Principles of Mechanics - Fundamental University Physics
Principles of Mechanics - Fundamental University Physics
Series Editor
Mourad Amer
Enrichment and Knowledge Exchange, International Experts for
Research, Cairo, Egypt
Editorial Board
Anna Laura Pisello
Department of Engineering, University of Perugia, Italy
Dean Hawkes
Cardiff University, UK
Hocine Bougdah
University for the Creative Arts, Farnham, UK
Federica Rosso
Sapienza University of Rome, Rome, Italy
Hassan Abdalla
University of East London, London, UK
So ia-Natalia Boemi
Aristotle University of Thessaloniki, Greece
Nabil Mohareb
Beirut Arab University, Beirut, Lebanon
Saleh Mesbah Elkaffas
Arab Academy for Science, Technology, Egypt
Emmanuel Bozonnet
University of la Rochelle, La Rochelle, France
Gloria Pignatta
University of Perugia, Italy
Yasser Mahgoub
Qatar University, Qatar
Luciano De Bonis
University of Molise, Italy
Stella Kostopoulou
Regional and Tourism Development, University of Thessaloniki,
Thessaloniki, Greece
Biswajeet Pradhan
Faculty of Engineering and IT, University of Technology Sydney, Sydney,
Australia
Chaham Alalouch
Sultan Qaboos University, Muscat, Oman
Iman O. Gawad
Helwan University, Egypt
Principles of Mechanics
Fundamental University Physics
Salma Alrasheed
Thuwal, Saudi Arabia
The publisher, the authors and the editors are safe to assume that the
advice and information in this book are believed to be true and accurate
at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the
material contained herein or for any errors or omissions that may have
been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional af iliations.
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
1.1 Introduction
Physics is an exciting adventure that is concerned with unraveling the
secrets of nature based on observations and measurements and also on
intuition and imagination. Its beauty lies in having few fundamental
principles being able to reach out to incorporate many phenomena from the
atomic to the cosmic scale. It is a science that depends heavily on
mathematics to prove and express theories and laws and is considered to be
the most fundamental of physical sciences. Astronomy, geology, and
chemistry all involve applications of physics’ principles and concepts.
Physics doesn’t only provide theories, but it also provides techniques that
are used in every area of life. Modern physical techniques were the major
contributors to the wealth of mankind’s knowledge in the past century.
A simple law in physics can be used to explain a wide range of complex
phenomena that may appear to be not related. When studying a complex
physical system, a simpli ied model of the system is usually used, where the
minor effects are neglected and the main features of the system are
concentrated upon. For example, when dealing with an object falling near
the earth’s surface, air resistance can be neglected. In addition, the earth is
usually assumed to be spherical and homogeneous. However, in reality, the
earth is an ellipsoid and is not homogeneous. The difference between the
calculations of these different models can be assumed to be insigni icant.
Physics can be divided into two branches namely: classical physics and
modern physics. This book focuses on mechanics, which is a branch of
classical physics. Other branches of classical physics are: light and optics,
sound, electromagnetism, and thermodynamics. Mechanics is the science of
motion of objects and is the core of classical physics. On the other hand,
modern branches of physics include theories that have been developed
during the past twentieth century. Two main theories are the theory of
relativity and the theory of quantum mechanics. Modern physics explains
many physical phenomena that cannot be explained by classical physics.
zepto z
atto a
femto f
pico p
nano n
micro
milli m
centi c
deci d
deka da
Factor Pre ix Symbol
hecto h
kilo k
mega M
giga G
tera T
peta P
exa E
zetta Z
slug
yr
Solution 1.1
Example 1.2 If a volume of a room is , what is the volume in cubic
inches?
Solution 1.2
of acceleration is
Solution 1.3
Each term in the equation has the same dimension and therefore it is
dimensionally correct.
1.5 Vectors
When exploring physical quantities in nature, it is found that some
quantities can be completely described by giving a number along with its
unit, such as the mass of an object or the time between two events. These
quantities are called scalar quantities. It is also found that other quantities
are fully described by giving a number along with its unit in addition to a
speci ied direction, such as the force on an object. These quantities are
called vector quantities.
Scalar quantities have magnitude but don’t have a direction and obey
the rules of ordinary arithmetic. Some examples are mass, volume,
temperature, energy, pressure, and time intervals by a letter such as m, t, E
, etc. Vector quantities have both magnitude and direction and obey the
rules of vector algebra. Examples are displacement, force, velocity, and
acceleration. Analytically, a vector is speci ied by a bold face letter such as
. This notation (as used in this book) is usually used in printed material.
In handwriting, the designation is used. The magnitude of is written
as or A in print or as in handwriting.
Fig. 1.2 The two vectors and are said to be equal ( = ) only if they have the same
magnitude and direction
1.6.2 Addition
There are two ways to add vectors, geometrically and algebraically. Here, we
will discuss the geometric method which is useful for solving problems
without using a coordinate system. The algebraic method will be discussed
later. To add two vectors and using the geometric method, place the
head of at the tail of and draw a vector from the tail of to the head
of as shown in Fig. 1.3. This method is known as the triangle method. An
extension to sum up more than two vectors is shown in Fig. 1.4. An
alternative procedure of vector addition using the geometric method is
shown in Fig. 1.5. This is known as the parallelogram method, where is
the diagonal of a parallelogram with sides A and B. To ind analytically,
Fig. 1.6 shows that
(1.1)
and that
Fig. 1.4 Geometric method for summing more than two vectors
or
The direction of is
Note that only when and are parallel, the magnitude of the resultant
vector is equal to (unlike the addition of scalar quantities, the
magnitude of the resultant vector is not necessarily equal to ).
Example 1.4 A jogger runs from her home a distance of 0.5 km due south
and then 1 km to the west. Find the magnitude and direction of her
resultant displacement.
Solution 1.4
From Fig. 1.7, we can see that the magnitude of the resultant displacement
is given by
The direction of is
south of west.
Fig. 1.8 The negative vector of is a vector of the same magnitude of but in the opposite direction
(Distributive law).
, where q is a scalar.
The area of a parallelogram that has sides A and as shown in
Fig. 1.15.
Fig. 1.15 The magnitude of the vector product is the area of a parallelogram with sides
and
1.8.2.1 Addition
The resultant vector is given by
with a direction
in three dimensions
the magnitude of is
Fig. 1.24 The displacements are drawn to scale with the head of placed at the tail of and the
head of placed at the tail of .The resultant vector is the vector that extends from the tail of
to the head of
Solution 1.5
Graphically, in Fig. 1.24 the displacements are drawn to scale with the head
of placed at the tail of and the head of placed at the tail of .The
resultant vector is the vector that extends from the tail of to the head
of . By using graph paper and a protractor, the magnitude of is
measured to have the value of 34.8 km and a direction of from the
positive axis. Analytically, from Fig. 1.24, we have
north of east.
1.8.2.2 Subtraction
The magnitude and direction of are as in the case of addition except that
the plus sign is replaced by the minus sign.
Solution 1.6
or
Fig. 1.25 If we write the unit vectors around a circle, then reading counter clockwise gives the positive
products and reading clockwise gives the negative products
If we write the unit vectors around a circle as shown in Fig. 1.25, then
reading counterclockwise gives the positive products and reading clockwise
gives the negative products. Note that these results are for a right-handed
coordinate system. We have
using the distributive law and the above relations of unit vectors we get
A determinant of order 3 is
Note that this is not a determinant since the elements in the irst row are
vectors and not scalars, but it is a convenient way to represent the cross
product.
Solution 1.7
(a)
(b)
(c)
Solution 1.8
By the de inition of the unit vector, we have
Solution 1.9
(a)
(b)
(c)
Example 1.10 Using vectors method, ind the area of a triangle if the
coordinates of its three vertices are , , .
Solution 1.10
Area
1.8.2.7 Triple Product
Scalar Triple Product
The triple scalar product is a scalar quantity de ined as .
