremental change in the current is given by


50 mV
id = = 50 µA.
R

228d
e x a m p l e 5 .14 simplifying another logic
e x p r e s s i o n (a) Find the minimum sum-of-products representation for the
boolean expression in Equation 5.8, namely

Output = A B C D + A B C D + A B C D. (5.30)

(b) Further, show that the expression in Equation 5.30 is equivalent to the logic
expression in the caption of Table 5.7, namely

AB + C + D.

As directed in part (a), we will simplify the expression in Equation 5.30 as follows:

Output = A B C D + A B C D + A B C D
=ABCD+ABCD+ABCD+ABCD
= (A B C D + A B C D) + (A B C D + A B C D)

= A C D(B + B) + B C D (A + A)
=ACD·1+BCD·1

= A C D + B C D. (5.31)

To answer part (b), recall that we have previously shown that AB + C + D can be sim-
plified to A C D + B C D (see Equation 5.29). Since the expressions in Equations 5.29
and 5.31 are identical, it follows that AB + C + D and A B C D + A B C D + A B C D
are equivalent.

267a
e x a m p l e 5 . 16 y e t a n o t h e r i m p l e m e n t a t i o n u s i n g
n o r s Let us derive an implementation based on two-input NOR gates for the
function AB + C + D. Assume that both the true and complement version of each
of the inputs is available:

AB + C + D = A B + (C + D) (5.35)

=A+B+C+D+C+D (5.36)

= ((A + B) + ((C + D) + C + D)). (5.37)

Implementing each of the expressions within parentheses using two-input NOR gates,
we get the circuit shown in Figure 5.24.
Notice that the algebraic simplification process was quite cumbersome. We can actu-
ally perform the same transformation directly on a gate-level circuit with greater ease.
Figure 5.25 shows how the original circuit for AB + C + D from one of the implemen-
tations in Figure 5.18 can be transformed into a two-input NOR implementation. The
transformations exploit the fact that two inverters (or circles) in series cancel each other.

A
B
F I G U R E 5.24 NOR implemen-
C tation of AB + C + D.

A A
B B

C C
D D

F I G U R E 5.25 NOR trans-


formations for AB + C + D.

A A
B B

C C
D D

267b
6.11 A C T I V E P U L L U P S
Large valued resistors are difficult to fabricate in VLSI technology. For example,
R2 is usually on the order of a few tens of ohms for polysilicon, few hundreds
of ohms for diffusion, and few hundredths of an ohm for metal. Fabricating a
10-k resistor using polysilicon would require an area hundreds of times larger
than that of a minimum sized transistor. Fortunately, MOSFETs themselves
make good high-valued resistors for the same area, the resistance RON of a
minimum sized MOSFET is significantly higher than that of a resistance made
out of other materials, such as polysilicon.
Figure 6.55 shows an inverter constructed out of MOSFETs with Mpu
serving as an active pullup. The pullup MOSFET has its drain tied to the power
supply connection, and thus the drain has a voltage VS applied with respect to
ground. To keep the pullup MOSFET permanently in its ON state, its gate is
connected to a second voltage VA , where VA is at least one threshold voltage
higher that the supply voltage. In other words,

VA > VS + VT .

VS

VA W
-----
Mpu  L  pu

vOUT

Mpd W
-----
 L  pd
vIN

F I G U R E 6.55 Logic gate with


active pullup. In the circuit,
VA > VS + VT , so that the pullup
MOSFET is always in its ON state. VS VS

RONpu RONpu

vOUT vOUT

vIN = High RONpd vIN = Low

321a
Let the W/L ratios of the pullup and the pulldown MOSFETs be (W/L)pu and
(W/L)pd , respectively. Let the corresponding ON-state resistances (according
to the SR model) be RONpu and RONpd . We also know that

L
RON ∝
W

where the constant of proportionality is Rn .25


Let us now choose the respective (W/L) ratios so that the inverter satisfies
the relationship derived in Equation 6.6, and repeated below for convenience:

RON
VS < VT
RON + RL

This relationship between the output low voltage of the inverter and the thresh-
old voltage of a MOSFET is necessary for the inverter to be able to drive the
MOSFET in another inverter into its OFF state. In the preceding equation, RL
is the resistance of the pullup device, and RON is the resistance of the pulldown
device.
With both an active pullup and an active pulldown,

1
VT > VS (6.12)
RL
1+
RON
1
> VS (6.13)
(L/W)pu
1+
(L/W)pd

where we have substituted the L/W ratios in place of the resistance values.

25. As mentioned earlier, the MOSFET displays resistive behavior between its drain and its source
only when the drain voltage is much smaller than the gate voltage (specifically, vDS  vGS − VT ).
Furthermore, the resistance Rn , and therefore RON , depends on the value of the applied gate
voltage. We will see more appropriate models for MOSFETs in other regions of operation in later
chapters. But for now, let us go ahead and use the SR model with a single value for Rn to analyze
the active pullup.

321b
For our typical parameters: VS = 5 V and VT = 1 V. Therefore,
we get

1
5 <1 (6.14)
(L/W)pu
1+
(L/W)pd
 
L
W pu
5<1+   (6.15)
L
W pd
 
L
W pu
4<   . (6.16)
L
W pd

In other words, we can choose the size of the pullup so its (L/W) ratio is four
times that of the pulldown.

321c
VS
e x a m p l e 6. 9 s i z i n g p u l l u p d e v i c e s For a 5-V supply volt-
age, suppose our static discipline prescribes a VOL = 0.5 V. How do we size the pullup
MOSFET in Figure 6.56 relative to the pulldown MOSFET to meet the valid output VA
RONpu ∝  -----
L
low threshold?  W pu

When the pulldown device is on, we know that the output voltage is given by vOUT
RONpd vIN
vOUT = VS . RONpd ∝  -----
L
RONpd + RONpu  W pd

To satisfy the static discipline, we must have VOL > vOUT when the input is high. Recall
that the on-state resistance is proportional to the ratio of the device gate length L and F I G U R E 6.56 An inverter with
its width W. Thus we have, an active pullup.

VOL > vOUT (6.17)


RONpd
> VS (6.18)
RONpd + RONpu
 
L/W pd
> VS     (6.19)
L/W pd + L/W pu

1
>5   . (6.20)
L/W pu
1+  
L/W pd

(L/W)pu
For VOL = 0.5 V, it is easy to see that if we choose > 9 we will satisfy the static
(L/W)pd
discipline. In other words, if both devices are of the same width W, the pullup device
must be sized so its length is nine times that of the pulldown device.

321d
e x a m p l e 6.10 c o m b i n a t i o n a l l o g i c u s i n g m o s f e t
s w i t c h e s Let us now rework some of our previous examples using all-MOSFET
designs and the SR model. Assume that we need to design our gates such that they satisfy
a static discipline with the low output voltage threshold VOL = VT− V, where VT is given
to be 1 V. Let us design all-MOSFET circuits and let us attempt to make them as small
as possible. Assume that the area of the circuit is proportional to the area of the gates
(W × L) of the individual MOSFETs. Let us also compute the power dissipated by the
circuits. Assume that Rn for the MOSFETs is 1 k.

Let us first consider the expression: AB + C + D. Figure 6.57 shows a compound gate
comprising only MOSFETs that implements this expression. This gate design replaces
the load resistor in Figure 6.23 with an active pullup. Our task is to determine the sizes
of both the pulldown and the pullup MOSFETs so this gate satisfies the static discipline
for VOL = 1− V. Note that the gate must satisfy the static discipline for any combination
of inputs.

As we have seen before, the key issue in designing a NMOS logic gate is to choose the
relative values of the pullup and the pulldown resistances so that even the highest value
for the gate’s output low voltage satisfies the VOL constraint. Since we are asked to
design the circuit that occupies the least area, and there are more pulldown transistors
than pullups, let us start by choosing minimum-sized transistors ((L/W )pd = 1) for the
pulldown circuit. Therefore the on resistance of an individual pulldown MOSFET is

 
L
RONpd = Rn = Rn .
W pd

The highest value for the output low voltage occurs when the pulldown circuit has its
highest resistance. Notice that the pulldown circuit has its largest on-state resistance
for an output low when A and B are on, and C and D are off. This largest pulldown
resistance is given by the sum of the on-state resistances of the MOSFETs with the A

VS

VA W -
----
L  pu
OUT
F I G U R E 6.57 Transistor-level
implementation of AB + C + D
using an active pullup. In the circuit,
VA > VS + VT , so that the active A
W -
pullup is in its ON state at all times. B C D ----
L  pd

321e
and B inputs. That is,

Rpdmax = 2RONpd = 2Rn .

To satisfy the static discipline, the output voltage of the gate for a logical 0 must be less
than VOL for any combination of inputs that can result in a logical 0 at the gate’s output.
In other words, the highest value for the low output voltage of the gate must be less
than VOL , which is given to be VT− .

Since the output voltage of the gate is given by

RONpd
VS ,
RONpu + RONpd

We can write the following constraint so that the gate satisfies the static discipline for
VOL = VT− :

RONpd
VT > VS
RONpu + RONpd
2Rn
> VS
RONpu + 2Rn
2Rn
> VS  
L
Rn + 2Rn
W pu
2
> VS   .
L
+2
W pu

For VS = 5 V and VT = 1 V, the previous constraint simplifies to


 
L
> 8.
W pu
In other words, the L/W ratio of the pullup must be chosen to be greater than 8. Thus
the resistance of the pullup is 8Rn .

Let us now compute the power dissipated by the circuit. The maximum amount of
power is dissipated when the resistance of the pulldown circuit is a minimum. This
happens when A = 1, B = 1, C = 1, and D = 1. Recalling that the resistances of each
of the pulldowns is Rn and that of the pullup is 8Rn ,

V2S
Pmax =
8Rn + 2Rn Rn Rn
= 3 × 10−3 W.

321f
We can also design a circuit for the expression (A + B)CD in like manner as depicted in
Figure 6.58. In this design, the maximum on-state resistance of the pulldown circuit for
VS an output low is achieved when both C and D is high and only one of A and B is high.
The corresponding maximum on-state resistance of the pulldown (assuming minimum
sized transistors) is 3Rn .
VA W
---- 
 L  pu As before, the pullup must be designed to have four times the resistance of the pulldown.
Since the pulldown circuit has resistance 3Rn , the L/W ratio of the pullup transistor must
Out be chosen as

(L/W )pu = 4 × (L/W )pd = 4 × 3 = 12.


W
---- 
 L  pd We can also calculate the maximum power dissipated by computing the minimum
A B resistance in the current path. The minimum resistance occurs when all inputs are high.
Thus the total resistance in the current path is given by
C
Rpu + Rpd = 12Rn + 2Rn + (Rn Rn ) = 14.5Rn .

D
The corresponding power dissipation is26

V2s
F I G U R E 6.58 Transistor-level Pmax = = 1.7 × 10−3 W.
14.5Rn
implementation of (A + B)CD
using an active pullup. In the circuit,
VA > VS + VT .

26. We note that a milliwatt of power per gate would cause today’s million-gate circuits on a VLSI
chip to dissipate a thousand watts of power! Because VLSI chips cannot dissipate more than few
tens of watts without esoteric packaging technologies, modern VLSI chips use another form of
logic called CMOS involving both n-channel and the complementary p-channel MOSFETs. We
will study this technology in Chapter 11.

321g
e x a m p l e 7. 18 b e t t e r b j t m o d e l s Since the base-emitter junc-
tion of a BJT functions like a diode, we can build more accurate models for the BJT
by using more sophisticated models for the BJT’s base-to-emitter diode. Figure 7.61
shows a pair of models for the BJT that provide better accuracy than the ideal-diode-
voltage-source model shown in Figure 7.49c. Notice that we are ignoring the presence
of the base-to-collector diode (shown in a faint outline form in Figures 7.61b and 7.61c)
by assuming that the BJTs are constrained to operate in their active region (that is, we
assume vCE > vBE −0.4). In this example, we will use each of these two models to com-
pute vBE , iC , and iE for the BJT in the circuit shown in Figure 7.53. Assume RD = 10 ,
VTH = 0.025 V, and Is = 10−12 A.

First, let us compute the parameters based on the model in Figure 7.61b. We will start
by making our calculations assuming that the BJT is operating in its active region, and
then verify that the results satisfy the conditions for active region operation. Under
active-region operation, we can obtain iC directly from the value of iB as

iC = βiB = 1 mA.

The emitter current is the sum of the base and collector currents. Thus

iE = iC + iB = 1.01 mA.

We can now determine vBE by summing the source voltage and the voltage drop across
RD as

vBE = 0.6 + iE RD = 0.6101 V.

This completes our calculations based on the model in Figure 7.61b. To verify that the
BJT is operating in its active region, we need to check that the following two conditions
are met: iB > 0 and vCE > vBE − 0.4 V. Substituting iB = 0.01 mA, vCE = 5 V, and
vBE = 0.6101 V, we can see that both conditions are indeed met.

Next, let us compute the parameters based on the model in Figure 7.61c. As with the
previous model, we will start by making our calculations assuming that the BJT is oper-
ating in its active region, and then verify that the results satisfy the conditions for active
region operation. Under active-region operation, we can obtain iC from the value of iB as

iC = βiB = 1 mA.

The emitter current is the sum of the base and collector currents. Thus

iE = iC + iB = 1.01 mA.

381a
C C
+ +
iC iC
C +
iC βiB βiB
iB iB vCE
B iB B B
vCE vCE
+ + v BE
+ ---------
iE + 0.6 V V TH
vBE vBE - vBE iE = I s e –1
- E -
RD iE
iE
- - - E -
E
v
For CE > v BE – 0.4 V, iC = βiB For v CE > v BE – 0.4 V, iC = βiB
Otherwise, iC = 0 Otherwise, iC = 0
(a) (b) (c)

F I G U R E 7.61 More accurate


models for a bipolar junction
transistor. We can now determine vBE from
 vBE 
iE = Is e VTH − 1 .

Solving by trial and error, we find that iE = 1.01 mA results in vBE ≈ 0.52 V.

This completes our calculations based on the model in Figure 7.61c. The computed
values once again confirm that the BJT is operating in its active region.

381b
9.4.1 S I N U S O I D A L I N P U T S *
Sinusoidal signals are an important class of inputs to electronic circuits. So, as
a first example of specific inputs to the circuits shown in Figures 9.31 through
9.34, consider the special cases of


0 t≤0
I(t) = (9.67)
I◦ sin(ωt) t > 0

0 t≤0
V(t) = (9.68)
V◦ sin(ωt) t > 0.

