Вы находитесь на странице: 1из 30

C H A R

Particle Dynamics in
One Dimension

2.1 INTRODUCTION

Suppose a particle of mass m is acted on by several forces F 1; F2, . . ., Fn. The net force F act-
ing on the particle is given by the superposition principle as

(2.1)

and the motion of the particle is described by Newton's second law as

F = (2.2a)
dt
where p is the linear momentum of the particle. Only when the mass m remains constant may
we write
d\
F = m y = ma (2.2b)
dt
If we describe the motion in rectangular coordinates, Eq. (2.2b) may be written in the form of
three components as

= mx = max (2.3)

with similar expressions for Fy and Fz. If the acceleration a or its components ax, ay, az are
known, then Eq. (2.2b) may be used to solve for the force F. In general, the situation in particle
dynamics is just the reverse; that is, we know the net force F acting on a particle and we want
28
Sec. 2.1 Introduction 29

:o solve Eq. (2.2b) to find the position of the particle as a function of time t. In this chapter, since
ve are limiting our motion of the particle to one dimension, the only equation of interest is
Eq. (2.3), which after dropping the subscript x may be written as

dh
(2.4)

To be more explicit, we may write this equation as

d2A
F(x, = m dt2 (2.5)

vhere x = dxldt = v is the velocity of the particle and Eq. (2.5) states that the force acting on
:he particle is a function of position, velocity, and time. Such a problem in which the applied
force is a function of all three variables simultaneously is difficult to solve. On the other hand,
if the applied force is a function of only one variable, the problem is much simplified. Hence we
ihall divide our discussion into the following four cases:

1. The applied force is constant; that is, F = constant, such as freely falling bodies and every-
day motion.
2. The applied force is time dependent; that is, F = (Ft), such as in the case of electromag-
netic waves.
3. The applied force is velocity dependent; that is, F = F(v), such as air resistance to falling
or rising objects.
4. The applied force is position dependent; that is, F = F(x), such as restoring force to vi-
brating springs.

Before we start solving Eq. (2.4) for these different cases, we may remind ourselves that since

d x dv dv dx dv
(2.6)
dt1
dt dx dt dx
Eq. (2.4) may be written in the following different forms:

d2x
F = m (2.7a)
dt2
dv
F = m (2.7b)
dt
dv
F = mv (2.7c)
dx
Also, since momentum p is defined as p = mv = m{dxldt), we may write Eq. (2.7a) as [or di-
rectly from Eq. (2.2a) as applied to the one-dimensional case]

dp
F = (2.8)
dt
30 Particle Dynamics in One Dimension Chap. 2

That is, the applied force is equal to the rate of change of momentum. If the applied force acts
between the time interval tx and t2, then, by integrating Eq. (2.8), we get

Pi - Pi = Fdt (2.9)

which is the integral form of Newton's second law, while Eqs. (2.7) are the differential forms.
The integral on the right side of Eq. (2.9) is the impulse delivered by a force F during a short
time interval (t2 t{); that is, the change in the linear momentum is equal to the impulse deliv-
ered. Thus Eq. (2.9) is a statement of the impulse-momentum theorem.

2.2 CONSTANT APPLIED FORCE: F= CONSTANT

We are interested in studying the motion of a particle when the applied force acting on the par-
ticle is constant in time. Since F is constant, so will be the acceleration a, and we may write
Newton's second law as

dv F
2 = - - = a = constant (2.10)
dt dt m
The equation may be solved by direct integration provided we know the initial conditions. Solv-
ing Eq. (2.10) gives us the familiar results obtained in elementary mechanics, as we will show
now. Let us assume that at t = 0, the initial velocity is v0, and at time / the velocity is v. Thus,
from Eq. (2.10),

dv = dt

which on integration yields

v = vQ + at (2.11)

Substituting v = dx/dt in Eq. (2.11) and again assuming the initial condition that x = x0 at t
0, we get by direct integration

x = vQt w (2.12)

By eliminating t between Eqs. (2.11) and (2.12), we get

v2 = VQ + 2a(x - x0) (2.13)

Equations (2.11), (2.12), and (2.13) are the familiar equations that describe the translational mo-
tion of a particle in one dimension.
One of the most familiar examples of motion with constant force, hence constant acceler-
ation, is the motion of freely falling bodies. In this case, a is replaced by g, the acceleration due
to gravity, having the value g = 9.8 m/s2 = 32.2 ft/s2. The magnitude of the force of gravity act-
ing downward is mg.
Sec. 2.3 Time-Dependent Force: F = F(t) 31

2.3 TIME-DEPENDENT FORCE: F= F(fl

In this case, the force being given by F = F(t) implies that it is an explicit function of time;
hence Newton's second law may be written as

(2.14)

which on integration gives, assuming that v = v0 at t = t0,

1 f'
v = v0 + F(t) dt (2.15)

Since v - v(t) = dx{t)ldt, Eq. (2.15) takes the form

dt
dt m
or, integrating again,

x = x0 + vo(t ~ t0) + F{t)dt\dt (2.16)

Since there are two integrations, we may use two variables t' and t" and write Eq. (2.16) as

x = xn vo(t -to)+ - f dt' \ F(f) df (2.17)

We will illustrate this discussion by applying it to the interaction of radio waves with elec-
trons in the ionosphere, resulting in the reflection of radio waves from the ionosphere. The
ionosphere is a region that surrounds Earth at a height of approximately 200 km (about
125 miles) from the surface of Earth. The ionosphere consists of positively charged ions and
negatively charged electrons forming a neutral gas. When a radio wave, which is an electro-
magnetic wave, passes through the ionosphere, it interacts with the charged particles and ac-
celerates them. We are interested in the motion of an electron of mass m and charge e initially
at rest when it interacts with the incoming electromagnetic wave of electric field intensity E,
given by

E = Eo sin(a)t (2.18)

where w is the oscillation frequency in radians per second of the incident electromagnetic wave
and (f) is the initial phase. The interaction results in a force F on the electron given by

F = eE= eE0 (2.19)


32 Particle Dynamics in One Dimension Chap. 2

while the acceleration of the electron is given by

F eEn
a = = + (2.20)
m m
Let a0 = eEJm be the maximum acceleration so that Eq. (2.20) becomes

a = a0 sin(tot + <j>) (2.21)