This quantity can be represented by a determinant that involves the
components of the vectors,
Fig. 1.26 The triple scalar product is equal to the volume of a parallepiped with sides , and
where , and
Furthermore, the triple scalar product is equal to the
volume of a parallepiped with sides , and as shown in Fig. 1.26.
Because any edges can be used, the triple scalar product can be written as
or as . These products are positive and negative for
a right-handed coordinate system respectively. Therefore, there are 6 equal
triple scalar products or 12 if you include the terms of the form
. A. Three of these six products are positive and the rest are negative. By
expanding the determinant, you can prove that
Solution 1.11
Then
1.9.1 Some Rules
If and are vector functions and is a scalar function then
Solution 1.12
(a)
At
(b)
At
1.9.2.1 Del
The vector differential operator del is de ined as
1.9.2.2 Gradient
1.9.2.3 Divergence
is called the divergence of (written div ).
1.9.2.4 Curl
(c) .
Solution 1.13
(a)
at ,
(b)
at ,
(c)
at ,
If , then
one and it will be used throughout this book. Suppose the position vector of
any point on the curve (see Fig. 1.27) that extends from
at to at is given by
where t is a scalar variable, and suppose that
is a vector ield, then the line integral of
is given by
(1.2)
Solution 1.14
(a)
(b)
Hence
Example 1.15 A vector is given by . Compute
(a) The straight lines from (0, 0) to (0, 1) and then to (1, 1).
(b) Along the straight line (c) Along the curve
Solution 1.15
(a) Along the straight line from (0,0) to (0,1) we have , and
therefore
Solution 1.16
By using the polar coordinates, we have and (since
) , and , also , therefore
we have
Fig. 1.28 The line integral along the curve using polar coordinates
Problems
1.
Check if the relation is dimensionally correct, where
v represents the escape speed of a body, and are the mass and
radius of the earth, respectively, and G is the universal gravitational
constant.
2.
If the speed of a car is 180 , ind its speed in
3.
How many micrometers are there in an area of 3
4.
Figure 1.29 shows vectors , and D. Find graphically the
following vectors (a) show that
.
5.
A car travels a distance of 1 km due east and then a distance of 0.5 km
north of east. Find the magnitude and direction of the resultant
displacement of the car using the algebraic method.
6. Prove that .
7.
A parallelogram has sides and . Prove that its area is equal to
8.
If and , ind (a) (b)
(c) (d) the length of and the length of (e) the angle between
and (f) the scalar projection of on and the scalar projection
of on .
9.
Show that is perpendicular to if
10.
Given that and ,
determine which vectors are perpendicular and which are parallel.
11.
Use the vectors and to prove
that
12.
If and
, ind at (−1,1,1)(a) (b) (c)
.
13.
Evaluate where and
14.
If , show that
15.
A force ield is given by , ind (a) (b) a scalar
ield such that Calculate the line integral along the
straight lines from (0, 0) to (1, 0) to (1, 1) and from (0, 0) to (0, 1) to
(1, 1). Is the line integral independent of path?
Fig. 1.29 Vectors and
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not included in
the chapter's Creative Commons license and your intended use is not permitted by statutory regulation
or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_2
2. Kinematics
Salma Alrasheed1
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
2.1 Introduction
Mechanics is the science that studies the motion of objects and can be divided
into the following:
1.
Kinematics: Describes how objects move in terms of space and time.
2.
Dynamics: Describes the cause of the object’s motion.
3.
Statics: Deals with the conditions under which an object subjected to
various forces is in equilibrium.
This chapter is considered with kinematics which answers many
questions such as: How long it takes for an apple to reach the ground when it
falls from a tree? What is the maximum height reached by a baseball when
thrown into air? What is the distance it takes an airplane to take off?
In physics, there are three types of motion: translational, rotational, and
vibrational. A block sliding on a surface is in translational motion, (Merry-
go-Round) is an example of rotational motion, and a mass–spring system
when stretched and released is in vibrational motion. From here until Chap. 7,
the object studied will be treated as a particle (i.e., a point mass with no size).
This assumption is possible only if the object moves in translational motion
without rotating and by neglecting any internal motions that might exist in
the object.
That is, an object can be treated as a particle only if all of its parts move in
exactly the same way.
For example, if a man jumps into a pool without rotating by doing a
somersault (freezing his body), he can be treated as a particle since all
particles in his body will move in exactly the same way. Another example of an
object that can be treated as a particle is the Earth in its motion about the
Sun. Since the dimensions of the Earth are small compared to the dimensions
of its path, it can be considered as a particle. The motion of an object is
described either by equations or by graphs. Both ways provide information
about the motion; however, equations provide precise information while
graphs give greater insight about the motion.
2.2.1 Displacement
Consider a car that is treated as a particle moving along the straight-line path
shown in Fig. 2.1. The -axis of a coordinate system is used to describe the
position of the car with respect to the origin , where the points and
correspond to the positions at and at , respectively. The position–
time graph of this motion is shown in Fig. 2.2. The displacement of the truck is
a vector quantity de ined as the change in its position during the time interval
from to and is given by
Hence displacement is a quantity that depends only on the initial and inal
positions of the object. The direction of the displacement in one dimension is
speci ied by a plus or minus sign. It is positive if the particle is moving in the
positive direction and negative if the particle is moving in the negative
direction. In two or three dimensions, the displacement is represented by a
vector. The SI unit of the displacement is the meter (m).
Fig. 2.1 A car that is treated as a particle moving along the straight-line path
2.2.3 Velocity
The average velocity of an object is a vector quantity de ined in terms of
displacement rather than the total distance traveled:
is positive if the motion is in the positive -direction and negative if it is in
the negative -direction. On the position–time graph in Fig. 2.2, is the slope
of the straight line connecting the points and Q. The average velocity helps
in describing the overall motion of the particle in a certain time interval. To
describe the motion in more detail, the instantaneous velocity is de ined. This
velocity corresponds to the velocity of a particle at a particular time. That
involves allowing to approach zero:
Fig. 2.3 Geometrically, the instantaneous velocity of a particle at a particular time on the position-time
curve is the slope (the tangent) to the position-time curve at that point or instance
2.2.4 Speed
The speed of the particle is de ined as the magnitude of its velocity. Note that
speed and average speed are different since speed is de ined in terms of
displacement, whereas average speed is de ined in terms of the total distance
traveled.
2.2.5 Acceleration
If the particle’s velocity changes with time, it is said to be accelerating. The
average acceleration of the particle is de ined as the ratio of the change of
its velocity to the time interval :
as
The average acceleration is the slope of the line joining the points and
on the velocity–time graph, whereas the instantaneous acceleration is the
slope of the curve at a particular point (see Fig. 2.4). Figure 2.5 shows the
position, velocity, and acceleration for a particle simultaneously.
Fig. 2.4 The average acceleration is the slope of the line joining the points and on the velocity-time
graph, whereas the instantaneous acceleration is the slope of the curve at a particular point
Fig. 2.5 This igure shows the position, velocity and acceleration as a function of time of a particle moving
in one direction. The particle starts from rest, accelerates to a certain speed, is maintained at that speed for
some time, then it decelerates back to rest
Fig. 2.6 A car moving along the curved path where it is located at km at , and at
km at hr
Example 2.1 A car travels along the path shown in Fig. 2.6, where it is
located at km at , and at km at h. Find the
displacement, average velocity, and average speed of the car during this time
interval if the total distance traveled is 20 km.