Note that both sources are zero for t ≤ 0, but nonzero for t > 0, so that they
effectively turn on at t = 0. A sketch of I(t) is shown in Figure 9.35a.
To complete the analysis of the circuits, we substitute the correspond-
ing source function from either Equation 9.67 or 9.68 into Equations 9.63
through 9.66 and carry out the indicated integration or differentiation. This
results in

0 t≤0
v(t) = I◦   (9.69)
 1 − cos(ωt) t > 0
ωC

I(t)

v(t)
Io 2Io
--------
(a) 2π
---- (b) ωC
ω--
π- 2π
----
--ω t ω--
-Io F I G U R E 9.35 The current I, the
π-
--ω t voltage v, the power vI, and the
energy wE stored in the capacitor,
for the circuit shown in Figure 9.31
v(t)I(t) given the sinusoidal source current
ωE (t) from Equation 9.67.
2I2o
----------
(d) ω C
2
(c) 2π
----
ω--

π-
--ω
t ----
ω--
π-
--ω t

482a
for the capacitor circuit shown in Figure 9.31,


0 t≤0
i(t) = (9.70)
ωCV◦ cos(ωt) t > 0

for the capacitor circuit shown in Figure 9.32,



0 t≤0
i(t) = V   (9.71)

 ◦ 1 − cos(ωt) t>0
ωL

for the inductor circuit shown in Figure 9.33, and


0 t≤0
v(t) = (9.72)
ωLI◦ cos(ωt) t>0

for the inductor circuit shown in Figure 9.34. Note that for these equations to
make sense, the units of ωC must be conductance and the units of ωL must be
resistance; they are. We will encounter these products again in future chapters.
A comparison of the circuit inputs given in Equations 9.67 and 9.68 to
the circuit responses given in Equations 9.69 through 9.72 shows that the
sinusoidal components of the current and voltage in each circuit are π /2 radians
out of phase with each other. This is in keeping with the observation that
the circuits perform integration or differentiation from current to voltage or
voltage to current. In the case of the capacitor circuits, the current leads the
voltage because the current must be present first to build up the charge to
which voltage is proportional. In the case of the inductor circuits, the voltage
leads the current because the voltage must be present first to build up the flux
linkage to which the current is proportional.
The operation of the circuits in Figures 9.31 through 9.34 demonstrates
that inductors and capacitors are capable of reversible energy storage. To see
this, let us examine the circuit shown in Figure 9.31 in detail; an examina-
tion of the three remaining circuits would yield identical observations. For this
circuit, the power delivered by the source to the capacitor is given by



0 t≤0
v(t)I(t) = 2   (9.73)

 I◦ sin(ωt) 1 − cos(ωt) t > 0.
ωC

482b
Integration of this power, or rate of energy delivery to the capacitor, yields

 0 t≤0

wE (t) =   (9.74)
 I2 3 1
 ◦ − cos(ωt) + cos(2ωt) t>0
ω2 C 4 4

as the energy stored in the capacitor. The current I, the voltage v, the power
vI into the capacitor, and the energy wE stored in the capacitor are all shown
in Figure 9.35. From the figure we see that the power can be both positive
and negative indicating that energy can be delivered to and retrieved from
the capacitor. In fact, during odd intervals of π /ω in time, energy is deliv-
ered to the capacitor. It is then retrieved without loss during the following even
interval of π/ω in time. Thus, ideal capacitors are lossless energy reservoirs.
The same is true for inductors.

482c
9.4.4 R O L E R E V E R S A L *
In each example in this section, a single capacitor or inductor was driven by a
source. When that element was driven by a current source its branch current
was imposed, and its branch voltage evolved in response. Alternatively, when
the element was driven by a voltage source, its branch voltage was imposed, and
its branch current evolved in response. However, because the branch variables
of a capacitor are self-consistently related by Equations 9.9 and 9.12, their roles
as the sourced and the responding branch variable may be reversed. Similarly,
because the branch variables of an inductor are related by Equations 9.28 and
9.30, their roles may also be reversed. This allows us to use one circuit response
to derive its converse. Specifically, we will derive the circuit responses to impulse
inputs using the role reversal argument.
As an example of role reversal consider the circuit shown in Figure 9.32
with the source voltage V given by Equation 9.80. The current i that circulates
through its source and capacitor in response to the step in source voltage is
the current impulse given by Equation 9.86. Now suppose instead that it is the
current i in Equation 9.86 that is imposed by a source, as in Figure 9.31 with
I ≡ i. What would be the voltage response v across the source and capacitor?
The answer is that it would be V from Equation 9.80 so that v = V. This can
be verified by substituting i in Equation 9.86 for I in Equation 9.63 and carrying
out the indicated integration with the help of Equation 9.82 to derive v. Thus
the current and voltage in Equations 9.86 and 9.80 are a self-consistent pair of
branch variables for a capacitor. They can be either the source and response, or
the response and the source. In this way we are able to find the circuit response
to a current impulse from the circuit response to a voltage step.
In the same way, we can use Equations 9.90 and 9.91, which apply to the
circuit shown in Figure 9.34, to determine i in Figure 9.33 for the case in which
V is an impulse. For example, suppose that the voltage v in Equation 9.91 is
imposed by the source in Figure 9.33 with V ≡ v. What would be the current
response i through the source and inductor? The answer is that it would be
I from Equation 9.90 so that i = I. This can be verified by substituting v in
Equation 9.91 for V in Equation 9.65 and carrying out the indicated integra-
tion with the help of Equation 9.82 to derive i. Thus the current and voltage
in Equations 9.90 and 9.91 are a self-consistent pair of branch variables for an
inductor. They can be either the source and response, or the response and the
source. In this way we are able to find the circuit response to a voltage impulse
from the circuit response to a current step.

489a
10.5.4 S O L U T I O N B Y I N T E G R A T I N G F A C T O R S *
Another approach to solution of first-order differential equations is via integrat-
ing factors. To illustrate, we return to the simple RC circuit driven by a current
source, Figure 10.2a. The corresponding differential equation, slightly rewritten
from Equation 10.2 is
dvC vC 1
+ = i(t). (10.103)
dt τ C
We now assume i(t) is some arbitrary input waveform, and that there is an initial
charge on the capacitor:
vC (t = 0− ) = Vo . (10.104)
To solve Equation 10.103, we look for an integrating factor f such that the
left-hand side of the equation becomes the derivative of a product;

dvC fvC d
f + = (fvC ) (10.105)
dt τ dt
dvC df
=f + vC . (10.106)
dt dt

Equating corresponding terms we find

f df
= (10.107)
τ dt
f = et/τ . (10.108)

After multiplying both sides of Equation 10.103 by this factor, we obtain

d 1
(vC et/τ ) = et/τ i(t). (10.109)
dt C

Now integrate from zero to t, and use corresponding limits on both sides of
the equation:
 vC (t)et/τ 
t/τ 1 t t /τ
d(vC e ) = e i(t )dt (10.110)
vC (0)e0/τ C 0
where t is a dummy variable of the integration. Performing the integration on
the left, and evaluating, we obtain
 t
1
vC (t)e t/τ
− V0 = et /τ i(t )dt . (10.111)
C 0

544a
Hence we find an explicit closed form solution for vC (t):
 t
e−t/τ
vC (t) = V0 e−t/τ + et /τ i(t )dt . (10.112)
C 0

Because the first term on the right depends only on the initial voltage on the
capacitor, this term must be the zero-input response. Similarly, the second term
is the zero-state response. Thus Equation 10.112 validates our initial assumption
(Section 10.5.3) that the total response is the sum of the ZIR and ZSR.
Equation 10.112 can be applied to any first-order linear system with
arbitrary input waveform. Examples will be presented in the next section. Unfor-
tunately, the extension to second- and higher-order systems is beyond the scope
of this text, so we will continue to rely on the homogeneous solution-particular
solution approach in dealing with such systems.

544b
e x a m p l e 10 . 3 solution by integrating factors
We can also solve Equation 10.114 by integrating factors. To do so, we note that the
Norton equivalent source is vI /R. Then the ZSR, from Equation 10.112, is
 t S t
e−t/τ
dt .
1
vC = et /τ (10.131)
C 0 R

A helpful integral at this point is



xex dx = xex − ex .

So equating x to t /τ , so that dx = dt /τ we find

 t
t
vC = S1 τ e−t/τ et /τ − et /τ (10.132)
τ 0
 
t
= S1 τ e−t/τ et/τ − et/τ + 1 (10.133)
τ
= S1 (t − RC) + S1 RCe−t/RC . (10.134)

This function is identical to that in Equation 10.124 and is plotted in Figure 10.33e.

550a
10.6.6 R C R E S P O N S E T O D E C A Y I N G E X P O N E N T I A L *
To illustrate the application of Equation 10.148, suppose we now apply a short
decaying exponential pulse to the RC circuit, as in Figure 10.47. Specifically we
assume that the input driving signal is

vI = Ae−t/τ1 t > 0. (10.155)

If the ‘‘short pulse’’ concept is correct, and τ1 is much less that the circuit time
constant RC = τ2 , then the output response to this exponential pulse should

vI R
(a)
+
Ae –t⁄ τ1 vI + C vC
-
-

vC
(b)
Aτ1
- e–t⁄RC
------------------
RC – τ1

F I G U R E 10.47 Response to
decaying exponential pulse. –Aτ1 –t⁄ τ1 t
-----------------
-e
RC – τ1

vC
(c)

558a
be proportional to its area. The pulse area is
 ∞
Area = Ae−t/τ1 dt (10.156)
0
 ∞
= −Aτ1 e−t/τ1 (10.157)
0

= Aτ1 . (10.158)

Hence the zero-state response of the circuit should be, from Equation 10.148,

Aτ1
vC  e−t/RC . (10.159)
τ2

To check this answer, we solve for the ZSR by using integrating factors. The
differential equation describing the circuit is

dvC
vI = RC + vC . (10.160)
dt

From Equation 10.112, assuming an exponential drive as given by Equa-


tion 10.155, the ZSR is

e−t/τ2 t
vC = et /τ2 A e−t /τ1 dt . (10.161)
τ2 0

The solution has two distinct forms, depending on the relative size of the two
time constants τ1 and τ2 .
We first assume that the drive pulse does not have the same time constant
as the circuit. Then, from Equation 10.161,
   1
−t/τ2 t
Ae 1 −t
vC =   e τ2 τ1  (10.162)
τ2 1
− 1
τ2 τ1 0
A  
= e−t/τ1 − e−t/τ2 . (10.163)
1 − τ2 /τ1

The first term of Equation 10.163 is the forced response to the exponential
input, with a time dependence the same as the input, but scaled in magnitude
by a factor related to the circuit time constant. The second term is the natural
response, (the homogeneous solution) with a time dependence characteristic of
the circuit rather than the drive.
This solution is completely general (except τ1  = RC). To match the ‘‘short
pulse’’ solution, we must assume that τ1 is much smaller than τ2 = RC. For this

558b
case, the two terms are shown in Figure 10.47b, and the complete response
is shown in 10.47c. If we make the pulse drive very short, then τ1 becomes
negligible compared to τ2 , the first transient becomes shorter and shorter, and
except very near t = 0 the capacitor voltage becomes

Aτ1
vC  e−t/RC (10.164)
RC

as we found from the area calculation.


The results of this discussion can be generalized to state that whenever the
characteristic time of the input pulse is much shorter than the time constants
of a linear circuit, the capacitor voltages and inductor currents in the circuit
respond to the area of the input pulse, and are almost independent of the shape
of the pulse.
In solving Equation 10.161, we set aside a special case which is of some
interest in a broader context. The question is, does the circuit behave in some
bizarre fashion if the drive pulse has the same time constant as the circuit? One
might be led to think so from Equation 10.163, because for τ1 = RC, the
denominator goes to zero. To find the correct answer, assume τ2 = τ1 = τ in
Equation 10.161 and solve:

e−t/τ t
vC = Adt (10.165)
τ 0

te−t/τ
=A . (10.166)
τ

This waveform looks much like that shown in Figure 10.47c.

558c
e x a m p l e 12. 5 g r a p h i c a l i n t e r p r e t√a t i o n This example
studies an interesting
 graphical interpretation of ωo = 1/ LC and the characteristic
impedance L/C.
√ 
Figure 12.13 shows contours of constant ω◦ ≡ 1/ LC and constant L/C in the L C
plane over practical ranges for L and C. These contours are straight lines in the figure
owing tothe logarithmic scales of the figure. This figure is particularly useful for finding
ω◦ and L/C for a given L and C, and vice versa, for example.

1 L
------------ ----
... LC C
... ... ...
rad rad rad rad rad rad
107 -------- 7 106 ---------- 6 105 ---------- 5 104 ---------- 4 103 ---------- 102 ----------
s 10 Ω s 10 Ω s 10 Ω s 10 Ω s 1000 Ω s 100 Ω
1H
rad 10 Ω
108 --------
s

1Ω
1 mH

rad 0.1 Ω
1010 --------
s

L
1 mH
rad 0.01 Ω
1011 --------
s

10–3 Ω
1 nH

rad 10–4 Ω
1013 --------
s

1 pH
1 fF 1 pF 1 nF 1 mF 1 mF
C

√ 
F I G U R E 12.13 Contours of constant ω◦ ≡ 1/ LC and constant L/C in the L–C plane.

640a
12.4 U N D R I V E N , P A R A L L E L R L C C I R C U I T *
v We will now analyze the undriven parallel RLC circuit shown in Figure 12.24,
+
i
L which is copied from Figure 2.14a. To analyze the behavior of this circuit
we can again employ the node method, and this analysis closely parallels
vC C R L
that of Section 12.1. As in Figure 12.6, a ground node is already selected
-
in Figure 12.24, and the unknown node voltage v is already labeled. So, we
may again proceed immediately to Step 3 of the node method. Here, we write
KCL in terms of v for the node at which v is defined. This yields
F I G U R E 12.24 The parallel
second-order RLC circuit shown in  t
dv(t) 1 1
Figure 2.14a. C + v(t) + v(t̃)dt̃ = 0. (12.81)
dt R L −∞

The first term in Equation 12.81 is the capacitor current, the second term is the
resistor current, and the third term is the inductor current. Because the circuit
contains an inductor, Equation 12.81 contains a time integral. To remove this
integral, we differentiate Equation 12.81 with respect to time, and also divide
by C, to obtain

d2 v(t) 1 dv(t) 1
+ + v(t) = 0, (12.82)
dt2 RC dt LC
which is easier to work with.
To complete the node analysis, we complete Steps 4 and 5 by solving
Equation 12.82 for v. Then, we use v to determine the other branch variables
of interest, for example, iL and vC . Like Equation 12.4, Equation 12.82 is
an ordinary second-order linear differential equation with constant coefficients.
Since the circuit does not have a drive, its homogeneous solution is also the
complete solution. Therefore, as with Equation 12.4, we expect its solution to
be a superposition of two terms of the form

Ae st .

