Вы находитесь на странице: 1из 11

September 11, 2008 11:18 boyce-9e-bvp Sheet number 452 Page number 432 cyan black

432 Chapter 7. Systems of First Order Linear Equations

(a) Find J2 , J3 , and J4 .


(b) Use an inductive argument to show that
⎛ n ⎞
λ nλn−1 [n(n − 1)/2]λn−2
⎜ ⎟
J =⎝0
n
λn nλn−1 ⎠.
0 0 λn
(c) Determine exp(Jt).
(d) Observe that if you choose λ = 2, then the matrix J in this problem is the same as
the matrix J in Problem 17(f). Using the matrix T from Problem 17(f), form the product
T exp(Jt) with λ = 2. Observe that the resulting matrix is the same as the fundamental
matrix (t) in Problem 17(e).

7.9 Nonhomogeneous Linear Systems


In this section we turn to the nonhomogeneous system
x = P(t)x + g(t), (1)
where the n × n matrix P(t) and n × 1 vector g(t) are continuous for α < t < β. By
the same argument as in Section 3.5 (see also Problem 16 in this section), the general
solution of Eq. (1) can be expressed as
x = c1 x(1) (t) + · · · + cn x(n) (t) + v(t), (2)
where c1 x (t) + · · · + cn x (t) is the general solution of the homogeneous system
(1) (n)

x = P(t)x, and v(t) is a particular solution of the nonhomogeneous system (1). We


will briefly describe several methods for determining v(t).

Diagonalization. We begin with systems of the form


x = Ax + g(t), (3)
where A is an n × n diagonalizable constant matrix. By diagonalizing the coefficient
matrix A, as indicated in Section 7.7, we can transform Eq. (3) into a system of
equations that is readily solvable.
Let T be the matrix whose columns are the eigenvectors ξ (1) , . . . , ξ (n) of A, and
define a new dependent variable y by
x = Ty. (4)
Then, substituting for x in Eq. (3), we obtain
Ty = ATy + g(t).
When we multiply by T−1 , it follows that
y = (T−1AT)y + T−1 g(t) = Dy + h(t), (5)
−1
where h(t) = T g(t) and where D is the diagonal matrix whose diagonal entries
are the eigenvalues r1 , . . . , rn of A, arranged in the same order as the corresponding
September 11, 2008 11:18 boyce-9e-bvp Sheet number 453 Page number 433 cyan black

7.9 Nonhomogeneous Linear Systems 433

eigenvectors ξ (1) , . . . , ξ (n) that appear as columns of T. Equation (5) is a system of


n uncoupled equations for y1 (t), . . . , yn (t); as a consequence, the equations can be
solved separately. In scalar form Eq. (5) has the form

yj (t) = rj yj (t) + hj (t), j = 1, . . . , n, (6)

where hj (t) is a certain linear combination of g1 (t), . . . , gn (t). Equation (6) is a first
order linear equation and can be solved by the methods of Section 2.1. In fact, we
have
 t
yj (t) = e rj t
e−rj s hj (s) ds + cj erj t , j = 1, . . . , n, (7)
t0

where the cj are arbitrary constants. Finally, the solution x of Eq. (3) is obtained
from Eq. (4). When multiplied by the transformation matrix T, the second term on
the right side of Eq. (7) produces the general solution of the homogeneous equation
x = Ax, while the first term on the right side of Eq. (7) yields a particular solution
of the nonhomogeneous system (3).

Find the general solution of the system


EXAMPLE    
1  −2 1 2e−t
x = x+ = Ax + g(t). (8)
1 −2 3t

Proceeding as in Section 7.5, we find that the eigenvalues of the coefficient matrix are
r1 = −3 and r2 = −1 and that the corresponding eigenvectors are
   
(1) 1 (2) 1
ξ = , ξ = . (9)
−1 1

Thus the general solution of the homogeneous system is


   
1 −3t 1 −t
x = c1 e + c2 e . (10)
−1 1

Before writing down the matrix T of eigenvectors, we recall that eventually we must find T−1 .
The coefficient matrix A is real and symmetric, so we can use the result stated at the end of
Section 7.3: T−1 is simply the adjoint or (since T is real) the transpose of T, provided that the
eigenvectors of A are normalized so that (ξ , ξ ) = 1. Hence, upon normalizing ξ (1) and ξ (2) ,
we have
   
1 1 1 −1 1 1 −1
T= √ , T = √ . (11)
2 −1 1 2 1 1

Letting x = Ty and substituting for x in Eq. (8), we obtain the following system of equations
for the new dependent variable y:
   