Solution 2.1
The displacement of the car is
Fig. 2.7 A particle moves along the -axis according to the expression
Example 2.2 A particle moves along the -axis according to the expression
plot of this equation is shown in Fig. 2.7. the
displacement and average velocity of the particle during the time interval
between and the instantaneous velocity of the particle as
a function of time and at and
Solution 2.2
(a)
at , and at
Example 2.3 A particle is moving along the -axis. The position–time graph
of its motion is shown in Fig. 2.8. Find: (a) the average velocity between a and
the instantaneous velocity at the points and
Solution 2.3
(a)
(b)
Fig. 2.8 The position-time graph of a particle moving along the x-axis
Solution 2.4
(a)
(2.1)
(b) When the object reaches its maximum speed and hence
where ds is the in initesimal arc length along the path and comes from the
fact that as approaches zero, the distance traveled by the particle along the
path becomes equal to the vector displacement . Figure 2.10 shows the
instantaneous velocities along the path. The average acceleration is
In terms of components
Fig. 2.11 If is changed in magnitude only (motion along a straight line) then is parallel to if is
increasing, and antiparallel if is decreasing. If is changed in direction only (motion along a curved
Fig. 2.12 If is changed in both magnitude and direction then will be directed at some angle to
Fig. 2.13 The parallel (or tangential) component of the acceleration is always tangent to the path while
the perpendicular (or normal) component is normal to the path at each point
Fig. 2.14 At A the acceleration of a car is in the same direction of the velocity since the latter changes only
in magnitude. As it moves its velocity is changed in both magnitude and direction. Therefore at the
direction of the acceleration is at some angle to the velocity. At the speed reaches a maximum and
therefore the instantaneous change of speed is zero at this point and the acceleration has only a
perpendicular component. As the car moves up its velocity decreases and changes in direction also, thus
the acceleration has both parallel and perpendicular components. Finally at , the acceleration is in the
opposite direction of the velocity since the velocity is decreasing but its direction is the same
Fig. 2.15 The TNB frame moves with the particle
or
(2.3)
Example 2.5
A particle is moving in space according to the expression
Solution 2.5
Hence
Fig. 2.16 A car moving with a constant tangential acceleration down a ramp
Solution 2.6
Since the tangential acceleration of the car is constant, it can be found from
acceleration is
(2.4)
(2.5)
Furthermore,
(2.6)
Finally,
(2.7)
Equations 2.4, 2.5, 2.6, and 2.7 are called the kinematic equations for motion
in a straight line under constant acceleration. The motion graphs for an object
moving with constant acceleration in the positive -direction are shown in
Fig. 2.17.
Fig. 2.17 The motion graphs for an object moving with constant acceleration in the positive -direction
Example 2.7 A train accelerates uniformly from rest and travels a distance of
200 in the irst 8 s. Determine: (a) the acceleration of the train; (b) the
time it takes the train to reach a velocity of 70 , the distance traveled
during that time; (d) the velocity of the train 5 s later from the time calculated
in (b).
Solution 2.7
(a)
Since , we have
(b)
and therefore
(c)
(d)
before taking off. If it is to take off at a speed of 100 how much time
is required for it to take off; (b) what distance will it have traveled before
taking off?
Solution 2.8
(a)
car to catch the other car? (b) What is the distance traveled by both during
that time? (c) How much time has passed from where the car passed the
police car to where it was caught?
Solution 2.9
Let’s assume that at where the car passed the police car and that
at the instant the police car begins to accelerate. The velocity of the car is
equal to 38.9 , and the initial velocity and acceleration of the police car
are 22.2 and 1.1 , respectively The police will catch the car when
both their displacements from are equal. (a) From the expression
, the displacement of the car at any time is
That gives
(b)
(c)
Fig. 2.18 The displacement and velocity graph for a free-falling object
Fig. 2.19 The important features of a free falling object that is dropped from rest
Solution 2.10
(a) First we take at the position where the ball is thrown and positive y
to be upwards. At the maximum height the velocity of the ball is zero,
and
and
Solution 2.11
(a) Taking and at we have
at
(b)
(c)
Solution 2.12
The time it takes the irst ball to hit the ground is found from
The second ball must fall the same distance during a time of
and therefore
2.4.3 Motion in Two Dimensions with Constant Acceleration
The position vector can be written as
(2.9)
(2.10)
(2.11)
(2.12)
(2.13)
Example 2.13 If the motion of a particle in a plane is described by
and plot the component of the
Solution 2.13
(a)The -component of position is
since at , then
The plot of y against t is shown in Fig. 2.20.
(b) The -components of velocity and acceleration is
or
and
Therefore, the acceleration of the particle is constant at any time and is given
by
(2.15)
(2.16)
(2.17)
Combining and eliminating t from Eqs. 2.16 and 2.17 we ind that
This equation which is of the form ax–bx2 (a and b are constants), is the
equation of a parabola. Therefore, when air resistance is neglected (when
using the simpli ied model of the system), the trajectory of the projectile is
always a parabola. At any instant, the velocity of the object is tangent to its
trajectory Its magnitude and direction with respect to the positive -
direction are given by
and
Solution 2.14
(a) The maximum height reached by the ball is
(d) The -component of the velocity of the ball just before it hits the ground
is
The -component is
Solution 2.15
Let the origin of the reference frame be the releasing point. Since we
have
and
Hence, when the ball reaches the ground, the elapsed time is
and
The subscript rad is for radial. Thus, this radial or centripetal acceleration
is always directed toward the center of the circle. Therefore, the
directions of and a change continuously with time but their magnitudes are
constant (see Fig. 2.23). The time required for the particle to complete one
revolution around the circle is called the period of revolution and is given by
Thus
Fig. 2.23 The directions of and a change continuously with time but their magnitudes are constant
Example 2.16 In a fun fair ride, the passengers rotate in a circle with a
constant speed of 3 . If the period of revolution is 1.5 , ind the total
acceleration of the passenger.
Solution 2.16
Since the speed of the passenger is constant, it follows that the passenger’s
total acceleration is just the centripetal acceleration given by
In Chap. 8, the concepts of angular velocity and acceleration and their vector
relationship with the normal and tangential accelerations are introduced.
Figure 2.24 shows the velocity and total acceleration vectors of a particle
moving in a circular path with increasing speed (clockwise) until it reaches
the maximum speed at the bottom, and then slows down as it goes back up.
An example of this motion is in a roller coaster ride in a vertical circle.
Fig. 2.24 The velocity and total acceleration vectors of a particle moving in a circular path with
increasing speed (clockwise) until it reaches the maximum speed at the bottom, and then slows down as it
goes back up. An example of this motion is in a roller coaster ride in a vertical circle
Solution 2.17
Since both the direction and the magnitude of the car’s velocity change, its
total acceleration is the vector sum of its tangential and radial accelerations.
The tangential acceleration is
(2.18)
or
We will extend this to three dimensions in the case where the velocity of
with respect to is constant in both magnitude and direction (see
Fig. 2.26). The position vector of the particle relative to is given by
(2.19)
Fig. 2.26 The velocity of with respect to is constant in both magnitude and direction
Example 2.18 Two motor cyclists and are driving along the same road
(See Fig. 2.27) with speeds 90 and 50 , respectively. Determine:
(a) the velocity of motorcyclist A relative to and of relative to and
(b) if the two motor cyclists approach each other along two parallel roads,
(See Fig. 2.28), A moving at 80 , and moving at 60 , what is the
velocity of motorcyclist A relative to and of relative to A.
Fig. 2.27 Two motor cyclists and driving with speeds 90 and 50 respectively
Solution 2.18 Using the above discussion, consider as the Earth’s frame
of reference denoted E, as the frame of reference of motorcyclist B and the
point P as the motor cyclist A
(a) The velocity of A relative to is found from
(b)
Solution 2.19
Using Fig. 2.26, consider the Earth as S (denoted E), the waves as , and the
boat as the point P. As we can see from Fig. 2.29, the velocity of the boat
relative to the earth is given by , where and are the
velocities of the boat relative to the waves and the velocity of the waves
relative to the earth respectively With the east as the direction of the positive
-axis we get
Hence
The direction of is
Fig. 2.29 A boat is traveling at 8 north relative to the sea’s waves, and the waves are traveling
and
Unlike the rectangular unit vectors, the polar unit vectors are not ixed in
direction. Their direction changes as the particle moves along some path.
