Vous êtes sur la page 1sur 32

6.

4 Systems of Equations 487


6.4 Systems of Equations
6.4.1 Linear Systems
We now consider rst-order linear systems of differential equations:
X

= A(t)X +F(t), (6.28)


where
X(t) =

x
1
(t)
x
2
(t)
.
.
.
x
n
(t)

, A(t) =

a
11
(t) a
12
(t) . . . a
1n
(t)
a
21
(t) a
22
(t) . . . a
2n
(t)
.
.
.
.
.
. . . .
.
.
.
a
n1
(t) a
n2
(t) . . . a
nn
(t)

, and F(t) =

f
1
(t)
f
2
(t)
.
.
.
f
n
(t)

.
Homogeneous Linear Systems
The corresponding homogeneous system of equation (6.28) is
X

= AX. (6.29)
In the same way as with the previously discussed linear equations, a general solu-
tion of (6.28) is X = X
h
+ X
p
where X
h
is a general solution of (6.29) and X
p
is a
particular solution of the nonhomogeneous system (6.28). A particular solution to a
system of ordinary
differential equations is a set
of functions that satisfy the
system but do not contain
any arbitrary constants. That
is, a particular solution to a
system is a set of specic
functions, containing no
arbitrary constants, that satisfy
the system.
If
1
,
2
, . . . ,
n
are n linearly independent solutions of (6.29), a general
solution of (6.29) is
X = c
1

1
+c
2

2
+ +c
n

n
=

1

2

n

c
1
c
2
.
.
.
c
n

= C,
where
=

1

2

n

and C =

c
1
c
2
.
.
.
c
n

.
is called a fundamental matrix for (6.29). If is a fundamental matrix for (6.29),

= A or

A = 0.
After loading the DEtools package and dening A, the command
matrixDE(A,t)
attempts to nd a fundamental matrix for (6.29).
488 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
A(t) constant
Suppose that A(t) = A has constant real entries. Let be an eigenvalue of A with
corresponding eigenvector v. Then, ve
t
is a solution of X

= AX.
If = + i, = 0, is an eigenvalue of A and has corresponding eigenvector
v = a +bi, two linearly independent solutions of X

= AX are
e
t
_
a cos t bsin t
_
and e
t
_
a sin t +bcos t
_
. (6.30)
EXAMPLE 6.4.1: Solve each of the following systems: (a) X

=
_
1/2 1/3
1/3 1/2
_
X; (b)
_
x

=
1
2
y
y

=
1
8
x
; (c)
_
dx/dt =
1
4
x + 2y
dy/dt = 8x
1
4
y
.
SOLUTION: (a) With eigenvects, which is contained in the linalg
package, we see that the eigenvalues and eigenvectors of A =
_
1/2 1/3
1/3 1/2
_
are
1
= 1/6 and
2
= 5/6 and v
1
=
_
1
1
_
and
v
2
=
_
1
1
_
, respectively.
> with(linalg):
> with(DEtools):
> A:=matrix(2,2,[-1/2,-1/3,-1/3,-1/2]);
A := matrix
_
[[1/2, 1/3], [1/3, 1/2]]
_
> eigenvects(A);
[1/6, 1, {vector ([1, 1])}], [5/6, 1, {vector ([1, 1])}]
Then X
1
=
_
1
1
_
e
t/6
and X
2
=
_
1
1
_
e
5t/6
are two lin-
early independent solutions of the system so a general solution
is X =
_
e
t/6
e
5t/6
e
t/6
e
5t/6
__
c
1
c
2
_
; a fundamental matrix is =
_
e
t/6
e
5t/6
e
t/6
e
5t/6
_
, which we conrm using matrixDE.
> matrixDE(A,t);
_
matrix
_
[[e
1/6 t
, e
5/6 t
], [e
1/6 t
, e
5/6 t
]]
_
, vector ([0, 0])
_
6.4 Systems of Equations 489
y(t)
1
0.5
0
-0.5
-1
x(t)
1 0.5 0 -0.5 -1
Figure 6-27 Direction eld for X

= AX
We use dsolve to nd a general solution of the system by entering
> gensol:=dsolve(diff(x(t),t)=-1/2*x(t)-1/3*y(t),
diff(y(t),t)=-1/3*x(t)-1/2*y(t),x(t),y(t));
gensol :=

x(t) =_C1e
5/6t
+_C2e
1/6t
, y(t) =_C1e
5/6t
_C2e
1/6t

We graph the direction eld with DEplot, which is contained in the


DEtools package, in Figure 6-27.
Remark. After you have loaded the DEplot package,
DEplot([diff(x(t),t)=f(x(t),y(t)),diff(y(t),t)=g(x(t),
y(t))],x=a..b,y=c..d,scene=[x(t),y(t)])
generates a basic direction eld for the system {x

= f (x, y), y

= g(x, y)}
for a x b and c y d.
> DEplot([diff(x(t),t)=-1/2*x(t)-1/3*y(t),diff(y(t),t)
=-1/3*x(t)-1/2*y(t)],
> [x(t),y(t)],t=-1..1,x=-1..1,y=-1..1,scene=[x(t),
> y(t)],scaling=CONSTRAINED,color=BLACK);
Several solutions are also graphed with DEplot and shown together
with the direction eld in Figure 6-28.
> ivals:=seq(-1+.25*i,i=0..8):
> i1:=seq([x(0)=1,y(0)=i],i=ivals):
> i2:=seq([x(0)=i,y(0)=1],i=ivals):
> i3:=seq([x(0)=-1,y(0)=i],i=ivals):
> i4:=seq([x(0)=i,y(0)=-1],i=ivals):
490 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
y(t)
1
0.5
0
-0.5
-1
x(t)
1 0.5 0 -0.5 -1
Figure 6-28 Direction eld for X

= AX along with various solution curves


> DEplot([diff(x(t),t)=-1/2*x(t)-1/3*y(t),
> diff(y(t),t)=-1/3*x(t)-1/2*y(t)],[x(t),y(t)],t=0..10,
> [i1,i2,i3,i4],x=-1..1,y=-1..1,scene=[x(t),y(t)],
scaling=CONSTRAINED,
> color=BLACK,linecolor=GRAY,thickness=1);
(b) In matrix form the system is equivalent to the system X

