You are on page 1of 48

Chapter 3.

Particle Kinetics

34

Chapter 3

Particle Kinetics
Chapter Objectives

To state Newton's Laws of Motion and Gravitational Attraction


and to define mass and weight

To analyze the accelerated motion of a particle using the


equation of motion with different coordinate systems.

To develop the principle of work and energy and apply it to


solve problems that involve force, velocity. and displacement.

To study problems that involve power and efficiency.

To introduce the concept of a conservative force and apply the


theorem of conservation of energy to solve kinetic problems.

To develop the principle of linear impulse and momentum for a


particle

To study the conservation of linear momentum for particles

To introduce the concept of angular impulse and momentum

3.1 Kinetics of a Particle: Force and Acceleration


3.1.1 Basic concepts
The basic axioms of dynamics were give in Chapter1. Newtons first and third laws of motion
were used extensively in statics to study bodies at rest and the forces acting at them. These
two laws are also used in dynamics, in fact, they are sufficient for the study of the motion
which have no acceleration. However when bodies are accelerated, i.e., when the magnitude
or the direction of their velocity changes, it is necessary to use the second law of motion in
order to relate the motion of the body with the forces acting on it.
Measurements of force and acceleration can be recorded in a laboratory so that in accordance
with the second law, if a known unbalanced force F is applied to a particle, the acceleration a
of the particle may be measured. Since the force and acceleration are directly proportional, the
constant of proportionality, m, may be determined from the ratio to F/a. The positive scalar m
is called the mass of the particle. Being constant during any acceleration, m provides a
quantitative measure of the resistance of the particle to a change in its velocity.
Newton's second law of motion may be written in mathematical form as
F = ma

(3.1)

Chapter 3. Particle Kinetics

35

This equation, which is referred to as the equation of motion, is one of the most important
formulations in mechanics. As previously stated, its validity is based solely on experimental
evidence. In 1905, however, Albert Einstein developed the theory of relativity and placed
limitations on the use of Newton's second law for describing general particle motion. Through
experiments it was proven that time is not an absolute quantity as assumed by Newton; as a
result, the equation of motion fails to predict the exact behavior of a particle, especially when
the particle's speed approaches the speed of light (0.3 Gm/s). Developments of the theory of
quantum mechanics by Erwin Schrdinger and others indicate further that conclusions drawn
from using this equation are also invalid when particles are the size of an atom and move
close to one another. For the most part, however, these requirements regarding particle speed
and size are not encountered in engineering problems, so their effects will not be considered
in this course.
It is worthy to note that the system of axes with respect to which the acceleration a is
determined is not arbitrary. These axes must have a constant orientation with respect to the
stars, and their origin must either be attached to the sun or move with constant velocity with
respect to the sun. Such a system of axes is called a Newtonian frame of reference or inertial
system. In most engineering problems, the acceleration a may be determined with respect to
axes attached to the earth and Eq. (3.1) used without any appreciable error. On the other hand,
this equation does not hold if a represents a relative acceleration measured with respect to
moving axes, such as axes attached to an accelerated car or to a rotating piece of machinery.
We can consider a inertial system as the one which does not rotate and is either fixed or
translates in a given direction with a constant velocity (zero acceleration). This definition
ensures that the particle's acceleration measured by observers in two different inertial frames
of reference will always be the same.
Since m is constant, we can also write
F

d(mv)
dt

(3.2)

where mv, is the particles linear momentum. This equation expresses that the force acting on
the particle is equal to the rate of change of the linear momentum of the particle.
When more than one force acts on a particle, the resultant force is determined by a vector
summation of all the forces; i.e., FR

F.

For this more general case, the equation of

motion may be written as

F ma

(3.3)

Chapter 3. Particle Kinetics

36

To illustrate application of this equation, consider the particle P shown in Fig. 3.1a, which has
a mass m and is subjected to the action of two forces, F1 and F2. We can graphically account
for the magnitude and direction of each force acting on the particle by drawing the particle's
free-body diagram, Fig. 3.1b. Since the resultant of these forces produces the vector ma, its
magnitude and direction can be represented graphically on the kinetic diagram, shown in
Fig. 3.1c. Recall the free-body diagram considers the particle to be free of its surroundings
and shows all the forces acting on the particle. The kinetic diagram pertains to the particle's
motion as caused by the forces
We may observe that if the resultant FR of the forces acting on the particle is zero, it follows
from Eq. (3.3) thus the acceleration is also zero, so that the particle will either remain at rest
or move along a straight line path with constant velocity. Such are the conditions of static
equilibrium, Newton's first law of motion.

Free-body diagram
(a)

(b)

ma

Kinetic diagram
(c)

Fig. 3.1
Here it is worthy to note that the equation of motion can also be rewritten in the form

F ma 0.

The vector ma is referred to as the inertia force vector. If it is treated in the

same way as a "force vector;" then the state of "equilibrium" created is referred to as dynamic
equilibrium. This method for application is often referred to as the d'Alembert principle,
named after the French mathematician Jean Le Rond d'Alembert.
We can now extend the equation of motion to include a system of n particles isolated within
an enclosed region in space, as shown in Fig. 3.2a. In particular, there is no restriction in the
way the particles are connected, and as a result the following analysis will apply equally well
to the motion of a solid, liquid, or gas system. At the instant considered. the arbitrary ith
particle, having a mass m, is subjected to a system of internal forces and a resultant external
force. The resultant internal force, represented symbolically as fi, is determined from the
forces which the other particles exert on the ith particle. Usually these forces are developed by
direct contact, although the summation extends over all n particles within the dashed

Chapter 3. Particle Kinetics

37

boundary. The resultant external force Fi represents. for example. the effect of gravitational,
electrical. magnetic. or contact forces between the ith particle and adjacent bodies or particles
not included within the system.
The free-body and kinetic diagrams for the ith particle are shown in Fig. 3.2b. Applying the
equation of motion yields
Fi +fi =mi ai

(3.4)

When the equation of motion is applied to each of the other particles of the system, similar
equations will result. If all these equations are added together vectorially, we obtain
Fi +fi =mi ai

Fi
fi
Free-body
diagram
(a)

Fig.3.2

Kinetic
diagram
(b)

Since internal forces between particles all occur in equal but opposite collinear pairs, the
summation of the internal forces, will equal zero. Consequently, only the sum of the external
forces will remain, and therefore the equation of motion, written for the system of particles,
becomes
Fi =mi ai

(3.5)

If we denote rG as a position vector which locates the center of mass G of the particles, Fig.
3.2a, then by definition of the center of mass
rG =

mi ri
mi

Differentiating this equation twice with respect to time, assuming that no mass is entering or
leaving the system, and m = mi is the total mass of all the particles, yields
maG =mi ai
Substituting this result into Eq. (3.5), we obtain
F=maG

(3.6)

Chapter 3. Particle Kinetics

38

Hence, the sum of the external forces acting on the system of particles is equal to the total
mass of the particles times the acceleration of its center of mass G.
Note that when applying the equation of motion of the particle we can distinguish two
problems of dynamics. In the first type the acceleration is either specified or can be
determined directly from known kinematic conditions. The corresponding forces which act on
the particle are the determined by direct substitution into Eq. (3.3). This problem is generally
quite straightforward.
In the second type of problem the forces are specified and the resultant motion is to be
determined. If the forces are constant the acceleration is constant and is easily found from Eq.
(3.3). When the forces are functions of time, position, velocity or acceleration, Eq. (3.3)
becomes a differential equation which must be integrated to determine the velocity and
displacement. Problems of this second type are often more formidable, as the integration may
be difficult to carry out, particularly, when the force is a function of two or more motion
variables. The numerical methods could be used for these cases.
Important Points

The equation of motion is based on experimental evidence and is valid only when
applied from an inertial frame of reference.

The equation of motion states that the unbalanced force on a particle causes it to
accelerate.

The force acting on the particle is equal to the rate of change of the linear momentum of
the particle.

An inertial frame of reference has axes that either translate with constant velocity or are
at rest.

The sum of the external forces acting on the system of particles is equal to the total
mass of the particles times the acceleration of its center of mass.

The vector ma is referred to as the inertia force vector and the state of dynamic
equilibrium of the particle is created. This method for application is often referred to as
the D'Alembert principle

When applying the equation of motion of the particle we encounter two problems of
dynamics.

Chapter 3. Particle Kinetics

39

3.1.2 Equations of Motion: Rectangular Coordinates


Consider a particle of mass m moving relatively to an inertial x, y, z frame of reference. The
forces acting on the particle, as well as its acceleration, may be expressed in terms of their i, j,
k components. Applying the equation of motion, we have
(Fxi Fy j Fz k)=m(axi+ay j+az k)
For this equation to be satisfied, the respective i, j, k components on the left side must equal
the corresponding components on the right side. Consequently, we may write the following
three scalar equations:
Fx =max
Fy =may

(3.7)

Fz =maz
In particular, if the particle is constrained to move only in the x-y plane, then the first two of
these equations are used to specify the motion.
Recalling from Chap. 2 that the components of the acceleration are equal to the second
derivatives of the coordinates of the particle, we have
&
&
Fx =mx
&
&
Fy =my

(3.8)

&
Fz =mz&
Keep in mind that all accelerations should be measured with respect to a newtonian frame of
reference.
Procedure for Analysis
The equations of motion are used to solve problems which require a relationship between the
forces acting on a particle and the accelerated motion they cause.
Free-Body Diagram

Select the inertial coordinate system. Most often, rectangular or x, y, z coordinates are
chosen to analyze problems for which the particle has rectilinear motion.