Therefore, when inding the velocity and acceleration of a particle the
derivatives of the polar unit vectors must be considered. The position vector
of the particle is given by
To ind the velocity in terms of the polar unit vectors let us differentiate
and with respect to time. That gives
Fig. 2.31 Unlike the rectangular unit vectors, the polar unit vectors are not ixed in direction. Their
direction changes as the particle moves along some path
(2.22)
or
where
and
and
Solution 2.20
At rad, and
. Also and
. Therefore
and
Fig. 2.32 An object moving in one dimension along the -axis
Fig. 2.33 The position-time graph of a particle moving along the -axis
Problems
1.
A sports car moves around a circular track of radius of 100 . If the car
makes one round in 75 , ind the car’s (a) average speed (b) average
velocity.
2.
An object is moving in one dimension along the -axis according to
Fig. 2.32. Describe the motion of the object.
3.
The position–time graph of a particle moving along the -axis is shown
in Fig. 2.33. Find (a) the average velocity between a and the
instantaneous velocity at , and
4. A motorist drives along a straight-line road. His speed varies with time
according to Fig. 2.34. Sketch the position versus time and acceleration
versus time graphs of the motorist.
5.
A particle moves along the curve de ined by and
Find the position, velocity and acceleration of the particle at any time.
6.
A car moves at constant speed of 40 along the road shown in
Fig. 2.35. If the radius of curvature at A is 350 and the total
acceleration of the car at is , ind (a) the total acceleration of the
reaches a speed of 3
8.
A stone is thrown downwards from a height of 10 . Find its initial
speed if it reaches the ground after l s.
9.
A block is thrown horizontally from the top of a cliff that is 30 high
with a speed of 10 . Find (a) the block’s magnitude of displacement
from the origin and its velocity after 1.5 the horizontal distance
from the releasing point to where the block hits the ground.(Hint: the
magnitude of displacement from the origin is ).
10.
A river has a uniform speed of 0.5 due east. If a boat travels east at a
speed of 3 relative to the water, ind the time it takes the boat to
travel a distance of 1100 km and return to its starting point.
11. An aircraft is tracked by a radar (see Fig. 2.36). If at a certain instant the
radar measurements give
, and . Find the velocity and acceleration of the airplane
at that instant.
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/),
which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons
license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's
Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the
permitted use, you will need to obtain permission directly from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_3
3. Newton’s Laws
Salma Alrasheed1
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
3.1 Introduction
In this chapter, dynamics which is a branch of mechanics will be
discussed. Dynamics is concerned with the cause behind the motion of
objects and answers questions such as: Why does a skydiver loat in
air? What makes an apple fall from a tree? Why a block connected to a
spring oscillates when the spring is stretched? We will ind that these
motions occur when objects interact with each other, i.e., the apple is
interacting with earth, the skydiver is interacting with air, and the block
is interacting with the spring.
That is, the acceleration of the particle measured from both inertial
reference frames and is the same. To show that Newton’s irst law
is only valid when applied with respect to an inertial frame of
reference; consider a girl named Mia that is at rest while watching her
friend Lea driving a car moving at constant velocity. Lea has her
seatbelt fastened and put her suitcase in the seat right next to her
without restraining it. Now, suppose that Lea steps on the brakes,
which would cause her vehicle to decelerate, her suitcase will start to
move forward. According to Lea, who is in an accelerated frame, the
suitcase moved from rest even though there was no apparent net
external force acting on it. Therefore, in Lea’s frame, Newton’s irst law
seems to be incorrect.
The situation, however, is different to Mia, who is in an inertial
frame of reference. In her perspective, the suitcase was initially moving
with constant velocity and the net force on it was zero. When the car
started to decelerate, the net force on the suitcase is still equal to zero
and thus the suitcase must continue to move forward with constant
velocity and stop by friction or impact with the inside of the car.
Therefore, it is apparent to Mia that Newton’s irst law is valid. From
the previous example, we conclude that Newton’s irst law (and in
general Newton’s laws) is not valid in all kinds of reference frames; it is
only valid when applied with respect to inertial frames. That is, Lea
must not apply Newton’s irst law in her reference frame.
The same situation would be observed by Lea if she were to turn
her car while moving. When the car turns, the suitcase will start to
move in the direction opposite to the turn. Once again, Lea observes
that the suitcase has moved from rest without any apparent force acting
on it which contradicts Newton’s irst law in her opinion. Mia sees no
contradiction with Newton’s irst law because when the car turns, the
suitcase tends to continue its initial uniform straight line motion, and
thus it moves toward the direction opposite to the turn. Therefore
Newton’s laws are obeyed by objects when observed from inertial
frames of reference (see Fig. 3.1).
To apply classical (Newtonian) mechanics with respect to a
noninertial reference frame, new forces named as pseudo forces are
introduced. In this book, only inertial frames are used, and all laws are
stated with respect to those frames. One convenient inertial frame of
reference, used throughout this book, is the surface of the earth. The
earth can be considered as an inertial frame since its motion about its
axis and about the sun has a small effect on calculations and thus can be
neglected.
Fig. 3.1 The boy is throwing the water out by pitching the bucket forward. If he stops, the water
will continue its motion along a straight line. However, because of the force of gravity, it follows a
parabolic path
3.2.3 Mass
As mentioned earlier, the tendency of an object to resist any change in
its motion (i.e., to remain at rest or maintain uniform motion along a
straight line) is called inertia. From experiments and everyday
experience, it is observed that a certain force produces different
accelerations when applied to different bodies. This variation in the
produced acceleration depends upon the quantity of matter contained
in the body Such quantity is known as the mass of the body Therefore,
mass is a measure of inertia. Objects with large masses have less
acceleration when exposed to the same force. Thus, mass is a quantity
that relates the acceleration of the body to the force acting on it. The SI
unit of mass is the kilogram (kg). Experimentally, it is found that the
ratio of the masses of any two bodies (say and ) is equal to the
inverse of the ratio of the magnitudes of their accelerations if both are
acted upon by the same force. That is, we have
The mass of any body can be found by comparing its acceleration to the
acceleration of a l kg mass when both bodies are acted upon by the
same force. This leads to the conclusion that mass is independent of
force; it is an inherent characteristic of matter. Furthermore, it has been
experimentally proved that when two masses and are attached
together, the combined body behaves as a single body of mass .
Thus, mass is a scalar quantity and obeys the rules of ordinary
arithmetic.
we have
and
system is the Newton (N). One Newton (1N) is de ined as the force that
gives a l kg mass an acceleration of 1 .
Example 3.1 A body is exposed to three forces acting in different
directions as shown in Fig. 3.2. Find the magnitude and direction of the
resultant force acting on the body and the corresponding acceleration.
Solution 3.1
The net force in the -direction is
Solution 3.2
(a) The acceleration of the container is given by
That gives
(b) If the force is at to the horizontal, the acceleration is
and
(c) In situation (a) we have
and
In (b) we have
and
Example 3.3 A particle of mass 0.5 kg is moving along the curve given
by where t is time. Determine the force acting on
the particle.
Solution 3.3
The force acting on the particle is
Example 3.4
A particle of mass of 1 kg is moving under the in luence of a force given
by
Solution 3.4
Fig. 3.4 The gravitational force exerted by the Earth on the apple and that exerted by the apple to
the Earth form an action-reaction pair
Solution 3.6
The free-body diagram of each block is shown in Fig. 3.5, where is
the force exerted on by . (a) Applying Newton’s second law for
the ,
(b) Applying Newton’s second law for each block, we have
and . From Newton’s law
of action and reaction we have , and , and
therefore
Fig. 3.5 Three blocks of masses , and are placed on a frictionless surface and pushed
by a horizontal force F
directed toward the center of earth. Using Newton’s second law, we can
calculate the force that caused this acceleration. If an object has a mass
m then the gravitational force is given by , and is denoted by , i.e.,
. is known as the weight of an object and is de ined as the
gravitational force exerted on it by earth (or any other astronomical
body, where g is different than that of earth). In Chap. 9, we will see that
a gravitational force exists between any two bodies. When one of the
bodies is an astronomical body, such as the earth or moon, and the
other body is relatively smaller in size and mass the gravitational force
is then called the weight of the body We will also see that the
gravitational force varies with the distance between objects, and that
the value of becomes less at greater altitudes. Thus, weight is not an
intrinsic property of an object. In everyday life, it is common to use the
word weight when measuring the mass of a body mass and weight
represent different quantities but they are proportional for a given
value of . For two masses at the same location, the ratio of their
weights is equal to the ratio of their masses.