0 1/2
1/8 0

X. As in (a), we use eigenvects to see that the eigen-


values and eigenvectors of A =

0 1/2
1/8 0

are
1,2
= 0
1
4
i and
v
1,2
=

1
0

0
1/2

i.
> A:=matrix(2,2,[0,1/2,-1/8,0]);
A := matrix

[[0, 1/2], [1/8, 0]]

> eigenvects(A);
[1/4 i, 1,

vector

[1, 1/2 i]

], [1/4 i, 1,

vector

[1, 1/2 i]

]
Two linearly independent solutions are then X
1
=

1
0

cos
1
4
t

0
1/2

sin
1
4
t =

cos
1
4
t

1
2
sin
1
4
t

and X
2
=

1
0

sin
1
4
t +

0
1/2

cos
1
4
t
=

sin
1
4
t
1
2
cos
1
4
t

and a general solution is X = c


1
X
1
+ c
2
X
2
6.4 Systems of Equations 491
=

cos
1
4
t sin
1
4
t

1
2
sin
1
4
t
1
2
cos
1
4
t

c
1
c
2

or x = c
1
cos
1
4
t + c
2
sin
1
4
t and y
= c
1
1
2
sin
1
4
t +
1
2
c
2
cos
1
4
t.
> matrixDE(A,t);
[matrix

[[cos

1/4 t

, sin

1/4 t

], [1/2 sin

1/4 t

,
1/2 cos

1/4 t

]]

, vector ([0, 0])]


As before, we use dsolve to nd a general solution.
> gensol:=dsolve(diff(x(t),t)=1/2*y(t),diff(y(t),t)
=-1/8*x(t),x(t),y(t));
gensol :=

y (t) = 1/2 _C1 cos

1/4 t

1/2 _C2 sin

1/4 t

, x (t)
= _C1 sin

1/4 t

+ _C2 cos

1/4 t

Initial-value problems for systems are solved in the same way as for
other equations. For example, entering
> partsol:=dsolve(diff(x(t),t)=1/2*y(t),diff(y(t),t)
=-1/8*x(t),x(0)=1,y(0)=-1,x(t),y(t));
partsol :=

y (t) = cos

1/4 t

1/2 sin

1/4 t

, x (t) = 2 sin

1/4 t

+ cos

1/4 t

nds the solution that satises x(0) = 1 and y(0) = 1.


We graph x(t) and y(t) together as well as parametrically with plot
in Figure 6-29.
> assign(partsol):
> plot([x(t),y(t)],t=0..8*Pi,color=[BLACK,GRAY]);
> plot([x(t),y(t),t=0..8*Pi],color=BLACK,scaling=
CONSTRAINED);
We can also use DEplot to graph the direction eld and/or various
solutions as we do next in Figure 6-30.
> x:=x:y:=y:
> i1:=seq([x(0)=0,y(0)=-1+.25*i],i=0..8):
> DEplot([diff(x(t),t)=1/2*y(t),diff(y(t),t)
=-1/8*x(t)],[x(t),y(t)],
> t=0..8*Pi,[i1],x=-1..1,y=-1..1,scene=[x(t),y(t)],
scaling=CONSTRAINED,
> color=GRAY,linecolor=BLACK,thickness=1,arrows=LARGE);
492 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
2
1
(a) (b)
0
-1
-2
t
25 20 15 10 5 0
1
0.5
0
-0.5
-1
2 1 0 -1 -2
Figure 6-29 (a) Graph of x(t) and y(t). (b) Parametric plot of x(t) versus y(t)
-1 -0.5
y(t)
-1
0 1
-0.5
0
x(t)
1
0.5
0.5
Figure 6-30 Notice that all nontrivial solutions are periodic
(c) In matrix form, the system is equivalent to the system X

1
4
2
8
1
4

X. The eigenvalues and corresponding eigenvectors of A =

1
4
2
8
1
4

are found to be
1,2
=
1
4
4i and v
1,2
=

0
2

1
0

i with
eigenvects.
> A:=matrix(2,2,[-1/4,2,-8,-1/4]):
> eigenvects(A);
[1/4 + 4 i, 1, {vector ([1, 2 i])}], [1/4 4 i, 1, {vector ([1, 2 i])}]
6.4 Systems of Equations 493
A general solution is then
X = c
1
X
1
+c
2
X
2
= c
1
e
t/4
__
1
0
_
cos 4t
_
0
2
_
sin 4t
_
+c
2
e
t/4
__
1
0
_
sin 4t +
_
0
2
_
cos 4t
_
= e
t/4
_
c
1
_
cos 4t
2 sin 4t
_
+c
2
_
sin 4t
2 cos 4t
__
= e
t/4
_
cos 4t sin 4t
2 sin 4t 2 cos 4t
__
c
1
c
2
_
or x = e
t/4
(c
1
cos 4t +c
2
sin 4t) and y = e
t/4
(2c
2
cos 4t 2c
1
sin 4t).
> matrixDE(A,t);
_
matrix
_
[[e
1/4 t
cos
_
4 t
_
, e
1/4 t
sin
_
4 t
_
_
,
_
2 e
1/4 t
sin
_
4 t
_
, 2 e
1/4 t
cos
_
4 t
_
]]
_
, vector
_
[0, 0]
_
_
We conrm this result using dsolve.
> dsolve(diff(x(t),t)=-1/4*x(t)+2*y(t),diff(y(t),t)
=-8*x(t)-1/4*y(t),
> x(t),y(t));
_
y (t) = 2 e
1/4 t
(_C1 cos (4 t) +_C2 sin (4 t)) ,
x (t) = e
1/4 t
(_C1 sin (4 t) +_C2 cos (4 t))
_
We use DEplot to graph the direction eld associated with the system
along with various solutions in Figure 6-31.
> ivals:=seq(-1+.25*i,i=0..8):
> i1:=seq([x(0)=1,y(0)=i],i=ivals):
> i2:=seq([x(0)=i,y(0)=1],i=ivals):
> i3:=seq([x(0)=-1,y(0)=i],i=ivals):
> i4:=seq([x(0)=i,y(0)=-1],i=ivals):
> DEplot([diff(x(t),t)=-1/4*x(t)+2*y(t),diff(y(t),t)
=-8*x(t)-1/4*y(t)],
> [x(t),y(t)],t=0..10,[i1,i2,i3,i4],x=-1..1,y=-1..1,
scene=[x(t),y(t)],
> scaling=CONSTRAINED,color=GRAY,linecolor=BLACK,
thickness=1,
> stepsize=.05);
494 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
y(t)
1
0.5
0
-0.5
-1
x(t)
1 0.5 0 -0.5 -1
Figure 6-31 Various solutions and direction eld associated with the system
Last, we illustrate how to solve an initial-value problem and graph the
resulting solutions by nding the solution that satises the initial condi-
tions x(0) = 100 and y(0) = 10 and then graphing the results with plot
in Figure 6-32.
> x:=x:y:=y:
> partsol:=dsolve(diff(x(t),t)=-1/4*x(t)+2*y(t),
> diff(y(t),t)=-8*x(t)-1/4*y(t),x(0)=100,y(0)=10,x(t),
y(t));
> assign(partsol):
> plot([x(t),y(t)],t=0..20,color=[BLACK,GRAY]);
> plot([x(t),y(t),t=0..20],color=BLACK);
partsol :=