Once the coordinates are established, draw the particle's free-body diagram. Drawing
this diagram is very important since it provides a graphical representation that accounts
for all the forces which act on the particle, and thereby makes it possible to resolve
these forces into their x, y, z components.

Chapter 3. Particle Kinetics

40

The direction and sense of the particle's acceleration a should also be established. If the
senses of its components are unknown, for mathematical convenience assume that they
are in the same direction as the positive inertial coordinate axes.

The acceleration may be represented as the ma vector on the kinetic diagram.

Identify the unknowns in the problem.

Equations of Motion

If the forces can be resolved directly from the free-body diagram, apply the equations of
motion in their scalar component form.

If the geometry of the problem appears complicated, which often occurs in three
dimensions, Cartesian vector analysis can be used for the solution.

If the particle contacts a rough surface, it may be necessary to use the frictional
equation, which relates the coefficient of kinetic friction k to the magnitudes of the
frictional and normal forces Ff and N acting at the surfaces of contact, i.e., Ff = k N.
Remember that Ff, always acts on the free-body diagram such that it opposes the motion
of the particle relative to the surface it contacts.

If the particle is connected to an elastic spring having negligible mass, the spring force
Fs can be related to the deformation of the spring by the equation Fs = ks. Here k is the
spring's stiffness measured as a force per unit length, and s is the stretch or compression
defined as the difference between the deformed length 1 and the undeformed length l0
i.e., s = l l0.

Kinematics

If the velocity or position of the particle is to be found, it will be necessary to apply the
proper kinematic equations once the particle's acceleration is determined from F = ma.

If acceleration is a function of time, use a = dv/dt and v = ds/dt which, when integrated,
yield the particle's velocity and position.

If acceleration is a function of displacement, integrate ads = v dv to obtain the velocity


as a function of position.

If acceleration is constant, use v = v0 + act, s = s0 v0t +

2
at to determine the

velocity or position of the particle.

In all cases, make sure the positive inertial coordinate directions used for writing the
kinematic equations are the same as those used for writing the equations of motion.
Otherwise, simultaneous solution of the equations will result in errors.

Chapter 3. Particle Kinetics

41

If the solution for an unknown vector component yields a negative scalar, it indicates
that the component acts in the direction opposite to that which was assumed.

EXAMPLE 3.1
The 50-kg crate shown in Fig. 3.3a rests on a horizontal plane for which the coefficient of
kinetic friction is k = 0.3. If the crate is subjected to a 400-N towing force as shown,
determine the velocity of the crate in 3 s starting from rest.
490.5 N

400 N
300

(a)

F=0.3 NC

(b)
NC
Fig. 3.3

Solution
Using the equations of motion, we can relate the crate's acceleration to the force causing the
motion. The crate's velocity can then be determined using kinematics.
Free-Body Diagram.
The weight of the crate is W = mg = 50 kg (9.81 m/s2) = 490.5 N. As shown in Fig. 3.3b, the
frictional force has a magnitude F = k NC and acts to the left, since it opposes the motion of
the crate. The acceleration a is assumed to act horizontally, in the positive x direction. There
are two unknowns, namely NC and a.
Equations of Motion.
Using the data shown on the free-body diagram. we have
400 cos30o 0.3 NC = 50a

(1)

NC 490.5 + 400 sin300 = 0

(2)

Solving Eq. (2) for NC, substituting the result into Eq. (1), and solving for a yields
NC = 290.5 N
a = 5.19 m/s2
Kinematics.
Note that the acceleration is constant since the applied force P is constant. Since the initial
velocity is zero, the velocity of the crate in 3 s is

Chapter 3. Particle Kinetics

42

v = v0 + ac t = 0 +5.19 (3) = 15.6 m/s.

EXAMPLE 3.2
The 100-kg block A shown in Fig. 3.4a is released from rest. If the masses of the pulleys and
the cord are neglected, determine the speed of the 20-kg block B in 2 s.

2T

(a)

(b)

(c)

Fig. 3.4
Solution
Free-Body Diagrams.
Since the mass of the pulleys is neglected, then for pulley C, ma = 0 and we can apply
Fy = 0 as shown in Fig. 3.4b. Reader can easily draw the free-body diagrams for blocks A
and B as shown Fig. 3.4c. One can see that for A to remain static requires T = 490.5 N,
whereas for B to remain static requires T = 196.2 N. Hence A will move down while B moves
up. Here we will assume both blocks accelerate downward, in the direction of +sA and +sB.
The three unknowns are T, aA and aB.
Equation of motion.
Block A:
981 2T = 0

(1)

Block B (Fig. 3.5c):


196.2 - T = 20aB
Kinematics.

(2)

Chapter 3. Particle Kinetics

43

The necessary third equation is obtained by relating aA to aB using a dependent motion


analysis. Since the coordinates sA and sB measure the positions of A and B from the fixed
datum, Fig. 3.5a, we can write
2 sA+ sB = l
where l is constant and represents the total vertical length of cord. Differentiating this
expression twice with respect to time yields
2aA = - aB

(3)

Notice that in writing Eqs. (1) to (3), the positive direction was always assumed downward. It
is very important to be consistent in this assumption since we are seeking a simultaneous
solution of equations. The solution yields
T = 327.0 N, aA = 3.27 m/s2, aB = - 6.54 m/s2
Hence when block A accelerates downward, block B accelerates upward. Since aB is constant,
the velocity of block B in 2 s is thus
v = v0 + aB t
= 0 + (-6.54) (2) = -13.1 m/s
The negative sign indicates that block B is moving upward.

3.1.3 Equations of Motion: Normal and Tangential Coordinates


Now we turn our attention to the kinetics of particles moving over a curved path which is
known. In this case the equation of motion for the particle may be written in the tangential,
normal and binormal directions. We have
Ft =mat
Fn =man

(3.9)

Fb= 0
Here Ft, Fn, Fb represent file sums of all the force components acting on the particle in
the tangential, normal and binormal directions, respectively, Fig 3.5. Note that there is no
motion of the particle in the binormal direction, since the particle is constrained to move
along the path. The above equation is satisfied provided

Fig. 3.5

Chapter 3. Particle Kinetics

Recall that at =

dv
dt

44

represents the time rate of change in the magnitude of velocity.

Consequently. if Ft acts in the direction of motion, the particle's speed will increase, whereas
if it acts in the opposite direction, the particle will slow down. Likewise. an =

v2

represents

the time rate of change in the velocity's direction. Since this vector always acts in the positive
n direction, i.e., toward the path's center of curvature, then Fn which causes an also acts in
this direction. For example. when the particle is constrained to travel in a circular path with a
constant speed, there is a normal force exerted on file particle by the constraint in order to
change the direction of the particle's velocity (an). Since this force is always directed toward
the center of the path, it is often referred to as the centripetal force.
Procedure for Analysis
When a problem involves the motion of a particle along a known curved path, normal and
tangential coordinates should be considered for the analysis since the acceleration components
can be readily formulated. The method for applying the equations of motion, which relate the
forces to the acceleration, has been outlined in the procedure given in Sec. 3.1.2. Specifically
for t, n, b coordinates it may be stated as follows:
Free-body Diagram

Establish the inertial t, n, b coordinate system at the particle and draw the particle's freebody diagram.

The particle's normal acceleration an always acts in the positive n direction.

If the tangential acceleration at is unknown, assume it acts in the positive t direction.

Identify the unknowns in the problem.

Equations of motion

Apply the equations of motion, Eqs. (3.9).

Kinematics

Formulate the tangential and normal components of acceleration, i.e., at = dv/dt or


at = vdv/ds and an = v2 /.

If the path is defined as y = f(x), the radius of curvature at the point where the particle is
located can be obtained from

= 1+(dy / dx)2
EXAMPLE 3.3

3/ 2

/ d2y / dx
2 .

Chapter 3. Particle Kinetics

45

Determine the banking angle for the race track so that the wheels of the racing cars shown
in Fig. 3.6a will not have to depend upon friction to prevent any car from sliding up or down
the track. Assume the cars have negligible size, a mass m, and travel around the curve of
radius with a speed v.

W=mg

(a)

(b)

Fig. 3.6
Solution
Before looking at the following solution. give some thought as to why it should be solved
using t, n, b coordinates.
Free-Body Diagram
As shown in Fig. 3.6a, and as stated in the problem, no frictional force acts on the car. Here
NC represents the resistant of the ground on all four wheels. Since an can be calculated, the
unknowns are NC and .
Equations of Motion
Using the n, b axes we have
NC cos mg = 0

(1)

v2

(2)

NC sin =m

Eliminating NC and m from these equations by dividing Eq. 2 by Eq. 1, we obtain


tan =

v2
g

v2
=tan-1

g
Notice that the result is independent of the mass of the car. Also, a force summation in the
tangential direction is of no consequence to the solution. If it were considered, then
at=dv/dt=0, since the car moves with constant speed.
EXAMPLE 3.4

Chapter 3. Particle Kinetics

46

Packages, each having a mass of 2 kg, are delivered from a conveyor to a smooth circular
ramp with a velocity of v0 = 1 m/s as shown in Fig. 3.7a. If the radius of the ramp is 0.5 m.
determine the angle max at which each package begins to leave the surface.