The substitution of this candidate term into Equation 12.82 yields


 
1 1
A s2 + s+ e st = 0 (12.83)
RC LC

from which it follows that

1 1
s2 + s+ = 0. (12.84)
RC LC

Equation 12.84 is the characteristic equation of the circuit. To simplify Equa-


tion 12.84, and to put it in a form that is more standard for the characteristic

654a
equation in second-order circuits, we write it as

s 2 + 2αs + ω◦2 = 0 (12.85)


where
1
α≡ (12.86)
2RC

1
ω◦ ≡ ; (12.87)
LC
note that Equation 12.87 is the same as Equation 12.9. Equation 12.85 is
a quadratic equation having two roots. Those roots are
!
s1 = −α + α 2 − ω◦2 (12.88)
!
s2 = −α − α 2 − ω◦2 . (12.89)

Therefore, the solution for v is a linear combination of the two functions e s1 t


and e s2 t , and takes the form

v(t) = A1 e s1 t + A2 e s2 t (12.90)

where A1 and A2 are as yet unknown constants that are equivalent to the two
constants of integration encountered when integrating Equation 12.82 twice to
find v.
To complete the solution to Equation 12.82 we must again determine A1
and A2 from initial conditions vC and iL specified at t = 0, To do so, note that

vC (t) = v(t). (12.91)

Further, from KCL applied to either node, that is, from Equation 12.81,

1 dv
iL (t) = − v(t) − C . (12.92)
R dt

Equations 12.91 and 12.92 can be solved to determine v and dv/dt in terms of
iL and vC . Doing so, and evaluating the result at t = 0, then yields

v(0) = vC (0) (12.93)


dv 1 1
(0) = − iL (0) − vC (0). (12.94)
dt C RC

654b
Now, to find A1 and A2 , we evaluate Equation 12.90 and its derivative at
t = 0, and set the results equal to Equations 12.93 and 12.94, respectively.
This results in

v(0) = A1 + A2 = vC (0) (12.95)


dv
(0) = s1 A1 + s2 A2
dt (12.96)
1 1
= − iL (0) − vC (0).
C RC

Equations 12.95 and 12.96 can be jointly solved for A1 and A2 to obtain

(1 + RCs2 )vC (0) + RiL (0) s1 vC (0) − iL (0)/C


A1 = = (12.97)
RC(s2 − s1 ) (s1 − s2 )

(1 + RCs1 )vC (0) + RiL (0) s2 vC (0) − iL (0)/C


A2 = = (12.98)
RC(s1 − s2 ) (s2 − s1 )
where we have used the fact that both s1 and s2 satisfy Equation 12.84, and the
fact that LCs1 s2 = 1, to obtain the second equalities. Finally, Equations 12.97
and 12.98 can now be substituted into Equation 12.90 to obtain

s1 vC (0) − iL (0)/C s2 vC (0) − iL (0)/C


v(t) = e s1 t + e s2 t . (12.99)
(s1 − s2 ) (s2 − s1 )

Further, the substitution of Equation 12.99 into Equations 12.91 and 12.92
yields

s1 vC (0) − iL (0)/C s2 vC (0) − iL (0)/C


vC (t) = e s1 t + e s2 t (12.100)
(s1 − s2 ) (s2 − s1 )

 
1 + RCs1 s1 vC (0) − iL (0)/C
iL (t) = − e s1 t
R (s1 − s2 )
 
1 + RCs2 s2 vC (0) − iL (0)/C
− e s2 t
R (s2 − s1 )
vC (0)/L − s2 iL (0) vC (0)/L − s1 iL (0)
= e s1 t + e s2 t (12.101)
(s1 − s2 ) (s2 − s1 )

as the states of the parallel circuit. To obtain the second equality in


Equation 12.101 we have again used the fact that both s1 and s2 satisfy

654c
Equation 12.84, and the fact that LCs1 s2 = 1. This completes the formal
node analysis of the circuit shown in Figure 12.24.
We will now close this subsection by examining the dynamic behavior of
vC and iL for the same three cases defined in Section 12.2. Those are the cases
of under-damped, critically-damped, and over-damped dynamics. As we shall
see, the dynamics of the series circuit are essentially identical to those of the
parallel circuit for all three cases, except for the details of the role of R. In the
series circuit, small R caused light damping while large R caused heavy damping.
This role reverses for the parallel circuit because it is in the limit of large R that
Figure 12.24 reduces to Figure 12.6.

12.4.1 U N D E R - D A M P E D D Y N A M I C S
The case of under-damped dynamics is characterized by

α < ω◦

or, after substitution of Equations 12.86 and 12.87, by


2R > L/C.

As R becomes large, the corresponding resistor approaches an open circuit,


and so the circuit shown in Figure 12.24 approaches the LC circuit shown in
Figure 12.6. Therefore, we should expect the under-damped dynamics to be
oscillatory in nature. As we shall see shortly, this is indeed the case.
With α < ω◦ , the quantity inside the radicals in Equations 12.88 and 12.89
is negative, and so the natural frequencies s1 and s2 are complex numbers. To
simplify matters, let us again define ωd according to
!
ωd ≡ ωo2 − α 2 . (12.102)

With this definition, s1 and s2 from Equations 12.88 and 12.89 can be
written as

s1 = −α + jωd (12.103)
s2 = −α − jωd . (12.104)

The real and imaginary parts of s1 and s2 are now more apparent.
Since s1 and s2 are now complex, the exponentials in Equations 12.100
and 12.101 are also complex. Thus, iL and vC will exhibit both oscillatory and
decaying behavior. To see this, we substitute Equations 12.103 and 12.104 into

654d
Equations 12.100 and 12.101, and use the Euler relation to obtain
 
αCvC (0) + iL (0)
vC (t) = vC (0)e−αt cos(ωd t) − e−αt sin(ωd t)
Cωd
 
αCvC (0) + iL (0) 2 −αt
= vC2 (0) + e
Cωd
  
αCvC (0) + iL (0)
× cos ωd t + tan−1 (12.105)
Cωd vC (0)

 
−αt vC (0) + αLiL (0)
iL (t) = iL (0)e cos(ωd t) + e−αt sin(ωd t)
Lωd
 
vC (0) + αLiL (0) 2 −αt
= iL2 (0) + e
Lωd
  
Lωd iL (0)
× sin ωd t + tan−1 . (12.106)
vC (0) + αLiL (0)

Sketches of iL and vC are shown in Figure 12.25 for the special case of

iL (0) = 0.

As was the case for the series circuit, the states in the parallel circuit display
oscillatory and decaying behavior. It is also the case that Equations 12.105
and 12.106 reduce to Equations 12.21 and 12.22, respectively, as the circuit
damping characterized by α vanishes. The difference here is that this occurs as
R → ∞ because it is in this limit that Figure 12.24 reduces to Figure 12.6.
A comparison of Equations 12.105 and 12.106 with Equations 12.63 and
12.64 shows that the under-damped dynamics of the parallel and series circuits
are quite similar. This is to be expected because their characteristic equations are
identical. It is for this reason that α, ω◦ , ωd , and Q have the same interpretations
for the two circuits. Our comments concerning stored energy also hold for both
circuits. Therefore, we will not repeat the details here. Rather, we will identify
three important differences. The first difference, which has been mentioned
already, is the reversed role of R. A large R in the series circuit corresponds
to a small R in the parallel circuit and vice versa. The second difference is the
evaluation of the quality factor Q. While Equation 12.65 still holds for the
parallel circuit, that is,
ω◦
Q≡ , (12.107)

654e
vC

vC(0)
-αt
e

| | | | | | | | | | | | | | | |
0 π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd

t
|

F I G U R E 12.25 Waveforms of
vC and iL in undriven, parallel RLC
circuit for the case of iL (0) = 0.
iL

Io

0
|

π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd


t

Equation 12.66 does not. Rather, for the parallel circuit shown in Figure 12.24,
the substitution of Equations 12.86 and 12.87 into Equation 12.65 yields

1 L
Q= (12.108)
R C

or
ωo L
Q= . (12.109)
R

654f
The third difference is the role of φ in Figure 12.17. For the parallel circuit,
assuming iL (0) = 0 in Equations 12.105 and 12.106, vC is advanced with
respect to iL by quadrature plus the additional angle φ, where φ = tan−1 (α/ωd ).

12.4.2 O V E R - D A M P E D D Y N A M I C S
As with the case of the series circuit, the case of over-damped dynamics is
characterized by
α > ω◦
or, after substitution of Equations 12.86 and 12.87, by

2R < L/C.

In this case, the quantity inside the radicals in Equations 12.88 and 12.89 is
positive, and so both s1 and s2 are real. For this reason, the dynamic behavior
of iL and vC , as expressed by Equations 12.100 and 12.101, does not exhibit
oscillation. Rather, it involves two real exponential functions that decay at
different rates, as the two equations show. The expressions for vC and iL for
the case of iL (0) = 0 with over-damping are obtained from Equations 12.100
and 12.101, and are shown here:

s1 vC (0) s2 vC (0)
vC (t) = e s1 t + e s2 t (12.110)
(s1 − s2 ) (s2 − s1 )

vC (0) vC (0)
iL (t) = e s1 t + e s2 t . (12.111)
L(s1 − s2 ) L(s2 − s1 )
Since α > ω◦ for over-damped circuits, note that s1 and s2 are both real in these
two equations. 
As R becomes small, in particular smaller than 1/2 L/C, it becomes a
significant short circuit across the capacitor and inductor. In this way it diverts
the oscillating current that the capacitor and inductor share for larger values of R.
As a consequence, the energy exchange between the capacitor and inductor is
interrupted, and the circuit ceases to oscillate. Instead, its behavior is more
like that of an independent capacitor and an independent inductor discharging
through the resistor. To see this, let us determine the asymptotic values of s1
and s2 as R becomes small and hence as α becomes large. They are
 
!  2
ω◦  ≈ α −ω◦ = −R
2
s1 = −α + α 2 − ω◦2 = α −1 + 1−
α 2α 2 L
(12.112)

654g
 
!  2
ω◦  ≈ −2α = −1 . (12.113)
s2 = −α − α 2 − ω◦2 = α −1 − 1−
α RC

As expected the corresponding time constants approach RC and L/R, the


time constants of an independent capacitor-resistor circuit and an independent
inductor-resistor circuit. Note that, for over-damped dynamics, α > ω◦ from
which it follows that RC is the faster time constant and L/R is the slower time
constant.

12.4.3 C R I T I C A L L Y - D A M P E D D Y N A M I C S
The case of critically-damped dynamics is characterized by

α = ω◦ .

In this case, it follows from Equations 12.88 and 12.89 that

s1 = s2 = −α

and that the characteristic equation, Equation 12.85, has a repeated root.
Because of this, e s1 t and e s2 t are no longer independent functions, and so the
general solution for v is no longer the superposition of these two functions as
given by Equation 12.90. Rather, it is again the superposition of the repeated
exponential function

e s1 t = e s2 t = e−αt and te−αt .

From this observation, and Equations 12.91 and 12.92, it follows that vC and
iL will exhibit similar behavior.
Perhaps the easiest way to determine vC and iL for the case of critical
damping is to evaluate Equations 12.105 and 12.106 under the conditions of
that case. To do so, observe from Equation 12.102 that, for critical damping
ω◦ = α, and so ωd = 0. Therefore, we can obtain vC and iL for the case
of critical-damping by evaluating Equations 12.105 and 12.106 in the limit
ωd → 0. This results in

αCvC (0) + iL (0)


vC (t) = vC (0)e−αt − te−αt (12.114)
C

vC (0) + αLiL (0)


iL (t) = iL (0)e−αt + te−αt . (12.115)
L
From Equations 12.114 and 12.115 we see that vC and iL contain both the
decaying exponential function e−αt and the function te−αt , as expected.

654h
12.6 D R I V E N , P A R A L L E L R L C C I R C U I T *
Consider now the circuit shown in Figure 12.50. As in previous sections of
this chapter, we will analyze the behavior of this circuit using the node method
beginning at Step 3. In doing so, we will follow the analysis presented in
Section 12.4 very closely.
We begin by completing Step 3, of the node method. To do so, we write
KCL in terms of vC for the node at which vC is defined to obtain
 t
dvC (t) 1 1
C + vC (t) + vC (t̃)dt̃ = iIN (t), (12.198)
dt R L −∞

which upon differentiation and division by C becomes

d2 vC (t) 1 dvC (t) 1 1 diIN (t)


+ + vC (t) = . (12.199)
dt2 RC dt LC C dt

Unlike Equations 12.4, 12.82 and 12.40, Equation 12.199 is an inhomogeneous


differential equation because it is driven by the external signal iIN . Unfortunately,
iIN enters Equation 12.199 through a derivative, which poses an unnecessary
complication. To eliminate this complication, we substitute the constitutive law
for the inductor,
diL (t)
vC (t) = L
, (12.200)
dt
into Equation 12.198 and divide by LC. This yields

d2 iL (t) 1 diL (t) 1 1


+ + iL (t) = iIN (t), (12.201)
dt2 RC dt LC LC

which is easier to work with.


Equation 12.201 is an inhomogeneous differential equation, which unlike
our previous undriven examples (for example, Equation 12.4 for the undriven

vC

iL
F I G U R E 12.50 The parallel +
second-order circuit with a resistor, vC
iIN C R L
capacitor, inductor, and current -
source.

678a
LC circuit) has an additional term for the input drive. Furthermore, notice the
term proportional to diL /dt. As we saw in Section 12.5, this term modifies
the homogeneous response to include damping. Therefore we now expect the
oscillations in the step and impulse responses to decay in time.
To complete the node analysis, we complete Steps 4 and 5 by solving
Equation 12.201 for iL , and using it to determine vC and any other variables of
interest. To do so we employ our usual method of solving differential equations:

1. Find the homogeneous solution iLH (t).


2. Find the particular solution iLP (t).
3. The total solution is then the sum of the homogeneous solution and the
particular solution as follows:

iL (t) = iLH (t) + iLP (t).

Use the initial conditions to solve for the remaining constants.


The homogeneous solution iLH (t) to Equation 12.201 is obtained by solving
the differential equation with the drive iIN ≡ 0. With iIN ≡ 0, the circuit
shown in Figure 12.50 is identical to the parallel, undriven RLC circuit shown
in Figure 12.24, and so the two circuits have the same homogeneous equation.
The homogeneous equation in terms of the current is given by

d2 iLH (t) 1 diLH (t) 1


+ + iLH (t) = 0. (12.202)
dt2 RC dt LC

Note the similarity between this homogeneous equation and Equation 12.4
for the undriven, parallel RLC circuit. Following the solution (Equation 12.10) of
the homogeneous equation for the undriven, parallel RLC circuit, we can write
the form of the homogeneous solution for our driven, parallel RLC circuit as

iLH (t) = K1 e s1 t + K2 e s2 t (12.203)

where K1 and K2 are as yet unknown constants that will be determined from
the initial conditions after the total solution has been formed. s1 and s2 , the
roots of the characteristic equation,

s 2 + 2αs + ω◦2 = 0 (12.204)


1
α≡ (12.205)
2RC
1
ω◦ ≡ . (12.206)
LC

678b
The roots are given by
!
s1 = −α + α 2 − ω◦2 (12.207)
!
s2 = −α − α 2 − ω◦2 . (12.208)

As observed with other second-order circuits, the circuit exhibits under-damped,


over-damped, or critically-damped behavior depending on the relative values of
α and ω◦ :
α < ω◦ ⇒ under-damped dynamics;
α = ω◦ ⇒ critically-damped dynamics;
α > ω◦ ⇒ over-damped dynamics.
For brevity, the rest of the section will assume that

α < ω◦

so that the circuit displays under-damped dynamics. For the under-damped


case, since s1 and s2 are now complex, they can be written explicitly in complex
form as

s1 = −α + jωd
s2 = −α − jωd (12.209)

where
!
ωd ≡ ω◦2 − α 2 . (12.210)
As we did with the series RLC circuit, we shall rewrite the homogeneous
solution in Equation 12.121 into a more intuitive form using the Euler relation
as follows:
iLH (t) = A1 e−αt cos(ωd t) + A2 e−αt sin(ωd t) (12.211)
where A1 and A2 are unknown constants we will evaluate later depending on
the initial conditions of the circuit.
Next, we need to find iLP (t). Knowing it, we can write the total solution as

iL (t) = iLP (t) + iLH (t) = iLP (t) + A1 e−αt cos(ωd t) + A2 e−αt sin(ωd t).
(12.212)

At this point, only iLP , and A1 and A2 , remain as unknowns.


We will now proceed to find the iLP , and then A1 and A2 . iLP depends
on the input drive. We will find iLP for two cases of iIN , namely a step and

678c
an impulse. That is, we will proceed to find the step response and the impulse
response of the circuit. To simplify matters, we will assume that the circuit is
under-damped, that both the step and the impulse occur at t = 0, and that the
circuit is initially at rest prior to that time. The latter assumption implies that
we are seeking the zero-state response for which

iL (0) = 0 (12.213)

and
vC (0) = 0. (12.214)
The zero-state response is the response of the circuit for zero initial state.
Equations 12.213 and 12.214 provide the initial conditions for the solution
of Equation 12.201 after the step and impulse occur, that is, for t > 0.

12.6.1 S T E P R E S P O N S E
Let iIN be the current step given by iIN
Io
iIN (t) = I◦ u(t) (12.215)
t
and shown in Figure 12.51. With the substitution of Equation 12.215, 0
Equation 12.201 becomes
F I G U R E 12.51 A current step
d2 iL (t) 1 diL (t) 1 1 input.
+ + iL (t) = I◦ (12.216)
dt2 RC dt LC LC

for t > 0. Any function that satisfies Equation 12.216 for t > 0 is an acceptable
iLP . One such function is
iLP (t) = I◦ . (12.217)
Thus, we have the particular solution for a step input.
The total solution is given by summing the homogeneous solution (Equa-
tion 12.211) and the particular solution (Equation 12.217) as

iL (t) = I◦ + A1 e−αt cos(ωd t) + A2 e−αt sin(ωd t), (12.218)

again for t > 0. Additionally, the substitution of Equation 12.218 into


Equation 12.200 yields

vC (t) = (ωd LA2 − αLA1 )e−αt cos(ωd t) − (ωd LA1 + αLA2 )e−αt sin(ωd t),
(12.219)

also for t > 0. Now only A1 and A2 remain as unknowns.