y (t) = 2 e
1/4 t
(5 cos (4 t) +100 sin (4 t)) , x (t)
= e
1/4 t
(5 sin (4 t) +100 cos (4 t))

Application: The Double Pendulum


The motion of a double pendulum is modeled by the system of differential
equations

(m
1
+m
2
) l
1
2
d
2

1
dt
2
+m
2
l
1
l
2
d
2

2
dt
2
+(m
1
+m
2
) l
1
g
1
= 0
m
2
l
2
2
d
2

2
dt
2
+m
2
l
1
l
2
d
2

1
dt
2
+m
2
l
2
g
2
= 0
using the approximation sin for small displacements.
1
represents the dis-
placement of the upper pendulumand
2
that of the lower pendulum. Also, m
1
and
6.4 Systems of Equations 495
t
20 15 10 5 0
150
(a) (b)
100
50
0
-50
-100
-150
150
100
50
0
-50
-100
-150
80 40 0 -40 -80
Figure 6-32 (a) Graph of x(t) and y(t). (b) Parametric plot of x(t) versus y(t)
m
2
represent the mass attached to the upper and lower pendulums, respectively,
while the length of each is given by l
1
and l
2
.
EXAMPLE 6.4.2: Suppose that m
1
= 3, m
2
= 1, and each pendulumhas
length 16. If
1
(0) = 1,
1

(0) = 0,
2
(0) = 1, and
2

(0) = 0, solve the


double pendulum problem using g = 32. Plot the solution.
SOLUTION: In this case, the system to be solved is

4 16
2
d
2

1
dt
2
+16
2
d
2

2
dt
2
+4 16 32
1
= 0
16
2
d
2

2
dt
2
+16
2
d
2

1
dt
2
+16 32
2
= 0
,
which we simplify to obtain

4
d
2

1
dt
2
+
d
2

2
dt
2
+8
1
= 0
d
2

2
dt
2
+
d
2

1
dt
2
+2
2
= 0
.
First, we use dsolve to solve the initial value problem.
> Eq1:=4*diff(theta[1](t),t$2)+diff(theta[2](t),t$2)+
> 8*theta[1](t)=0:
496 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
2
(a) (b)
0
1
20
t
15 5
-2
-1
0 10
2
1
0
-2
1 0.5
-1
-0.5 0 -1
Figure 6-33 (a)
1
(t) (in black) and
2
(t) (in gray) as functions of t. (b) Parametric plot of

1
(t) versus
2
(t)
> Eq2:=diff(theta[1](t),t$2)+diff(theta[2](t),t$2)+
> 2*theta[2](t)=0:
To solve the initial-value problem using traditional methods, we use
the method of Laplace transforms. To do so, we dene sys to be the The Laplace transform of
y = f (t) is F(s)=L

f (t)

0
e
st
f (t) dt.
system of equations and then use dsolve together with the option
method=laplace to compute the Laplace transform of each equation.
> sola:=dsolve(Eq1,Eq2,theta[1](0)=1,D(theta[1])(0)=0,
theta[2](0)=1,
> D(theta[2])(0)=0,theta[1](t),theta[2](t),
method=laplace);
sola :=

1
(t) = 1/4 cos (2 t) +3/4 cos

2/3

3t

,
2
(t)
= 3/2 cos

2/3

3t

1/2 cos (2 t)

These two functions are graphed together in Figure 6-33(a) and para-
metrically in Figure 6-33(b).
> assign(sola):
> plot([theta[1](t),theta[2](t)],t=0..20,
color=[BLACK,GRAY]);
> plot([theta[1](t),theta[2](t),t=0..20],color=BLACK);
We can illustrate the motion of the pendulum as follows. First, we
dene the function pen2.
> pen2:=proc(t0,len1,len2)
> local pt1,pt2,xt0,yt0;
6.4 Systems of Equations 497
> xt0:=evalf(subs(t=t0,theta[1](t)));
> yt0:=evalf(subs(t=t0,theta[2](t)));
> pt1:=[len1*cos(3*Pi/2+xt0),len1*sin(3*Pi/2+xt0)];
> pt2:=[len1*cos(3*Pi/2+xt0)+len2*cos(3*Pi/2+yt0),
> len1*sin(3*Pi/2+xt0)+len2*sin(3*Pi/2+yt0)];
> plot([[0,0],pt1,pt2],xtickmarks=2,ytickmarks=2,
> view=[-32..32,-32..0]);
> end:
Next, we dene ivals to be a list of 16 evenly spaced numbers between
0 and10. seqis then usedto apply pen2to the list of numbers inivals.
The resulting set of graphics is displayed as an array using display
with the option insequence=true in Figure 6-34.
> with(plots):
> ivals:=[seq(10*i/15,i=0..15)]:
0.
-.2e2
0. -2.
0.
-.1e2
0.
0.
.1e2
-.2e2
0.
-.1e2
0. -.2e2
0.
0.
-.2e2
-.2e2
0.
-.2e2 4.
0.
0.
-.2e2
0. -.1e2
0.
0.
-.2e2
-.1e2
-.2e2
2. 0.
0.
-2.
-.1e2
0. -.2e2
0.
-.2e2
0. -.1e2
0.
0.
-.2e2
-.2e2
.2e2
4. 2.
0.
0.
0.
-.2e2
.2e2 0.
0.
-.1e2
0.
.2e2
-.1e2
0.
.2e2 0.
0.
-.1e2
.1e2 0.
Figure 6-34 The double pendulum for 16 equally spaced values of t between 0 and 10
498 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
> toshow:=[seq(pen2(i,16,16),i=ivals)]:
> nops(toshow);
16
> anarray:=display(toshow,insequence=true):
> display(anarray);
We can also use displayto generate an animation. We showone frame
from the animation that results from the following command.
> display(toshow,insequence=true);
0
-20
0