 −1 −3 0 1 2e−t − 3t
y = Dy + T g(t) = y+ √ −t
. (12)
0 −1 2 2e + 3t
September 11, 2008 11:18 boyce-9e-bvp Sheet number 454 Page number 434 cyan black

434 Chapter 7. Systems of First Order Linear Equations

Thus
√ −t 3
y1 + 3y1 = 2e − √ t,
2
(13)
√ 3
y2 + y2 = 2e−t + √ t.
2
Each of Eqs. (13) is a first order linear equation and so can be solved by the methods of Section
2.1. In this way we obtain
√  
2 −t 3 t 1
y1 = e −√ − + c1 e−3t ,
2 2 3 9
(14)
√ −t 3 −t
y2 = 2te + √ (t − 1) + c2 e .
2
Finally, we write the solution in terms of the original variables:
 
1 y1 + y2
x = Ty = √
2 −y1 + y2
⎛ √ ) √ * ⎞
(c1 / 2)e−3t + (c2 / 2) + 21 e−t + t − 43 + te−t
⎜ ⎟
=⎝ √ −3t ) √ * ⎠
−t −t
−(c1 / 2)e + (c2 / 2) − 2 e + 2t − 3 + te
1 5

           
1 −3t 1 −t 1 1 −t 1 1 1 4
= k1 e + k2 e + e + te−t + t− , (15)
−1 1 2 −1 1 2 3 5
√ √
where k1 = c1 / 2 and k2 = c2 / 2. The first two terms on the right side of Eq. (15) form the
general solution of the homogeneous system corresponding to Eq. (8). The remaining terms
are a particular solution of the nonhomogeneous system.

If the coefficient matrix A in Eq. (3) is not diagonalizable (because of repeated


eigenvalues and a shortage of eigenvectors), it can nevertheless be reduced to a
Jordan form J by a suitable transformation matrix T involving both eigenvectors and
generalized eigenvectors. In this case the differential equations for y1 , . . . , yn are not
totally uncoupled since some rows of J have two nonzero elements: an eigenvalue
in the diagonal position and a 1 in the adjacent position to the right. However, the
equations for y1 , . . . , yn can still be solved consecutively, starting with yn . Then the
solution of the original system (3) can be found by the relation x = Ty.

Undetermined Coefficients. A second way to find a particular solution of the nonhomo-


geneous system (1) is the method of undetermined coefficients. To make use of this
method, we assume the form of the solution with some or all of the coefficients un-
specified, and then seek to determine these coefficients so as to satisfy the differential
equation. As a practical matter, this method is applicable only if the coefficient ma-
trix P is a constant matrix, and if the components of g are polynomial, exponential, or
sinusoidal functions, or sums or products of these. In these cases the correct form of
the solution can be predicted in a simple and systematic manner. The procedure for
choosing the form of the solution is substantially the same as that given in Section 3.5
September 11, 2008 11:18 boyce-9e-bvp Sheet number 455 Page number 435 cyan black

7.9 Nonhomogeneous Linear Systems 435

for linear second order equations. The main difference is illustrated by the case of a
nonhomogeneous term of the form ueλt , where λ is a simple root of the characteristic
equation. In this situation, rather than assuming a solution of the form ateλt , it is
necessary to use ateλt + beλt , where a and b are determined by substituting into the
differential equation.

Use the method of undetermined coefficients to find a particular solution of


EXAMPLE    
−2 1 2e−t
2 
x =
1 −2
x+
3t
= Ax + g(t). (16)

This is the same system of equations as in Example 1. To use the method of undetermined
coefficients, we write g(t) in the form
   
2 −t 0
g(t) = e + t. (17)
0 3

Then we assume that


x = v(t) = ate−t + be−t + ct + d, (18)

where a, b, c, and d are vectors to be determined. Observe that r = −1 is an eigenvalue of the


coefficient matrix, and therefore we must include both ate−t and be−t in the assumed solution.
By substituting Eq. (18) into Eq. (16) and collecting terms, we obtain the following algebraic
equations for a, b, c, and d:

Aa = −a,
 
2
Ab = a − b − ,
0
  (19)
0
Ac = − ,
3
Ad = c.

From the first of Eqs. (19) we see that a is an eigenvector of A corresponding to the eigenvalue
r = −1. Thus aT = (α, α), where α is any nonzero constant. Then we find that the second of
Eqs. (19) can be solved only if α = 1 and that in this case
   
1 0
b=k − (20)
1 1

for any constant k. The simplest choice is k = 0, from which bT = (0, −1). Then the third and
fourth of Eqs. (19) yield cT = (1, 2) and dT = (− 43 , − 53 ), respectively. Finally, from Eq. (18) we
obtain the particular solution
       