(a)

(b)
Fig. 3.7

Solution
Free-Body Diagram
The free-body diagram for a package, when it is located at the general position , is shown in
Fig. 3.7b. The package must have a tangential acceleration at. since its speed is always
increasing as it slides downward. The weight is W = 2(9.81) = 19.62 N. Specify the three
unknowns.
Equations of Motion
- NB + 19.62 cos = 2 v2/0.5

(1)

19.62 sin = 2at

(2)

At the instant = max, the package leaves the surface of the ramp so that NB = 0. Therefore,
there are three unknowns v, at and .
Kinematics
The third equation for the solution is obtained by noting that the magnitude of tangential
acceleration at may he related to the speed of the package v and the angle . Since at ds = v dv
and ds = r d = 0.5 d, Fig. 3.7a, we have
at =

vdv
0.5d

To solve, substitute Eq. (3) into Eq. (2) and separate the variables. This gives
vdv = 4.905 sin d
Integrate both sides, realizing that when = 0, v0 = 1 m/s

(3)

Chapter 3. Particle Kinetics

47

vdv=4.905 sind
v2 = 9.81(1-cos )+1
Substituting into Eq. (1) with NB = 0 and solving for cosmax yields
19.62cos max

2
9.81(1 cosmax ) 1
0.5

max = 42.7o.

3.1.4 Equations of Motion: Cylindrical Coordinates


When all the forces acting on a particle are resolved into cylindrical components, i.e., along
the unit-vector directions ur, u, uz, Fig. 3.8, the equation of motion may be expressed as
F = ma
Fr ur F u Fz uz mar ur ma u mazuz
We get easily the following three scalar equations of motion:
Fr =mar
F =ma

(3.10)

Fz = 0
If the particle is constrained to move only in the
r- plane, then only the first two of Eqs. (3.10)

F u F u
z

are used to specify the motion.


The

most

straightforward

type

of

problem

Fu

involving cylindrical coordinates requires the

determination of the resultant force components


Fr, F, Fz causing a particle to move with a
known acceleration. If, however, the particle's
accelerated motion is not completely specified at
the given instant, then some information regarding

Fig. 3.8

the directions or magnitudes of the forces acting on the particle must be known or computed
in order to solve Eqs. (3.10). For example, the force P causes the particle in Fig. 3.9a to move
along a path r = f(). The normal force N which the path exerts on the particle is always
perpendicular to the tangent of the path, whereas the frictional force F always acts along the
tangent in the opposite direction of motion. The directions of N and F can be specified relative

Chapter 3. Particle Kinetics

48

to the radial coordinate by using the angle , Fig. 3.9b, which is defined between the extended
radial line and the tangent to the curve.
This angle can be obtained by noting that when the particle is displaced a distance ds along
the path Fig. 3.9c, the component of displacement in the radial direction is dr and the
component of displacement in the transverse direction is rd.
Since these two components are mutually perpendicular, the angle can be determined from
r
tan
dr / d

(a)

(b)

(c)

Fig. 3.9
If is calculated as a positive quantity, it is measured from the extended radial line to the
tangent in a counterclockwise sense or in the positive direction of . If it is negative, it is
measured in the opposite direction to positive .
Procedure for Analysis
Cylindrical or polar coordinates are a suitable choice for the analysis of a problem for which
data regarding the angular motion of the radial line r are given, or in cases where the path can
be conveniently expressed in terms of these coordinates. Once these coordinates have been
established, the equations of motion can be applied in order to relate the forces acting on the
particle to its acceleration components. The method for doing this has been outlined in the
procedure for analysis given in Sec. 3.1.2.The following is a summary of this procedure.
Free-Body Diagram

Establish the r, , z inertial coordinate system and draw the particle's free-body diagram.

Assume that ar, a , az act in the positive directions of r, , z, if they are unknown.

Identify all the unknowns in the problem.

Equations of motion

Apply the equations of motion, Eqs. (3.10).

Kinematics

Chapter 3. Particle Kinetics

49

&, z&
&, &, &
&
Use the methods of Sec. 2.6 to determine r and the time derivatives r&, r&
& 2r&
&.
&and az z&
& r&2 , a r&
and evaluate the acceleration components ar r&

If any of the acceleration components is computed as a negative quantity, it indicates


that it acts in its negative coordinate direction.

When taking the time derivatives of' r = f(), it is very important to use the chain rule of
calculus.

EXAMPLE 3.5
The smooth 2-kg cylinder C in Fig. 3.10a has a peg P through its center which passes through
the slot in arm OA. If the arm rotates in the vertical plane at a constant rate & = 0.5 rad/s,
determine the force that the arm exerts on the peg at the instant 600.

(a)

(b)
Fig. 3.10

Solution
We will use polar coordinates to solve this problem.
Free-Body Diagram
The free-body diagram for the cylinder is shown in Fig. 3.10b. The force on the peg, FP, acts
perpendicular to the slot in the arm. As usual, ar and a are assumed to act in the directions of
positive r and , respectively. In this case the four unknowns are ar , a, FP and NC.
Equation of Motion
According to Fig. 3.10b
19.62 sin NC sin =2ar

(1)

19.62 cos +FP NC cos =2a

(2)

Kinematics
From Fig. 3.10b, r can be related to by the equation
r

0.4
sin

Chapter 3. Particle Kinetics

50

&=0. We get
&with =60o, &= 0 and &
We can calculate r, r&, r&
&=0.192
r = 0.462 , r&=-0.133 , r&
Hence
& r&2 = 0.192 0.462(0.5)2 = 0.0770
ar r&
& 2r&
a r&
&= 0+2(-0.133)(0.5) =-0.133
Substituting these results into Eqs. (1) and (2) with = 60O and solving yields
NC = 19.4 N, FP = -0.356 N

The negative sign indicates that FP acts opposite to that shown in Fig.3.10b.

3.2 Kinetics of a Particle: Work and Energy


3.2.1 Introduction
In the previous parts, most problems dealing with the motion of particles were solved through
the use of Newtons second law. Given a particle acted upon by a force, we could determine
the acceleration, then by applying the principles of kinematics we could calculate the velocity
and position of the particle at any time.
We may incorporate the principles of kinematics into the second law of dynamics so that it
becomes unnecessary to solve directly for the acceleration. Integrations with respect to
displacement leads to the equation of work and energy. Integrations with respect to time leads
to the equations of impulse momentum. The method of work and energy will be considered
first. We will discuss the principle of impulse and momentum later in this chapter.

3.2.2 The Work of a Force


We shall first define the terms work and displacement as they are used in mechanics. In
mechanics a force F does work on a particle only when the particle undergoes a displacement
in the direction of the force. Consider the force F acting on the particle in Fig. 3.11. If the
particle moves along the path s from position r to a new position r', the displacement is then
dr = r'- r. The magnitude of dr is represented by ds, the differential segment along the path. If
the angle between the tails of dr and F is , then the work dU which is done by F is a scalar
quantity, defined the dot product
dU = F dr

(3.11)

This equation may also be written as


dU = F ds cos

(3.12)

Chapter 3. Particle Kinetics

51

This result may be interpreted in one of two ways: either as the product of F and the
component of displacement in the direction of the force, i.e., ds cos, or as the product of ds
and the component of force in the direction of displacement. i.e., Fcos. Note that if
0o<90o, then the force component and the displacement have the same sense so that the
work is positive, whereas if 90o< 180o, these vectors have an opposite sense, and therefore
the work is negative. Also, dU = 0 if the force is perpendicular to displacement, since cos
90= 0, or if the force is applied at a fixed point, in which case the displacement is zero.

Fig. 3.11
We may also express the work dU in term of rectangular components of the force and of the
displacement
dU = Fxdx +Fydy +Fzdz

(3.13)

The basic unit for work in the SI system is called a joule (J). This unit combines the units of
force and displacement. Specifically, 1 joule of work is done when a force of 1 newton moves
1 meter along its line of action (1J = 1 Nm). The moment of a force has this same combination
of units (Nm), however, the concepts of moment and work are in no way related. A moment is
a vector quantity, whereas work is a scalar.
The work of the force F during a finite displacement of the particle along its path from r1 to
r2 or s1 to s2. Fig 3.12, the work is obtained by integration. If F is expressed as a function of
position, F = F(s), we have
r2

s2

r1

s1

U 12 Fdr F cosds

(3.14)

Chapter 3. Particle Kinetics

52

Fig. 3.12
When the force F is determined by its rectangular components, the expression (3.13) may be
used for elementary work. We write then
U 12

A2

(F dx F dy F dz)
x

(3.15)

A1

where the integration is to be performed along the path described by the particle.
Now we will consider some special cases. Firstly, if the force FC has a constant magnitude and
acts at a constant angle from its straight-line path, we can get easily from Eq. (3.14):
U1-2 = FC cos (s2 s1)

(3.16)

We can also determine the work of a weight. Consider a particle which moves up along the
path s shown in Fig. 3.13 from position s1 to position s2. Note that y-direction is vertical, we
can get from Eq. (3.15):
U1-2 = -W(y2 y1) = -W y

(3.17)

Thus, the work dome is equal to the magnitude of the particle's weight times its vertical
displacement. In the case shown in Fig. 3.13 the work is negative, since W is downward, i.e.,
W < 0, and y is upward, i.e., y > 0. It is clear, however, that if the particle is displaced
downward, the work of the weight is positive.