6.4.2 Nonhomogeneous Linear Systems


Generally, the method of undetermined coefcients is difcult to implement for
nonhomogeneous linear systems as the choice for the particular solution must be
very carefully made.
Variation of parameters is implemented in much the same way as for rst-order
linear equations.
Let X
h
be a general solutionto the corresponding homogeneous systemof (6.28),
X a general solution of (6.28), and X
p
a particular solution of (6.28). It then follows
that XX
p
is a solution to the corresponding homogeneous system so XX
p
= X
h
and, consequently, X = X
h
+X
p
.
Aparticular solution of (6.28) is found in much the same way as with rst-order
linear equations. Let be a fundamental matrix for the corresponding homoge-
neous system. We assume that a particular solution has the form X
p
= U(t).
6.4 Systems of Equations 499
Differentiating X
p
gives us
X
p

U+U

.
Substituting into (6.28) results in

U+U

= AU+F
U

= F
U

=
1
F
U =


1
F dt,
where we have used the fact that

UAU =

U = 0. It follows that
X
p
=


1
F dt. (6.31)
A general solution is then
X = X
h
+X
p
= C +


1
F dt
=

C +


1
F dt


1
F dt,
where we have incorporated the constant vector C into the indenite integral


1
F dt.
EXAMPLE 6.4.3: Solve the initial-value problem
X

1 1
10 1

t cos 3t
t sin t +t cos 3t

, X(0) =

1
1

.
Remark. In traditional form, the system is equivalent to

= x y t cos 3t
y

= 10x y t sin t t cos 3t


, x(0) = 1, y(0) = 1.
500 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
SOLUTION: The corresponding homogeneous system is X

h
=

1 1
10 1

X
h
. The eigenvalues and corresponding eigenvectors of A =

1 1
10 1

are
1,2
= 3i and v
1,2
=

1
10

3
0

i, respectively.
> with(linalg):
> with(DEtools):
> A:=matrix(2,2,[1,-1,10,-1]):
> eigenvects(A);
[3 i, 1, {vector ([1, 1 3 i])}] , [3 i, 1, {vector ([1, 1 + 3 i])}]
A fundamental matrix is =

sin 3t cos 3t
sin 3t 3 cos 3t cos 3t + 3 sin 3t

> fm:=matrixDE(A,t);
fm := [matrix ([[sin (3 t) , cos (3 t)], [sin (3 t) 3 cos (3 t) , cos (3 t) ,
+3 sin (3 t)]]) vector ([0, 0])]
> fm[1];
matrix ([[sin (3 t) , cos (3 t)], [sin (3 t) 3 cos (3 t) , cos (3 t) + 3 sin (3 t)]])
> fminv:=simplify(inverse(fm[1]));

1/3 cos (3 t) + sin (3 t) 1/3 cos (3 t)


1/3 sin (3 t) + cos (3 t) 1/3 sin (3 t)

We now compute
1
F(t)
> ft:=matrix(2,1,[-t*cos(3*t),-t*sin(t)-t*cos(3*t)]);

t cos (3 t)
t sin (t) t cos (3 t)

> step1:=evalm(fminv &* ft);

(1/3 cos(3t)+sin(3t))t cos(3t)1/3 cos(3t)(t sin(t)t cos(3t))


(1/3 sin(3t)+cos(3t))t cos(3t)+1/3 sin(3t)(t sin(t)t cos(3t))

and


1
F(t) dt.
For length considerations, we
display only the nal results.
To see each result as it is
generated, replace the colons
with semi-colons. > step2:=map(int,step1,t):
6.4 Systems of Equations 501
A general solution of the nonhomogeneous system is then

__

1
F(t) dt +C
_
.
> simplify(evalm(fm[1] &* step2)):
It is easiest to use matrixDE
> check1:=matrixDE(A,ft,t):
> check1[1];
_
cos (3t) sin (3t)
cos (3t) +3 sin (3t) sin (3t) 3 cos (3t)
_
> check1[2];
_

1
72
cos (3t) 1/32 cos (t) 1/12 sin (3t) t 1/4 cos (3t) t
2
+1/8t sin (t) ,

1
72
cos (3t) 1/32 cos (t) 1/12 sin (3t) t 1/4 cos (3t) t
2
+1/8t sin (t) +1/24 sin (3t)

5
32
sin (t) +3/4t cos (3t) 3/4 sin (3t) t
2
+
23
8
t cos (t) 4t (cos (t))
3
_
or dsolve to solve the initial-value problem directly.
> check2:=dsolve(diff(x(t),t)=x(t)-y(t)-t*cos(3*t),
> diff(y(t),t)=10*x(t)-y(t)-t*sin(t)-t*cos(3*t),x(0)=1,y(0)=-1,
> x(t),y(t));
check2 :=
_
y (t) =
31
32
cos (3 t) +
123
32
sin (3 t)
5
32
sin (t) 1/4 t cos (3 t) 1/8 t cos (t)
3/4 sin (3 t) t
2
1/12 sin (3 t) t +1/8 t sin (t) 1/32 cos (t) 1/4 cos (3 t) t
2
,
x (t) = 2/3 sin (3 t) +
33
32
cos (3 t) 1/12 sin (3 t) t +1/8 t sin (t)
1/32 cos (t) 1/4 cos
_
3 t
_
t
2
_
The solutions are graphed with plot in Figure 6-35.
> assign(check2):
> plot([x(t),y(t)],t=0..8*Pi,color=[BLACK,GRAY]);
> plot([x(t),y(t),t=0..8*Pi],color=BLACK,scaling=CONSTRAINED);