1 −t 0 −t 1 1 4
v(t) = te − e + t− . (21)
1 1 2 3 5

The particular solution (21) is not identical to the one contained in Eq. (15) of Example 1
because the term in e−t is different. However, if we choose k = 21 in Eq. (20), then bT = ( 21 , − 21 )
and the two particular solutions agree.
September 11, 2008 11:18 boyce-9e-bvp Sheet number 456 Page number 436 cyan black

436 Chapter 7. Systems of First Order Linear Equations

Variation of Parameters. Now let us turn to more general problems in which the coef-
ficient matrix is not constant or not diagonalizable. Let

x = P(t)x + g(t), (22)

where P(t) and g(t) are continuous on α < t < β. Assume that a fundamental matrix
(t) for the corresponding homogeneous system

x = P(t)x (23)

has been found. We use the method of variation of parameters to construct a partic-
ular solution, and hence the general solution, of the nonhomogeneous system (22).
Since the general solution of the homogeneous system (23) is (t)c, it is natural to
proceed as in Section 3.6 and to seek a solution of the nonhomogeneous system (22)
by replacing the constant vector c by a vector function u(t). Thus we assume that

x = (t)u(t), (24)

where u(t) is a vector to be found. Upon differentiating x as given by Eq. (24) and
requiring that Eq. (22) be satisfied, we obtain

  (t)u(t) + (t)u (t) = P(t)(t)u(t) + g(t). (25)

Since (t) is a fundamental matrix,   (t) = P(t)(t); hence Eq. (25) reduces to

(t)u (t) = g(t). (26)

Recall that (t) is nonsingular on any interval where P is continuous. Hence  −1 (t)
exists, and therefore
u (t) =  −1 (t)g(t). (27)

Thus for u(t) we can select any vector from the class of vectors that satisfy Eq. (27);
these vectors are determined only up to an arbitrary additive constant vector; there-
fore we denote u(t) by

u(t) =  −1 (t)g(t) dt + c, (28)

where the constant vector c is arbitrary. If the integrals in Eq. (28) can be evaluated,
then the general solution of the system (22) is found by substituting for u(t) from
Eq. (28) in Eq. (24). However, even if the integrals cannot be evaluated, we can still
write the general solution of Eq. (22) in the form
 t
x = (t)c + (t)  −1 (s)g(s) ds, (29)
t1

where t1 is any point in the interval (α, β). Observe that the first term on the right
side of Eq. (29) is the general solution of the corresponding homogeneous system
(23), and the second term is a particular solution of Eq. (22).
Now let us consider the initial value problem consisting of the differential equation
(22) and the initial condition
x(t0 ) = x0 . (30)
September 11, 2008 11:18 boyce-9e-bvp Sheet number 457 Page number 437 cyan black

7.9 Nonhomogeneous Linear Systems 437

We can find the solution of this problem most conveniently if we choose the lower
limit of integration in Eq. (29) to be the initial point t0 . Then the general solution of
the differential equation is
 t
x = (t)c + (t)  −1 (s)g(s) ds. (31)
t0

For t = t0 the integral in Eq. (31) is zero, so the initial condition (30) is also satisfied
if we choose
c =  −1 (t0 )x0 . (32)

Therefore
 t
−1
x = (t) (t0 )x + (t)
0
 −1 (s)g(s) ds (33)
t0

is the solution of the given initial value problem. Again, although it is helpful to use
 −1 to write the solutions (29) and (33), it is usually better in particular cases to solve
the necessary equations by row reduction than to calculate  −1 and substitute into
Eqs. (29) and (33).
The solution (33) takes a slightly simpler form if we use the fundamental matrix
(t) satisfying (t0 ) = I. In this case we have
 t
x = (t)x0 + (t) −1 (s)g(s) ds. (34)
t0

Equation (34) can be simplified further if the coefficient matrix P(t) is a constant
matrix (see Problem 17).

Use the method of variation of parameters to find the general solution of the system
EXAMPLE    
−2 2e−t
3 
x =
1
x+ = Ax + g(t). (35)
1 −2 3t

This is the same system of equations as in Examples 1 and 2.