Since a = dvldt, the equation of motion of the electron may be written as

dv eE(l
= sm(cot + <p) (2.22)
dt m
Assuming initially the electron to be at rest, that is, t = t0 = 0, u0 = 0, the integration of
Eq. (2.22) yields

e
eEn En
v = cos (j> + cos(cot + <p) (2.23)
mco mco

Since i; = dxldt, and assuming that* = xQ at t0 = 0, the integration of Eq. (2.23) yields

eE eEn
x = sin (/> c o s </> }t + sin(tot + 4") (2.24)
mco moj mco
The first two terms indicate that the electron is drifting with a uniform velocity and this veloc-
ity is a function of the initial conditions only. Superimposed on this drifting motion of the elec-
tron is an oscillating motion represented by the last term. The oscillating frequency to of the
electron is independent of the initial conditions and is the same as the frequency of the incident
electromagnetic waves. In the following, we want to investigate how such coherent oscillations
offree electrons can modify the propagation characteristics of incident electromagnetic waves.
Comparing Eq. (2.19) for F with Eq. (2.24) for x, it becomes quite clear that the oscillat-
ing part of the displacement x is 180 out of phase with the applied force that results from the
electric field of incident electromagnetic waves. Ordinarily, in a dielectric at low frequencies,
the charges are displaced in the direction of the applied force, resulting in the polarization of the
charges in phase with the applied force. In such situations, the resulting dielectric coefficient of
the material is greater than 1. In the case of the ionosphere, it can be shown that the resulting
polarization is 180 out of phase with the electric field; hence the dielectric coefficient of the
ionosphere is less than 1. This result has two consequences.

1. The phase velocity v of electromagnetic waves in the ionosphere is greater than the speed
of light c.
2. The refractive index of the ionosphere for incoming electromagnetic waves is less than the
refractive index of the free space from where the waves are coming (the incident medium,
which is a vacuum in this case).
Sec. 2.3 Time-Dependent Force: F= F(t) 33

Ionosphere

Empty space
Figure 2.1 Reflection of radiowaves
from the ionosphere. The total internal
reflection of electromagnetic waves
from the ionosphere.

This leads to the possibility of total internal reflection, that is, the reflection of incident electro-
magnetic waves from the ionosphere back to Earth, as illustrated in Fig. 2.1.

y Example 2.1
A block of mass m is initially at rest on a frictionless surface at the origin. At time
t = 0, a decreasing force given by F = F0exp(-A,t), where % = 0.5 is positive and less than
1, is applied. Calculate x(t) and v(t). Graph x, v, and F versus t.

Solution
From Newton's second law, after
dt
rearranging and integrating, we get the
velocity vl at time tl to be vl rtl
F0
1 dv= V^dt
0

1)
(km)

xl tl
Once again rearranging and ldx=
integrating, we get the displacement xl
to be

X -m
34 Particle Dynamics in One Dimension Chap. 2

N:=100 i:=0..N t. :=i


Now we can write expressions for
Fj, vj, and xj, keeping in mind that
at t = 0, F = Fo. These calculations are m:=2 Fo:=l X\=.5
made for N = 100 values even though
only 15 values are shown in the graph. -X-t.
Fo ,, -x-t.
R : = Fo-(e v. : = -\1 - e
The values of F, x, and v at four m-X
different times are given below.
Fo i) Fo-X,
V= 2
1 2
m-X m-X

= 0.607 x, =0.213 = 0.393

F 5 = 0.082 x 5 = 3.164 . =0.918

F 25 =3.727-10 x 2 5 =23

F 50 = 1.389-10 X
50=48

(a) Look at the variation in F, v, and x


versus t and explain the conclusions (
you draw from such variations. 4 1
X.
1 I
-X-
(b) What does the leveling of the V.

values of F and v for high t mean? /


F.
J
(c) What changes in (b) will be _Ht1t t t t t
observed if X is 0.01, 1.0, and 5?
1 n o p n
(You can explain by regraphing for 10 15
different values of A,.) t.
Time
Motion due to a decreasing force

Exercise 2.1: A particle of mass m is at rest at the origin of the coordinate system. At t = 0, a force

F= - te-kt)
is applied to the particle. Find the acceleration, velocity, and position of the particle as a function of time.
Graph these values and answer (a), (b), and (c) in the example.
Sec. 2.4 Velocity-Dependent Force: F= F(v) 35

Z4 VELOCITY-DEPENDENT FORCE: F=F{v)

There are many situations of common everyday occurrence where, in addition to constant ap-
r ad forces, forces are present that are a function of velocity. For example, when a body is in a
rra\ itational field, in addition to the gravitational force, there exists a force of air resistance on
-j-.e falling or rising body, and this resisting force is some complicated function of velocity. The
-dme is true of objects moving through fluids (gases and liquids). Such opposing forces to the
-.otion of objects through fluids are called viscous forces or viscous resistance. In these cases,
Newton's second law may be written in the following form:

dv
F(v) = m (2.25)
dt
dv dx dv
F(v) = m = mv (2.26)
dx dt dx
Knowing the form of the force F(v), either of these two equations may be solved to analyze the
~otion, that is, to calculate x as a function of t. Starting with Eq. (2.25), we may write
si,,
dt = m
F{v)
- hich on integration yields

t = t(v) = m\ (2.27)
J F(v)
Solving this gives v as a function of t; that is v = v(t). Thus, knowing v(t), we can solve for x.
dx
V = ~ = V(t) (2.28)

:r dx - rit) dt

- hich on integration gives

x = x{t) = I v{t)dt (2.29)


Thus the problem is solved. Similarly, if we start with Eq. (2.26), we get

vdv
dx = m (2.30)
F(v)
- hich on integration yields
v dv
x = x(t) = m (2.31)

Equations (2.29) and (2.31), which describe the displacement x as a function of t, may appear
:o be quite different, but when evaluated, they yield the same relationships as can be demon-
-j-ated. We shall divide our discussion into two parts. First, we shall discuss those cases in which
36 Particle Dynamics in One Dimension Chap. 2

there is no externally applied force besides the viscous resistance opposing the motion of the
body. Later, we shall investigate more practical situations in which both types of forces, fric-
tional as well as applied, are present.