502 Chapter 6 Applications Related to Ordinary and Partial Differential Equations


400
(a) (b)
0
200
10
200
t
25 5 0
400
15 20
400
0
200
200
100 50 15010050 0
400
Figure 6-35 (a) Graph of x(t) (in black) and y(t) (in gray). (b) Parametric plot of x(t)
versus y(t)
6.4.3 Nonlinear Systems
Nonlinear systems of differential equations arise in numerous situations. Rigorous
analysis of the behavior of solutions to nonlinear systems is usually very difcult,
if not impossible.
To generate numerical solutions of equations, use dsolve with the numeric
option.
6.4 Systems of Equations 503
Also see Example 6.4.7.
EXAMPLE 6.4.4 (Van-der-Pols equation): Van-der-Pols equationx

x
2
1

+x = 0 can be written as the system


x

= y
y

= x

x
2
1

y.
(6.32)
If = 2/3, x(0) = 1, and y(0) = 0, (a) nd x(1) and y(1). (b) Graph the
solution that satises these initial conditions.
SOLUTION: We use dsolve with the numeric option to solve (6.32)
with = 2/3 subject to x(0) = 1 and y(0) = 0. We name the resulting
numerical solution numsol.
> with(plots):
> numsol:=dsolve([diff(x(t),t)=y(t),diff(y(t),t)=-x(t)
-2/3*(x(t)2-1)*y(t),
> x(0)=1,y(0)=0],[x(t),y(t)],numeric);
Warning, the name changecoords has been redefined
numsol := proc(x
r
kf 45)...endproc
We evaluate numsol if t = 1 to see that x(1) .5128 and y(1) .9692.
> numsol(1);
[t = 1.0, x (t) = 0.512847902997304538, y (t) = 0.969203620640395002]
odeplotis usedtographx(t) andy(t) together inFigure 6-36(a); a three-
dimensional plot, (t, x(t), y(t)), is shown in Figure 6-36(b); a parametric
plot is shown in Figure 6-36(c); and the limit cycle is shown more clearly
in Figure 6-36(d) by graphing the solution for 20 t 30.
> odeplot(numsol,[[t,x(t)],[t,y(t)]],0..15,numpoints
=200);
> odeplot(numsol,[t,x(t),y(t)],0..15,axes=BOXED,
numpoints=200);
> odeplot(numsol,[x(t),y(t)],0..15,numpoints=200);
> odeplot(numsol,[x(t),y(t)],20..30,numpoints=200);

504 Chapter 6 Applications Related to Ordinary and Partial Differential Equations


x, y
1
2
0
-2
t
12 10 8
-1
6 0 14 2 4
-2
-1
x
0
1
2
0
2
4
6
8
t 10
12
14
y
1
2
0
-2
x
2 1
(c)
(a) (b)
(d)
-2
-1
0 -1
y
1
2
0
-2
x
2 1 -2
-1
0 -1
Figure 6-36 (a) x(t) and y(t). (b) Athree-dimensional plot. (c) x(t) versus y(t). (d) x(t) versus
y(t) for 20 t 30
Linearization
Consider the autonomous system of the form An autonomous system
does not explicitly depend on
the independent variable, t.
That is, if you write the
system omitting all
arguments, the independent
variable (typically t) does not
appear.
x
1

= f
1
(x
1
, x
2
, . . . , x
n
)
x
2

= f
2
(x
1
, x
2
, . . . , x
n
)
.
.
.
x
n

= f
n
(x
1
, x
2
, . . . , x
n
) .
(6.33)
An equilibrium (or rest) point, E = (x
1

, x
2

, . . . , x
n

), of (6.33) is a solution of the


system
f
1
(x
1
, x
2
, . . . , x
n
) = 0
f
2
(x
1
, x
2
, . . . , x
n
) = 0
.
.
.
f
n
(x
1
, x
2
, . . . , x
n
) = 0.
(6.34)
6.4 Systems of Equations 505
The Jacobian of (6.33) is
J (x
1
, x
2
, . . . , x
n
) =

f
1
x
1
f
1
x
2
. . .
f
1
x
n
f
2
x
1
f
2
x
2
. . .
f
2
x
n
.
.
.
.
.
. . . .
.
.
.
f
n
x
1
f
n
x
2
. . .
f
n
x
n

.
Use the jacobian function, which is contained in the linalg package, to
compute the Jacobian matrix for a set of functions.
The rest point, E, is locally stable if and only if all the eigenvalues of J(E) have
negative real part. If E is not locally stable, E is unstable.
EXAMPLE 6.4.5 (Dufngs Equation): Consider the forced pendu-
lum equation with damping,
x

+kx

+sin x = F(t). (6.35)


Recall the Maclaurin series for sin x: sin x = x
1
3!
x
3
+
1
5!
x
5

1
7!
x
7
+ .
Using sin x x, (6.35) reduces to the linear equation x

+kx

+x = F(t).
On the other hand, using the approximation sin x x
1
6
x
3
, we
obtain x

+kx

x
1
6
x
3

= F(t). Adjusting the coefcients of x and


506 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
x
3
and assuming that F(t) = F cos t gives us Dufngs equation:
x

+kx

+cx +x
3
= F cos t, (6.36)
where k and c are positive constants.
Let y = x

. Then, y

= x

= F cos t kx

cx x
3
= F cos t ky
cx x
3
and we can write (6.36) as the system
x

= y
y

= F cos t ky cx x
3
.
(6.37)
Assuming that F = 0 results in the autonomous system
x

= y
y

= cx x
3
ky.
(6.38)
The rest points of system (6.38) are found by solving
x

= y
y

= cx x
3
ky,
resulting in E
0
= (0, 0).
> with(DEtools):
> with(linalg):
> solve(y=0,-c*x-epsilon*x3-k*y=0,x,y);

y = 0, x = 0

y = 0, x = RootOf

c + _Z
2
, label = _L1

We nd the Jacobian of (6.38) in s1, evaluate the Jacobian at E


0
,
> s1:=jacobian([y,-c*x-epsilon*x3-k*y],[x,y]);

0 1
c 3 x
2
k

> s2:=subs(x=0,eval(s1));

0 1
c k

and then compute the eigenvalues with eigenvalues.