Fig. 3.13
Now consider the work of a spring force. If a particle (or body) is attached to a linear elastic
spring and the spring is displaced a distance s from its unstretched position, e.g. in x-direction,
then the force Fs exerted on the particle is opposite to the direction of displacement and has
the magnitude Fs = ks, where k is the spring stiffness. Consequently, the force will do
negative work on the particle when the particle is moving so as to further elongate (or
compress) the spring. Hence, we get from Eq. (3.15)
1
U 12 k(s22 s12)
2

(3.18)

Chapter 3. Particle Kinetics

53

When this equation is used, a mistake in sign can be eliminated if one simply notes the
direction of the spring force acting on the particle and compares it with the direction of
displacement of the particle s = s2 s1 , if both are in the same direction, positive work results;
if they are opposite to one another, the work is negative.

3.2.3 Kinetic Energy of a Particle and Principle of Work and Energy


Consider a particle P of mass m acted upon by the force resultant F = F, Fig. 3.14. At the
instant considered the particle is located on the path, either rectilinear or curved, as measured
from an inertial coordinate system. We can write the equation of motion for the particle in the
tangential direction as
Ft = mat

Fig. 3.14
Applying the kinematic equation
at

vdv
ds

and integrating and integrating both sides, assuming initially that the particle has a position
s = s1 and a speed v = v1 and later at s = s2, v = v2 yields
s2

v2

s1

v1

s2

1 2 1 2
mv2 mv1
2
2

Ftds mvdv
Ftds
s1

(3.19)

Using the definition of work of a force the final result may be written as
U 12

1 2 1 2
mv2 mv1
2
2

(3.20)

This equation represents the principle of work and energy for the particle. The term on the left
is the sum of the work done by all the forces acting on the particle as the particle moves from
point 1 to point 2. The two terms on the right side. which are of the form

Chapter 3. Particle Kinetics

1 2
mv
2

54

(3.21)

define the particle's final and initial kinetic energy, respectively. These terms are always
positive scalars. The kinetic energy is measured in the same units as work, i.e., in joule. We
check that in SI units:
T

1 2
mv kg(m/ s)2 (kg m/ s2)=Nm= J
2

Substituting Eq. (3.21) into Eq. (3.20) we have


U1-2 = T2 T1

(3.22)

which states that the work done by all the forces acting on the particle as it moves from its
initial to its final position is equal to the change in particle's kinetic energy. Rearranging the
terms in Eq. (3.22) we write
T1 + U1-2 = T2

(3.23)

Thus the particles initial kinetic energy plus the work done by all forces acting on the particle
as it moves from its initial to its final position is equal to the particles final kinetic energy.
As noted from the derivation, the principle of work and energy represents an integrated form
of Ft = mat obtained by using the kinematic equation at = vdv/ds. As a result, this principle
will provide a convenient substitution for Ft = mat, when solving those types of kinetic
problems which involve force, velocity, and displacement, since these variables are involved
in the terms of Eq. (3.22). For example, if a particle's initial speed is known and the work of
all the forces acting on the particle can be determined, then Eq. (3.22) provides a direct means
of obtaining the final speed v2, of the particle after it undergoes a specified displacement. If
instead v2 is determined by means of the equation of motion, a two-step process is necessary;
i.e., apply Ft = mat to obtain at, then integrate at=vdv/ds to obtain v2. Note that the principle
of work and energy cannot be used, for example, to determine forces directed normal to the
path of motion, since these forces do no work on the particle. Instead Ft = mat must be
applied. For curved paths, however, the magnitude of the normal force is a function of speed.
Hence, it may be easier to obtain this speed using the principle of work and energy, and then
substitute this quantity into the equation of motion Fn = mv2/ to obtain the normal force.
Procedure for Analysis
The principle of work and energy is used to solve kinetic problems that involve velocity, force
and displacement, since these terms are involved in the equation. For application it is
suggested that the following procedure be used.
Work (Free-Body Diagram)

Chapter 3. Particle Kinetics

55

Establish the inertial coordinate system and draw a free-body diagram of the particle in

order to account for all the forces that do work cm the particle as it moves along its path.
Principle of Work and Energy

Apply the principle of work and energy, U1-2 = T2 T1

The kinetic energy at the initial and final points is always positive, since it involves the
speed squared T mv2 / 2.

A force does work when it moves through a displacement in the direction of the force.

Work is positive when the force component is in the same direction as its
displacement, otherwise it is negative.

Forces that are functions of displacement must be integrated to obtain the work.

The work of a weight is the product of the weight magnitude and the vertical
displacement, U1-2 = -W(y2 y1). It is positive when the weight moves downwards.

The work of a spring is of the form U 12 k(s22 s12)/ 2 where k is the spring
stiffness and s is the stretch or compression of the spring.

3.2.4 Principle of Work and Energy for a System of Particles


In this section we will extend the principle of work and energy to include a system of n
particles isolated within an enclosed region of space as shown in Fig. 3.15. Here the arbitrary
ith particle, having a mass mi , is subjected to a resultant external force Fi and a resultant
internal force fi which each of the other particles exerts on the ith particle.

Fig. 3.15
Using Eq. (3.19), which applies in the tangential direction, the principle of work and energy
written for the ith particle is thus
s

i2
1 2 i2
1
mvi 1 (Fi )t ds ( fi )t ds mvi22
2
2
si 1
si 1

Chapter 3. Particle Kinetics

56

Similar equations result if the principle of work and energy, is applied to each of the other
particles of the system. Since both work and kinetic energy are scalars, the results may be
added together algebraically, so that
s

i2
i2
1
1
mvi21 (Fi )t ds ( fi )t ds mvi22
2
2
si 1
si 1

We can write this equation symbolically as


T1 + U1-2 = T2

(3.24)

This equation states that the system's initial kinetic energy (T1) plus the work done by all the
external and internal forces acting on the particles of the system (U1-2) is equal to the
system's final kinetic energy (T2). To maintain this balance of energy, strict accountability of
the work done by all the forces must be made. In this regard, note that although the internal
forces on adjacent particles occur in equal but opposite collinear pairs, the total work done by
each of these forces will, in general, not cancel out since the paths over which corresponding
particles travel will be different.
However, there are two important exceptions to this rule which often occur in practice. If the
particles are contained within the boundary of a translating rigid body, the internal forces all
undergo the same displacement, and therefore the internal work will be zero. Also, particles
connected by inextensible cables make up a system that has internal forces which are
displaced by an equal amount. In this case, adjacent particles exert equal but opposite internal
forces that have components which undergo the same displacement, and therefore the work of
these forces cancels. On the other hand, note that if the body is assumed to be non-rigid, the
particles of the body are displaced along different paths, and some of the energy due to force
interactions would be given off and lost as heat or stored in the body if permanent
deformations occur.
It is worthy also to note that the procedure for analysis outlined in the previous section
provides a method for applying Eq. (3.24); however, only one equation applies for the entire
system. If the particles are connected by cords, other equations can generally be obtained by
using the kinematic principles in order to relate the particle's speeds.
Note also that Eq. (3.24) can be applied to problems involving sliding friction of a body,
however, it should be fully realized that the work of the resultant frictional force is not
represented by KNs, where K is the kinetic friction coefficient, N is the normal component
of the contact force and s is the displacement of the applied force acting on the body. Instead,
this term represents both the external work of friction and internal work which is converted
into various forms of internal energy, such as heat, deformation etc.

Chapter 3. Particle Kinetics

57

EXAMPLE 3.6
The platform P, shown in Fig. 3.16a, has negligible mass and is tied down so that the 0.4-m
long cords keep a 1-m long spring compressed 0.6 m when nothing is on the platform. If a 2kg block is placed on the platform and released from rest after the platform is pushed down
0.1 m, Fig. 3.16b, determine the maximum height h the block rises in the air, measured from
the ground.

(a)

(b)

Fig.3.16
Solution
Work (Free-Body Diagram)
Since the block is released from rest and later reaches its maximum height, the initial and final
velocities are zero. When the block is still in contact with the platform) one can draw easily
the free-body diagram of the block with two forces: spring force and weight of the block. It is
clear that the weight does negative work and the spring force does positive work. In
particular, the initial compression in the spring is s1 = 0.6 m + 0.1 m = 0.7 m. Due to the
cords, the spring's final compression is s2 = 0.6 m (after the block leaves the platform). The
bottom of the block rises from a height of (0.4 m - 0.1 m) = 0.3 m to a final height h.
Principle of Work and Energy
T1 + U1-2 = T2

1 2
1
1
1
mv1 ( ks22 ks12) W y mv22
2
2
2
2
Note that here s1 = 0.7 m > s2 = 0.6 m and so the work of as determined from Eq. (3.18) will
indeed be positive once the calculation is made. Thus,
1
1

0 [ (200 N / m)(0.6 m) (200 N / m)(0.7 m)] (19.62 N )[ h (0.3m)] 0


2
2

Solving yields
h = 0.963 m

EXAMPLE 3.7

Fig. 3.17

Chapter 3. Particle Kinetics

58

Packages having a mass of 2 kg are delivered from a


conveyor to a smooth circular ramp with a velocity of v = 1
m/s as shown in Fig. 3.17. If the radius of the ramp is 0.5 m,
determine the angle max at which each package begins to
leave the surface.
Solution
Work (Free-Body Diagram)
The free-body diagram of the block is shown at the intermediate location . It is clear that the
weight W = 2(9.81) = 19.62 N does positive work during the displacement. If a package is
assumed to leave the surface when = max, then the weight moves through a vertical
displacement of [0.5-0.5cosmax] m.
Principle of Work and Energy
T1 + U1-2 = T2
1
1
(2kg)(1m/ s)2 19.62N(0.5-0.5cos max )m (2kg)v22
2
2
v22 9.81(1 cos max ) 1

(1)

Equation of Motion
There are two unknowns in Eq. (1), max and v2. A second equation relating these two
variables may be obtained by applying the equation of motion in the normal direction to the
forces on the free-body diagram. Therefore

v2
N B 19.62N cos (2kg)

0.5m
When the package leaves the ramp at = max , NB = 0 and v = v2 hence, this equation becomes
cos max

v22
4.905m

(2)

Eliminating the unknown v22 between Eqs. (1) and (2) gives
4.905cos max 9.81(1 cos max ) 1
Solving, we have
cos max 0.735

max = 42.7o
This problem has also been solved in Sec.3.1.3. If the two methods of solution are compared,
it will be apparent that a work-energy approach yields a more direct solution.