678d
In Chapter 9, we saw that the voltage across a capacitor is continuous
unless the current through it contains an impulse. We also saw that the cur-
rent through an inductor is continuous unless the voltage across it contains an
impulse. Since iIN contains no impulses, we can therefore assume that both vC
and iL are continuous across the step at t = 0. Consequently, since both states
are zero for t ≤ 0, Equations 12.218 and 12.219 must both evaluate to zero as
t → 0. This observation allows us to use the initial conditions to determine A1
and A2 . Evaluation of both equations as t → 0, followed by the substitution of
the initial conditions, yields

iL (0) = I◦ + A1 = 0 (12.220)
vC (0) = ωd LA2 − αLA1 = 0. (12.221)

Equations 12.220 and 12.221 can be solved to yield

A1 = −I◦ (12.222)
α
A2 = − I◦ . (12.223)
ωd

Finally, the substitution of Equations 12.222 and 12.223 into Equations 12.218
and 12.219 yields
   
ω◦ −αt −1 α
iL (t) = I◦ 1 − e cos ωd t − tan u(t) (12.224)
ωd ωd
I◦
vC (t) = e−αt sin(ωd t)u(t); (12.225)
ωd C

Equations 12.210 and 12.206 have also been used to simplify the results.
Note that the unit step function u has been introduced into Equations 12.224
and 12.225 so that they are valid for all time. The validity of Equations
12.224 and 12.225 can be demonstrated by observing that they satisfy the
initial conditions, and Equations 12.201 and 12.200, respectively, for all time.
Because they do, our assumption that the states are continuous at t = 0 is
justified.
Figure 12.52 shows iL and vC as given by Equations 12.224 and 12.225.
As expected, the ringing in both states now decays as t → ∞. This decay is
well characterized by the quality factor Q, as defined in Equation 12.66 and
discussed shortly thereafter. In fact, because the circuits shown in Figures 12.24
and 12.50 have the same homogeneous response, the entire discussion of α,
ωd , and ω◦ given in Section 12.4 applies here as well. In fact, the series and
parallel circuits are duals. This can be observed by comparing the evolution of
their branch variables. For example, like the capacitor voltage vC in the series
circuit, iL undergoes nearly a two-fold overshoot during the initial transient.

678e
vC

e-αt

0
π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd

t
|

F I G U R E 12.52 iL and vC for the


parallel RLC circuit shown in
Figure 12.50 for the case of a step
iL

input through iIN .

Io

0
π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd

Another observation concerns the short-time behavior of the circuit. We


have seen in Chapter 10 that the transient behavior of an uncharged capacitor
is to act as a short circuit during the early part of a transient, while the cor-
responding transient behavior of an uncharged inductor is to act as an open
circuit. This behavior is observed in Figure 12.52 since iIN is carried entirely

678f
by the capacitor (and iL is 0) at the start of the transient, and vC ramps up
correspondingly.
A related observation concerns the long-time behavior of the circuit. We
have also seen in Chapter 10 that the transient behavior of a capacitor is to
act as an open circuit as t → ∞, while the corresponding transient behavior
of an inductor is to act as a short circuit. This behavior is also observed in
Figure 12.52, since iIN is carried entirely by iL as t → ∞.
We also note the overshoot of iL above the input current step of I◦ during
the transient. Although the average value of iL is close to I◦ during the transient,
the peak value is closer to 2I◦ .
Finally, note that as t → ∞, iIN is carried entirely by the inductor since
iL → I◦ . This is consistent with the relative long-time transient behavior of the
inductor, resistor, and capacitor.

12.6.2 I M P U L S E R E S P O N S E
Let iIN be the impulse given by

iIN = Q◦ δ(t) (12.226)

iIN as shown in Figure 12.53. Because iIN is an impulse, it vanishes for t > 0.
Therefore, Equation 12.201 reduces to a homogeneous equation for t > 0, and
Qo so the simplest acceptable particular solution is

iLP (t) = 0. (12.227)

0
The substitution of Equation 12.227 into Equation 12.212 now yields
t

F I G U R E 12.53 The current


iL (t) = A1 e−αt cos(ωd t) + A2 e−αt sin(ωd t) (12.228)
impulse iIN .
again for t > 0. Additionally, we can obtain vC (t) by using

diL (t)
vC (t) = L
dt
as

vC (t) = (LA2 ωd − LαA1 )e−αt cos(ωd t) − (LA1 ωd + LαA2 )e−αt sin(ωd t)


(12.229)

also for t > 0. Now only A1 and A2 remain as unknowns.


From this discussion, it is apparent that the role of the impulse in iIN is to
establish the initial conditions for a subsequent homogeneous response. This,

678g
incidentally, might explain how the circuit shown in Figure 12.24 began its
operation.
As mentioned earlier during our discussion of the step response, the tran-
sient behavior of an uncharged capacitor is to act as a short circuit during
the early part of a transient, while the corresponding transient behavior of an
uncharged inductor is to act as an open circuit. Because of this the impulse in
iIN passes entirely through the capacitor while iL remains zero at t = 0. An
important consequence of this is that the charge Q◦ delivered by iIN is delivered
entirely to the capacitor, and so vC steps to Q◦ /C at t = 0. This establishes the
initial conditions after the impulse needed to determine A1 and A2 . The evalu-
ation of Equations 12.228 and 12.229 as t → 0, followed by the substitution
of these initial conditions (vC (0) = Q◦ /C and iL (0) = 0), yields

iL (0) = A1 = 0 (12.230)
Q◦
vC (0) = ω◦ LA2 = . (12.231)
C

Equations 12.230 and 12.231 can be rearranged to yield

A1 = 0 (12.232)
Q◦
A2 = . (12.233)
LCωd

Finally, the substitution of Equations 12.232 and 12.233 into Equations 12.228
and 12.229 yields

Q◦ ωo2
iL (t) = e−αt sin(ωd t) (12.234)
ωd
  
Q◦ ω◦ −αt −1 α
vC (t) = e cos ωd t + tan u(t), (12.235)
C ωd ωd

where the unit step function u has been introduced into Equations 12.234 and
12.235 so they are valid for all time.
Note that our solution in Equation 12.234 satisfies the initial conditions
established by the impulse, and that it satisfies Equation 12.201. Because it does,
it justifies our interpretation of the circuit behavior at t = 0. The waveforms for
vC and iL are as shown in Figure 12.54.
It is interesting to note that the impulse response of the circuit can also be
obtained from the step response. The circuit shown in Figure 12.50 is a linear
circuit. Therefore, since the impulse iIN given in Equation 12.226 is a scaled
derivative of the step iIN given in Equation 12.215, it follows that the impulse
response is the same scaled derivative of the step response. (See Section 10.6.2

678h
vC
Qo/C
e-αt

0
π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd

t
|

F I G U R E 12.54 Waveforms of iL
and vC for the parallel RLC circuit
iL

for a short pulse input.


e-αt

0
π/ωd 2π/ωd 3π/ωd 4π/ωd 5π/ωd 6π/ωd 7π/ωd

for a more detailed discussion on the use of linearity to obtain responses to


the derivative or the integral of an input, once the response to the input is
known.)
To be more specific, iIN as given in Equation 12.226, can be constructed
by applying (Q◦ /I◦ )d/dt to iIN as given in Equation 12.215. In other words,

d
Q◦ δ(t) = (Q◦ /I◦ ) I◦ u(t).
dt

Therefore, the same operator may be applied to Equations 12.234 and 12.235
to determine iL and vC respectively, for the impulse response. Thus, we can

678i
obtain the impulse response by differentiating the step response. Thus, applying
the operator (Q◦ /I◦ )d/dt to the step response, we obtain
     
Q◦ d ω◦ −αt α
iL (t) = I◦ 1 − e cos ωd t − tan−1 u(t)
I◦ dt ωd ωd
  
α
= ω◦ Q◦ e−αt sin ωd t − tan−1 u(t)
ωd
  
αω◦ Q◦ −αt α
+ e cos ωd t − tan−1 u(t)
ωd ωd
   
ω◦ −αt α
+ Q◦ 1 − e cos ωd t − tan−1 δ(t)
ωd ωd
ω◦2
= Q◦ e−αt sin(ωd t)u(t) (12.236)
ωd

 
Q◦ d I◦ −αt
vC (t) = e sin(ωd t)u(t)
I◦ dt ωd C
Q◦ αQ◦ Q◦
= e−αt cos(ωd t)u(t) − e−αt sin(ωd t)u(t) + e−αt sin(ωd t)δ(t)
C ωd C ωd C
  
Q◦ ω◦ α
= e−αt cos ωd t + tan−1 u(t) (12.237)
C ωd ωd

as the impulse response of the circuit. Note that terms involving the impulse
δ vanish in Equations 12.236 and 12.237 because δ is itself zero everywhere
except t = 0, and the coefficients of the impulse are both zero at t = 0.
From this experience with the impulse, we can see that the impulse response
of the circuit is essentially a homogeneous response. Thus this response is
identical to that studied in Section 12.4. In fact, the role of the impulse is to
establish initial conditions for the subsequent homogeneous response. As we
argued in Section 12.6, the current impulse passes entirely through the capacitor
delivering its charge in the process. Therefore, vC steps to Q◦ /C as iL remains
zero. As a result, for t > 0, the impulse response described by Equations
12.236 and 12.237 are identical to Equations 12.105 and 12.106, respectively,
with vC (0) replaced by Q◦ /C, and iL (0) replaced by zero. Therefore, the entire
discussion of the circuit shown in Figure 12.24 is applicable. Not surprisingly,
note that vC and iL for the impulse as shown in Figure 12.54 are the same as
those in Figure 12.25 with vC (0) replaced by Q◦ /C.

678j
12.10 S T A T E - S P A C E A N A L Y S I S *
A state-variable analysis naturally results in a set of coupled first-order differ-
ential equations. The results of a node analysis can also be expressed in this
way. Therefore, it is worth exploring the direct solution of coupled first-order
differential equations. In the case of a state-variable analysis, this approach to
solving the equations eliminates the need to perform their back substitution in
order to obtain a single high-order differential equation.
When working with a set of coupled first-order differential equations it
is common to present them in the vector format referred to as a state-space
format, and to solve them with the corresponding mathematical mechanics.
We will refer to this method of solution as a state-space analysis of the dif-
ferential equations. It is beyond our scope to present a detailed treatment of
state space analysis.10 Rather, we will summarize its mechanics through an
example.
To illustrate the mechanics of a state-space analysis, consider again the
circuit shown in Figure 12.50. In Section 12.9 we completed a state-variable
analysis of this circuit. The resulting state equations are given in Equations
12.253 and 12.254. Those equations may be summarized in state-space format
according to
      
1
d  vC (t)   − RC −1 vC (t)
1
= C   +  C  [iIN (t)] . (12.255)
dt iL (t) 1 iL (t)
0 0
L

Equation 12.255 is an example of the more general linear time-invariant state-


space format
dx(t)
= Ax(t) + Bz(t). (12.256)
dt
Here, x is referred to as the state vector, z is referred to as the input vector, A
is referred to as the state matrix, and B is referred to as the input matrix. In the
case of Equation 12.255,

vC (t)
x(t) = (12.257)
iL (t)
z(t) = iIN (t) (12.258)

10. For a detailed treatment of this analysis, see G. Strang, Linear Algebra and Its Applications,
Au:Please provide dates Third Edition, Chapter 5, Academic Press; or Finizio and Ladas, Ordinary Differential Equations,
of publication for these with Modern Applications, Second Edition, Section 3.3.

691a
 
1
− −1
A= RC C  (12.259)
1
0
L
1

B= C . (12.260)
0

It is interesting to note that the results of a node analysis can also be expressed in
state-space format. This is important at least because most commercial numer-
ical analysis packages focus on the solution of differential equations presented
in this format. The circuit shown in Figure 12.50 was analyzed by the node
method in Section 12.6, and the resulting differential equation is given in Equa-
tion 12.201. To put Equation 12.201, which is of order two, into state-space
format, we use iL and diL /dt as the two states within x, and write
   
d iL (t) 0 1 iL (t) 0
diL (t) = 1 1 diL (t) + 1 [iIN (t)] . (12.261)
dt − −
dt LC RC dt LC

Thus, the results of both a state-variable analysis and a node analysis can be put
into the standard state-space format. However, note that Equations 12.255
and 12.261 have different values for x, A and B. Because Equations 12.255 and
12.261 describe the dynamics of the same circuit, it is also apparent that there
exists more than one state-space representation of the dynamics of a given
circuit.
Since Equations 12.255 and 12.261 describe the dynamics of the same cir-
cuit, their solutions will ultimately yield the same branch variables. Therefore,
for the sake of brevity, we will now consider the state-space analysis of only
Equation 12.255; the analysis of Equation 12.261 proceeds in an identical man-
ner. To be specific about the solution of Equation 12.255 we will consider its
response to a step input at t = 0. That is, we will assume that

iIN (t) = I◦ u(t) (12.262)

as shown in Figure 12.51. To simplify matters, we will also assume that the
circuit is at rest prior to the step. That is, we will assume that both vC and iL are
zero for t ≤ 0. This information provides the initial conditions for the solution
of Equation 12.255 after the step occurs, that is, for t > 0. To solve Equa-
tion 12.255 we now proceed essentially as we did in the earlier sections of this
chapter. That is, we break the solution into two parts, namely a particular solu-
tion, xP , and a homogeneous solution, xH . The particular solution will satisfy
Equation 12.255 without regard for initial conditions, and the homogeneous
solution will match the initial conditions.

691b
Consider first the homogeneous solution. In general, the homogeneous
solution is the solution to Equation 12.256 with z ≡ 0. Therefore, xH satisfies

dxH (t)
= AxH (t). (12.263)
dt

Since Equation 12.263 represents a set of ordinary first-order homogeneous


linear differential equations with constant coefficients, we expect its solution to
be a superposition of terms of the form re st where r is a constant vector having
the same dimension as x. The substitution of this candidate term into Equation
12.263 yields

(sI − A)re st = 0 (12.264)

where I is the identity matrix of appropriate dimension. Since e st is never zero


for finite st, it follows that

(sI − A)r = 0. (12.265)

Further, since r = 0 is a trivial solution that leads to x = 0, we are interested


only in solutions to Equation 12.265 for which r  = 0. For Equation 12.265 to
be satisfied for a nonzero r, the matrix (sI − A) must be singular, hence

det(sI − A) = 0. (12.266)

Equation 12.266 is a polynomial in s, referred to as the characteristic equation


of A, and its roots are the eigenvalues of A. For each root, there is a corre-
sponding r that is the corresponding right eigenvector of A.11 To determine
each eigenvector, the corresponding root of Equation 12.266 is substituted
into Equation 12.265, and the resulting equation is solved for r to within an
arbitrary scale factor.
With the substitution of Equation 12.259, Equation 12.266 becomes

 
1 1
s+
det(sI − A) = det  RC C  = s 2 + 1 s + 1 = 0, (12.267)
−1 s RC LC
L

which is the same as Equation 12.84. Thus, we see that the characteristic
equation of A is the same as the characteristic equation of the circuit, and that

11. Here, we ignore the degenerate case in which repeated eigenvalues share a single eigenvector.

691c
the eigenvalues of A are the natural frequencies of the circuit. Equation 12.267
has two roots and they are given by
!
s1 = −α + α 2 − ω◦2 (12.268)
!
s2 = −α − α 2 − ω◦2 (12.269)
1
α≡ (12.270)
2RC
1
ω◦ ≡ √ (12.271)
LC

just as given in Equations 12.125 and 12.126. This is expected because the
homogeneous version of the circuit shown in Figure 12.50 is the circuit shown
in Figure 12.24. Finally, the corresponding eigenvectors, determined from
Equation 12.265 for each eigenvalue, are

1
r1 = 1 (12.272)
s1 L

1
r2 = 1 . (12.273)
s2 L

With these results we can now assemble the homogeneous solution of Equation
12.255. It is given by
 
1 1
s1 t s2 t s1 t
xH (t) = A1 r1 e + A2 r2 e = A1 1 e + A2 1 e s2 t (12.274)
s1 L s2 L

where A1 and A2 are two coefficients that depend on initial conditions.