The general solution of the corresponding homogeneous system was given in Eq. (10). Thus
 
e−3t e−t
(t) = (36)
−e−3t e−t

is a fundamental matrix. Then the solution x of Eq. (35) is given by x = (t)u(t), where u(t)
satisfies (t)u (t) = g(t), or
    
e−3t e−t u1 2e−t
= . (37)
−e−3t e−t u2 3t

Solving Eq. (37) by row reduction, we obtain

u1 = e2t − 23 te3t ,


u2 = 1 + 23 tet .
September 11, 2008 11:18 boyce-9e-bvp Sheet number 458 Page number 438 cyan black

438 Chapter 7. Systems of First Order Linear Equations

Hence

u1 (t) = 21 e2t − 21 te3t + 16 e3t + c1 ,


u2 (t) = t + 23 tet − 23 et + c2 ,

and

x = (t)u(t)
           
1 −3t 1 −t 1 −t 1 1 −t 1 1 4
= c1 e + c2 e + te + e + t− , (38)
−1 1 1 2 −1 2 3 5

which is the same as the solution obtained previously.

Laplace Transforms. We used the Laplace transform in Chapter 6 to solve linear equa-
tions of arbitrary order. It can also be used in very much the same way to solve
systems of equations. Since the transform is an integral, the transform of a vector is
computed component by component. Thus L{x(t)} is the vector whose components
are the transforms of the respective components of x(t), and similarly for L{x (t)}.
We will denote L{x(t)} by X(s). Then, by an extension of Theorem 6.2.1 to vectors,
we also have
L{x (t)} = sX(s) − x(0). (39)

Use the method of Laplace transforms to solve the system


EXAMPLE    
−2 1 2e−t
4 
x =
1 −2
x+
3t
= Ax + g(t). (40)

This is the same system of equations as in Examples 1 through 3.


We take the Laplace transform of each term in Eq. (40), obtaining

sX(s) − x(0) = AX(s) + G(s), (41)

where G(s) is the transform of g(t). The transform G(s) is given by


 
2/(s + 1)
G(s) = . (42)
3/s2

We will simplify the remaining calculations by assuming that x(t) satisfies the initial condition
x(0) = 0. Then Eq. (41) becomes

(sI − A)X(s) = G(s), (43)

where, as usual, I is the identity matrix. Consequently, X(s) is given by

X(s) = (sI − A)−1 G(s). (44)

The matrix (sI − A)−1 is called the transfer matrix because multiplying it by the transform of
the input vector g(t) yields the transform of the output vector x(t). In this example we have
 
s+2 −1
sI − A = , (45)
−1 s+2
September 11, 2008 11:18 boyce-9e-bvp Sheet number 459 Page number 439 cyan black

7.9 Nonhomogeneous Linear Systems 439

and by a straightforward calculation we obtain


 
−1 1 s+2 1
(sI − A) = . (46)
(s + 1)(s + 3) 1 s+2

Then, substituting from Eqs. (42) and (46) in Eq. (44) and carrying out the indicated multipli-
cation, we find that
⎛ ⎞
2(s + 2) 3
+
⎜ (s + 1)2 (s + 3) s2 (s + 1)(s + 3) ⎟
⎜ ⎟
X(s) = ⎜ ⎟. (47)
⎝ 2 3(s + 2) ⎠
+
(s + 1)2 (s + 3) s2 (s + 1)(s + 3)
Finally, we need to obtain the solution x(t) from its transform X(s). This can be done by
expanding the expressions in Eq. (47) in partial fractions and using Table 6.2.1, or (more
efficiently) by using appropriate computer software. In any case, after some simplification the
result is          
2 −t 2 1 −3t 1 1 1 4
x(t) = e − e + te−t + t− . (48)
1 3 −1 1 2 3 5

Equation (48) gives the particular solution of the system (40) that satisfies the initial condition
x(0) = 0. As a result, it differs slightly from the particular solutions obtained in the preceding
three examples. To obtain the general solution of Eq. (40), you must add to the expression in
Eq. (48) the general solution (10) of the homogeneous system corresponding to Eq. (40).

Each of the methods for solving nonhomogeneous equations has some advantages
and disadvantages. The method of undetermined coefficients requires no integra-
tion, but it is limited in scope and may entail the solution of several sets of algebraic
equations. The method of diagonalization requires finding the inverse of the trans-
formation matrix and the solution of a set of uncoupled first order linear equations,
followed by a matrix multiplication. Its main advantage is that for Hermitian coeffi-
cient matrices, the inverse of the transformation matrix can be written down without
calculation—a feature that is more important for large systems. The method of
Laplace transforms involves a matrix inversion to find the transfer matrix, followed
by a multiplication, and finally by the determination of the inverse transform of each
term in the resulting expression. It is particularly useful in problems with forcing
functions that involve discontinuous or impulsive terms. Variation of parameters is
the most general method. On the other hand, it involves the solution of a set of
linear algebraic equations with variable coefficients, followed by an integration and
a matrix multiplication, so it may also be the most complicated from a computational
viewpoint. For many small systems with constant coefficients, such as the one in
the examples in this section, all of these methods work well, and there may be little
reason to select one over another.