Chapter 3. Particle Kinetics

59

3.2.5 Power and Efficiency


The capacity of a machine is measure by the time rate at which it can do work or deliver
energy. The total work of energy output is not a measure of this capacity since a motor, no
matter how small can deliver a large amount of energy if given sufficient time. On the other
hand a large and powerful machine is required to deliver a large amount of energy in a short
period. Hence the capacity of a machine is rated by its power which is defined as the amount
of work performed per unit of time. Therefore the power generated by a machine or engine
that performs an amount of work dU within the time interval dt is
P

dU
dt

(3.25)

Provided the work dU is expressed by dU = Fdr, then it is also possible to write


P

dU
Fdr
dr

F
dt
dt
dt

or
P Fv

(3.26)

It is clear that power is a scalar. Note also that in the formulation v represents the velocity of
the point which is acted upon by the force F.
In the SI systems the basic unit of power used is the watt (W) . We have 1 W =1 J/s =1 Nm/s.
The mechanical efficiency of a machine is defined as the ratio of the output of useful power
produced by the machine to the input of power supplied to the machine. Hence,
=

power output
power input

(3.27)

It is clear that if energy applied to the machine occurs during the same time interval at which
it is removed, then according to the definition of power the efficiency may also be expressed
in terms of the ratio of output energy to input energy; i.e.
=

energy output
energy input

(3.28)

Since machines consist of a series of moving parts, frictional forces will always be developed
within the machine, and as a result, extra energy or power is needed to overcome these forces.
Consequently, the efficiency of a machine is always less than l. When a machine is used to
transform mechanical energy into electrical energy or thermal energy into mechanical energy,
its overall efficiency may be obtained from Eq. (3.28). The overall efficiency provide a
measure of all the various energy losses involved(losses of electrical or thermal energy as well

Chapter 3. Particle Kinetics

60

as frictional losses). Note that we should express the power output and the power input in the
same units.
Procedure for Analysis

The power supplied to a body can be computed using the following procedure.

First determine the external force F acting on the body which causes the motion. This
force is usually developed by a machine or engine placed either within or external to the
body.

If the body is accelerating, it may be necessary to draw its free-body diagram and apply
the equation of motion (F = ma) to determine F.

Once F and the velocity v of the point where F is applied have been found, the power is
determined by multiplying the force magnitude by the component of velocity acting in
the direction of F, i.e., P = Fv = Fv cos.

In some problems the power may be found by calculating the work done by F per unit
of time (Pavg = U/t, or P = dU/dt).

EXAMPLE 3.8
The sports car shown in Fig. 3.18 has a mass of 2 Mg and an engine running efficiency of

0.63 . As it moves forward, the wind creates a drag resistance on the car of Ft = 1.2v2 N,
where v is the velocity in m/s. If the car is traveling at a constant speed of 50 m/s, determine
the maximum power supplied by the engine.

Fig. 3.18
Solution
One can draw easily the free-body diagram: the normal force NC and frictional force FC
represent the resultant forces of all four wheels. In particular, the unbalanced frictional force
drives or pushes the car forward. This effect is, of course, created by the rotating motion of
the rear wheels on the pavement and is developed by the power of the engine.
Applying the equation of motion in the vertical direction, we have
FC 1.2v2 (2000kg)

dv
dt

Since the car is traveling with constant velocity, dv/dt = 0. Hence with v = 50 m/s.
FC = 1.2 (50 m/s)2 = 3000 N

Chapter 3. Particle Kinetics

61

The power output of the car is manifested by the driving (frictional) force FC Thus
P = FC v = (3000 N)(50 m/s) = 150 kW
The power supplied by the engine (power input) is therefore
power input =

1
(power output) = 0.63 (150 kV )= 238kW.

3.2.6 Conservative Forces, Potential Energy and Conservation of Energy


So far a particle or a combination of joined particles are isolated and the work done by
weights, spring forces and other externally applied forces acting on the particles is determined
when evaluating U in the work-energy equation. In the present section we will treat the work
done by gravity forces and spring forces by introducing the concept of potential energy.
When the work done by a force in moving a particle from one point to another is independent
of the path followed by the particle, then this force is called a conservative force. The weight
of a particle and the force of an elastic spring are two examples of conservative forces often
encountered in mechanics. The work done by the weight of a particle is independent of the
path since it depends only on the particle's vertical displacement. The work done by a spring
force acting on a particle is independent of the path of the particle, rather it depends only on
the extension or compression of the spring. In contrast to a conservative force, consider the
force of friction exerted on a moving object by a fixed surface. The work done by the
frictional force depends on the path: the longer the path, the greater the work. Consequently,
frictional forces are non-conservative. The work is dissipated from the body in the form of
heat.
As we know energy may be defined as the capacity for doing work. When energy comes from
the motion of the particle, it is referred to as kinetic energy. When it comes from the position
of the particle, measured from a fixed datum or reference plane, it is called potential energy.
Thus, potential energy is a measure of the amount of work a conservative force will do when
it moves from a given position to the datum.
In mechanics, the potential energy due to gravity (weight) or an elastic spring is important. If
a particle is located a distance y above an arbitrarily selected datum the particle's weight W
has positive gravitational potential energy, Vg, since W has the capacity of doing positive
work when the particle is moved back down to the datum. Likewise, if the particle is located a
distance y below the datum, Vg, is negative since the weight does negative work when the
particle is moved back up to the datum. Clearly, at the datum Vg = 0. In general, if y is positive
upward, the gravitational potential energy of the particle of weight W is thus
Vg = Wy

(3.29)

Chapter 3. Particle Kinetics

62

In going from one level y = y1 to a higher level at y = y2 the change in potential energy
becomes
Vg = W(y2-y1)= Wy.
The corresponding work done by the gravitational force on the particle U = Wy. Thus the
work done by the gravitational force is the negative of the change in potential energy. It is
also clear that the level, or datum, from which the elevation y is measured may be chosen
arbitrarily. Of course, here the weight is assumed to be constant. This assumption is suitable
for small differences in elevation y. If the elevation change is significant. however. a
variation of weight with elevation must be taken into account.
The second important potential energy is the one due to an elastic spring. When an elastic
spring is elongated or compressed a distance s from its unstretched position, the elastic
potential energy Ve due to the spring's configuration can he expressed as
1
Ve ks2
2

(3.30)

Here Ve is always positive since, in the deformed position, the force of the spring has the
capacity for always doing positive work on the particle when the spring is returned to its
unstretched position, Fig. 3.19.

Fig. 3.19
In the general case, if a particle is subjected to both gravitational and elastic forces. the
particle's potential energy can be expressed as a potential function, which is the algebraic
sums
V = Vg + Ve

(3.31)

Measurement of V depends on the location of the particle with respect to a selected datum in
accordance with Eqs. (3.29) and (3.30). If the particle is located at an arbitrary point (x, y, z)
in space, this potential function is then V = V(x, y, z).The work done by a conservative force in
moving the particle from point (x1, y1 , z1 ) to point (x2, y2 , z2) is measured by the difference of
this function, i.e.,
U1-2 = V1 V2

(3.32)

Chapter 3. Particle Kinetics

63

When the displacement along the path is infinitesimal, i.e., from point (x, y, z) to (x+dx, y+dy,
z+dz), Eq. (3.32) becomes
dU = V(x,y,z) - V(x+dx,y+dv,z+dz) = - dV(x, y, z)

(3.33)

Provided both the force and displacement are defined using rectangular coordinates, then the
work can also be expressed as
dU = Fdr = (Fxi + Fyj+ Fzk) (dxi + dyj + dzk) = Fx dx+Fy dy+Fz dz
Substituting this result into Eq. (3.33) and expressing the differential dV(x, y, z) in terms of its
partial derivatives yields
Fx dx+Fy dy+Fz dz =

V
V
V
dx
dy
dz
x
y
z

Since changes in x, y and z are all independent of one another, this equation is satisfied
provided
Fx

V
V
V
; Fy
; Fz
y
x
z

(3.34)