3.3.3 Tension
The tension force is the force that a cord, rope, cable, or any other
similar object exerts on an object attached to it. This force is directed
along the rope away from the object at the point where the rope is
attached. In solving problems, ropes are usually assumed to be
massless (referred to as light ropes) and unstretchable. For any light
rope, the magnitude of the tension force T is the same at all points
along the rope.
3.3.4 Friction
Imagine that everything around you is coated with an extremely good
lubricant. Simple activities such as walking, sitting, driving a car, or
holding objects would become extremely dif icult or impossible.
Therefore, friction plays a very important role in our everyday life. The
frictional force is due to the interaction between the surface atoms of
any two bodies in contact. The direction of this force is always parallel
to the surface of contact, opposing the motion or the planned motion of
one object relative to the other. Hence, the normal and frictional forces
are both contact forces and they are always perpendicular to each
other.
Consider a block resting on a table. If the block is pushed with a
horizontal force and remains stationary, it is because that the applied
force is balanced by an equal and opposite force. This opposing force is
known as the statistical frictional force and it has the value .
The statistical frictional force increases with increasing (see Fig. 3.6).
The name statistical comes from the fact that the block remains
stationary.
Fig. 3.6 The opposing force is known as the statistical frictional force and it has the value
. Its maximum value represents the applied force when the block is at the
verge of slipping i.e. it is the minimum force necessary to initiate motion. When the block moves,
the retarding frictional force is then called the kinetic frictional force
Fig. 3.7 A graph of the friction force versus the applied force
1.
, where is the coef icient of static friction and n is
the normal force acting on the block. As long as the block is at rest
where
2.
where is the coef icient of the kinetic friction.
3. The directions of and are always parallel to the surface. is
opposite to the component of the applied force that is parallel to
the surface and is opposite to the instantaneous velocity of the
body relative to the surface.
The dimensionless coef icients and depend on the nature
of the surfaces in contact and are independent of the area of
p
Fig. 3.9 An object falling through air experiences a drag force and a gravitational force
Find: (a) the coef icient of kinetic friction between the box and the
surface; (b) the maximum angle the box would be at the verge of
slipping if the angle of the incline is changeable.
Solution 3.7
(a) The free-body diagram is shown in Fig. 3.10. Applying Newton’s
second law to the box gives
The coef icient of kinetic friction is
Solution 3.8
The free-body diagram of each mass is shown in Fig. 3.11. Applying
Newton’s second law to each block (taking positive y to be upwards)
gives
Fig. 3.12 The free-body diagram showing the forces on each block
Solution 3.9
(a) The acceleration value is the same for both masses since they are
connected by a string. Figure 3.12 shows the free-body diagram for
each mass. Applying Newton’s law to and in the direction of
motion we have for
and for
from this we have
and
and therefore
(b) If the system moves with constant speed, its acceleration is zero
and
that gives
and
hence
Example 3.10 A 3 kg block is hanged from the ceiling as in Fig. 3.13.
Find the magnitude of , and
Solution 3.10
The free-body diagrams of the block and the knot are shown in
Fig. 3.14. From Newton’s second law is equal to the weight of the
block, i.e.,
(3.1)
(3.2)
Example 3.11 Figure 3.15 shows a weight of 200 that is lifted with a
constant speed. Find the tension in each part of the rope and the force
of lift.
Solution 3.11
Since the pulleys are massless and frictionless we have ( is
the tension in each rope)and , thus
we also have
Fig. 3.15 Using two pulleys to reduce the force necessary to lift a weight
where r is the radius of the circle. Figure 3.16 shows an object attached
to a string in uniform circular motion (the plane of motion is parallel to
the Earth’s surface). From Newton’s second law, the centripetal
acceleration is caused by a force or net force directed towards the
center of the circle. Therefore, as , the centripetal (or radial) force
has a constant magnitude but its direction changes continuously
The magnitude of this centripetal force is given by
If at some instant the radial force becomes zero, the object would then
move along a straight line path tangent to the circle. Hence, the
centripetal force is necessary to keep the object in its circular path. The
centripetal force may be any kind of force such as friction, gravity, or
tension.
Fig. 3.16 An object attached to a string in uniform circular motion
Solution 3.12
The horizontal component of the tension force supplies the required
centripetal force to keep the bob in its circular path while the vertical
component balances the weight of the bob. (a) Applying Newton’s
second law in both the - and -directions we have
(3.3)
and
(3.4)
Since , we have
Example 3.13 (a) A car needs to turn on a level road without skidding
as in Fig. 3.18. Find the maximum speed for which the car can take the
curved path of the level road safely (b) If the road is banked, i.e., the
outer edge is raised relative to the inner edge as in Fig. 3.19, ind the
maximum speed for which the car can take the curved path of the level
road safely without depending on friction.
Solution 3.13
(a) The centripetal force required for the car to remain in its circular
path, is in this case, is the force of static friction. The maximum speed
for which the car can take the curve without skidding is when the static
frictional force is a maximum. That is,
Hence
(b) If the road is banked the car can take the turn without depending on
friction as the required centripetal force. In that case, the horizontal
component of the normal force supplies the necessary centripetal force.
Thus we have
and
where v is the speed of the car. Dividing these two equations gives
If the angle and the curvature r are known, then the safe speed limit
can be found. If the car moves at a speed lower or higher than that
speed then the frictional force must supply the additional centripetal
force for the car to stay in its circular path.
and
Fig. 3.20 The radial and tangential forces in nonuniform circular motion
Solution 3.14
(a) Applying Newton’s second law in both the tangential and radial
directions gives
and
(b) For the object to remain in its circular path, the string must remain
taut, i.e., must be positive . If then . Hence,
the velocity must satisfy
Fig. 3.24 Two masses connected by a light string over a frictionless pulley of negligible mass
Problems
1.
A 4 kg object is exposed to two forces (see Fig. 3.22). Find the
magnitude and direction of the acceleration of the object.
2.
A 0.2 kg block is released from the top of an inclined plane of
angle as in Fig. 3.23. Find the speed of the block just as it
reaches the bottom.
3.
Two masses are connected by a light string that is connected to a
frictionless pulley of negligible mass as in Fig. 3.24. Find the
magnitude of the acceleration of the masses and the tension in the
string.
4.
A block is pushed up along a smooth inclined plane of angle of
where it is given an initial velocity of 8 . Determine the
time it takes the block to return to its initial position.
5.
Find the tension in each string in the system shown in Fig. 3.25.
6.
A 5 kg block is held in equilibrium as in Fig. 3.26. Find the normal
force acting on the block.
7.
Find the normal force exerted on a 70 kg man standing inside an
elevator that is accelerating upwards at a rate of 2
8.
Find the acceleration of the system shown in Fig. 3.27 and the
tension in the string if kg and kg (assume
massless string and frictionless surface).
Two blocks of masses 3 and 5 kg are placed on top of each other
Two blocks of masses 3 and 5 kg are placed on top of each other
9. as in Fig. 3.28. If the coef icient of static friction between the
blocks is 0.2 and assuming there is no friction between the lower
block and the surface on which it rests, ind the maximum
horizontal force that can be applied to the lower block such that
the blocks move together.
10.
A 1000 kg car move along the track shown in Fig. 3.29. Find (a)
the maximum speed the car can have at point A such that it does
not leave the track (b) the normal force exerted on the car at if
its speed there is 15 .
11.
A block of mass m on a frictionless table is attached to light string
that passes through the center of the table and is connected to a
larger block of mass M (see Fig. 3.30). If m moves in uniform
circular motion of radius r and speed v, ind v such that M remains
at rest when released.