Consider now the particular solution. In general, it is the solution to Equa-
tion 12.256 without regard for the initial conditions. Thus, any function xP that
satisfies
dxP (t)
= AxP (t) + Bz(t) (12.275)
dt
is an acceptable particular solution. To treat the general case of the step input,
we let
z(t) = Z◦ u(t). (12.276)
The substitution of Equation 12.276 into Equation 12.275 yields

dxP (t)
= AxP (t) + BZ◦ (12.277)
dt

691d
for t > 0. Assuming that A−1 exists, one solution to Equation 12.277 is

xP (t) = −A−1 BZ◦ , (12.278)

again for t > 0. With the substitution of Equations 12.262 and 12.276,
Equation 12.258 becomes
Z◦ = I◦ . (12.279)
Then, with the substitution of Equations 12.259, 12.260, and 12.279, Equa-
tion Equation 12.278 becomes
 −1  1

1
− −1 C
xP (t) = −  RC C    I◦
1
0 0
L
  
0 L 1 0
=− C I◦ = , (12.280)
−C −L 0 I◦
R

also for t > 0.


We can now combine the particular and homogeneous solutions with the
initial conditions to solve Equation 12.255. To begin, the superposition of
Equations 12.274 and 12.280 yields

vC (t)
= x(t) = xP (t) + xH (t)
iL (t)
  
0 1 1
s1 t
= + A1 1 e + A2 1 e s2 t . (12.281)
I◦ s1 L s2 L

Now, A1 and A2 are the only unknowns. To find A1 and A2 , we use the initial
conditions. Since iIN contains no impulses, we can assume that both vC and iL
are continuous at t = 0. Consequently, since both states are zero for t ≤ 0,
Equation 12.281 must evaluate to zero as t → 0. This yields
    
vC (0) 0 1 1 0
= + A1 1 + A2 1 = . (12.282)
iL (0) I◦ s1 L s2 L
0

The two rows of Equation 12.282 are two equations that can be solved for
A1 and A2 . Doing so yields

s1 s2 LI◦ I◦
A1 = = (12.283)
s1 − s 2 C(s1 − s2 )

691e
s1 s2 LI◦ I◦
A2 = = ; (12.284)
s2 − s 1 C(s2 − s1 )
note that the second equalities in Equations 12.283 and 12.284 follow from
the substitution of Equations 12.268 and 12.269. Finally, the substitution of
Equations 12.283 and 12.284 into Equation 12.281 yields
   
vC (t) 0 I◦ 1 I◦ 1
= u(t) + 1 e s1 t u(t) + 1 e s2 t u(t)
iL (t) I◦ C(s1 − s2 ) s1 L
C(s2 − s1 ) s2 L
(12.285)
as the solution of Equation 12.255. Note that the unit step function u has been
introduced into Equation 12.285 to extend the range of its validity. With a
little effort it can be shown that the two rows of Equation 12.285 are identical
to Equations 12.224 and 12.225. This is to be expected since both sets of
equations are the results of analyses of the same circuit.

12.10.1 N U M E R I C A L S O L U T I O N *
The state equation (specifically, Equations 12.255 and in a more general form,
Equation 12.256) can also be solved using numeric methods. Our goal here is to
demonstrate that the initial state contains all the information that is necessary
to determine the entire future behavior of the system given the subsequent
input. To help our intuition, we will describe an extremely simple method here.
However, we note that other more efficient, but less intuitive, methods are
employed in practice.
Suppose that the input signal vector z(t) is known for all time t ≥ t0 . Also
suppose that the initial value of the state vector at time t = t0 (denoted as x(t0 )
and called the initial state) is also known. Then the slope of the state vector
(that is, dx/dt) at time t0 can also be found from Equation 12.256 as follows:

dx
(t0 ) = Ax(t0 ) + Bz(t0 ). (12.286)
dt

From this slope and the initial state x(t0 ), the value of the state vector at time
t0 + t can now be estimated by standard numerical methods. Using Euler’s
method, for example, we can approximate the value of the state vector at time

t = t0 + t

as

dx
x(t0 + t) = x(t0 ) + (t0 )t
dt
= x(t0 ) + Ax(t0 )t + Bz(t0 )t. (12.287)

691f
Proceeding in the same manner, the value of x at time

t = t0 + 2t

can then be determined from the value of x(t0 + t) and z(t0 + t). Subsequent
values of x(t) can be determined using the same process. By choosing small
enough values of t, a computer can determine the waveform for the x(t)
vector to an arbitrary degree of accuracy. This process illustrates the fact that
the initial state contains all the information that is necessary to determine the
entire future behavior of the system from the initial state and the subsequent
input.
Notice the similarity between the numerical solution process for the
second-order system (Equations 12.256 and 12.287) and the first-order sys-
tem (Equations 10.82 and 10.83). The major difference is that we are dealing
with vectors and vector-matrix operations in the second-order system, while
the first-order system dealt with scalar operations. Higher-order systems with
many capacitors and inductors will result in a larger set of first-order state equa-
tions, which can be gathered into a single-vector state equation that is identical
in form to Equation 12.256.

691g
12.11 H I G H E R - O R D E R C I R C U I T S *
To close this chapter, we briefly consider the analysis of circuits having an order
higher than two. The important message here is that the methods of analysis
developed earlier in this chapter for second-order circuits are perfectly applica-
ble to the analysis of higher-order circuits. We will illustrate this through the
analysis of the circuit shown in Figure 12.63. Since the circuit has two inde-
pendent capacitors and two independent inductors, it is a fourth-order circuit.
Despite this, it readily submits to both a node analysis with the node voltages
as the primary unknowns, and a state-variable analysis with the states as the
primary unknowns.
Consider first the node analysis of the circuit shown in Figure 12.63, carried
out using v1 and v2 and as the two unknown node voltages. To begin, we write
KCL at Nodes #1 and #2 in terms of these voltages. This yields
 t
dv1 (t) 1 1
C1 + (v1 (t) − v2 (t)) + (v1 (t̃) − vIN (t̃))dt̃ = 0 (12.288)
dt R L1 −∞

for Node #1, and


   t
dv2 (t) dvIN (t) 1 1
C2 − + (v2 (t)−v1 (t))+ v2 (t̃)dt̃−iIN = 0 (12.289)
dt dt R L2 −∞

for Node #2. To treat these equations simultaneously, we use Equa-


tion 12.288 to determine v2 in terms of v1 , and then substitute the result

vIN

iL1
+
L1 vC2 C2
-
Node 1 Node 2
+ F I G U R E 12.63 A fourth-order
vIN
- v1 v2 circuit.
R iL2
+
vC1 C1 iIN
L2
-

691h
into Equation 12.289. This yields
 t
dv1 (t) R
v2 (t) = RC1 + v1 (t) + (v1 (t̃) − vIN (t)(t̃))dt̃ (12.290)
dt L1 −∞

   
d4 v1 (t) 1 1 d3 v1 (t) 1 1 d2 v1 (t)
+ + + +
dt4 RC1 RC2 dt3 L1 C 1 L2 C2 dt2
   
1 1 dv1 (t) 1
+ + + v1 (t) =
RC1 L2 C2 RC2 L1 C1 dt L1 C 1 L2 C2
     
1 d3 vIN (t) 1 d2 vIN (t) 1 dvIN (t)
+ + +
RC1 dt3 L1 C1 dt2 RC2 L1 C1 dt
   
1 1 d2 iIN (t)
vIN (t) + .
L1 C1 L2 C 2 RC1 C2 dt

(12.291)

Note that in deriving Equation 12.291 we differentiated Equation 12.289 twice


and divided it by RC1 C2 prior to the substitution of Equation 12.290. Finally,
to complete the node analysis, we solve Equation 12.291 for v1 , substitute the
result into Equation 12.290 to determine v2 , and then use the two node voltages
to determine any other branch variables of interest. For brevity, however, we
will not carry out these remaining steps. Instead, we note that to do so requires
initial conditions for v1 , and its first, second, and third derivatives. Most likely,
this information will be determined from the state variables specified at the
initial time, which takes additional work.
Consider next a state-variable analysis of the circuit shown in Figure 12.63.
To carry out this analysis we determine the state equation for each capacitor
and inductor. This yields,

dvC1 (t) 1
C1 = iC1 (t) = (vIN (t) − vC1 (t) − vC2 (t)) + iL1 (t) (12.292)
dt R
dvC2 (t) 1
C2 = iC2 (t) = (vIN (t) − vC1 (t) − vC2 (t)) + iL2 (t) − iIN (t) (12.293)
dt R
diL1 (t)
L1 = vL1 (t) = vIN (t) − vC1 (t) (12.294)
dt
diL2 (t)
L2 = vL2 (t) = vIN (t) − vC2 (t). (12.295)
dt

691i
These equations may be summarized in state-space form as
   1 1 1
 
vC1 (t) − − 0 vC1 (t)
RC1 RC1 C1
    
 v (t)   − 1 − 1 0 1   v (t) 
d  C2     C2 
= RC2 RC2 C2 
dt  iL1 (t) 
 
  −1 0 0 0   
  iL1 (t) 
   L1  
iL2 (t) 0 −1 0 0 iL2 (t)
L2

 1

0
RC1
  
 1 1 
  vIN (t)
+ 
RC2 C2
 . (12.296)

1
0 
 iIN (t)
 L1 
1
0
L2

Finally, to complete the state-variable analysis, we solve Equation 12.296 for the
states, and then use them to determine any other branch variables of interest.
For brevity, however, we will not carry out these remaining steps. Instead, we
note that to do so requires an initial condition for each state.
To close this section, it is again worth mentioning that both analyses predict
the same behavior for the circuit shown in Figure 12.63. The only difference
is that they do so in terms of different sets of variables, and through differ-
ent mathematical mechanics. Thus, the important message here is that both
analyses are applicable to higher-order systems.

691j
13.4.3 T H E B O D E P L O T : S K E T C H I N G T H E
FREQUENCY RESPONSE OF GENERAL
FUNCTIONS *
Sections 13.4.1 and 13.4.2 demonstrated the ease with which we can sketch
the frequency response of simple circuits by observing their behavior at low
frequencies and high frequencies. Things get more complicated for a network
with several inductors or capacitors. This section discusses a simple and intu-
itive method called Bode plots for sketching the frequency response of more
general circuits. The Bode method uses the insight gained from Sections 13.4.1
and 13.4.2 that the frequency response plots can be closely approximated by
straight line segments derived from the asymptotic behavior of the transfer
functions.
The method proceeds as follows: First, write the relationship (Equa-
tion 13.46, for example) in the form of a system function, the ratio of the
complex amplitude of the response to the complex amplitude of the input:

Response
H(s) = . (13.91)
Input

In general, the system function H(s) will be the ratio of two polynomials:

am sm + am−1 sm−1 + · · · + a1 s + a0
H(s) = (13.92)
bn sn + bn−1 sn−1 + · · · + b1 s + b0

where the coefficients ai and bi are real numbers since our circuit parameters
are real numbers. We saw one example of this in Equation 13.64. We can
factor the numerator and denominator polynomials and write

K1 (s − z1 )(s − z2 ) · · · (s − zm )
H(s) = (13.93)
(s − p1 )(s − p2 ) · · · (s − pn )

where K1 is a constant and z1 , z2 , . . . , zm are the roots of the numer-


ator polynomial, and p1 , p2 , . . . , pn are the roots of the denominator
polynomial.9

9. Because the system function goes to zero when s = zi , the roots of the numerator, z1 , z2 , . . . , zm ,
are called the zeros, definition of the system function. Similarly, the roots of the denominator,
p1 , p2 , . . . , pn are called the poles of the system function. The system function goes to infinity when
s takes on the value of one of the poles (in other words, when s = pi ). When one or more of the
z i s or p i s is zero, the system is said to have zeros or poles at the origin. The poles and zeros of a
system function are important system parameters because they characterize the general behavior

741a
In general, some of the roots of the numerator or the denominator polyno-
mials can be zero. Furthermore, the roots of the numerator and denominator
polynomials can also be complex. If any of the roots are complex, then they
must appear in complex conjugate pairs, so that the overall system function
remains real. We will rewrite Equation 13.93 into the following standard form
to reflect these facts:

 
Ko sl (s + a1 )(s + a2 ) · · · s2 + 2α1 s + ω12 · · ·
H(s) =   . (13.94)
(s + a3 )(s + a4 ) · · · s2 + 2α2 s + ω22 · · ·

In Equation 13.94, we have combined complex conjugate pairs into quadratic


terms of the form (s2 + 2αs + ω2 ). Thus all the remaining ai values are real. The
sl term, where l can be positive or negative, reflects the case where the roots in
Equation 13.93 are zero.
We will now show that it is possible to sketch without formal calculation
the general shape of H(s) as a function of frequency. More precisely, we can
make an approximate sketch of the magnitude and phase of H(s) as a func-
tion of the input frequency ω. The resulting pair of graphics representing an
approximate sketch of the frequency response is called a Bode plot, in honor
of the Bell Laboratories engineer who devised it to study stability in feedback
amplifiers.10
The Bode plot is an approximation of the frequency response and accord-
ingly has two parts: a sketch of the log magnitude of H( jω) versus logω and a
sketch of the angle of H( jω) versus logω. These coordinates are chosen because
they facilitate straightforward construction of the frequency response graphs
even for complicated functions without the use of a computer. Taking the
magnitude and log on both sides of Equation 13.94,

log|H(s)| = log Ko +
log |s| + log |s| + · · · (l terms)+
log |s + a1 | + log |s + a2 | + · · · − log |s + a3 | − log |s + a4 | + · · ·

log |s2 + 2α1 s + ω12 | + · · · − log |s2 + 2α2 s + ω22 | + · · · (13.95)

of the system. A detailed discussion of system analysis using poles and zeros is beyond the scope
of this book.
10. Bode, H.W., Network Analysis and Feedback Amplifier Designs, Van Nostrand, New York,
1945, Chapter 15.

741b
and for the phase

∠H(s) = ∠Ko +
∠s + ∠s + · · · (l terms)+
∠(s + a1 ) + ∠(s + a2 ) + · · · − ∠(s + a3 ) − ∠(s + a4 ) − · · ·

+ ∠(s2 + 2α1 s + ω12 ) + · · · − ∠(s2 + 2α2 s + ω22 ) − · · · (13.96)

Notice that there are four types of terms in the magnitude and phase equations:

1. The Ko constant term,


2. the s terms,
3. terms of the form (s + a), and
4. quadratic terms of the form (s2 + 2αs + ω2 ), which have complex roots.