> s3:=eigenvalues(s2);
6.4 Systems of Equations 507
s3 := 1/2 k +1/2

k
2
4 c, 1/2 k 1/2

k
2
4 c
Because k and c are positive, k
2
4c < k
2
so the real part of each
eigenvalue is always negative if k
2
4c = 0. Thus, E
0
is locally stable.
For the autonomous system
x

= f (x, y)
y

= g(x, y),
(6.39)
Bendixsons theorem states that if f
x
(x, y)+g
y
(x, y) is a continuous func-
tion that is either always positive or always negative in a particular
region R of the plane, then system (6.39) has no limit cycles in R. For
(6.38) we have
d
dx

+
d
dy

cx x
3
ky

= k,
which is always negative. Hence, (6.38) has no limit cycles andit follows
that E
0
is globally, asymptotically stable.
> diff(y,x)+diff(-c*x-epsilon*x3-k*y,y);
k
We use DEplot to illustrate two situations that occur. In Figure 6-37(a),
we use c = 1, = 1/2, and k = 3. In this case, E
0
is a stable node. On the
other hand, in Figure 6-37(b), we use c = 10, = 1/2, and k = 3. In this
case, E
0
is a stable spiral.
> ivals:=seq(-2.5+.5*i,i=0..10):
> i1:=seq([x(0)=2.5,y(0)=i],i=ivals):
> i2:=seq([x(0)=i,y(0)=2.5],i=ivals):
> i3:=seq([x(0)=-2.5,y(0)=i],i=ivals):
> i4:=seq([x(0)=i,y(0)=-2.5],i=ivals):
> DEplot([diff(x(t),t)=y(t),diff(y(t),t)
=-1*x(t)-1/2*x(t)3-3*y(t)],
> [x(t),y(t)],t=0..10,[i1,i2,i3,i4],x=-2.5..2.5,
y=-2.5..2.5,
> scene=[x(t),y(t)],scaling=CONSTRAINED,color=GRAY,
linecolor=BLACK,
> thickness=1,stepsize=.05);
> ivals:=seq(-1+.25*i,i=0..8): i1:=seq([x(0)=1,y(0)=i],
i=ivals):
> i2:=seq([x(0)=i,y(0)=1],i=ivals):
> i3:=seq([x(0)=-1,y(0)=i],i=ivals):
508 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
y(t)
2
(a) (b)
1
0
-1
-2
x(t)
2 1 0 -1 -2
y(t)
1
0.5
0
-0.5
-1
x(t)
1
0.5
0
-0.5 -1
Figure 6-37 (a) The origin is a stable node. (b) The origin is a stable spiral
> i4:=seq([x(0)=i,y(0)=-1],i=ivals):
> DEplot([diff(x(t),t)=y(t),diff(y(t),t)
=-10*x(t)-1/2*x(t)3-3*y(t)],
> [x(t),y(t)],t=0..10,[i1,i2,i3,i4],x=-1..1,y=-1..1,
> scene=[x(t),y(t)],scaling=CONSTRAINED,color=GRAY,
linecolor=BLACK,
> thickness=1,stepsize=.01);
EXAMPLE 6.4.6 (Predator Prey): The predatorprey equations take
the form
dx
dt
= ax bxy
dy
dt
= dxy cy
where a, b, c, and d are positive constants. x represents the size of the
prey population at time t while y represents the size of the predator
population at time t. We use solve to calculate the rest points. In this
case, there is one boundary rest point, E
0
= (0, 0), and one interior rest
point, E
1
= (c/d, a/b).
> with(linalg):
> with(DEtools):
> rps:=solve(a*x-b*x*y=0,d*x*y-c*y=0,x,y);
rps :=

x = 0, y = 0

x =
c
d
, y =
a
b

6.4 Systems of Equations 509


y(t)
2
1.5
1
0.5
0
x(t)
2 1.5 1 0.5 0
Figure 6-38 Multiple limit cycles about the interior rest point
The Jacobian is then found using jacobian.
> jac:=jacobian([a*x-b*x*y,d*x*y-c*y],[x,y]);
jac := matrix

[[a by, xb], [yd, dx c]]

E
0
is unstable because one eigenvalue of J(E
0
) is positive. For the lin-
earized system, E
1
is a center because the eigenvalues of J(E
1
) are
complex conjugates.
> eigenvalues(subs(rps[1],eval(jac)));
a, c
> eigenvalues(subs(rps[2],eval(jac)));

ca,

ca
In fact, E
1
is a center for the nonlinear system as illustrated in
Figure 6-38, where we have used a = 1, b = 2, c = 2, and d = 1.
Notice that there are multiple limit cycles around E
1
= (1/2, 1/2).
> inits:=seq([x(0)=i/20,y(0)=i/20],i=3..10):
> DEplot([diff(x(t),t)=x(t)-2*x(t)*y(t),
> diff(y(t),t)=2*x(t)*y(t)-y(t)],[x(t),y(t)],
t=0..50,[inits],
> x=0..2,y=0..2,scene=[x(t),y(t)],scaling=CONSTRAINED,
color=GRAY,
> linecolor=BLACK,thickness=1,stepsize=.1);
In this model, a stable interior rest state is not possible.
510 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
The complexity of the behavior of solutions to the system increases
based on the assumptions made. Typical assumptions include adding
satiation terms for the predator (y) and/or limiting the growth of the
prey (x). The standard predator prey equations of Kolmogorov type,
x

= x
_
1
1
K
x
_

mxy
a +x
y

= y
_
mx
a +x
s
_
,
(6.40)
incorporate both of these assumptions.
We use solve to nd the three rest points of system (6.40). Let E
0
=
(0, 0) and E
1
= (k, 0) denote the two boundary rest points, and let E
2
represent the interior rest point.
> with(linalg):
> with(DEtools):
> rps:=solve(alpha*x*(1-1/k*x)-m*x*y/(a+x)=0,
y*(m*x/(a+x)-s)=0,x,y);
rps :=
_
x = 0, y = 0
_
,
_
y = 0, x = k
_
,
_
y =
a (mk +ks +sa)
k
_
m
2
2 ms +s
2
_ , x =
sa
m+s
_
The Jacobian, J, is calculated next in s1 with jacobian.
> s1:=jacobian([alpha*x*(1-1/k*x)-m*x*y/(a+x),
y*(m*x/(a+x)-s)],[x,y]);

1 2 x
y
1/10+x
+
xy
(1/10+x)
2

x
1/10+x
y
_
_
1/10 +x
_
1

x
(1/10+x)
2
_
x
1/10+x
s

Because J(E
0
) has a positive eigenvalue, E
0
is unstable.
> e0:=subs(rps[1],eval(s1));
_
1 0
0 s
_
> eigenvalues(e0);
, s
The stability of E
1
is determined by the sign of ms am/(a +k).
> e1:=subs(rps[2],eval(s1));
6.4 Systems of Equations 511