The force may also be written as the vector


F = - V

(3.35)

where the symbol (del) stands for the vector operator


i

j
k
x
y
z

The expression V is known as the gradient of the potential function. Eq. (3.35) relates a
force F to its potential function V and thereby provides a mathematical criterion for proving
that F is conservative. For example, the gravitational potential function for a weight located a
distance y above a datum is Vg = Wy. To prove that W is conservative, it is necessary to show
that it satisfies Eq. (3.35), in which case
Fy

V
(Wy)
W
;F
y
y

The negative sign indicates that W acts downward, opposite to positive y, which is upward.
When a particle is acted upon by a system of both conservative and non-conservative forces.
the portion of the work done by the conservative forces can be written in terms of the
difference in their potential energies using Eq. (3.32). As a result, the principle of work and
energy can be written as
T1 + V1 + (U1-2)noncons = T2 + V2

(3.36)

where (U1-2)noncons represents the work of the non-conservative forces acting on the particle.
If only conservative forces are applied to the body, this term is zero and then we have

Chapter 3. Particle Kinetics

T1 + V1

64

= T2 + V 2

(3.37)

Formula (3.37) indicates that when a particle moves under the action of conservative forces
the sum of the particle's kinetic and potential energies remains constant. For this to occur,
kinetic energy must be transformed into potential energy, and vice versa. The sum T+V is
called the total mechanical energy of the particle and is denote by E. Hence, we can write
E1 = E2

(3.38)

This equation is referred to as the conservation of mechanical energy or simply the


conservation of energy.
We can extend the law to the system of particles. If a system of particles is subjected only to
conservative forces, then an equation similar to Eq. (3.37) can be written for the particles.
Applying the ideas of the preceding discussion, Eq. (3.24) becomes
T1 + V1

= T2 + V2

(3.39)

Therefore, the sum of the system's initial kinetic and potential energies is equal to the sum of
the system's final kinetic and potential energies. Or in other words:
T + V = const

(3.40)

It is important to remember that only problems involving conservative force systems (weights
and springs) may be solved by using the conservation of energy theorem. As stated
previously, friction or other drag-resistant forces, which depend upon velocity or acceleration,
are non-conservative. A portion of the work done by such forces is transformed into thermal
energy, and consequently this energy dissipates into the surroundings and may not be
recovered.
Procedure for Analysis
The conservation of energy equation is used to solve problems involving velocity,
displacement, and conservative force systems. It is generally easier to apply than the principle
of work and energy because the energy equation just requires specifying the particle's kinetic
and potential energies at only two points along the path, rather than determining the work
when the particle moves through a displacement. For application it is suggested that the
following procedure be used.
Potential Energy

Draw two diagrams showing the particle located at its initial and final points along the
path.

If the particle is subjected to a vertical displacement, establish the fixed horizontal


datum from which to measure the particle's gravitational potential energy Vg.

Chapter 3. Particle Kinetics

65

Data pertaining to the elevation y of the particle from the datum and the extension or
compression s of any connecting springs can be determined from the geometry
associated with the two diagrams.

Recall Vg = Wy. where y is positive upward from the datum and negative downward
from the datum; also Ve ks2 / 2, which is always positive.

Conservation of Energy

Apply the equation T1 + V1 = T2 + V2.

When determining the kinetic energy, T mv2 / 2. the particle's speed v must be
measured from an inertial reference frame.

EXAMPLE 3.9
The gantry structure in Fig 3.20a is used to test the response of an airplane during a crash. As
shown in Fig. 3.20b the plane, having a mass of 8 Mg is hoisted back until = 60, and then
the pull-back cable AC is released when the plane is at rest. Determine the speed of the plane
just before crashing into the ground, = 15o. Also, what is the maximum tension developed in
the supporting cable during the motion? Neglect the effect of lift caused by the wings during
the motion and the size of the airplane.

(a)

(b)
Fig. 3.20

Solution
Since the force of the cable does no work on the plane, it must be obtained using the equation
of motion. First, however, we must determine the plane's speed at B.
Potential Energy
For convenience, the datum has been established at the top of the gantry.
Conservation of Energy
According to the law of conservation of energy we have
TA+VA =TB+VB

Chapter 3. Particle Kinetics

66

0 - 8000 kg (9.81 m/s2)(20 cos 60 m)


=

1
(8000 kg) vB2 - 8000 kg (9.81 m/s2 )(20 cos 15o m)
2

Hence
vB = 13.5 m/s
Equation of Motion.
One can draw easily the free-body diagram when the plane is at B (with two forces: the force
of cable T and the weight W), we have
Fn = man
T - 8000 (9.81)N cos 15 = (8000) kg) (13.5 m/s2) /20 m
Hence
T = 149 kN.
EXAMPLE 3.10
The ram R shown in Fig. 3.21a has a mass of 100 kg and is released from rest 0.75m from the
top of a spring, A, that has a stiffness kA = 15 kN/m. If a second spring B, having a stiffness
kB = 15 kN/m, is "nested" in A, determine the maximum displacement of A needed to stop the
downward motion of the ram. The unstretched length of each spring is indicated in the figure.
Neglect the mass of the springs.
Solution
Potential Energy.
We will assume that the ram compresses both springs at the instant it comes to rest. The
datum is located through the center of gravity of the ram at its initial position, Fig. 3.21b.
Then the kinetic energy is reduced to zero (v2 = 0), A is compressed a distance sA and B
compresses sB = sA 0.l m.

(a)

(b)

Chapter 3. Particle Kinetics

67

Fig. 3.21
Conservation of Energy
According to the law of conservation of energy we have
T1+V1 =T2+V2
or
0+0=0+[

1
1
kA sA2 + kB (sA 0.1)2 Wh]
2
2

1
1
0 + 0 = 0 + [ (12 000 N/m) sA2 + (15 000 N/m) (sA 0.1)2 981N (0.75m +sA)]
2
2
Rearranging the terms
13500 sA2 - 2481sA - 660.75 = 0,
we have the quadratic formula. It is clear that negative root does not represent the physical
situation since positive s is measured downward, the negative sign indicates that spring A
would have to be "extended" to stop the ram. Solving for the positive root we have
sA = 0.331 m
Since sB = 0.331m - 0.1m = 0.231 m, which is positive, the assumption that both springs are
compressed by the ram is correct.

3.3 Kinetics of a Particle: Impulse and Momentum


3.3.1 Principle of Linear Impulse and Momentum
In the previous articles we focused our attention on the equation of work and energy which are
obtained by integrating the equation of motion F = ma with respect to the displacement of the
particle. In consequence we found that the velocity changes could be expressed directly in
terms of the work done or in terms of the overall changes in the energy. In this section we will
integrate the equation of motion with respect to time rather than displacement and thereby
obtain the principle of impulse and momentum. It will then be shown that the resulting
equation will be useful for solving problems involving force, velocity, and time.
Consider a particle of mass m acted upon by the resultant of all forces. The equation of
motion can be written as
F ma m

dv
dt

Chapter 3. Particle Kinetics

68

where a and v are both measured from an inertial frame of reference. Rearranging the terms
and integrating between the limits v = v1, at t = t1 , and v = v2 at t = t2 we have
t2

v2

t1

v1

Fdt m dv
or
t2

Fdt mv2 mv1

(3.41)

t1

This equation is referred to as the principle of linear impulse and momentum. From the
derivation it can be seen that it is simply a time integration of the equation of motion. It
provides a direct means of obtaining the particle's final velocity v2 after a specified time
period when the particle's initial velocity is known and the forces acting on the particle are
either constant or can be expressed as functions of time. By comparison, if v2 was determined
using the equation of motion, a two step process would be necessary: i.e., apply F = ma to
obtain a, then integrate a= dv/dt to obtain v2.
In Eq. (3.41) the product of the mass and velocity is defined as the particle's linear momentum
L= mv

(3.42)

Since m is a positive scalar, the linear-momentum vector has the same direction as v, and its
magnitude mv has units of mass-velocity, i.e., kgm/s.
Also, by definition the integral
t2

I Fdt

(3.43)

t1

is called the linear impulse. This term is a vector quantity which measures the effect of a force
during the time the force acts. Since time is a positive scalar, the impulse acts in the same
linear impulse direction as the force, and its magnitude has units of force-time, i.e., Ns. If the
force is expressed as a function of time. the impulse may be determined by direct evaluation
of the integral. Note that L and I have the same units, hence Eq. (3.41) is dimensionally
homogenous.
The principle of linear impulse and momentum can be rewritten in the form
t2

mv1 Fdt mv2

(3.44)

t1

which states that the initial momentum of the particle at t1 plus the sum of all the impulses
applied to the particle from t1 to t2 is equivalent to the final momentum of the particle at t2. It

Chapter 3. Particle Kinetics

69

is clear that if each of the vectors in Eq. (3.44) is resolved into its x, y, and z components, we
can write symbolically the following three scalar equations:
t2

m(vx) )1 F xdt m(vx ) 2


t1

t2

m(vy) )1 F ydt m(vy ) 2

(3.45)

t1

t2

m(vz) )1 F zdt m(vz ) 2


t1

Keep in mind that while kinetic energy and work are scalar quantities, momentum and
impulse are vector quantities and Eqs. (3.45) represent the principle of linear impulse and
momentum for the particle in the x, y and z direction.
Procedure for Analysis
The principle of linear impulse and momentum is used to solve problems involving force,
time and velocity, since these terms are involved in the formulation. For application it is
suggested that the following procedure be used.
Free-Body Diagram

Establish the x, y, z inertial frame of reference and draw the particle's free-body diagram
in order to account for all the forces that produce impulses on the particle.