PROBLEMS In each of Problems 1 through 12 find the general solution of the given system of equations.
     √   
2 −1 et 1 3 et

1. x = x+ 
2. x = √ x + √ −t
3 −2 t 3 −1 3e
September 11, 2008 11:18 boyce-9e-bvp Sheet number 460 Page number 440 cyan black

440 Chapter 7. Systems of First Order Linear Equations

       

2 −5 − cos t 
1 1 e−2t
3. x = x+ 4. x = x+
1 −2 sin t 4 −2 −2et
   
4 −2 t −3
5. x = x+ , t>0
8 −4 −t −2
   
−4 2 t −1
6. x = x+ , t>0
2 −1 2t −1 + 4
       
1 1 2 t 2 −1 1 t
7. x = x+ e 
8. x = x+ e
4 1 −1 3 −2 −1
 5     √   
−4 3
2t −3 2 1 −t
9. x = 4
x+ t 
10. x = √ x+ e
3
4
−45 e 2 −2 −1
   
2 −5 0
11. x = x+ , 0<t<π
1 −2 cos t
   
2 −5 csc t π
12. x = x+ , <t<π
1 −2 sec t 2
13. The electric circuit shown in Figure 7.9.1 is described by the system of differential equations
 1  1
dx −2 − 18 2
= x+ I(t), (i)
dt 2 − 21 0

where x1 is the current through the inductor, x2 is the voltage drop across the capacitor,
and I(t) is the current supplied by the external source.

L = 8 henrys

I(t)
1
C= farad
2

R = 4 ohms R = 4 ohms
FIGURE 7.9.1 The circuit in Problem 13.

(a) Determine a fundamental matrix (t) for the homogeneous system corresponding to
Eq. (i). Refer to Problem 25 of Section 7.6.
(b) If I(t) = e−t/2 , determine the solution of the system (i) that also satisfies the initial
conditions x(0) = 0.

In each of Problems 14 and 15 verify that the given vector is the general solution of the
corresponding homogeneous system, and then solve the nonhomogeneous system. Assume
that t > 0.
       

2 −1 1 − t2 1 1 −1
14. tx = x+ , x = c1
(c)
t + c2 t
3 −2 2t 1 3
September 11, 2008 11:18 boyce-9e-bvp Sheet number 461 Page number 441 cyan black

7.9 Nonhomogeneous Linear Systems 441

       

3 −2 −2t 1 −1 2 2
15. tx = x+ 4 , x (c)
= c1 t + c2 t
2 −2 t −1 2 1
16. Let x = φ(t) be the general solution of x = P(t)x + g(t), and let x = v(t) be some
particular solution of the same system. By considering the difference φ(t) − v(t), show
that φ(t) = u(t) + v(t), where u(t) is the general solution of the homogeneous system
x = P(t)x.
17. Consider the initial value problem

x = Ax + g(t), x(0) = x0 .

(a) By referring to Problem 15(c) in Section 7.7, show that


 t
x = (t)x +0
(t − s)g(s) ds.
0

(b) Show also that


 t
x = exp(At)x0 + exp[A(t − s)]g(s) ds.
0

Compare these results with those of Problem 27 in Section 3.6.

REFERENCES Further information on matrices and linear algebra is available in any introductory book on the subject.
The following is a representative sample:
Anton, H. and Rorres, C., Elementary Linear Algebra (8th ed.) (New York: Wiley, 2000).
Johnson, L. W., Riess, R. D., and Arnold, J. T., Introduction to Linear Algebra (5th ed.) (Reading, MA:
Addison-Wesley, 2001).
Kolman, B., Elementary Linear Algebra (8th ed.) (Upper Saddle River, NJ: Prentice-Hall, 2003).
Leon, S. J., Linear Algebra with Applications (6th ed.) (New York: Macmillan, 2002).
Strang, G., Linear Algebra and Its Applications (4th ed.)(New York: Academic Press, 1998).

The following book treats elementary differential equations with a particular emphasis on systems of
first order equations:
Brannan, J. R. and Boyce, W. E., Differential Equations: An Introduction to Modern Methods and
Applications (New York: Wiley, 2007).
September 11, 2008 11:18 boyce-9e-bvp Sheet number 462 Page number 442 cyan black

Вам также может понравиться