1
10
11
0
10
11
s

> eigs1:=eigenvalues(e1);
eigs1 := ,
mk +ks +sa
a +k
The eigenvalues of J(E
2
) are quite complicated.
> e2:=subs(rps[3],eval(s1)):
> eigenvalues(e2);
1/2
s
2
a +s
2
k s mk + sam+

s
4

2
a
2
+2 s
4

2
ak +2 s
3

2
a
2
m+s
4

2
k
2
2 s
3

2
k
2
m+s
2

2
m
2
k
2
2 s
2

2
m
2
ka +
2
s
2
a
2
m
2
+4 mks
4
a +4 mk
2
s
4
12 m
2
k
2
s
3

8 m
2
ks
3
a +12 m
3
k
2
s
2
+4 m
3
ks
2
a 4 m
4
k
2
s
mks m
2
k
,
1/2
s
2
a +s
2
k s mk + sam

s
4

2
a
2
+2 s
4

2
ak +2 s
3

2
a
2
m+s
4

2
k
2
2 s
3

2
k
2
m+s
2

2
m
2
k
2
2 s
2

2
m
2
ka +
2
s
2
a
2
m
2
+4 mks
4
a +4 mk
2
s
4
12 m
2
k
2
s
3

8 m
2
ks
3
a +12 m
3
k
2
s
2
+4 m
3
ks
2
a 4 m
4
k
2
s
mks m
2
k
Instead of using the eigenvalues to classify E
2
, we compute the char-
acteristic polynomial of J(E
2
), p() = c
2

2
+ c
1
+ c
0
, and examine the
coefcients. Notice that c
2
is always positive.
> cpe2:=charpoly(e2,lambda);
cpe2 :=
s
3
k+s
3
a+s
2
a+s
2
k2s
2
mks
2
ams
2
kmsmk+sm
2
k+sma+
2
km
2
m(m+s)k
> c0:=simplify(subs(lambda=0,eval(cpe2)));
c0 :=
(mk +ks +sa) s
mk
> c1:=simplify(coeff(cpe2,lambda));
c1 :=
s (sa +ks mk +am)
m(m+s) k
> c2:=simplify(coeff(cpe2,lambda2));
c2 := 1
On the other hand, c
0
and ms am/(a+k) have the same sign because
> simplify(c0/eigs1);
512 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
(mk +ks +sa) s
mk
is always positive. In particular, if ms am/(a +k) < 0, E
1
is stable.
Because c
0
is negative, by Descartes rule of signs, it follows that p()
will have one positive root and hence E
2
will be unstable.
On the other hand, if m s am/(a + k) > 0 so that E
1
is unstable,
E
2
may be either stable or unstable. To illustrate these two possibilities
let = K = m = 1 and a = 1/10. We recalculate.
> alpha:=1:k:=1:m:=1:a:=1/10:
> rps:=solve(alpha*x*(1-1/k*x)-m*x*y/(a+x)=0,
y*(m*x/(a+x)-s)=0,x,y);
rps :=

x = 0, y = 0

,

x = 1, y = 0

x = 1/10
s
1 +s
, y =
1
100
10 +11 s
(1 +s)
2

> s1:=jacobian([alpha*x*(1-1/k*x)-m*x*y/(a+x),
y*(m*x/(a+x)-s)],[x,y]);

1 2 x
y
1/10+x
+
xy
(1/10+x)
2

x
1/10+x
y

1/10 +x

x
(1/10+x)
2

x
1/10+x
s

> e2:=subs(rps[3],eval(s1)):
> cpe2:=charpoly(e2,lambda);
cpe2 := 1/10
10
2
10
2
s 9 s +11 s
2
+10 s 21 s
2
+11 s
3
1 +s
> c0:=simplify(subs(lambda=0,cpe2));
c0 := 1/10 s (10 +11 s)
> c1:=simplify(coeff(cpe2,lambda));
c1 := 1/10
s (9 +11 s)
1 +s
> c2:=simplify(coeff(cpe2,lambda2));
c2 := 1
Using solve, we see that
1. c
0
, c
1
, and c
2
are positive if 9/11 < s < 10/11, and
2. c
0
and c
2
are positive and c
1
is negative if 0 < s < 9/11.
6.4 Systems of Equations 513
> solve(c0>0 and c1>0,s);

9
11
< s, s <
10
11

> solve(c0>0 and c1 <0,s);

0 < s, s <
9
11

In the rst situation, E


2
is stable; in the second, E
2
is unstable.
Using s = 19/22, we graph the direction eld associated with the
system as well as various solutions in Figure 6-39. In the plot, notice
that all nontrivial solutions approach E
2
(.63, .27); E
2
is stable a
situation that cannot occur with the standard predatorprey equations.
> subs(s=19/22,rps[3]);

x =
19
30
, y =
121
450

> ivals:=seq(i/14,i=0..14):
> i1:=seq([x(0)=1,y(0)=i],i=ivals):
> i2:=seq([x(0)=i,y(0)=1],i=ivals):
> DEplot([diff(x(t),t)=alpha*x(t)*(1-1/k*x(t))
-m*x(t)*y(t)/(a+x(t)),
> diff(y(t),t)=y(t)*(m*x(t)/(a+x(t))-19/22)],
> [x(t),y(t)],t=0..25,[i1,i2],x=0..1,y=0..1,
scene=[x(t),y(t)],
> scaling=CONSTRAINED,color=GRAY,linecolor=BLACK,
> thickness=1,stepsize=.075);
On the other hand, using s = 8/11 (so that E
2
is unstable) in Figure 6-40
we see that all nontrivial solutions appear to approach a limit cycle.
> DEplot([diff(x(t),t)=alpha*x(t)*(1-1/k*x(t))
-m*x(t)*y(t)/(a+x(t)),
> diff(y(t),t)=y(t)*(m*x(t)/(a+x(t))-8/11)],
> [x(t),y(t)],t=0..50,[i1,i2],x=0..1,y=0..1,
scene=[x(t),y(t)], > scaling=CONSTRAINED,color=GRAY,
linecolor=BLACK,thickness=1,stepsize=.075);
The limit cycle is shown more clearly in Figure 6-41.
> DEplot([diff(x(t),t)=alpha*x(t)*(1-1/k*x(t))
-m*x(t)*y(t)/(a+x(t)),
> diff(y(t),t)=y(t)*(m*x(t)/(a+x(t))-8/11)],
> [x(t),y(t)],t=0..50,[[x(0)=.759,y(0)=.262]],x=0..1,y=0..1,
scene=[x(t),y(t)],
514 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
> scaling=CONSTRAINED,color=GRAY,linecolor=BLACK,
thickness=1,
> arrows=NONE,stepsize=.075);
y(t)
1
0.8
0.6
0.4
0.2
0
x(t)
1 0.8 0.6 0.4 0.2 0
Figure 6-39 s = 19/22
y(t)
1
0.8
0.6
0.4
0.2
0
x(t)
1 0.8 0.6 0.4 0.2 0
Figure 6-40 s = 8/11
6.4 Systems of Equations 515
y(t)
1
0.8
0.6
0.4
0.2
0
x(t)
1 0.8 0.6 0.4 0.2 0
Figure 6-41 A better view of the limit cycle without the direction eld
Also see Example 6.4.4.
EXAMPLE 6.4.7 (Van-der-Pols equation): In Example 6.4.4 we saw
that Van-der-Pols equation x