Special Case

Suppose an automobile is moving with velocity v0 on a smooth frictionless surface when its
engine is suddenly shut off. Let us assume that the air resistance is proportional to velocity;
that is,

Fr = Fr(v) = ~kv (2.32)

Assuming that at t = 0, v = % calculate v and x as functions of t. We may write the differen-


tial equation of motion as

dv
Fr(v) = kv = m (2.33)
dt
m r" dv
That is, dt =
k v
\
which on integration gives (J dv/v = In v)

m v
t= --in - (2.34)
k \v0
After rearranging
-(k/m)t
v = vne (2.35)

That is, the velocity decreases exponentially with time.


Substituting v = dxldt in Eq. (2.35) and rearranging, we get

= voe-(k/m)'dt
which on integration, setting the limits as t = 0 when x = 0, and allowing the displacement to
be v at time t, gives

mvn
x = [1 (2.36)

It is clear from Eqs. (2.35) and (2.36) that when t = 0, v = % x = 0, as it should be. We note
from Eq. (2.35) that v = 0 only when t = o , and then from Eq. (2.36), x = mvjk = xb where
A. is the limiting distance. The body never goes beyond this distance. (But it takes infinite time
to reach there! We shall discuss this shortly.)
Reconsidering Eqs. (2.35) and (2.36), we know that the motion cannot continue forever
and the automobile must come to rest long before the infinite time as calculated earlier. Let us
assume that when the automobile reaches a certain minimum velocity ve it has almost reached
the f.nal distance xe. This will be true as long as v is less than the certain minimum value ve.
Sec. 2.4 Velocity-Dependent Force: F= F(v) 37

By substituting v = v( in Eq. (2.35), we can calculate the time t( it takes to reach velocity ve;
that is,
_ p-{k/m)tc
u( - vOi (2.37)

or tt = - l n M (2.38)

Another interesting fact is revealed if we consider the motion in a short time interval when
the retarding or the resistive force just begins acting on the moving body. To discuss this, let us
expand the right sides of Eqs. (2.35) and (2.36) by using a Taylor series (ex = 1 + x + x2/2\ +
xV3! + --O-Thatis,

kvn 1
rO
v = v0 1 + = v0 + t +
m
= vo arOt (2.39)
where a^ = F^m is the acceleration at t = 0. Similarly,

1 jtwa( .2 1 Fo
X =
^ " 2 m 2 m
(2.40)

If we ignore the higher terms, Eqs. (2.39) and (2.40) reveal that they describe the motion
of a particle acted on by a constant force provided t is very small. Note that kv0 = Fr0 = mar0,
which is the force acting on the particle initially when t = 0; that is, these equations are simply
the equations of motion of a particle under a constant force.

General Case

The preceding situation was limited to a simple case in which the retarding force Fr(v) was pro-
portional to velocity. In actual situations, Fr is a complicated function of velocity and the solu-
tions cannot be obtained by simply using the integration tables. Instead it becomes necessary to
do numerical integrations. But, in many cases over a wide range of velocities, it is possible, in
practice, to use the following approximation in which the retarding force or frictional force is
proportional to some power of velocity. That is,

Fr = (+)kvn (2.41)

where k is the positive constant of proportionality for the strength of the retarding force and n
is a positive integer. If n is an odd integer, the negative sign in Eq. (2.41) must be used. If n is
an even integer, if will be positive, and the sign + or in Eq. (2.41) is chosen in such a way
that it gives the direction of Fr to be opposite to that of v (Fr in the direction of v will be adding
energy to the system instead of retarding it!). For small objects moving in air with velocities less
than 25 m/s, it is found that, if we take n 1, we get good agreement with experimental results,
38 Particle Dynamics in One Dimension Chap. 2

while for velocities greater than 25 m/s but less than 32 m/s, the use of n = 2 gives good
agreement with experimental values.
Let us apply these ideas to the case of a freely falling body, that is, to the vertical motion
of an object in a resisting medium, the medium being air in this case. Let us assume that the air
resistance is proportional to v, which can be written as kv, independent of the sign of v. Thus
the net force acting on the body is
F = Fg + Fr = mg kv (2.42)
and the differential equation describing the motion of the freely falling body is
dv
m = mg kv (2.43)
at
Taking v = v0 at t = 0, we may write
m dv m
dt= - r r kv)
/ Vo mg + kv k
m mg + kv
or t = -In (2.44)
k mg + kv0
Solving this equation for v, we get
mg mi ,-(k/m)t
(2.45)
vo\e
k
Note that if the initial velocity is zero, that is, if at t = 0, vQ = 0, we get

_ e-(k/m)t. (2.46)

In Eq. (2.45), we may substitute v = v(t) = dxldt and the limits x = x0 when t = 0 and x = x
when t t. After integrating, we get the following result:

mi nfi mv, [1 _ .-(k/m)t


(2.47)
k k2
Terminal Velocity

Let us consider Eq. (2.45) once again. As t increases, the exponential term decreases and drops
to zero when t is very large as compared to mlk; that is, for t > mlk,
e-(k/m)t _ e-[
= 0
And for such large values of t, the velocity v reaches a limiting value, which from Eq. (2.45) is
vt= (mg/k). This limiting velocity v, is called the terminal velocity of a falling body. The ter-
minal velocity of a body is defined as the velocity of the body when the retarding force (the force
of air resistance) is equal to the weight of the body, and the net force acting on the body is zero
(hence there will be no acceleration of the body); that is,
Fr + Fg = 0 or kv, mg = 0
mg
or (2.48)
Sec. 2.4 Velocity-Dependent Force: F= F(v) 39

The magnitude of the terminal velocity, which is equal to mglk, is called the terminal speed. As
an example, the terminal speed of raindrops varies anywhere from 3 to 7 m/s. Different bodies
starting with different initial velocities will approach terminal velocities in different times.
There are three different possibilities for initial velocities: |vo| = 0, |v0 < vj, and v0 > v,|. Fig-
ure 2.2 illustrates the time taken by the bodies to reach terminal velocity for the three cases.