12.
A l kg particle moves in the force ield given by
Find the position of the particle at any
time if at , and .
Fig. 3.28 Two blocks placed on top of each other, where a horizontal force is applied to the lower
block
Fig. 3.29 A car moving on a curved path
Fig. 3.30 A block of mass m on a frictionless table is attached to light string that passes through
the center of the table and is connected to a larger block of mass M
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_4
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
4.1 Introduction
Energy is a very important concept that is heavily used in everyday life.
Everything around us, including ourselves, needs energy to function. For
example, electricity provides home appliances with the energy they
require, food gives us energy to survive, and the sun provides earth with
the energy needed for the existence of life!
Experiments show that energy is a scalar quantity related to the
state of an object. Energy may exist in various forms: mechanical,
chemical, gravitational, electromagnetic, nuclear, and thermal.
Furthermore, energy cannot be created or destroyed; it can only be
transformed from one form to another. In other words, if energy were to
be exchanged between objects inside a system, then the total amount of
energy (the sum of all forms of energy) in the system will remain
constant.
A transformation of energy occurs due to the action of a force known
as work or due to heat exchange between objects (or between an object
and its environment). If energy is transferred due to work then it may be
de ined as the capacity of doing work. This book is concerned with
mechanical energy which involves kinetic energy (associated with the
object’s motion) and potential energy (associated with the position of
the object in space).
4.2 Work
Work may have many meanings. Sometimes, work is said to be done
when a muscular activity is performed. Work may also refer to mental
activity (mental work). In physics, the de inition of work is different.
Work is said to be done if a force is applied to an object while it is
moving, i.e., if there is no resulting displacement, no work is done.
Suppose that a person holds a heavy box for sometime and then starts to
feel tired. The reason he/she feels tired is because chemical energy in
his/her body is converted into internal microscopic motions of the
muscles. Since the energy is not transferred to the box being carried (the
box did not move), the work done on the box is equal to zero.
where
Solution 4.1
(a) The work done by the applied force is
The work done by the normal force and the force of gravity are both
zero since each force is perpendicular to the displacement.
(b) The total work done is
The total work done can also be found by computing the net force
acting on the block and calculating its work.
(c) For , If , then and therefore
, i.e., it is easier for the man to pull at an angle larger than
.
Solution 4.2
(a) The minimum work that the delivery man must do is the work done
against gravity The work done on the crate by the force of gravity is
Hence the minimum work that the delivery man must do is equal to
or
As mentioned in Sect. 1. 10. 1, this integral is called the line integral. Each
component of may be a function of x, y, and z, and the
curve can be determined by its equations that relates x, y, and z to each
other. The component form of the above equation is
(4.1)
Now consider the case in which the particle moves along a straight line
(for example the positive -axis) and in which the force acting on the
particle has a constant direction along the -axis and a magnitude that
varies with x. Equation 4.1 is then reduced to
(4.2)
This equation represents the area under the curve in Fig. 4.4. If F(x) is
constant then we have
The work is then equal to the rectangular area shown in Fig. 4.5.
Fig. 4.3 A particle moving along a curved path. While itõ s moving, a force that varies in both
magnitude and direction with the position of the particle acts on it
Fig. 4.4 The area under the curve represents the work
Fig. 4.5 The work is equal to the rectangular area
Example 4.3 In Example 3.3, ind the work done by the force in moving
the particle during the time interval from to
Solution 4.3
The work done from to is
Solution 4.4
The work done is equal to the area of the triangle under the curve
between to , i.e.
Fig. 4.6 A force acting on a particle is a function of position
Solution 4.5
Because the tension force is always perpendicular to the displacement,
the work done by it is zero at all times. The only component of the
gravitational force that does work is its tangential component.
Therefore,
Since , then , and we have
Fig. 4.7 A ball suspended by a light rope and displaced a small distance from the position of
equilibrium
Fig. 4.8 The center of mass of the system (man+skateboard) moves and the work-energy theorem
can be applied to that point
the net force in displacing the particle is equal to the change in the KE of
the particle
Similar to work, the SI unit of kinetic energy is the Joul. Note that the
work–energy theorem is applied only if the object is treated as a particle
(all of its parts move in exactly the same way). As an example of how the
theorem is applied only for particle-like objects consider a man standing
on a skateboard on a horizontal surface (see Fig. 4.8). If the man pushes
the bar then that would move him backwards along with his skateboard.
This motion is due to the reaction force exerted on him by the bar.
The work done by or is equal to zero since each force is
perpendicular to the displacement. Because the point of application of
did not move it follows that the work done by that force is zero. Thus,
from the work–energy theorem the man should not move. The question
is why did he move?
The fact here is that it is incorrect to treat the man as a particle, since
different parts of his body move in different ways as he pushes the bar.
Therefore, the work–energy theorem does not hold. The man must be
treated as a system of particles. In Chap. 6, we will see that the motion of
a system of particles can be represented by the motion of its center of
mass. The center of mass behaves as if all of the mass of the object (or
system) is concentrated there and as if the net external force is applied
there. In the case of the skateboarder, the center of mass of the system
(man skateboard) moves and the work–energy theorem can be
applied to that point.
The work–energy theorem is an alternative method for describing
motion without using Newton’s laws. It is especially useful in problems
involving a varying force. Note that the work and the kinetic energy are
not invariant quantities; they have different values when measured in
different inertial frames of reference. However, from the principle of
invariance, the equation still holds for any inertial frame.
Solution 4.6
As we will see later in Sect. 4.3.1, the change in the kinetic energy of the
block due to friction is , where s is the displacement of the
block.
Example 4.7 A 10 kg block is pushed on a frictionless horizontal
surface by a constant force of magnitude of 100 and that is at
below the horizontal. If the block starts from rest, ind its inal speed
after it has moved a distance of using work–energy theorem.
Solution 4.7
since we get
The work done on the block by the spring as it moves from an initial
position to a inal position is
Solution 4.8
Fig. 4.11 A 2 kg block attached to a light spring of force constant 300 on a horizontal
smooth surface
Note that unlike the spring force the reference point may be chosen
anywhere. If the object moved downwards from to , the
work done by the gravitational force is
This result is the same as if the object has followed a straight vertical
path. Therefore, the work done by the gravitational force depends only
on the initial and inal positions of the object.
Fig. 4.12 By taking at the hand level, in the work done by gravity a is and in b is
Fig. 4.13 a The total work done by the spring force on the block is zero since . b Along any
path the work done by the gravitational force is the same since the initial and inal positions are the
same
4.3.3 Power
Power is a quantity that de ines how much work is done over a period of
time, i.e., power is the time rate of doing work, or more generally, it is
the time rate of energy transfer. If an external force does work W on
an object for a time interval , then the average power during that time
is
The SI unit of power is joules per second and is called the watt
(W).
Fig. 4.14 The longer the path the more interaction between the block and the surface and the more
the force of friction will act and do work on the block
Furthermore
Thus, the total work done by a conservative force in moving a particle
from to (see Fig. 4.15) is
or
Fig. 4.15 The total work done by a conservative force in moving a particle from to
Example 4.10 A force acting on a particle is given by .
Determine: (a) whether or not the force is conservative; (b) the potential
energy associated with the force if it is conservative.
Solution 4.10
(a)
Solution 4.11
(a) Along path 1 we have and and along path 3 we have
and
(b) Since the total work done through the closed path is zero, the force is
conservative.
Fig. 4.16 The work done in moving the particle along a closed path
Example 4.12 Find the force acting on a particle if the potential energy
associated with it is
Solution 4.12
and therefore
The force of gravity near the surface of the earth can be found from the
gravitational potential energy In general we have here, since
the motion is in one direction we have
Since , we have
The spring force can be found from the elastic potential energy
4.5 Conservation of Mechanical Energy
The total mechanical energy of a system is de ined as the sum of all of
the kinetic energies of the objects within the system plus all of the
potential energies of the system.