The direction and sense of the particle's initial and final velocities should he established.

If a vector is unknown, assume that the sense of its components is in the direction of the
positive inertial coordinates.

Principle of Impulse and Momentum

In accordance with the established coordinate system apply the principle of linear
impulse and momentum, Eq. (3.44). If motion occurs in the x-y plane, the two scalar
component equations can be formulated.

Realize that all the forces acting on the particle's free-body diagram will create an
impulse, even though some of these forces will do no work.

Forces that are functions of time must be integrated to obtain the impulse.

If the problem involves the dependent motion of several particles, use the kinematic equations
to relate their velocities. Make sure the positive coordinate directions used for writing these
kinematic equations are the same as those used for writing the equations of impulse and
momentum.
EXAMPLE 3.11

Chapter 3. Particle Kinetics

70

Blocks A and B shown in Fig. 3.22 have a mass of 3 kg and 5 kg, respectively. If the system is
released from rest, determine the velocity of block B in 6 s. Neglect the mass of the pulleys
and cord.
Solution
Free-Body Diagram
One can easily draw free-body diagrams for pulley D and
blocks A and B. Since the weight of each block is constant, the
cord tensions will also be constant. Furthermore, since the
mass of pulley D is neglected, the cord tension TA = 2TB
Assume that the blocks are both to be traveling downward in
the positive coordinate directions sA and sB.
Principle of Impulse and Momentum
Fig. 3.22

Block A:
t2

m(vA )1 Fydt m(vA )2


t1

0 - 2TB (6 s) + 3(9.81) N (6 s) = (3 kg)(vA)2

(1)

Block B:
t2

m(vB )1 Fydt m(vB )2


t1

0 + 5(9.81) N (6 s) TB(6 s) = (5 kg)(vB)2

(2)

Kinematis
Since the blocks are subjected to dependent motion, the velocity of A may be related to that of
B by using the kinematic analysis. A horizontal datum is established through the fixed point at
C and the position coordinates, sA and sB are related to the constant total length l of the
vertical segments of the cord by the equation
2sA + sB = l
Taking the time derivative yields
2vA = - vB
As indicated by the negative sign, when B moves downward A moves upward. Substituting
this result into Eq. (1) and solving Eqs. (1) and (2) yields
(vB)2 = 35.8 m/s , TB = 19.2 N.

Chapter 3. Particle Kinetics

71

Now we can extend the principle of linear impulse and momentum to a system of particles
moving relative to an inertial reference, Fig. 3.23. The principle is obtained from the equation
of motion applied to all the particles in the system, i.e.,
Fi m

dvi
dt

(3.46)

The term on the left side represents only the sum of the external forces acting on the system of
particles.

Fig. 3.23
Recall that the internal forces fi acting between particles do not appear with this summation
since by Newton's third law they occur in equal but opposite collinear pairs and therefore
cancel out. Multiplying both sides of Eq. (3.46) by dt and integrating between the limits t = t1,
vi = (vi)1 and t = t2, vi = (vi)2 yields
t2

mi (vi )1 Fidt mi (vi )2

(3.47)

t1

This equation states that the initial linear momenta of the system plus the impulses of all the
external forces acting on the system from t1 to t2 are equal to the system's final linear
momenta.
Using the definition of the location of the mass center G of the system one get easily the
relation
m vG = mi vi

(3.48)

which states that the total linear momentum of the system of particles is equivalent to the
linear momentum of a "fictitious" aggregate particle of mass m = mi, moving with the
velocity of the mass center of the system. Substituting into Eq. (3.47) yields
t2

m(vG )1 Fidt m(vG )2


t1

(3.49)

Chapter 3. Particle Kinetics

72

Therefore we can state that the initial linear momentum of the aggregate particle plus the
external impulses acting on the system of particles from t1 to t2 is equal to the aggregate
particle's final linear momentum. Since in reality all particles must have finite size to possess
mass, the above equation justifies application of the principle of linear impulse and
momentum to a rigid body represented as a single particle.
When the sum of the external impulses acting on a system of particles zero, Eq. (3.47)
reduces to a simplified form, namely,
mi (vi )1 mi (vi )2

(3.50)

This equation is referred to as the conservation of linear momentum. It states that the linear
momenta for a system of particles remain constant during the time period t1 to t2. Substituting
Eq. (3.48) into Eq. (3.50), we can also write
m(vG )1 m(vG )2

(3.51)

which indicates that the velocity vG of the mass center for the system of particles does not
change when no external impulses are applied to the system.
The conservation of linear momentum is often applied when particles collide or interact. For
application, a careful study of the free-body diagram for the entire system of particles should
he made in order to identify the forces which create either external or internal impulses and
thereby determine in what direction(s) linear momentum is conserved. As stated earlier, the
internal impulses for the system will always cancel out, since they occur in equal but opposite
collinear pairs. If the time period over which the motion is studied is very short, some of the
external impulses may also be neglected or considered approximately equal to zero. The
forces causing these negligible impulses are called nonimpulsive tierces. By comparison,
forces which are very large and act for a very short period of time produce a significant
change in momentum and are called impulsive forces. They, of course, cannot be neglected in
the impulse-momentum analysis.
Impulsive forces normally occur due to an explosion or the striking of one body against
another, whereas nonimpulsive forces may include the weight of a body, the force imparted by
a slightly deformed spring having a relatively small stiffness, or for that matter, any force that
is very small compared to other larger (impulsive) forces. When making this distinction
between impulsive and nonimpulsive forces. It is important to realize that this only applies
during the time t1 to t2. To illustrate, consider the effect of striking a tennis ball with a racket
as shown in Fig. 3.24. During the very short time of interaction, the force of the racket on the
ball is impulsive since it changes the ball's momentum drastically. By comparison, the ball's

Chapter 3. Particle Kinetics

73

weight will have a negligible effect on the change in momentum, and therefore it is
nonimpulsive. Consequently, it can be neglected from an impulse-momentum analysis during
this time. If an impulse-momentum analysis is considered during the much longer time of
flight after the racket-ball interaction, then the impulse of the ball's weight is important since
it, along with air resistance, causes the change in the momentum of the ball.

Fig. 3.24
EXAMPLE 3.12
The 15-Mg boxcar A is coasting at 1.5 m/s on the horizontal track when it encounters a 12-Mg
tank car B coasting at 0.75 m/s toward it as shown in Fig. 3.25a. If the cars meet and couple
together, determine
(a) the speed of both cars just after the coupling
(b) the average force between them if the coupling takes place in 0.8 s.
A
(a)

(b)

(c)
Fig. 3.25

Solution
Part (a)
Free-Body Diagram.
We consider both cars as a single system. We can easily draw the free-body diagram with
horizontal forces in the x direction, Fig. 3.25b. By inspection, momentum is conserved in this
direction since the coupling force F is internal to the system and will therefore cancel out. It is
assumed both cars, when coupled, move at v2 in the positive x direction.

Chapter 3. Particle Kinetics

74

Conservation of Linear Momentum


mA(vA)1+ mB(vB)1=( mA + mB )v2
(151100 kg)(1.5 m/s) - 12 000 kg(0.75 m/s) = (27 000 kg)v2,
v2 = 0.5 m/s
Part (b)
The average (impulsive) coupling force. Favg can be determined by applying the principle of
linear momentum to either one of the cars.
Free-Body Diagram.
As shown in Fig. 3.25c, by isolating the boxcar the coupling force is external to the car.
Principle of Impulse and Momentum
Since

Fdt F

avg

t Favg 0.8 we have


t2

mA (vA )1 Fdt mA (vA )2


t1

(15 000 kg)(1.5 m/s) Favg (0.8 s) = (15 000 kg)(0.5 m/s)
Favg = 18.8 kN
Solution was possible here since the boxcar's final velocity was obtained in Part (a). Try
solving for Favg by applying the principle of impulse and momentum to the tank car.

3.3.2 Principle of Angular Impulse and Momentum


In addition to the equations of linear impulse and linear momentum there exists a parallel set
of equations for angular impulse and angular momentum. First we define the terms angular
momentum and angular impulse.
The angular momentum of a particle about point O is defined as the moment of the particle's
linear momentum about O. Since this concept is analogous to finding the moment of a force
about a point, the angular momentum HO is sometimes referred to as the moment of
momentum.