Characteristic Time

It should be clear from the previous discussion that the dimensions of mlk are time. We define
the quantity mlk as the characteristic time T; that is, from Eq. (2.48) (ignoring the sign),

m
T = (2.49)

Using the definitions of v, and T, we rewrite Eqs. (2.45) and (2.47) as

v = -v, + (vt + vQ)e~'/T (2.50)


,-t/r
x = x0 vtt + (gT2 + VOT)(1 (2.51)
Let us consider a special case in which the body starts from rest. Substituting v0 = 0 in
Eq. (2.50), we obtain

v= -vt(l - e~'/T) (2.52)


and, if t = T, Eq. (2.52) takes the form
v = -vt(l - \/e) = -0.63u (

That is, in one characteristic time the body reaches 0.63 of its terminal speed. This allows us to
define the characteristic time as that in which the body reaches 0.63 of its terminal speed. If
t = 2T, then v = -0.87ur, while for t = 10T, V = -0.99995u(.
Let us go back to Eqs. (2.45) and (2.47) again. Using a series expansion, we can show that
these equations reduce to the familiar equations of motion for the case of constant force. For
t < T(=m/k = v/g), we get

v = v0 - gt (2.53a)
and x = x0 + vot - \gt2 (2.53b)
That is, for small t, the effect of air resistance is negligible. On the other hand, for t > r (= mlk),
Eqs. (2.45) and (2.47) reduce to

v = -v, (2.54a)

m m
x = x0 VOT

= Xo+ VQT (r2g - grt) (2.54b)


40 Particle Dynamics in One Dimension Chap. 2

Figure 2.2

The graph illustrates the time it takes the bodies to reach terminal velocity for
different initial velocities for given values of m, g, and k. vt is the terminal velocity,
vOl, vO2, and vO3 are three different initial velocities
k:=5
n:=40 t :=0..n
vO = 0
v02 < vt (For absolute values) m:=2.5 g: = 9.8
v03 > vt
vt=-4.9
Note that the terminal velocity is
10
vt = -4.9 m/sec vol :=0 B
and, depending on the initial
" -(T
condition, the terminal velocity is
reached between 20 and 30 vo2: = -2 v2 .=_
seconds. (In the graphs, t is divided
by 10.) k\ t
vo3 :=-8 v3 :=.

vt=-4.9

vl 5 =-3.097 v25 =-3.833 v3 5 =-6.04

vll 11 00 =-4.237 v2,


2, 00 =-4.508 v3 1Q =-5.32
1Q
v2.

vl
20 = -4.81 v2
20 = -4 .847 v 3 2 0 =-4.957 v3

vl
30 = -4.888 v2
30 = -4 .893 v 3 30 =-4.908

vl
40 = -4.898 v2
40 = -4 .899 v 3. 0 =-4.901
-101
0.6 1.2 1.8 2.4
t

(a) What are the different factors that affect the To


time
terminal velocity? Time to reach terminal velocity

(b) If the initial velocity is positive and upward,


how will it affect the terminal velocity? How will
this graph differ from the others before terminal
velocity is reached? Graph this.
Sec. 2.4 Velocity-Dependent Force: F= F{v) 41

A Better Approximation

For small, compact, heavy bodies, a better approximation is that in which the retarding or vis-
cous force is proportional to v2. Thus the equation

takes the form

;,,,2
-r kv
dv
mg = m (2.55)
dt
where - kv2 is used for rising (or ascending) bodies, while +kv2 is used for falling bodies. The
terminal speed is given by
F+F=0 or kit = mg (2.56)

or (2.57)
and the characteristic time T is given by

(2.58)

Following the procedure outlined before, Eq. (2.55) can be solved for v and x. The results ob-
tained are
f m dv
t = \ ; , for rising objects (2.59)
J -mg - kxr
f m dv
t = ~c, for falling objects (2.60)
J -me + kxr
which give

t = rtan '( for rising objects (2.61)

, v\
t = -Ttanh [] + C2, for falling objects (2.62)

where C\ and C2 are constants to be determined from initial conditions. Solving these equations
for v, we get
C, - t
v - v. for rising objects (2.63)
T
t-C,
v = v, tanh for falling objects (2.64)

As compared to a retarding force kv, the terminal speed for the case of a retarding force
kv2 is reached much faster, as illustrated in Fig. 2.3. In this case, when t = 5T, V =
0.99991 vt; that is, it reaches the same speed in half the time.
42 Particle Dynamics in One Dimension Chap. 2

Figure 2.3

The graph shows the motion of a vertically falling object under a linear
force and a quadratic force with initial conditions to = 0, xo = 0 and vo = 0.
Linear retarding force: F=-kx Quadratic retarding force: F=-kv2 or kv2

The terminal velocity v2 = -0.808 for i N :=50 :=0..N _ I


quadratic retarding force ( kv2) is reached ' 50
much faster (in about 0.2 seconds) than m :=.1 ;:=9.8 k:=1.5
the terminal velocity vl = 0.653 for a
linear retarding force (-kv) is reached in m
x2 := - m _
about 0.3 seconds. To calculate vl and k
v2 use Eqs. (2.46) and (2.64).
Tl =0.067 x2 = 0.082

(a) In a given time interval, which object vl :=BJ v 2 :=


will travel a larger distance and why?
vl =0.653 v 2 = 0.808
(b) For the vertically falling objects
which of the two situations is more v2. :=-v2-tanh
desirable and why? vl. :=_(vl)-l l - e

(c) What do the values of vl and v2 at


time t = 20 and 30 indicate?
1

i a a [ l a a a -B( 1B- -B-H


vl 5 =-0.508 =-0.677 U.It)
' X X X- fX- X X -X>

v l | Q =-0.621 v 2 m =-0.796 g 0.5


>
v l | 5 =-0.646 v2 1 5 =-0.807 '
0.25
vl =-0.652 v2, n =-0.808
n:
vl 3 Q =-0.653 v2 3 Q =-0.808 0.1 0.2 0.3 0.5
t.
l
t3()=0.6 t15=0.3 time
Speed for quadratic and linear force

Note that the terminal velocity v2 is reached much faster (in about 0.3 second) than the
terminal velocity vl (in about 0.6 second). Explain why.
Sec. 2.4 Velocity-Dependent Force: F= F(v) 43

Example 2.2

A ball of mass m is thrown with velocity vo on a horizontal surface where the


retarding force is proportional to the square root of the instantaneous velocity. Calculate
the velocity and the position of the ball as a function of time and graph the results.