Thus,
or
or
If more than one conservative force acts, there will be a potential energy
associated with each force. That is
Therefore we have
or
or
Hence
or
Thus
(4.3)
This implies that the total mechanical energy has changed by an amount
of . Not that the work done by a nonconservative force cannot be
calculated generally but the change in the kinetic energy can be
observed.
4.5.2 Friction
Friction is a nonconservative force as seen in Sect. 4.2. If this force is
applied externally to a system in which only internal conservative forces
act, it will decrease (dissipate) the kinetic energy of the system by
transforming it into thermal energy The change in the mechanical
energy of the system is
Thus
or
Therefore
This quantity represents the magnitude of the loss in the kinetic energy
of the block due to friction. This loss of energy appears as thermal
energy of the block and of the surface on which it slides.
Solution 4.13
(a) Consider the system to be the earth the apple. By neglecting air
resistance (the apple is in free-fall), the only internal force that acts
within the earth–apple system is the force of gravity. Because the
gravitational force is a conservative force, the total mechanical energy of
the system is conserved. Therefore as the apple falls its gravitational
potential energy is converted into kinetic energy such that at any instant
the total mechanical energy of the system is constant. Applying the law
of conservation of energy to the system and by taking at the
earth’s surface and the gravitational potential energy to be zero at
, we have
At
(b)
Example 4.14 A roller coaster of mass 500 kg starts from rest at point
, and rolls down the track as shown in Fig. 4.18. Ignoring friction,
determine: (a) the roller coaster speed at and ; (b) the work done
by gravity as the rollercoaster moves from A to B.
Fig. 4.18 By ignoring friction, the total energy of the roller coaster can be considered to be
conserved
Solution 4.14
(a) Consider the system to consist of the rollercoaster the track the
earth. Taking the gravitational potential energy to be zero at the earth’s
surface and from the conservation of energy we have
Therefore,
You may also calculate the velocity at C by taking B as the initial point.
(b) As the car moves from A to the work done by gravity is
Solution 4.15
First, we divide the path into three parts. Let us consider the system as
the block only Along the irst part the change in the total mechanical
energy of the system is equal to the energy dissipated by friction. Thus,
That gives
and therefore
Finally, along the third path, we also have
and
but we have
and thus
That gives
Solution 4.16
If air resistance is neglected, the only force acting in the masses-earth
system is the gravitational force between them and hence the total
mechanical energy of the system is conserved, i.e.,
Because the two masses are connected by a rope, they have the same
speed at any instant. If descends a distance will rise through
the same distance and we have
and therefore
Example 4.17 A 0.25 kg ball is attached to alight string of length
as in Fig. 4.21. Find (a) the tension in the string at
if the ball is given an initial velocity at its
lowest position; (b) the velocity of the ball at A if the ball is released
from rest at B.
Solution 4.17
(a) At point some of the kinetic energy of the ball is converted into
potential energy By taking the origin of the x-y coordinates at the lowest
point , we have
(4.5)
thus
and hence
hence
Fig. 4.22 A 3 kg block compresses a spring of negligible mass a distance of 0.1 from its
equilibrium position
Solution 4.18
(a) The only force acting inside the spring–mass–earth system is the
spring force that acts on the block. This force is conservative and
therefore the total mechanical energy of the system is conserved. The
potential energy of the spring is transformed into kinetic energy of the
block,
and therefore
this gives
(b)
and hence
We can also take the initial position before the block is released.
(c)
thus
That gives
Example 4.19 A small stone of mass 0.1 kg is released from rest inside
a large hemispherical bowl of radius . It then slides along the
surface as in Fig. 4.23. (a) Find the speed of the stone at point and ;
(b) If the surface of the bowl is not frictionless, how much energy is
dissipated by friction as the stone moves from A to if the speed at
is 1.7 ?
Fig. 4.23 A small stone of mass 0.1 kg is released from rest inside a large hemispherical bowl of
radius
Solution 4.19
(a)
thus
and therefore
(b) If a force of kinetic friction exists between the stone and the
bowl, the total mechanical energy at point is given by
where the stone is considered as the system, therefore
Solution 4.20
From the conservation of energy the velocity when he leaves the track is
That gives
Solution 4.21
(a) From the conservation of energy, we have
(d) When the stone is at the verge of falling at , then the only force
acting on it is the force of gravity and we have .
or
and
4.5.5 Power
Expanding on the de inition of power, power is the rate of energy
transfer due to a force. If is the amount of energy transferred in an
amount of time the average power is
The instantaneous power is then
That is, it is the negative of the slope of the curve at that point. Because
this force is conservative it follows that the total mechanical energy of
the system is conserved. Therefore the kinetic energy of the particle as a
function of position is given by
On the U versus x curve, the kinetic energy at any point can be found by
subtracting the value of U (at that certain point) from E.
Fig. 4.26 The potential energy of the system as a function of the particle’s position
or
hence
Solution 4.23
or
Problems
1.
A force acting on a particle varies with position as in Fig. 4.29. Find
the work done by the force as the particle moves from to
2.
A force acts on a particle that undergoes a
displacement . Find the work done by the
force on the particle.
3.
A 5 block is pulled from rest on a rough surface by a constant
force of 10 that is at to the horizontal. If the coef icient of
kinetic friction between the block and the surface is 0.15, ind the
inal speed of the block as it moves through a displacement of
using the work–energy theorem.
4. Calculate the work done against gravity in moving a 30 kg box
through a height of 6
through a height of 6
5.
A 1600 kg car accelerates from rest at a rate of 1 . Find the
average power delivered to the car during the irst 5
6.
Determine whether or not the force is
conservative, where is constant and m is the mass of the
particle. If the force is conservative determine the potential energy
associated with it.
7.
A 5 block slides down an inclined plane of angle (see
Fig. 4.30). Using energy methods, ind the speed of the block just as
it reaches the bottom if the coef icient of kinetic friction is
8.
A block of mass of 2 kg is pressed against a light spring of force
constant 400 (see Fig. 4.31). If the compression of the spring
is 10 cm, ind the maximum height the block will reach when it is
released.
9.
A force acting on a particle is given by . Find the work
done in moving the particle along the path shown in Fig. 4.32.
10.
Two blocks are connected by a light rope that passes over a
massless frictionless pulley (see Fig. 4.33). If the system is released
from rest, ind the total kinetic energy of the blocks when the 5 kg
block descends a distance of 0.5 assuming that the surface is
frictionless.
11. A particle of mass 1.5 kg moves along the -axis where its
potential energy varies as in Fig. 4.34. Plot the force versus x
f
from to
12.
A block of mass m rests on a hemispherical mound of ice as shown
in Fig. 4.35. If it is given a very small push and start sliding, ind the
height of the point in which the block will lose contact with the
mound.
13.
A 3 kg block hangs from a spring as in Fig. 4.36. If the spring
stretches a distance of 10 cm, ind (a) the force constant of the
spring (b) the work done in expanding the spring a distance of
5 cm without accelerating it.
14. In Fig. 4.37, determine the Turning points and the positions of
stable and unstable equilibrium.
Fig. 4.32 The work done in moving the particle along a closed path
Fig. 4.33 Two blocks connected by a light rope that passes over a massless frictionless pulley
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_5
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
where is the velocity of the particle. A fast moving car has more
momentum than a slow moving car of the same mass. Another example
is that a bowling ball has more momentum than a basketball moving at
the same speed. The SI unit of linear momentum is . In terms of
components, we may write , and .
Newton’s second law can be expressed in terms of momentum for a
particle-like object of constant mass as
or
That is, the rate of change of the linear momentum of an object is equal
to the resultant force acting on the object and is in the same direction
as that force.
Fig. 5.1 An isolated system consisting of two particles where the only forces that act in the
system are internal forces
That is,
or
That is, the linear momentum of each particle may change, but the total
linear momentum of the system is the same at all times. This statement
is known as the law of conservation of linear momentum: If the net
external force on a system is zero, the total linear momentum of the
system remains unchanged (constant). In terms of components, we have
, and . In solving problems involving
collisions, and refers to the total momentum of the system
immediately before and immediately after the collision, respectively.