This gives us a simple way of approximating the magnitude and phase curves
of the frequency response plot. First, draw the individual magnitude and angle
curves for each of the four types of terms in the numerator and denominator
of Equation 13.94. Then, construct the overall magnitude and phase plots by
simply adding together the individual curves.
Let us now address each of the four terms:

1. The Ko constant term.


We saw how to draw the frequency response of constant terms in
Section 13.4.1. Essentially constant terms result in horizontal lines on the
magnitude plot and have a phase of zero.
2. The s terms.
Terms of the form s and 1/s (if l is negative) were also plotted in
Section 13.4.1. We saw that each of these terms result in lines of +1 or
−1 slope on the log magnitude plot and contribute to a phase of 90◦ or
−90◦ , respectively.
3. Terms of the form (s + a).
Section 13.4.2 addressed terms of the form (s + a). We showed that the
magnitude part of the frequency response of these terms is approximated
by two straight lines corresponding to the low and high frequency
asymptotes meeting at the break frequency a. Accordingly, Bode plots
result in a series of straight line segments attached together at the break
frequencies.
The phase plot also uses low- and high-frequency asymptotes and passes
through 45◦ at the break frequency a. For more accuracy, the phase curve
can be approximated by a straight line that passes through 45◦ at the

741c
break frequency a, and meets the low- and high-frequency asymptotes at
0.1 times the break frequency (0.1a) and 10 times the break frequency
(10a), respectively.
4. Quadratic terms of the form (s2 + 2αs + ω2 ) with complex roots.
Although not as straightforward, it is possible to sketch frequency
response plots for system functions of the form (s2 + 2αs + ω2 ), where
the roots are complex. However, we will defer a further discussion on
plotting Bode plots for complex roots to Section 14.4. For now, we will
focus on real roots.

741d
e x a m p l e 13 .5 b o d e p l o t f o r s e r i e s r l c i r c u i t Let
us sketch the Bode plot for the RL circuit of Figure 13.11. From Equation 13.47, the
system function here is a voltage ratio:

Vo R/L
H( jω) = = . (13.97)
Vi jω + R/L

To make the example specific, let us assume that the time constant L/R has a value of
50 msec. Thus the break frequency is at

R
a= = 20 rad/s
L

and the system function becomes

20
H( jω) = . (13.98)
jω + 20

The system function has two terms: a constant term and a term of the form (s + a).
Figures 13.26 and 13.27 show the construction of the magnitude and phase plots,
respectively. The dashed lines in Figures 13.26c and 13.27c form the composite Bode
plot, and are obtained by simple subtraction of Figures 13.26b from Figure 13.27a. For
reference, the solid lines show the true magnitude and phase functions. Note that at the
break frequency, the true magnitude is given by:

|H( jω)| = 1/1.41 (13.99)

= 0.707.

The principal advantage of the Bode plot is that the composite magnitude asymptotes
for system functions that can be written in the form of Equation 13.94 are always
lines of integer slope in log space. Further, any system function that can be written as
a ratio of polynomials in ω (regardless of whether the roots are real or complex) must
approach at both low and high frequencies ( jω)n , where n is some integer. Hence the
magnitude asymptotes on Bode plots for both small and large ω must be straight lines
of integer slope in log space, and the phase must approach a multiple of 90◦ .

742a
3
log |R/L
1000.00
|R/L| log scale

100.00 2

10.00 1

1.00 0

0.10 -1

0.01
10-1 100 101 102 103
(a) ω

1000.00 3
log |jω + R/
|jω + R/L| log scale

2
100.00 F I G U R E 13.26 Magnitude curve
1 of the Bode plot for RL circuit:
10.00
(a) the magnitude curve for R/L;
1.00 0 (b) the magnitude curve for
jω + R/L; (c) the composite
0.10 -1
magnitude curve obtained by
0.01 -1
subtracting (b) from (a).
10 100 101 102 103
(b) ω

1000.00 3
|H| log scale

|||| | | | |

log |H|

2
|

100.00
|||| | | | |

1
|

10.00
|||| | | | |

0
|

1.00
|||| | | | |

-1
|

0.10
|||| | | | |
|

0.01 | | | | | | | | || | | | | | | | || | | | | | | | || | | | | | | | ||
|

10-1 100 101 102 103


(c) ω

742b
90

<R/L
70
50
30
10
-10 -1
10 100 101 102 103
-30
ω
-50
-70
-90 (a)

90
<R/L + jω

70
F I G U R E 13.27 Phase curve of 50
the Bode plot for RL circuit: (a) the 30
10
phase curve curve for R/L; (b) the
-10 -1
phase curve for jω + R/L; (c) the 10 100 101 102 103
-30
composite phase curve obtained by ω
-50
subtracting (b) from (a). -70
-90 (b)

90
<H

70
50
30
10 vv
-10 -1
10 100 101 102 103
-30
ω
-50
-70
-90 (c)

742c
e x a m p l e 13 . 6 a n o t h e r bode plot e x a m p l e To
illustrate the Bode method for more general transfer functions, let us sketch the Bode
plot for the following transfer function:

0.025(1000 + jω)
H( jω) = . (13.100)
100 + jω

The specific circuit that results in this transfer function is not relevant to us right now,
but will be discussed later in Section 13.6.

The system function has three terms: a constant term, and two terms of the form (s + a).
The Bode construction of the magnitude curve of the frequency response for the above
transfer function is shown in Figure 13.28. The corresponding phase construction is
shown in Figure 13.29. For reference, the actual frequency response generated using a
computer is shown using solid curves.

742d
104 4

||||| | | |
|0.025|

|
103 3

||||| | | |
|
102 2

||||| | | |
|
10 1 1

||||| | | |
|
10 0 0

||||| | | |
|
10-1 -1

||||| | | |
|
10-2 | -2 | | | | | | | || | | | | | | | || | | | | | | | || | | | | | | | ||
|
100 101 102 103 104
ω

104 4
||||| | | |
|

103 3
||||| | | |
|

102 2
||||| | | |
| 1000 + jω |

10 1 1
||||| | | |
|

100 0
||||| | | |
|

10-1 -1
||||| | | |

F I G U R E 13.28 Construction of
|

10-2 | -2 | | | | | | | || | | | | | | | || | | | | | | | || | | | | | | | ||
|

1
the magnitude curve of the Bode 100 10 102 103 104
plot. The composite magnitude ω
curve for the transfer function is
obtained by subtracting the 104 4
magnitude curve of (1000 + jω) 3 3
10
from the sum of the magnitude
curves of 0.025 and (100 + jω). 10 2 2
| 100 + jω |

101 1

100 0

10-1 -1

10 -2 -2
0
10 101 102 103 104
ω
104 4
|H|

10 3 3

10 2 2

101 1

100 0

10-1 -1

10-2 -2
100 101 102 103 104
ω

742e
90
<0.025

60

30

-30

-60

-90
100 101 102 103 104
ω

90
<1000 + jω

60

30

-30

-60

-90 |
100 101 102 103 104
ω
F I G U R E 13.29 Construction of
90 the phase curve of the Bode plot.
<100 + jω

60

30

-30

-60

-90
100 101 102 103 104
ω

90
<H

60

30

-30

-60

-90
100 101 102 103 104
ω

742f
e x a m p l e 13 .7 maximizing power transfer using
a t r a n s f o r m e r One use of the transformer discussed in Section 9.3.4 is to
match impedances between two halves of a circuit, and in doing so to maximize the
power transfered from a source to a load. For example, consider connecting a source
having a 1-V-peak and 50- Thévenin equivalent operating in the sinusoidal steady state
to a 1800- load as shown in Figure 13.56.
In the case of this direct connection, the voltage across the load is

1800
1 V sin(ωt),
1850

and so the time average power delivered to the load is approximately 0.26 mW; note
that the time average of sin2 (ωt) is 0.5.

Next, consider the circuit shown in Figure 13.57. In this circuit, an ideal transformer
having N1 primary turns and N2 secondary turns is inserted between the source and
load; with help from Figure 9.30, an equivalent model of this circuit is shown in
Figure 13.58.

50 Ω

+
F I G U R E 13.56 A source 1 Vsin (ωt) RL = 1800 Ω
connected directly to a load. -

Source Load

50 Ω

F I G U R E 13.57 A source
connected to a load through a +
1 V sin (ωt) N1 N2 1800 Ω
transformer. -

764a
50 Ω i1 i2

+ +
F I G U R E 13.58 An equivalent
+ + v2 circuit model of the circuit in
1 V sin (ωt) v1 1800 Ω
- N - Figure 13.57.
------2 i2 N
- N1 -----2 v1 -
N1

To analyze this new circuit, consider first the secondary side of the transformer. There,

N2 v1
i2 = − , (13.175)
N1 RL

and so at the primary side of the transformer,

N2 N22 v1
i1 = − i2 = . (13.176)
N1 N21 RL

Thus, as viewed from the primary side of the transformer, the transformer and resistor
together behave as a resistor having resistance

(N1 /N2 )2 RL .

In other words, the transformer has transformed the resistance of the load resistor by
the ratio of (N1 /N2 )2 . It is straightforward to show that any secondary-side impedance
is transformed to the primary side by the same ratio. Similarly, a primary-side
impedance is transformed to the secondary side by a ratio of (N2 /N1 )2 .

Let us now determine the ratio N2 /N1 that maximizes the power delivered to the load.
To do so, we use the circuit in Figure 13.59, in which the transformer and load resistor

50 Ω i1

+ F I G U R E 13.59 An equivalent
2
+  N1 circuit model with the transformer
1 V sin (ωt) v1 ------ 1800 Ω
-  N2 and load resistor replaced by the
effective load resistor.
-

764b
are replaced by the effective load resistor having resistance (N1 /N2 )2 1800 . In this
case, the voltage across the effective load resistor is

1800N21 /N22
1 V sin(ωt),
50 + 1800N21 /N22

and the average power delivered to the effective load resistor is

1800N21 /N22
0.5 W.
(50 + 1800N21 /N22 )2

Since the power into the primary side of an ideal transformer instantaneously exits the
secondary side of the transformer, this is the power delivered to the actual load. This
power is maximized for

50 = 1800N21 /N22 ,

or N2 /N1 = 6, in which case the resistance of the effective load resistor is 50 , and
the power delivered to the load resistor is 2.5 mW.

Thus, to achieve maximum power transfer, the resistance of the load must match that
of the source, and the ideal transformer performs this matching.

764c
e x a m p l e 13 . 8 n o n - i d e a l t r a n s f o r m e r s Real transform-
ers are never ideal, and at times their nonidealities are important. Such nonidealities
include the resistance of the coils, the leakage inductance of the coils and the magne-
tizing inductance of the core. These non-idealities can be added to the model shown
in Figure 9.29 to arrive at the model shown in Figure 13.60; note that the transformer
symbol in Figure 13.60 represents the original ideal transformer from Figure 9.29.

In Figure 13.60, R1 and R2 represent the resistances of the two coils, and LL1 and LL2
represent the leakage inductances of the two coils. The inductance L0 represents the
magnetizing inductance of the core given a single-turn coil, and so must be multiplied
by N21 if placed on the primary side of the ideal transformer, or by N22 if placed on the
secondary side of the ideal transformer. In either case, it represents the effect of the
non-infinite permeability of the core.

Let us now examine the effect of the magnetizing inductance on the results of
Example 13.7. Consider the case of N1 = 100, N2 = 600, and L0 = 8 µH, as
might be the case for a small-signal transformer. This case is shown in Figure 13.61.

Following the results of Example 13.7, we can replace the combination of the ideal
transformer and the 1800- load resistor with a 50- resistor, and compute the
magnitude of v1 to be

" " " "


" ) " " ω "
( jω 80 mH)(50 " "
|v1 | = "" " 1V = " 
" " 1V.
50  + ( jω 80 mH)(50 ) " 4ω2 + (625 rad/s)2 "

Ideal Transformer

R1 LL1 LL2 R2
F I G U R E 13.60 A non-ideal
N1 N2
N12L0 transformer.

50 Ω

+ +
N12L0 N2 v F I G U R E 13.61 Power transfer
+ v1 N1
1 V sin (ωt) 2 1800 Ω with a non-ideal transformer.
- = 80 mH = 100 = 600
- -

764d
Thus, for ω  312.5 rad/s, that is, for source frequencies well above 50 Hz, the voltage
magnitude across the transformer primary is approximately 0.5 V peak. In this case,
the maximum power is transfered to the resistor load. However, as the frequency goes
below 50 Hz, the inductor behaves like a relative short circuit in comparison to the
50- resistance of the transformed load resistor, and so the magnitude of v1 drops.
The magnitude of v2 and the power delivered to the load drops accordingly. In general,
the time-average power delivered to the load resistor is that which flows into the primary
of the ideal transformer, namely:

ω2
10 mW.
4ω2 + (625 rad/s)2

764e
14.4 T H E B O D E P L O T F O R R E S O N A N T
FUNCTIONS *
In this section, we will extend the Bode method for plotting approximate fre-
quency responses (Section 13.4.3) to resonant system functions. Recall that
a Bode plot is an approximate sketch of the frequency response, which can be
drawn by intuition without the use of a computer. Section 13.4.3 discussed
a simple and intuitive method for sketching the Bode plots for general circuits.
The method is based on the intuition that a general system function can be
written in the form shown in Equation 13.94, which contains four types of
terms:
1. A constant term,
2. s terms,
3. real terms of the form (s + a), and
4. quadratic terms of the form (s2 + 2αs + ωo2 ) with complex roots.
The Bode method proceeded by drawing the individual magnitude and angle
curves for each of the four types of terms in the numerator and denominator
of Equation 13.94. The magnitude curves are drawn on log-log scales and the
phase curves on log-linear scales. Observing that log-magnitudes and phases
add (Equations 13.95 and 13.96), the method concluded by constructing the
overall magnitude and phase plots by simply adding together the individual
curves.
Section 13.4.3 discussed how the real terms (types 1, 2, and 3) could be
plotted. This section discusses how we can plot type 4 terms, namely, quadratic
terms with complex roots. Once we know how to sketch the plots for each
of the four types of terms, we can then sketch any general system function by
superposition (see Section 13.4.3).
The Bode plot for a quadratic term with complex roots is easily drawn from
the insight gained in Section 14.2. There we showed that the low- and high-
frequency magnitude and phase asymptotes of second-order system functions
yielded insight into the general form of response. It turns out that for a quadratic
term of the form (s2 + 2αs + ωo2 ), the low- and high-frequency asymptotes can
be combined to yield a good approximation of the actual curve.
Accordingly, the following is a procedure for sketching the form of the
frequency response for a quadratic term of the form s2 + 2αs + ωo2 , which has
complex roots:
 Magnitude Plot
1. Sketch the low-frequency asymptote. For our quadratic term, the
low-frequency asymptote is given by the horizontal line:

|H( jω)| ≈ ωo2 .

808a
1010 180

<H (degrees)
|H|

170
160
150
140
109 130
120
110
100
90 O
108 80
70
60
50
107 40
30
20
10
0
106 -10
102 103 104 105 10 6 102 103 104 105 10 6
Frequency (rad/s) Frequency (rad/s)

(a) (b)

F I G U R E 14.28 Sketching the


frequency response of the 2. Sketch the high-frequency asymptote. The high-frequency asymptote
resonant function s2 + 2αs + ωo2 . is given by:
|H( jω)| ≈ ω2 .
This asymptote appears as a line of slope 2 in log-log scales.
Figure 14.28a shows these two asymptotes in dashed lines, assuming

ωo = 104

α = 500.
For comparison, the actual magnitude is also shown as a solid curve.
The two straight line asymptotes intersecting at ωo are a good
approximation of the magnitude curve.
It is also clear from Figure 14.28a that our approximation and the
actual curve differ in the vicinity of ωo , and amount by which they
differ relates to the peakiness of the curve, which in turn relates to the
value of Q. For ωo = 104 and α = 500,
ωo
Q= = 10.