Solution
The retarding force is given by dv 2
Rearrange this equation and integrate F=m-=-kv
dt
assuming that initial velocity at t = 0 is
vl
vo and vl at time tl.
-dv= -Idt
m
simplifies to
Solving, we get the value of vl
at time tl as vO 2-A/vl-2-A/v0=-k-
m
k tl
Integrating vl we find the (i)
2 m
displacement xl at time tl.
xl rti
Now we may graph x and v as function of l dx= -- dt

t using Eqs. (i) and (ii), rewritten as (iii) 2 m


and (iv)
xl - xO=tl-vO + -k 2 -
12 m 2

i :=0..20 t. :=i xo:=0 vo:=20 k := 1.1 m:=.5

k 2 k / N
v. :=vo t.-vo -( (t. (iii)
m 2 VV
' 4-m

x. : = xo+- vo-t.--'---ft.)
1
(iv)
2 m VV T2~^
44 Particle Dynamics in One Dimension Chap. 2

Displacement and velocity versus time Displacement and velocity versus time
200
i
1
1500

150

/ 1\
.1000

100
-K-
500 /
50

10 20
2.5 7.5 10
Time
Time

Explain the decrease in v and then the increase in v as a function of time t. How does it
affect the value of x? (Refer to the zoomed graph on the right).

EXERCISE 2.2 Repeat the example for a retarding force that is proportional to the
cube root of the instantaneous velocity.

2.5 POSITION-DEPENDENT FORCES: F= F(x), CONSERVATIVE


FORCES, AND POTENTIAL ENERGY

This is one of the most important cases considered so far. There are many situations in which
motion depends on the position of the object. Examples of position-dependent forces are grav-
itational force, Coulomb force, and elastic (tension and compression) forces. The differential
equation that describes the rectilinear motion of an object under the influence of a position-
dependent force is

d2x
m y = F(x) (2.65)
at

which may also be written in such a manner that v is a function of position; that is,

dv
mv = F(x) (2.66)
dx

or -{-my* )=F(x) (2.67)


Sec. 2.5 Position-Dependent Forces 45

Since the kinetic energy of the particle is K = \m^, we may write Eq. (2.67) as

dK
= F(x) (2.68)
dx
which on integration gives

K- Ko= F(x) dx (2.69a)

or mtfo= f F(x)dx (2.69b)

The right side is equal to the work done when the particle is displayed from position x0 to x.
It is convenient at this point to introduce potential energy or a potential energy function
or simply & potential function) V(x) such that

dV(x)
= F(x) (2.70)
dx
We define V(x) as the work done by the force when the particle is displaced from x to some ar-
bitrary chosen standard point xs; that is,

V(x) = F(x) dx = - F(x) dx (2.71)


'x 'x,

which is consistent with Eq. (2.70). Thus the work done is going from x0 to x is

= -j'dV(x)- j dV(x)
= +V(x0) - V(x) = -V(x) + V(x0) (2.72)

Combining Eqs. (2.69) and (2.72), we get

K + V(x) = Ko + V(x0) = constant = E (2.73)

or (2.74)

This equation states that if a particle is moving under the action of a position-dependent force,
then the sum of its kinetic energy and potential energy remains constant throughout its motion.
Such forces are called conservative forces. For nonconservative forces, K + V = constant, and
a potential energy function does not exist for such forces. An example of a nonconservative force
46 Particle Dynamics in One Dimension Chap. 2

is frictional force. [It may be pointed out that if V(x) is replaced by V(x) + constant, the pre-
ceding discussion still holds true. In other words, the sum of the kinetic and potential energy
will still remain constant and will be equal to E.] E is the total energy, and Eq. (2.74) states the
law of conservation of energy, which holds only if F = F(x). A description of the motion of a
particle may be obtained by solving the energy equation, Eq. (2.74); that is,

dx
A - [E - V(x)] (2.75)
dt Im

which on integration yields

-,-f dx
V(2/m)[ - V(xJ]
(2.76)

and gives t as a function of x. [We shall not discuss the significance of the negative sign, which
deals with time reversal.]
In considering the solution of Eq. (2.76), it is essential to note that only those values of x
are possible for which the quantity E V(x) is positive. Negative values lead to imaginary solu-
tions and hence are unacceptable. Also, the motion is limited to those values of x for which
E V(x) 3= 0; that is, the roots of this equation give the region or regions to which the motion
is confined. This is demonstrated in Fig. 2.4. The function \mx2 + V(x) is called the energy in-
tegral of the equation of motion m{dvldt) = F{x), and such an integral is called a constant of
motion. (This is the first integral of a second-order differential equation.)
Before we give specific examples of solving the equation of motion for x(t), we shall show
that much can be learned about the motion by simply plotting V(x) versus x. Figure 2.5 shows
the plot of a potential energy function for one-dimensional motion. As mentioned earlier, the
motion of the particle is confined to those regions for which E - V(x) > 0 or V(x) < E. Let us
keep Eq. (2.75) in mind and discuss different cases.

E - V(x) I
Allowed limits of motion of a particle

Figure 2.4 Allowed regions of motion for a particle in a position-dependent


force field.
Sec. 2.5 Position-Dependent Forces 47

V(x)

Figure 2.5 The solid curve corresponds to a potential function V(x), and EO,
E l , . . ., are different energies of a particle moving in such a potential.

If E = Eo, as shown in Fig. 2.5, then Eo - V(x) = 0 and, according to Eq. (2.75), x = 0;
that is, the particle stays at rest in equilibrium at x = x0. Let us consider the case in which the
particle energy is slightly greater than Eo, say E1. For x < xx and x > x[, v will be imaginary;
hence the particle cannot exist in these regions. Thus a particle with energy Ex is constrained to
move in the potential well (or valley) between xx and x[. A particle moving to the right is re-
flected back at x[; and when traveling to the left, it is reflected back at xx. The points xx and x\
are called the turning points and are obtained by solving Ex - V(x) = 0. The velocity of the par-
ticle at these points is zero. Between xx and x[, the velocity of the particle will change as V(x)
changes. We briefly explain the motion of a particle corresponding to different energies and
moving in a potential V(x), as shown in Fig. 2.5.