Chapter 3. Particle Kinetics

75

Fig. 3.26
Similar to the scalar definition the moment of a force, if a particle is moving along a curve
lying in the x-y plane. Fig. 3.26, the angular momentum at any instant can be determined
about point O (actually the z axis) by using a scalar formulation. The magnitude of HO is
(HO)z = d. mv

(3.52)

Here d is the moment arm or perpendicular distance from O to the line of action of mv.
Common units for (HO)z are kgm2/s.The direction of HO is defined by the right-hand rule. As
shown, the curl of the fingers of the right hand indicates the sense of rotation of mv about O,
so that in this case the thumb (or HO ) is directed perpendicular to the x-y plane along the +z
axis.

x
Fig. 3.27
In general, if the particle is moving along a space curve, Fig. 3.27 the vector cross product can
be used to determine the angular momentum about O. In this case
HO = r x mv

(3.53)

Here r denotes a position vector drawn from point O to the particle P. As shown in the Fig.332, Ho is perpendicular to the plane containing r and mv. In order to evaluate the cross
product r and mv should be expressed in terms of their Cartesian components, so that the
angular momentum is determined by evaluating the determinant:
i
j
k
HO x
y
z
mvx mvy mvz

(3.54)

Hx= m(vz y - vy z), Hy = m(vx z - vz x), Hz = m(vy x - vx y)

(3.55)

or that

Chapter 3. Particle Kinetics

76

Now with the introduction of the angular momentum it will be shown that the equation of
motion of the particle can be written in another way. We can use Newton's second law of
motion F = ma to find the relation of the moments about point O of all the forces acting on
the particle and the particle's angular momentum. Hence, assume that the mass of the particle
is constant, we may perform a cross-product multiplication of each side of Newton's second
law of motion by the position vector r, which is measured in the x, y and z inertial frame of
reference. It yields
&
r F r mv
or
&
MO r& mv r mv

d
(r mv)
dt

The first term on the right side, r& mv=0, since the cross product of a vector (v) with itself
is zero. Therefore, the above equation becomes
&
MO H
O

(3.56)

The equation states that the resultant moment about point O of all the forces acting on the
particle is equal to the time rate of change of' the particle's angular momentum about point O.
This result is similar to equation
&
F L

(3.57)

Here L = mv, so the resultant force acting on the particle is equal to the time rate of change
of the particle's linear momentum.
It is clear from the derivations that Eqs. (3.56) and (3.57) are actually another way of stating
Newton's second law of motion. In other parts of this course it will be shown that these
equations have many practical applications when extended and applied to the solution of
problems involving either a system of particles or a rigid body.

Fig. 3.28
Now we will extend this result to the system of particles shown in Fig. 3.28. An equation
having the same form as Eq. (3.56) may be derived for this system of particles The forces

Chapter 3. Particle Kinetics

77

acting on the arbitrary ith particle of the system consist of a resultant external force Fi and a
resultant internal force fi. Expressing the moments of these forces about point O, using the
form of Eq. (3.56) we have
&)
(ri Fi ) (ri fi ) (H
i O
Here ri is the position vector drawn from the origin O of an inertial frame of reference to the
& ) is the time rate of change in the angular momentum of the ith particle
ith particle, and (H
i O
about O. Similar equations can be written for each of the other particles of the system. When
the results are summed vectorially, the result is
&)
(ri Fi ) (ri fi ) (H
i O
The second term is zero since the internal forces occur in equal but opposite collinear pairs,
and hence the moment of each pair about point O is zero. Dropping the index notation, the
above equation can be written in a simplified form as
&
MO H
O

(3.58)

which states that the sum of the moments about point O of all the external forces acting on the
system of particles is equal to the time rate of change of the total angular momentum of the
system about point O. Note that here O has been chosen as the origin of coordinates, however,
it actually can represent any fixed point in the inertial frame of reference.
EXAMPLE 3.13
The box shown in Fig. 3.29a has a mass to and is traveling down the smooth circular ramp
such that when it is at the angle it has a speed v. Determine its angular momentum about
point O at this instant and the rate of increase in its speed, i.e., at.

(a)

(b)
Fig. 3.29

Solution
Since v is tangent to the path. applying Eq. (3-53) the angular momentum is
HO = r m v

Chapter 3. Particle Kinetics

78

The rate of increase in its speed (dv/dt) can be found by applying Eq. (3.56). One can draw
easily the free-body diagram of the block Fig. 3.29b and it is clear that only the weight
W = mg contributes a moment about point O. We have
mg(r sin) =

d
(rmv)
dt

Since r and m are constant


mgr sin= rm

dv
dt

dv
g sin
dt
Note that this same result can, of course, be obtained from the equation of motion applied in
the tangential direction, Fig. 3.29b.
Angular Impulse and Momentum Principles. Now we can state the principle of angular
impulse and momentum. Eq. (3.58) gives the instantaneous relation between the moment and
the time rate of change of angular momentum. To obtain the effect of the moment MO on
the angular momentum of the particle

over a finite period of time Eq. (3.58) may be

integrated, assuming that at time t = t1, HO = (HO)1 and at time t = t2, HO = (HO)2
t2

MOdt (HO ) 2 (HO )1


t1

or
t2

(HO )1 MOdt (HO ) 2

(3.59)

t1

This equation is referred to as the principle of angular impulse and momentum. The initial and
final angular momenta (HO)1 and (HO)2 are defined as the moment of the linear momentum of
the particle, i.e., HO = r x mv at the instants t1 and t2 respectively. The second term on the left
t2

side, MOdt , is called the angular impulse. It is determined by integrating, with respect to
t1

time, the moments of all the forces acting on the particle over the time period t1 to t2, Since the
moment of a force about point O is MO = r x F , the angular impulse may be expressed in
t2

vector form as

(r F)dt

where r is a position vector which extends from point O to any

t1

point on the line of action of F.

Chapter 3. Particle Kinetics

79

For a system of particles, the principle of angular impulse and momentum may be written as
t2

(HO )1 MOdt (HO ) 2

(3.60)

t1

Here the first and third terms represent the angular momenta of the system of particles, i.e.,
HO = (ri x mivi) , at the instants t1 and t2, respectively. The second term is the sum of the
angular impulses given to all the particles from t1 to t2. Keep in mind that these impulses are
created only by the moments of the external forces acting on the system where, for the ith
particle MO = ri x Fi.
It is worthy to note that using impulse and momentum principles it is possible to write two
equations which define the particle's motion. These equation are vectoriall, hence, in general,
they may be expressed in x, y and z component form, yielding a total of six independent scalar
equations. If the particle is confined to move in the x-y plane, three independent scalar
equations may be written to express the motion, namely,
t2

m(vx )1 Fxdt m(vx ) 2


t1

t2

m(vy )1 Fydt m(vy ) 2

(3.61)

t1

t2

(HO )1 M Odt (HO ) 2


t1

The first two of these equations represent the principle of linear impulse and momentum in
the x and y directions and the third equation represents the principle of angular impulse and
momentum about the z axis.
When thee angular impulses acting on a particle are all zero during the time t, Eq. (3.59)
reduces to the following simplified form
(HO)1 = (HO)2

(3.62)

This equation is known as the conservation of angular momentum. It states that front t1 to t2
the particle's angular momentum remains constant. Obviously, if no external impulse is
applied to the particle, both linear and angular momentum will be conserved. In some cases,
however, the particle's angular momentum will be conserved and linear momentum may not.
An example of this occurs when the particle is subjected only to a central force.
We can also write the conservation of angular momentum for a system of particles, namely,
(HO)1 = (HO)2

(3.63)

In this case the summation must include the angular momenta of all particles in the system.

Chapter 3. Particle Kinetics

80

Procedure for Analysis


When applying the principles of angular impulse and momentum, or the conservation of
angular momentum it is suggested that the following procedure be used.
Free Body Diagram

Draw the particle's free-body diagram in order to determine any axis about which
angular momentum may be conserved. For this to occur, the moments of the forces (or
impulses) must be parallel or pass through the axis so as to create zero moment
throughout the time period t1 to t2.

The direction and sense of the particle's initial and final velocities should also be
established.

Momentum Equations

Apply the principle of angular impulse and momentum Eq. (3.59) or if appropriate, the
conservation of angular momentum, Eq. (3.62).

EXAMPLE 3.14
The 2-kg disk shown in Fig. 3.30a rests on a smooth horizontal surface and is attached to an
elastic cord that has a stiffness kC = 20 N/m and is initially unstretched. If the disk is given a
velocity (vD)1 = 1.5 m/s. perpendicular to the cord, determine the rate at which the cord is
being stretched and the speed of the disk at the instant the cord is stretched 0.2 m.
Solution
Free-Body Diagram
After the disk has been launched, it slides along
the path shown in Fig. 3.30b. By

inspection,

angular momentum about point O (or the z axis)


is conserved, since none of the forces produce an
angular impulse about this axis. Also, when the

(a)

distance is 0.7 m, only the component (vD' ) 2


produces angular momentum of the disk about O.
Conservation of Angular Momentum
The component (vD' ) 2 can be obtained by
applying the conservation of angular momentum
about O (the z axis), i.e.,

(b)

(HO)1 = (HO)2
r1mD (vD' )1 = r2mD (vD' ) 2

Fig. 3.30

Chapter 3. Particle Kinetics

81

0.5 m(2 kg)(1.5 m/s) = 0.7 m(2 kg) (vD' ) 2


(vD' ) 2 = 1.07 m/s
Conservation of Energy.
The speed of the disk may be obtained by applying the conservation of energy equation at the
point where the disk was launched and at the point where the cord is stretched 0.2 m.
T1+V1 = T2+V2
1
1
1
(2 kg)(1.5 m/s)2 + 0 =
(2 kg) (vD )22 + (20 N/m)(0.2 m)2
2
2
2
Hence
(vD )2 = 1.36 m/s
Having determined (vD )2 and its component (vD' ) 2 the rate of stretch of the cord (vD" ) 2 is
determined from the Pythagorean theorem,
(vD" )2 (vD )22 (vD' )22
We obtain
(vD" )2 (1.36)2 (1.07)2 = 0.838 m/s.

You might also like