Figure 14.29 plots the frequency response for several values of Q
(keeping ωo constant). It is easy to see that the difference between the

808b
180

<H (degrees)
170
160
1010 150
|H|

140
130
120
109 110
100
90 O
0.5
80
108 1 70
|

2 60
5
50
40
10 30
107
20 20 Q 0.5 1 2 510
Q 10 20
0
|

106 -10
102 103 104 105 106 102 103 104 105 106
Frequency (rad/s) Frequency (rad/s)

(a) (b)

actual magnitude (solid curves) and the approximate value from the F I G U R E 14.29 Frequency
response of s2 + 2αs + ωo2 for
Bode splot (dashed curve) at ωo becomes substantial for large values
different values of Q.
of Q. The exact difference is computed in Equation 14.74.

 Phase Plot
1. Sketch the low-frequency asymptote. The low-frequency asymptote
is given by:
∠H( jω) ≈= 0◦ .
2. Sketch the high-frequency asymptote. The high-frequency asymptote
is given by:
∠H( jω) ≈= 180◦ .
3. Mark ∠H( jωo ) = 90◦ , the angle of the system function at the
frequency ωo .
4. Draw a smooth line starting with the low-frequency asymptote,
passing through 90◦ at ωo , and finishing off at the high-frequency
asymptote.
Figure 14.28b shows these two asymptotes in dashed lines. For
comparison, the actual phase curve is also shown as a solid line.

808c
e x a m p l e 14 .6 b o d e p l o t e x a m p l e Let us sketch the frequency
response of the admittance of the second-order circuit in Figure 14.25 using the Bode
method.

From Equation 14.66, the desired system function is

Iz s2 + s RL + 1
LC
H(s) = = s R
.
Vz
C
+ LC

For

L = 1 mH
C = 10 µF

R=1

we get
ωo = 104 rad/s,
and

Q = 10.
Since Q > 0.5, the roots of the characteristic equation are complex and the circuit is
resonant.
Substituting the numerical quantities into our system function, we get

s2 + 1000s + 108
H(s) = .
105 (s + 103 )

The system function has three terms: a constant term, a term of the form (s + a), and
a quadratic term of the form (s2 + 2αs + ωo2 ). The Bode construction of the magni-
tude curve of the frequency response for the preceding transfer function is shown in
Figure 14.30. The corresponding phase construction is shown in Figure 14.31. For ref-
erence, the actual frequency response generated using a computer is shown using solid
curves.

808d
1012 12 1012 12
|105|

|s + 103|
1010 10 1010 10
108 8 108 8
106 6 106 6
104 4 104 4
102 2 102 2
100 0 100 0
10-2 -2 10-2 -2
102 103 104 105 106 102 103 104 105 106
ω ω
1012 12 101 1
|s2 + 1000s + 108|

|H|
1010 10
108 8 100 0
106 6
104 4 10-1 -1
102 2
100 0 -2
10-2
10-2 -2
102 103 104 105 106 102 103 104 105 106
ω ω

F I G U R E 14.30 Construction of the magnitude curve of the Bode plot. The composite magnitude curve for the transfer function is
obtained by subtracting the sum of the magnitude curves of 105 and (s + 103 ) from the magnitude curve of (s2 + 1000s + 108 ).

180 180
<105

<s + 103

150 150
120 120
90 90
60 60
30 30
0 0
|
|

-30 -30
-60 -60
-90 -90 |
101 102 103 104 105 106 101 102 103 104 105 106
ω ω

180 180
<s2 + 1000s + 108

<H

150 150
120 120
90 90
60 60
30 30
0 0
-30 -30
-60 -60
-90 -90
|

101 102 103 104 105 106 101 102 103 104 105 106
ω ω

F I G U R E 14.31 Construction of the phase curve of the Bode plot.

808e
15.4.4 G E N E R A L I Z A T I O N O N I N P U T R E S I S T A N C E *
It is obviously of some importance to the circuit designer to know whether
feedback is going to increase or decrease the effective input resistance of a circuit.
We can generalize from the two circuits we have examined to state that the effect
of feedback on input resistance depends on the circuit topology. If the source
current and the current through the feedback resistor and the current through
the Op Amp input resistor ri all sum at a common node as in Figure 15.12,
then the effective input resistance is very low, as shown in Equations 15.36 and
15.38. (Remember, here we are referring to Ri , the resistance of the Op Amp
circuit to the right of Rs .) Equation 15.36 is in fact a general result: The input
conductance for any feedback circuit with this input topology (neglecting ri ) is
the conductance without feedback, here 1/(Rf + rt ), multiplied by 1 + A.
If, on the other hand, the source and the Op Amp input resistor are in series,
forming a loop with the feedback resistor, as in Figure 15.14, the effective input
resistance of the circuit will be very high. In a word, if at the Op Amp input we
sum currents at a node, the circuit input resistance is low, if we sum voltages in
a loop, the input resistance is high.

855a
15.6.5 S A L L E N - K E Y F I L T E R
This section introduces a lowpass filter called the Sallen-Key filter. Its circuit and
impedance model are shown in Figure 15.25.
Let us focus on sinusoidal inputs and use the impedance method to obtain
its input-output relationship. First, notice that the portion of the circuit within
the dashed box is a non-inverting connection of the Op Amp with gain:

R1
G=1+ . (15.94)
R2

vi v1
+
v2 vo
R R
-

C
R1
R2

(a) Circuit
F I G U R E 15.25 The Sallen-Key
circuit.
1/Cs

Vi V1
R R +
V2 Vo

-
1/Cs

R1

R2

(b) Impedance model

866a
Thus, for the purpose of analysis, we can replace the circuit within the dashed
box with an amplifier whose gain is G. Therefore, we can write

Vo = GV1 . (15.95)

Applying KCL for node V1

V2 − V 1 V1
= ,
R 1/Cs

which simplifies to:

V2 = (RCs + 1)V1 .

Substituting for V1 in terms of Vo from Equation 15.95, we get

RCs + 1
V2 = Vo . (15.96)
G

Now, KCL for node V2 yields,

Vi − V 2 V2 − V1 V2 − Vo
= + . (15.97)
R R 1/Cs

Substituting for V1 and V2 in terms of Vo from Equations 15.95 and 15.96,


we get


RCs+1 1 RCs+1
Vi − RCs+1
Vo Vo − V Vo − Vo
G G G o G
= + 1
. (15.98)
R R
Cs

We can simplify Equation 15.98 and obtain the following expression relating
the output voltage to the input voltage:

Vo (s) G
H(s) = = . (15.99)
Vi (s) R 2 C 2 s2 + RCs(3 − G) + 1

As a specific example, let us draw the frequency response for the filter trans-
fer function for RC = 1 and R1 = R2 . For these values, G = 2 and the

866b
1

0.8

0.6

0.4
Imag Axis

0.2

0
F I G U R E 15.26 Pole-zero plot of
-0.2 the Sallen-Key filter.

-0.4

-0.6

-0.8
-1
-1.5 -1 -0.5 0 0.5
Real Axis

transfer function is given by:

2
H(s) = (15.100)
s2 +s+1

or, factoring the denominator,

2
H(s) =   . (15.101)
(s + 1/2 + j 3/4)(s + 1/2 − j 3/4)

The transfer function represents a second-order filter. The expression in the


denominator has a pair of complex conjugate roots: −1/2 + j 3/4 and
−1/2 − j 3/4. In terms of the pole-zero nomenclature introduced in
Section 13.4.3, the transfer function has two polesand no zeros. Thepoles
are a complex conjugate pair located at −1/2 + j 3/4 and −1/2 − j 3/4.
Figure 15.26 depicts the pole locations using X’s in the complex plane. (When
zeros exist, their locations are depicted using circles.)
We can now plot the frequency response as shown in Figure 15.27 by
substituting s = jω in the transfer function:

2
H( jω) = (15.102)
( jω)2 + jω + 1

866c
101

Magnitude
100

10-1

10-2
10-1 100 101
Frequency (radians)
F I G U R E 15.27 Frequency
response for the Sallen-Key filter.
Phase (degrees)

-50

-100

-150

-200
10-1 100 101
Frequency (radians)

or, in terms of the factored transfer function,

2
H( jω) =   . (15.103)
( jω + 1/2 + j 3/4)( jω + 1/2 − j 3/4)

As before, the frequency response in Figure 15.27 plots the magnitude and the
phase of H( jω) versus the frequency ω.

866d
15.9 T W O - P O R T S *
It should be obvious by now that circuits with dependent sources can perform
much more interesting and useful signal processing than those constructed
solely from two-terminal resistive elements. But inclusion of dependent sources
has brought about a modest increase in circuit complexity, so it is useful at i2
i1
this point to generalize some of the concepts introduced in previous chapters.
In particular, let us examine how to generalize the Thévenin calculations to + +
+
three-terminal or four-terminal systems. v1 - v2
We start with a linear network containing resistors, voltage sources, and - -
current sources as we did in Figure 3.55, but now we assume two pairs of
external terminals, as shown in Figure 15.36. This network is called a two-port,
or a two-terminal-pair network. For the purposes of the present discussion, it
doesn’t matter whether the two negative leads are tied together or go to some F I G U R E 15.36 Linear two-port
common ground, or both terminal pairs are floating with respect to ground. network.
We wish to find a two-port Thévenin equivalent of this network. The deriva-
tion is a simple extension of the method in Section 3.6.1. We apply current
sources at each of the ports, as in Figure 15.37a, then solve the problem by
superposition. We first set all the independent sources, both internal and exter-
nal, to zero except i2 , and measure the resulting v2a as in the subcircuit of
Figure 15.37b. Because there is nothing left of the network except resistors (and
possibly dependent sources), v2a must be linearly dependent on i2 without off-
sets. In other words, the ratio v2a /i2 is a pure resistance, a Thévenin-equivalent
output resistance:
v2a
RThout = . (15.114)
i2

Then we set i1 and i2 to zero, leave the internal sources active, as in


Figure 15.37c, and measure v2b = v2oc , (this is what we previously called
the open-circuit voltage). Finally we set i2 and the internal sources to zero,
leaving i1 active, and measure v2c , which must be linearly dependent on i1 , and
hence can be written as

v2c = i1 R21 . (15.115)

This is clearly a dependent source relationship: an output voltage dependent on


an input current. Now by superposition, the total output voltage is the sum of
these three terms:

v2 = i1 R21 + i2 RThout + v2oc . (15.116)

A completely analogous argument yields for the input terminals:

v1 = i1 RThin + i2 R12 + v1oc (15.117)

872a
+ +
+
i1 v1 - v2 i2
- -

(a)

i1 = 0

+
V=0 i2
v2a
I=0 -
v2a
(b) RThout = -------
i2
F I G U R E 15.37 Two-port
calculations.

+ + +
- v2b
i1 = 0
- -

(c)

+
V=0 v2c
i1
I=0 -

v2c
(d) R21 = -------
i1

872b
+
v1a
i1 V=0 i2 = 0
- I=0

v1a
RThin = -------
i1
(a)

+ +
V1b = v1oc - i2 = 0 F I G U R E 15.38 Two-port input
- calculations.

(b)

+
v1c V=0 i2
- I=0

v1c
(c) R12 = -------
i2

where RThin , v1oc , and R12 are measured or calculated using the subcircuits in
Figures 15.38a, 15.38b, and 15.38c, respectively.
Equations 15.116 and 15.117 taken together, are a complete representa-
tion of the network as viewed from the two terminal pairs or two ports. It is
common practice in linear network theory to assume that there are no indepen-
dent sources inside the network. In this case a rather simple generalization of the
Thévenin equivalent circuit emerges. Equations 15.116 and 15.117 simplify to

v1 = i1 RThin + i2 R12 (15.118)


v2 = i1 R21 + i2 RThout (15.119)

872c
and a simple circuit interpretation is now apparent. The term i2 R12 in the equa-
tion for the input port, Equation 15.118, is a voltage, dependent on the current
at the output. That is, it is a dependent voltage source, under the control of i2 .
The first term in Equation 15.118 is the Thévenin input resistance. Hence the
equation can be represented in circuit form by the left half of Figure 15.39a. The
expression for the output port, Equation 15.119 has similar structure, except
the role of input and output variables have been reversed. Hence the right half
of Figure 15.39a. The circuits and equations for calculating the four parameters
(called z parameters in linear network theory) are given in Figures 15.37b and
15.37d, and Figures 15.38a and 15.38c.
If we had chosen to drive the two-port with two voltage sources, rather
than two current sources as in Figure 15.37, then from Section 3.6.2, the two-
port version of the Norton equivalent would have emerged. The equations
analogous to Equations 15.118 and 15.119 are

i1 = yin v1 + y12 v2 (15.120)


i2 = y21 v1 + yout v2 (15.121)

i1 Rthin Rthout i2

+ +

+ + v2
v1 i2R12 i1R21
- -

- -

(a) z parameter model

F I G U R E 15.39 z and y
parameter models.
i1 i2

+ +
y21v1

v1 yin yout v2

y12v2
- -

(b) y parameter model

872d
where the Y terms are conductances for resistive circuits. The circuit equivalent,
called the y parameter model, is shown in Figure 15.39b. The expressions
for each of the y parameters are readily derivable from Equations 15.120
and 15.121, or from first principles, as in Figures 15.37 and 15.38, or by a
linear transformation on the z parameters.
Two other representations, the g parameters and the h parameters, arise if
one excites the two-port with a voltage source at one port and a current source
at the other. All four representations are related by linear transformations.
It is helpful to re-examine the calculation of Op Amp input and output
resistance in Section 15.4 from the more general two-port point of view of this
section. Because in Figures 15.11 and 15.12 we used a test voltage source at
the input and a test current source at the output, we in fact were calculating
the g-parameters, defined in Figure 15.40a. To complete the calculation, we
assume that in the Op Amp circuit, Figure 15.12, the reverse signal flow through
the circuit is negligibly small. Hence g12 in Figure 15.40a is zero. Also, from
Figure 15.12 the forward dependent source g21 is approximately A, if we neglect
the drop in rt caused by current through Rf . On this basis the g-parameter rep-
resentation for the inverting Op Amp connection is as shown in Figure 15.40b,
assuming Rs is external to this model.
It is now (finally) possible to justify the omission of the ±12 V power
supplies in all calculations in this chapter. In terms of a two-port model, the

i1 g22 i2
+ +
+
v1 g11 g12i2 g21v1 v2
-

- -
i1 = g11v1 + g12i2 v2 = g21v1 + g22i2
(a) g parameters
F I G U R E 15.40 g parameter
Ro model for inverting Op Amp.

v+
+

+
1/Ri A(v+ - v-) vo
-

-
v-
(b)

872e
power supplies would produce no measurable voltage or current at either the
input or the output of the circuit, because of the balanced nature of the circuit in
the active region. Hence inclusion of the power supplies would not change the
model parameters we have just derived, so it is correct to neglect these supplies
in all Op Amp active-region calculations.