Eo: The particle is in stable equilibrium.


Ex: The particle moves between the turning points xx and x[.
E2: The particle moves between the turning points x2 and x'2 with changing velocity. While
moving between the turning points x"2 and x"2, the particle has constant velocity and hence
is in the region of neutral equilibrium. The particle can also exist in the region for x > x"2".
E3: When a particle with this energy is at x3, it is at a position of unstable equilibrium. It can
also move in the valley on the left with a motion similar to that of a particle with energy
E2. Once it starts moving to the right, it keeps on moving, first with increasing velocity to
x"2 and then with constant velocity up to x"2.
E4: A particle with this energy can move anywhere. When passing over the hills, it slows
down, while over the valleys, it speeds up, as it should.
48 Particle Dynamics in One Dimension Chap. 2

Continuing our discussion of position-dependent forces, we shall examine two special


cases of interest in the next sections:
1. Motion under a linear restoring force
2. Variation of g in a gravitational field

Example 2.3

A particle of mass m is subjected to a force F = a - 2 b x , where a and b are constants.


Find the potential energy V = V(x). Then graph F(x) and V(x). Discuss the motion
of the particle for different values of energy.

Solution
Substitute for F in the expression for V and integrate to get the value of V.

The motion is limited to


the region x = 0 and x = a/b p_ a _ 2.t,-x V=- I F dx
For E < 0, the motion is limited to
the left of V(x) and cannot cross
the barrier. V=- a-2-b-xdx simplifies to V=-a-x-i-b-x2 (i)

For V(x)=-a-x+b-xzss0 x=-


b : = 0.. 15 a:=20 b:=2

For V(x)=-a+2-b-x=0 x=
dx 2-b V. :=- F. :=a- 2-b-x: (ii)

At x= Vmin=
2- b 4-b
200
F=0 at x=0 and
2-b > 150

Different values are calculated


ial Ene

100
below S l
B 50
o
2 ft. V.
-a-=5
'
1 = 10 --^=-50
2-b b 4b

vo = o V5 =-50 V
10= -100
5 10
FQ = 20 F5=0 F ]0 =-20 x.
l
Distance x
(min(V)) = -50 max(V) =150

max(F) = 20 min(F) =-40


Sec. 2.6 Motion under a Linear Restoring Force 49

(a) Looking at the variation in the values of V and F versus x, what do you
conclude from this variation?
(b) What are the values of F and V where these graphs intersect? What is the
significance of this? Explain.

(c) What is the significance of x = 5 where F = 0 and V is minimum? Explain.

EXERCISE 2.3 Repeat the example for F=a-2bx2

2.6 MOTION UNDER A LINEAR RESTORING FORCE

Let the motion of a particle subject to a linear force be given by

F(x) = -kx (2.77)

This equation is a statement of Hooke's law. A typical example of such a motion is that of a mass
fastened to a spring. The resulting motion is simple harmonic, as we shall discuss in detail in
Chapter 3. For the time being, we shall use the energy method discussed in the previous section
to obtain the solution. Taking the standard point to be at the origin (also the equilibrium point),
that is xs = 0, we may write the potential energy to be

V(x) = - [ F{x)dx= - f (-kx) dx


or V(x) = {kx2 (2.78)

Once again, the total energy is a constant of motion and may be arrived at in the same manner;
that is,

dv
mv = F(x) = -kx (2.79)
dx
mv dv = kx dx

which on integration gives

\mv2 = - \kx2 + constant

or \mv2 + {kx2 = E = total energy (2.80)


50 Particle Dynamics in One Dimension Chap. 2

We can now use this equation or Eq. (2.76) with V(x) given by Eq. (2.78) to solve for the dis-
placement x. For the conditions t0 = 0 at x = x0, Eq. (2.80) or (2.76) takes the form (keeping
only the positive sign)

dx dx
J Vf2 m)(E - \kx ) 2
-I (2.81)

Substituting

k k
; x = sin 0 and A| dx = cos 0 d0 (2.82)

we get

t =
m
(2.83)

As usual, let the angular velocity be denned as

Therefore, t=\ie- 0O)


or 8= cot + (2.84)
Combining Eq. (2.84) and (2.82), we get

sin ' l - v / x | =

or x = A sin(f + 0O) (2.85)


where A is the amplitude given by

(2.86)

Thus Eq. (2.85) states that the motion of the particle is simple harmonic with the coordinate x
oscillating harmonically in time with amplitude A and frequency w. A and 0 can be determined
from the initial conditions; that is, if E and x0 are given, from Eqs. (2.86) and (2.85)

(2.87)
and x0 = A sin 0O (2.88)
Sec. 2.7 Variations of g in a Gravitational Field 51

2.7 VARIATION OF g IN A GRAVITATIONAL FIELD

For small heights just above the surface of Earth, the value of g is almost constant and is equal
io 9.8 m/s2 = 32.2 ft/s2. But at large distances above the surface of Earth, the value of g varies
with distance and may be calculated in a simple manner. According to Newton's law of gravita-
tion, the force between an object of mass m at a distance x from the center of Earth of mass Mis

Mm
F(x) = -G (2.89)
If we neglect air resistance, the differential equation of motion of an object in a gravitional field
may be written as

dv Mm
~-
dx
Since v = x, we may write

dx
m I xdx = - GMm i
x~2
vhich on integration gives

.,L Mm
- G = constant = E (2.90a)
2
?r \ mx + V(x) = E (2.90b)
.vhere we have defined the gravitational potential energy V(x) to be

Mm
V(x) = ~G (2.91)

We can derive the same expression for the gravitational potential energy by the direct defini-
::on of the potential function. But it is convenient to take the initial conditions to be xs = o, in-
>tead of xs = 0, as we did previously. Thus

V(x) = GMm GMm


= -J F{x)dx=-\ -
. defined previously.]
We may rewrite Eq. (2.90a) as

(2.92)

/-here the positive sign corresponds to ascending motion, while the negative sign corresponds
: .> descending motion. Equation (2.92) may be solved for x(t) by integrating; that is,

dx
[ + (GMm/x)] 1/2 (2.93)
ro
52 Particle Dynamics in One Dimension Chap. 2

V(x)i

E
?