x
2
1

+ x = 0 is equivalent
to the system

= y
y

1 x
2

y x
.
Classify the equilibrium points, use dsolve with the numeric option,
toapproximate the solutions tothis nonlinear system, andplot the phase
plane.
SOLUTION: We nd the equilibrium points by solving

y = 0

1 x
2

y x = 0
. From the rst equation, we see that y = 0. Then,
substitution of y = 0 into the second equation yields x = 0. There-
fore, the only equilibrium point is (0, 0). The Jacobian matrix for this
system is
J(x, y) =

0 1
1 2xy

x
2
1

.
516 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
The eigenvalues of J(0, 0) are
1,2
=
1
2
_

_

2
4
_
.
> with(DEtools):
> with(linalg):
> f:=(x,y)->y:
> g:=(x,y)->-x-mu*(x2-1)*y:
> jac:=jacobian([f(x,y),g(x,y)],[x,y]);
_
0 1
1 2 xy
_
x
2
1
_
_
> eigenvalues(subs([x=0,y=0],jac));
1/2 x
2
+1/2 +1/2
_

2
x
4
2
2
x
2
+
2
4 8 xy,
1/2 x
2
+1/2 1/2
_

2
x
4
2
2
x
2
+
2
4 8 xy
Alternatively, the sequence of commands
> lin_mat:=array([[0,1],[-1,mu]]):
> with(linalg):
> eigs:=eigenvals(lin_mat);
eigs := 1/2 +1/2
_

2
4, 1/2 1/2
_

2
4
gives us the same result.
Notice that if > 2, then both eigenvalues are positive and real.
Hence, we classify (0, 0) as an unstable node. On the other hand, if
0 < < 2, then the eigenvalues are a complex conjugate pair with a
positive real part. Hence, (0, 0) is an unstable spiral. (We omit the case
= 2 because the eigenvalues are repeated.)
> sys:=mu->[diff(x(t),t)=y(t),diff(y(t),t)
=mu*(1-x(t)2)*y(t)-x(t)];
sys :=
_
d
dt
x (t) = y (t) ,
d
dt
y (t) =
_
1 (x (t))
2
_
y (t) x (t)
_
We now show several curves in the phase plane that begin at various
points for various values of by using seq to generate a list of ordered
pairs that will correspond to the initial conditions in the initial-value
problem.
> inits1:=seq([x(0)=0.1*cos(2*Pi*i/4),
y(0)=0.1*sin(2*Pi/4)],i=0..4);
6.4 Systems of Equations 517
inits1 :=

[x (0) = 0.1, y (0) = 0.1], [x (0) = 0.0, y (0) = 0.1],
[x (0) = 0.1, y (0) = 0.1]

> inits2:=seq([x(0)=-5,y(0)=-5+10*i/9],i=0..9);
inits2 :=

x (0) = 5, y (0) =
35
9

, [x (0) = 5, y (0) = 5], [x (0) = 5, y (0) = 5],

x (0) = 5, y (0) =
35
9

x (0) = 5, y (0) =
25
9

, [x (0) = 5, y (0) = 5/3],


[x (0) = 5, y (0) = 5/9], [x (0) = 5, y (0) = 5/9], [x (0) = 5, y (0) = 5/3],

x (0) = 5, y (0) =
25
9

> inits3:=seq([x(0)=5,y(0)=-5+10*i/9],i=0..9):
> inits4:=seq([x(0)=-5+10*i/9,y(0)=-5],i=0..9):
> inits5:=seq([x(0)=-5+10*i/9,y(0)=5],i=0..9):
> initconds:=union(inits1,inits2,inits3,inits4,inits5):
> nops(initconds);
39
We then use phaseportrait in the same way as we use DEplot to
graph various solutions.
> phaseportrait(sys(1/2),[x(t),y(t)],t=0..20,initconds,
x=-5..5,y=-5..5,
> arrows=NONE,linecolor=BLACK,stepsize=0.05);
> phaseportrait(sys(1),[x(t),y(t)],t=0..20,initconds,
x=-5..5,y=-5..5,
> arrows=NONE,linecolor=BLACK,stepsize=0.05);
> phaseportrait(sys(3/2),[x(t),y(t)],t=0..20,initconds,
x=-5..5,y=-5..5,
> arrows=NONE,linecolor=BLACK,stepsize=0.05);
> phaseportrait(sys(3),[x(t),y(t)],t=0..20,initconds,
x=-5..5,y=-5..5,
> arrows=NONE,linecolor=BLACK,stepsize=0.05);
We show all four graphs together in Figure 6-42. In each gure, we
see that all of the curves approach a curve called a limit cycle. Physically,
the fact that the systemhas a limit cycle indicates that for all oscillations,
518 Chapter 6 Applications Related to Ordinary and Partial Differential Equations
y
4
2
0
-2
-4
x
4 2 0 -2 -4
y
4
2
0
-2
-4
x
4 2 0 -2 -4
y
4
2
0
-2
-4
x
4 2 0 -2 -4
y
4
2
0
-2
-4
x
4 2 0 -2 -4
Figure 6-42 Solutions to the Van-der-Pol equation for various values of
the motion eventually becomes periodic, which is represented by a
closed curve in the phase plane.

6.5 Some Partial Differential Equations


We now turn our attention to several partial differential equations. Several exam-
ples in this section will take advantage of commands contained in the PDEtools
package. Information regarding the functions contained in the PDEtools package
is obtained with ?PDEtools.

Vous aimerez peut-être aussi