872f
16.4.3 A S W I T C H E D P O W E R S U P P L Y U S I N G A D I O D E
In this example, we will analyze the behavior of the diode-based switched power
supply circuit shown in Figure 16.15. Notice that this circuit is similar to that in
Figure 12.41, with the switch S2 replaced with a diode. As before, the purpose of
the circuit is to convert the DC input voltage V to a different DC output voltage
vOUT . The MOSFET in the circuit operates as a switch, and the square-wave
input to the MOSFET is shown in Figure 16.16. As before, we are interested in
determining the behavior of vOUT over time. As we will see shortly, the diode
in the circuit also acts a switch, and results in an output waveform that is largely
the same as that of the circuit in Figure 12.41.
We will assume that the switch S1 has zero resistance associated with its
ON state, and that the diode is ideal, so that the model in Figure 16.6 applies.
Specifically, this means that the diode turns on and behaves like a short circuit
when a positive current (iD ) flows through it. The diode turns off and behaves
like an open circuit when the voltage (vD ) across it is negative.
When the switch S1 is closed, it shorts the terminal connecting the diode
and the inductor to ground. Assuming that vOUT is non-negative, the diode
being reverse biased is off. The DC voltage V appears directly across the inductor
as illustrated in Figure 16.17, and the inductor current iL ramps up. Since S1 is
the on for time T, the inductor current builds up to

VT
iL = (16.34)
L

as shown in Figure 16.16. Meanwhile, if there is no applied load at vOUT , the


capacitor voltage vOUT remains constant.
Next, when S1 is opened, the inductor current cannot instantaneously go
to 0. Instead, the current finds a path through the diode (thereby turning it on)
and into the capacitor. In its ON state, the diode behaves like a short circuit,
and so the driven LC circuit shown in Figure 16.18 results. The current iL in the

iD
iL vOUT

L + vD -
vC F I G U R E 16.15 A switched
+ + power supply circuit with diode
S1 C
V - and a switch.
-

918a
S1 State TP
One cycle

CLOSED
(ON)
T
OPEN
(OFF)
t

Diode
State
Diode Diode Diode
OFF ON OFF

t
F I G U R E 16.16 Switched power
supply operation. iL

VT
------
L

0
t

vC

vC [n + 1]

vC [n]

0
t

LC circuit follows a sinusoidal pattern as illustrated in Figure 16.16. Because of


the flow of current into the capacitor, its voltage vOUT starts to increase, and it
too follows a sinusoidal pattern.
As iL follows its sinusoidal pattern, it soon reaches zero and the positive
voltage on the capacitor attempts to drive it negative. At this instant, the diode
turns off and disconnects the capacitor from the rest of the circuit, so in the
absence of a load, the capacitor maintains its voltage.

918b
iL

F I G U R E 16.17 The equivalent


+ S1 circuit when S1 is closed and the
V diode is open.
-

iL iD

vC
+ + F I G U R E 16.18 The equivalent
C LC circuit when S1 is open and the
V
- diode is ON.
-

This cycle repeats, dumping some amount of charge into the capacitor
each cycle. We can compute the increase in vOUT very quickly using an energy
argument similar to that used in Example 12.4 as follows: At the end of the
ramp, the inductor current is given by Equation 16.34, and so the energy stored
in the inductor is given by:
V2 T2
wM = .
2L
Since the capacitor is charged by the inductor until iL becomes zero, the energy
(wM ) stored in the inductor is transferred completely to the capacitor in each
cycle. After n cycles, the energy stored in the capacitor becomes n times the
energy transferred in a single cycle, plus any energy initially stored on the
capacitor (say wE [0]):
V2 T2
wE [n] = n + wE [0].
2L
Unlike Example 12.4, the capacitor must start with vC = V, since it is connected
by a diode instead of a switch to a voltage source. Unlike the switch, which can

918c
be forced to stay off, the (ideal) diode turns on if V is greater than vC . Therefore,

1
wE [0] = CV2 .
2

Since wE [n] = CvOUT [n]2 /2, we can derive the voltage after n cycles as:

nT2
vOUT [n] = V + 1.
LC

Substituting, ωo = 1/ LC, we have
!
vOUT [n] = V nT2 ωo2 + 1.

If nT2 ωo2  1, we get



vOUT [n] = VTωo n.
Finally, when a load is added to the circuit as shown in Figure 16.19, the
capacitor begins to discharge through the load. Suppose we wish to maintain
the voltage vOUT at a specified average value, say vREF , then in each cycle, we
must arrange to have the capacitor charged up by the same amount of charge
that it supplies to the load. This can be accomplished by using a feedback
system as shown in Figure 16.20.
In the circuit in Figure 16.20, the controller compares vOUT to vREF , and
if vOUT falls below vREF , it increases the duration T for which the switch S1
is kept ON, thereby increasing vOUT . Conversely, the controller decreases the
duration T if vOUT increases past the value of vREF . Thus, vOUT is kept close
to vREF throughout.

iL
+
L
vC RL
+ +
F I G U R E 16.19 Adding a load. S1 C vOUT
V -
-

918d
iL
+
L
vOUT
vC Control
+ + RL
V S1 C
-
- change T
vREF
F I G U R E 16.20 Feedback
- system to maintain a voltage vREF
at the load.

TP t

918e
16.5 A D D I T I O N A L E X A M P L E S
For review purposes, more examples of both piecewise linear and incremental
analysis are given in the following subsections. No new material is presented,
so readers who do not need additional practice can omit this section without
loss of continuity.

16.5.1 P I E C E W I S E L I N E A R E X A M P L E :
CLIPPING CIRCUIT
The output voltage vo in the diode clipper circuit, shown in Figure 16.21a, will
resemble the input voltage vi , except that the bottom of the waveform will
be clipped off. The circuit has only one diode, so that the Thévenin solution
method discussed in Chapter 3.6.1 can be used, but here we will use the method
of assumed diode states. Assume that the diode in Figure 16.21 is ideal, then
draw the two subcircuits, one with the diode OFF, and the other with the diode
ON, as shown in Figures 16.21b and 16.21c. By inspection, the output voltage
with the diode OFF is constant, because there is a fixed current IO flowing up
through R. Thus
vo1 = −IO R. (16.35)
The current source and the voltage source are in series, so the voltage source
has no effect on vo1 . Furthermore, when the diode is OFF,

vi = vo1 + vD . (16.36)

Because the diode is in the OFF state, vD must be negative. It follows that in
the OFF state vi must always be more negative than −IO R.
Next, when the diode is ON, the output is directly connected to the input:

vo2 = vi . (16.37)

In the ON state, vi must be more positive than −IO R. Hence the valid portions
of the waveforms in the subcircuits are the darkened segments, and the complete
output waveform is as shown in Figure 16.21d. As promised, the circuit has
clipped off the bottom of the input wave.

16.5.2 E X P O N E N T I A T I O N C I R C U I T
The circuit shown in Figure 16.22 produces an output voltage vOUT that is
proportional to the exponential of the input voltage vIN for sufficiently large
vIN . To analyze this circuit, assume that the Op Amp is ideal, the saturation
current of the diode is Is = 10−12 A, and the temperature of the diode is
approximately 29◦ C so that its thermal voltage is VTH = 26 mV. Because

918f
IO

+
+
vi R vo
-
-
vo1
(a)
vi
IO

iD = 0 t
+ vD - +
+
vi R vo1 -IOR
-
(b)
-
Valid for vi < -IOR
vo2
F I G U R E 16.21 Diode clipper.
IO vi = vo2

- t
+ +
vD = 0
+ -IOR
vi R vo2
-
-
(c)
Valid for vi > -IOR
vo

vo

-IOR
(d)
vi

918g
R = 100 kΩ

iD
-
F I G U R E 16.22 A diode-based + +
exponentiation circuit.
vIN
vOUT
+
-
-

the Op Amp is ideal, and used in a stable negative-feedback configuration, the


voltage at the inverting terminal of the Op Amp is zero. As a result,

iD = Is evIN /VTH − 1 = 10−12 A evIN /(26 mV) − 1 .

For sufficiently large vIN , for example for vIN ≥ 120 mV, the exponential term
dominates, and this relation simplifies to

iD ≈ 10−12 A evIN /(26 mV) .

Next, because the voltage at the inverting terminal of the Op Amp is


zero, and because the current into that terminal is zero, the output voltage
is given by
vOUT = −RiD ≈ −10−7 V evIN /(26 mV) ,
which exhibits an exponential dependence on vIN .
For example, for vIN = 200 mV, 300 mV, and 400 mV, vOUT =
−0.219 mV, −10.3 mV, and −480 mV, respectively.

16.5.3 P I E C E W I S E L I N E A R E X A M P L E : L I M I T E R
The circuit in Figure 16.23 is useful for making square waves out of sine waves,
and for limiting the amplitude of an output waveform when the input waveform
amplitude varies over a wide range. To analyze the circuit, we note that the
Thévenin approach is not helpful, and a graphical solution might be messy
because of the two diodes (not so, in fact, but that is not obvious yet). So resort
to analysis by assumed diode states. The subcircuits for the four states, assuming
an ideal-diode model, are shown in Figures 16.23b, 16.23c, 16.23d, and 16.23e

918h
R

+ D1 D2 +
vi + -
- V V vo
- +
-

(a) Limiter circuit


vo1
vi
R
+ + +V
+v
+ OFF v- D1 OFF D2
vi - vo1 t
- + - 0
V- V -V
+ -

(b) Both diodes OFF


vo2
R
+
+ OFF +V
vi +
- - vo2
V - V F I G U R E 16.23 Diode limiter.
+ - t
0

(c) D1 ON
vo3

R
+
+ OFF
vi + - vo3 t
- V 0
V- + -V
-

(d) D2 ON vo
vi
R
+ vo
+ ON ON t
vi + - vo4 0
- V V
- +
-

(e) Both diodes ON (f) Complete waveform

918i
along with the appropriate subcircuit output voltages, obtainable by inspection.
From Figure 16.23b,
vo1 = vi (16.38)
because there is no current through R. When either diode is ON, the output
voltage is independent of the source voltage vi . For D1 , ON, for example,
vo2 = V. The fourth diode state, Figure 16.23e, cannot be reached with this
topology, (assuming V is a positive quantity) because there is no value of vi that
will force both diodes ON at the same time. Now we must identify the valid
segments of these waveforms. In Figure 16.23b, both diodes are assumed OFF,
so vd1 and vd2 must both be less than zero. Hence, using KVL:

vi − V = vD1 < 0 (16.39)


vi < V (16.40)
−vi − V = vD2 < 0 (16.41)
vi > −V. (16.42)

Thus vi must be between −V and +V. Likewise vo1 , from Equation 16.38.
This range of validity is indicated by the darkened segments of the waveform
in Figure 16.23b. It follows that the complete output wave must be as shown
in Figure 16.23f. If the peak amplitude of vi is ten or twenty times V, then vo is
a reasonable approximation of a square wave.

16.5.4 E X A M P L E : F U L L - W A V E D I O D E B R I D G E
Figure 16.24 shows one of the most common rectifier circuits found in elec-
tronic equipment, the full-wave diode bridge. We assume therefore that vi is a
60-Hertz sinusoid with 10-volt peak amplitude, and we wish to find the wave-
form vo across the output resistor. A full-blown assault using assumed diode
states would yield 16 subcircuits, but it will turn out that only two of these
are possible, suggesting a more insightful approach. Suppose that vi is a small
positive voltage. Then current must flow down through the bridge. The only
available path is D1 , RL , and D4 , because of the orientation of D3 and D2 .
Similarly, for vi negative, current must flow up, and thus must follow the path
D3 , RL , D2 . The two corresponding subcircuits, assuming ideal diodes, are
shown in Figures 16.24b and 16.24c. Now, by inspection, for vi positive,

vo = vi (16.43)

and for vi negative,


vo = −vi . (16.44)

918j
D1 D2
+ + vo -
vi
- RL
(a) D3 D4

D1 OFF D1 OFF v
+ voA + + voC
vi + - vi + oB - vi + -
- - - F I G U R E 16.24 Full-wave diode
D3 OFF D4 OFF D3 OFF D4 OFF
bridge.

(b) Subcircuit (c) Subcircuit (d) D1 and D2 ON


for vi positive for vi negative

vi vo

t t

(e)

All other subcircuits are degenerate. Consider, for example, the subcircuit for
both D1 and D2 ON, as in Figure 16.24d. Clearly no current can flow in RL .
Also, current can’t flow down through D3 , or up through D4 , so all diode cur-
rents must be zero in this state. A similar argument holds for all adjacent diode
pairs, whether ON or OFF. Hence the two states depicted in Figures 16.24b
and 16.24c are the only ones we need to consider.
Note that the current always flows in the same direction through RL ,
regardless of the polarity of vi . For sinusoidal input the waveforms appear
as in Figure 16.24e. The circuit is called a full-wave rectifier because current
flows through RL on both halves of the input wave. Neglecting diode volt-
age drops, the average value of the output voltage, that is the DC voltage, is
0.637 times the peak of the input sinusoid. Further, by symmetry there is no
frequency component in the output waveform at the input frequency, 60 Hertz
in our example, or odd multiples thereof. Hence the circuit has a much higher
percentage of DC relative to harmonics compared to the half-wave rectifier
discussed in Section 4.3.

918k
R
i
+
i +
50 mV AC
-
F I G U R E 16.25 Zener-diode v vo
v
regulator. +
20 V DC
-
-

(a) (b)

16.5.5 I N C R E M E N T A L E X A M P L E : Z E N E R - D I O D E
REGULATOR
All semiconductor diodes will break down and conduct appreciable current
under reverse bias conditions if the reverse voltage across the diode is large
enough. This breakdown is non-destructive if the current is not excessive; the
diode returns to normal reverse-bias behavior if the voltage is reduced. A Zener
diode is a semiconductor diode in which this breakdown under reverse bias is
carefully controlled by the manufacturing process so that the breakdown occurs
at a specified voltage, the so-called Zener voltage of the diode. A typical v i
curve is shown in Figure 16.25a.
Because the breakdown voltage of a Zener diode can be carefully con-
trolled by the manufacturing process, and the incremental resistance in the
breakdown region is quite small (around 10  to 50 ), Zener diodes are
quite useful as voltage regulators. A simple example is shown in Figure 16.25b.
Equation 4.74 is clearly inappropriate for finding the incremental resistance in
breakdown, because this part of the characteristic is not an exponential. Hence
the value must be obtained from the data sheet for the Zener diode in question.
Integrated-circuit regulators, with transistors, Zener diodes, and resistors all
on a single chip, will certainly outperform either of the crude regulator circuits
discussed here.

16.5.6 I N C R E M E N T A L E X A M P L E :
DIODE ATTENUATOR
It should be clear from Section 4.5, and particularly from Equations 4.63
and 4.74, that for small increments of voltage or current, the semiconduc-
tor diode looks like a linear resistor whose value depends on the DC current
flowing through it. Thus it should be fairly easy to build an attenuator with
an attenuation constant that can be changed by means of an external voltage
or current. Figure 16.26 shows a simple example. Here a diode is used in the
shunt branch of a voltage divider on a small-signal source vi . The DC cur-
rent through the diode is controlled by the DC voltage VC through the large

918l
RC Rt = R1||RC

R1 R1
+
VC + - +
VOC = VC ------------------- +
- vi
R1 + RC - - 0.6 V
- Rd
(a) (c)
F I G U R E 16.26 Diode
attenuator.
RC RC
+
R1 R1
+
VC 0 + vo
- ID0
vi rd
- -
(b) (d)

resistor RC . If we assume that vi produces less than a 5 mV change in the


diode voltage, then the incremental analysis approach discussed in Section 4.5
can be applied.
First, draw the circuit for the calculation of the DC current ID0 is to form the
Thévenin equivalent of the linear part of the circuit, as shown in Figure 16.26b.
An easy way to solve for ID0 is to form the Thévenin equivalent of the linear
part of the circuit, as shown in Figure 16.26c. At the same time we replace the
diode by its piecewise linear model. Then

(VOC − 0.6)
ID0 = . (16.45)
Rt + R d

The incremental resistance rd of the diode will thus be a function of VC :

kT
rd = (16.46)
qID0
 
kT  (R1 RC ) + Rd 
= . (16.47)
q VC R1 − 0.6
R1 +RC

Now draw the subcircuit that relates the incremental variables. Replace the
diode in Figure 16.26a by the incremental resistance rd , and set all DC sources,
in this case VC , to zero, as indicated in Figure 16.26d. By inspection,

RC rd
vo = vi . (16.48)
R1 + (RC rd )

The attenuation is clearly dependent on the DC voltage VC , as desired.

918m

Vous aimerez peut-être aussi