V(x)=-G-^

Figure 2.6 Particle of energy E in a


gravitational potential V(x) versus x.

The integration of this equation is not as simple as in the case of a linear restoring force. We
shall not pursue this any further at the present time, but shall give a graphical interpretation and
discuss some simple cases.
The plot of V(x) versus x is shown in Fig. 2.6. It is quite clear that for negative values of
E the motion is bound with a turning point at xT. That is, when E Ex the body will go a height
x = xT, come to a stop, and turn around. For values of E greater than zero, there is no turning
point and the body will never return to Earth. Thus, for a minimum energy E = 0, the velocity
corresponds to v = ve, the escape velocity, which we shall calculate below. We can calculate the
turning point by substituting velocity v = x = 0 and x = xT in Eq. (2.92) (by definition v is zero
at the turning point), and we get

Mm Mm
G = E or xT = G - (2.94)
x t
Since E is negative, xT will be positive.
Let us consider Eq. (2.90a) once again:
1 ., Mm
-mx G = constant (2.95)
2 x
Let a body be dropped from a height x = x0 with zero initial velocity. That is, substituting v =
x = 0 at x = x0 in Eq. (2.95),
Mm
G = constant

Using this value for the constant in Eq. (2.95) and rearranging, we get

x2 = 2GM(~ - - (2.96)
\x x0
Sec. 2.7 Variations of g in a Gravitational Field
53

Let g0 be the value of the acceleration due to gravity on the surface of Earth, where x =
?o that

, Mm
or ~ mg0
is.

go = ~ or G = go~ (2.97)

. Y-ially the distance is measured from the surface of Earth; hence we may write

x = R + r, x = r and x0 = R + h

-.ere h = x0 R equals the height (as measured from the surface of Earth) from which the
>:dy is dropped. Using this notation and Eqs. (2.96) and (2.97), we may write

1 1
(2.98)
\R + r R + h)
hus, for r = 0, that is, when the body reaches the surface, v = v0; hence

R R+h (2.99)
_- d may be written as

R (2.100)

-jch reduces to v0 2g$h ifh<R, as it should. Equation (2.98) also applies to the case when
?ody is projected upward with a velocity % and it will reach a height h when v = 0.
We can arrive at an expression for the escape velocity by substituting h = <* in Eq. (2.99),
r-ulting in

2GM
= 28oR =
R
2GM
- 11 km/s = 7 miles/s (2.101)
R
we could say that at the surface of the Earth E = 0; hence potential energy must be equal to
e kinetic energy. That is,

GMm

v. = x =
54 Particle Dynamics in One Dimension Chap. 2

PROBLEMS
2.1. Force F acting on a particle of mass m has the following dependencies:
(a) F(x, t) = f(x)g(t)
(b) F(x,t)=f(x)g(t)
(c) F{x,x)=f(x)g{x)
Write the differential equations describing these situations. Which of these differential equations
can be solved to describe the motion of the particle? Explain.
2.2. A particle of mass m is acted on by the force (a), (b), (c), (d), or (e) as given below. Solve these
equations to describe the motion of the particle.
(a) F(x, t) = k{x + t2), for t = 0, x = x0, and v = v0 = 0
(b) F(JC, t) = kx2x, for t = 0, x = x0, and v = D0 = 0
(c) F(x, t) = k(ax + t), for t = 0, v = v0
(d) F{x, x) = axVx
(e) Fix, x,t) = k(x + xt)
2.3. A block of mass m is initially at rest on a frictionless surface. At time t = 0, an increasing force
given by by F = kt2 is applied to the block. Find the velocity and the displacement of the block as
a function of time and graph x and v versus t.
2.4. A block of mass m is initially at rest on a frictionless surface at the origin. At time t = 0, a force
given by F = Fote Ar is applied. Calculate x(f) and vit) and graph them. What are these values when
(a) t is very small, and (b) t is very large?
2.5. A particle of mass m is at rest at t = 0 when it is subjected to a force F = Fo sin (otf + <). (a) Cal-
culate the values of :c(f) and vit) (b) Make plots of x(0 and vit) versus t. What are the maximum
and minimum values of x and vl
2.6. A particle of mass m is subjected to a force given by

F = F 0 <r A 'sin(crf + (f>)

Calculate the values of v(t) and x(t) and graph them. What is the magnitude of the terminal veloc-
ity in this case?
2.7. A particle of mass m is at rest at t = 0 when it is subjected to a force F = Fo cos 2 cot.
(a) Calculate the values of x(t) and vit)
(b) Make plots of jc(f) and vit) versus t.
(c) Describe the outstanding characteristic of these graphs.
2.8. A ball of mass m is thrown with velocity t>0 on a horizontal surface where the retarding force is pro-
portional to the square root of the instantaneous velocity. Calculate the velocity and the position of
the ball as a function of time. Discuss any limitations.
2.9. An object of mass m is thrown up an inclined plane of an angle 6 with an initial velocity v0. If the
motion is resisted by the retarding force Fr = kv, how far will the mass travel before coming to
rest? Assuming the same retarding force, how long will it take the object to travel back to the ini-
tial position?
2.10. Repeat Problem 2.9 for the retarding force Fr = kvL.
2.11. A boat is slowed down by a frictional force Fiv). Its velocity decreases according to the
relation D = kit - ts)2, where ts is the time it takes to stop the boat and k is the constant. Calcu-
late Fiv).
Problems 55

2.12. The motor of a speed boat is shut off when it has attained a speed of v0. Now the boat is slowed
down by a retarding force Fr = Ce ~kv. Calculate v(i) and x(i). How long will it take for the boat to
stop, and how much distance will it travel before stopping?
2.13. A particle of mass m starting with an initial velocity u0 is acted on by a force F = m(kv + cv2),
where k and c are constants. Calculate the displacement as a function of time.
2.14. For the situation in Problem 2.12 graph F and x versus t.
2.15. For the situation in Problem 2.13 graph F and x versus t. Compare the results with those in Prob-
lem 2.14.
2.16. A body of mass m is dropped from a height h. Calculate the speed when it hits the ground if (a) there
is no air resistance, (b) air resistance is proportional to the instantaneous velocity, and (c) air re-
sistance is proportional to the square of the instantaneous velocity. Graph velocity versus time in
each case and compare the results.
2.17. A projectile is thrown vertically with a velocity J;0. Calculate and compare times and maximum
heights reached when air resistance is (a) zero, (b) proportional to the instantaneous velocity, and
(c) proportional to the square of the instantaneous velocity. Graph distance versus time in all cases
and compare the results.
2.18. A ball is thrown vertially upward with an initial velocity v0. The air resistance is proportional to
the square of the velocity. Show that the velocity with which the ball returns to the original posi-
tion is

VnV.

where vt is the terminal velocity.


2.19. Derive Eq. (2.47); that is,

2.20. Using Eqs. (2.45) and (2.47), show that for t < T, we get the familiar equations of motion,
Eqs. (2.53a) and (2.53b).
2.21. Show that for the case t < T, Eqs. (2.45) and (2.47) reduce to Eqs. (2.54a) and (2.54b).
2.22. Starting with Eq. (2.55), derive Eqs. (2.61) and (2.62).
2.23. Starting with Eq. (2.60) and initial conditions that v0 = 0, x0 = 0 at t = 0, find the expressions for
x and v.
2.24. Using the law of conservation of energy, derive the general equations of motion for freely falling
bodies.
2.25. A particle of mass m is attracted toward the origin by a force that is inversely proportional to the
square of the distance; that is, F = Kir2. If this mass is released from a distance L, show that it
will take time t to reach the origin, where t is given by

ImL'V'2
i
56 Particle Dynamics in One Dimension Chap. 2

2.26. The velocity of a particle of mass m subjected to a certain force varies with the distance according
to the relation v = K/x", where A" is a constant. Assuming at t = 0, x = x0, calculate the force act-
ing on the particle as a function of (a) distance x, and (b) time t. (c) Calculate the position of the
particle as a function of time /. (d) Graph in (a), (b) and (c) and discuss any outstanding
features.
2.27. A particle of mass m is subjected to a force given by F = ax + bx2, where a and b are constants.
(a) Find the potential energy V(x).
(b) Make plots of F(x) and V(x).
(c) Discuss the motion of the particle for different values of energy and also point out the regions
in which the motion is forbidden.
2.28. A particle of mass m is subjected to a force represented by a potential function V(x) = ax2 + bx4,
where a and b constants.
(a) Calculate F(x).
(b) Make plots of F{x) and V(x).
(c) Discuss the motion of the particle for different values of energy. Also discuss the restrictions
on the motion.
2.29. A particle at time t = 0 is at rest at a distance x0from the origin and is subjected to a force that is
inversely proportional to the distance from the origin. Solve the equation of motion for this parti-
cle; that is, find v(i) and x(i).
2.30. A particle of mass m is subjected to a force

K
F(x) = - Cx + -r
x
where C and K are constants.
(a) Find the potential function V(JC).
(b) Make plots of F(x) and V(x).
(c) Discuss the nature of the motion for different values of E and also find the regions where the
motion is not possible.
2.31. The force between two particles in a diatomic molecule is such that it may be represented by a po-
tential function of the form

x' x
where Cx and C2 are the positive constants and x is the distance between the two atoms.
(a) FindF(x).
(a) Make plots of F(x) and V(x).
(c) Assume that one of the atoms in the molecules is very heavy and remains at rest at the origin.
Discuss the possible motions of the other atom in the molecule.
2.32. An alpha particle when inside the nucleus is bound by the potential shown for R < x < R, while
outside the nucleus the interaction between the alpha particle and the nucleus is represented by the
coulomb potential, as shown in Fig. P2.32.
(a) Write a potential function for the different regions shown in the figure.
(b) Calculate F(x) and make a plot of F(x) versus x.
(c) Discuss the motion of the alpha particle for different values of energy and different regions.
Suggestions for Further Reading 57

-Vn
Figure P2.32
2.33. Show that for h<R, Eq. (2.100) reduces to v0 = 2gji.
2.34. Starting with Eq. (2.90a) and substituting

/ E \1/2
GMm
show that x = xT cos2 6, where

2GM

and x = x0 = xT = turning point.

SUGGESTIONS FOR FURTHER READING


ARTHUR, W., and FENSTER, S. K., Mechanics, Chapter 3, New York: Holt, Rinehart and Winston, Inc., 1969.
BARGER, V., and OLSSON, M., Classical Mechanics, Chapter 1. New York: McGraw-Hill Book Co., 1973.
BECKER, R. A., Introduction to Theoretical Mechanics, Chapter 6, New York: McGraw-Hill Book Co., 1954.
DAVIS, A. DOUGLAS, Classical Mechanics, Chapter 2, New York: Academic Press, Inc., 1986.
FOWLES, G. R., Analytical Mechanics, Chapter 3, New York: Holt, Rinehart and Winston, Inc., 1962.
FRENCH, A. P., Newtonian Mechanics, Chapter 7, New York: W. W. Norton and Co., Inc., 1971.
HAUSER, W., Introduction to the Principles of Mechanics, Chapter 4. Reading, Mass.: Addison-Wesley
Publishing Co., 1965.
KLEPPNER, D., and KOLENKOW, R. J., An Introduction to Mechanics, Chapter 4. New York: McGraw-Hill
Book Co., 1973.
MARION, J. B., Classical Dynamics, 2nd ed., Chapter 2. New York: Academic Press, Inc., 1970.
ROSSBERG, K., Analytical Mechanics, Chapter 4. New York: John Wiley & Sons, Inc., 1983.
STEPHENSON, R. J., Mechanics and Properties of Matter, Chapter 2. New York: John Wiley & Sons, Inc., 1962.
SYMON, K. R., Mechanics, 3rd. ed., Chapter 2. Reading, Mass.: Addison-Wesley Publishing Co., 1971.
TAYLOR, E. F., Introductory Mechanics, Chapter 5. New York: John Wiley & Sons, Inc., 1963.

Вам также может понравиться