For a two-particle system, we have
The right side of the equation is a vector quantity known as the impulse
I
Hence,
That is, is a constant force that gives the same impulse as F. In the
case of a collision between two bodies, the variation of the impulsive
force that each body exerts on the other during the collision time takes
the form as shown in Fig. 5.2.
Fig. 5.2 One example of the variation of over time
5.4 Collisions
As discussed previously, when two bodies collide, they exert large
forces on one another (during the time of the collision) called impulsive
forces. These forces are very large such that any other forces ( .,
friction or gravity) present during the short time of the collision can be
neglected. This approximation is known as the impulse approximation.
For example, if a golf ball was hit by a golf club, the change in the
momentum of the ball can be assumed to be only due to the impulsive
force exerted on it by the club. The change in its momentum due to any
other force present during the collision can be neglected. That is, the
force in the expression I can be assumed to be the
impulsive force only The neglected forces present during the collision
time are external to the two-body system, whereas the impulsive forces
are internal. The two-body system can therefore be considered to be
isolated during the short time of the collision (which is in the order of a
few milliseconds). Hence, the total linear momentum of the system is
conserved during the collision, which enables us to apply the law of
conservation of momentum immediately before and immediately after
the collision. In general, for any type of collision, the total linear
momentum is conserved during the time of the collision. That is,
., where and are the momenta immediately before and
after the collision. In the next sections, we will de ine various types of
two- body collisions, depending on whether or not the kinetic energy of
the system is conserved.
Solution 5.1
The impulse of the force is
Solution 5.2
Because there are no external horizontal forces acting on the cannon-
carriage-ball system, then the total momentum of the system is
constant (conserved) in the -direction
therefore,
Solution 5.3
Solution 5.4
Along the -direction, we have
and
along the -direction, we have
and
and
where is measured from the positive -axis. The average force acting
on the puck is
Solution 5.5
For the two-skater system, the sum of the vertical forces are zero
(weight and normal forces) and the forces exerted by one skater on the
other is internal to the system. That is, there are no external forces
acting on the system and the total momentum is conserved. Because
the motion takes place in a straight line, we have
and hence,
Example 5.6
A particle is moving in space under the in luence of a force. If its
momentum as a function of time is
(a) Find the force acting on the particle at any time; (b) Find the
impulse of the force from to
Solution 5.6
(a)
(b)
5.4.1 Elastic Collisions
An elastic collision is one in which the total kinetic energy, as well as
momentum, of the two-colliding-body system is conserved. These
collisions exist when the impulsive force exerted by one body on the
other is conservative. Such force converts the kinetic energy of the body
into elastic potential energy when the two bodies are in contact. It then
reconverts the elastic potential energy into kinetic energy when there is
no more contact. After collision, each body may have a different velocity
and therefore a different kinetic energy. However, the total energy as
well as the total momentum of the system is constant during the time of
the collision. An example of such collisions is those between billiard
balls.
Fig. 5.4 Two particles of masses and experiencing an elastic head-on collision
(5.2)
(5.3)
(5.4)
(5.5)
Solution 5.7
For an elastic head-on collision, we have
Solution 5.8
(a) Using the impulse approximation, the law of conservation of
momentum gives the velocities just before and after the collision when
the string is still nearly vertical. For a perfectly inelastic collision, the
total momentum is conserved but the total kinetic energy is not
conserved during the collision. Thus, we have
That gives
(b) The kinetic energy of the bullet before collision is
therefore,
That is, nearly, all the mechanical energy is dissipated and converted
into internal (thermal) energy of the (block bullet) system.
Solution 5.9
Solution 5.10
(5.6)
From the conservation of momentum, we have
That gives
(5.7)
Solving Eqs. 5.6 and 5.7 gives and
Solution 5.11
Applying the law of conservation of momentum immediately before
and after the collision gives
by taking the (block bullet) as the system after the collision until it
comes to rest, we have
that gives
Hence,
and
Furthermore,
Fig. 5.8 A two dimensional elastic collision between two particles where one particle is moving
and the other is at rest
Fig. 5.9 A ball sliding along a horizontal frictionless surface collides with a second ball that is
initially at rest
Solution 5.12
Applying the law of conservation of momentum immediately before
and after the collision in each direction gives and .
Thus,
That is, some of the energy of the system is lost and thus the collision is
inelastic.
(c) In a perfectly elastic collision, both the total momentum and the
total mechanical energy of the system are conserved. That is
(5.8)
(5.9)
From the conservation of kinetic energy, we have
or
(5.10)
Substituting Eq. 5.8 into Eq. 5.9 gives
or
(5.11)
Therefore,
Solution 5.13
The direction of is
from the positive -axis. The change in the kinetic energy of the system
is
5.5 Torque
Consider a force acting on a particle that has a position vector with
respect to some origin that is in an inertial frame. The torque is a
vector quantity that measures the tendency of that force to rotate the
particle about and is de ined as
Let us consider a particle in the x–y plane exposed to a force that lies in
that plane (see Fig. 5.11). The resulting torque is then perpendicular to
the x–y plane parallel to the -axis. can also be written as
Fig. 5.12 is called the moment arm of and it represents the perpendicular
Solution 5.14
Fig. 5.13 If the particle is moving in the x-y plane, then the direction of is perpendicular to the
plane containing and and is found by the right-hand rule
This implies that the torque acting on a particle is equal to the time rate
of change of the angular momentum for that particle. This equation is
valid only if and are evaluated with respect to the same origin or
any other ixed point in an inertial frame. If several forces act on the
particle, Eq. 5.12 can be written as
where is the net torque on the particle. This is the rotational analog
of Newton’s second law in linear form, which states that the net force
acting on a particle is equal to the time rate of change of its linear
momentum. In component form, we have
and
Solution 5.15
Suppose the plane is the x–y plane. Since , we have
Solution 5.16
Solution 5.17
Solution 5.18
Problems
1.
A tennis ball of mass of 0.06 kg is initially traveling at an angle of
to the horizontal at a speed of 45 . It then was shot by the
tennis player and return horizontally at a speed of 35 . Find
the impulse delivered to the ball.
2.
A force on a 0.5 kg particle varies with time according to Fig. 5.16.
Find (a) The impulse delivered to the particle, (b) the average
force exerted on the particle from to . The inal
velocity of the particle if its initial velocity is 2
3.
A 1 kg particle moves in a force ield given by
N. Find the impulse delivered to the
Fig. 5.20 A bullet of mass of m is ired with a horizontal velocity v into a block of mass M
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_6
6. System of Particles
Salma Alrasheed1
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
Fig. 6.1 Two particles of masses and moving in space. Their position vectors at a
Example 6.1 Find the center of mass of the system shown in Fig. 6.3
where the three particles have an equal mass of
Solution 6.1
Solution 6.2
The position vector of each particle is
and
That gives
and
Fig. 6.4 An extended object of mass M divided into small volume elements each of mass and
a vector position
Solution 6.3
Substituting gives
Fig. 6.5 A thin rod of length has a linear density that increases with x
Fig. 6.6 A uniform square sheet suspended by a uniform rod where they both lie in the same
plane
Solution 6.4
Because the sheet and the rod are homogeneous, the center of mass of
each is at its geometric center. Since the center of the sheet is at the
origin we have
Example 6.5 Find the center of mass of the rectangular plate shown in
Fig. 6.7. The plate has a uniform surface density
Solution 6.5
Method 1:
Hence
Method 2:
Dividing the plate into very thin rods each of mass gives
Similarly by dividing the plate into thin horizontal rods each of mass
gives
and
Solution 6.6
The equation of an ellipse is
therefore
or
By dividing the area into very thin rectangles each of mass gives
To obtain the y coordinate of the center of mass we divide the area into
very thin rectangles each of mass as in Fig. 6.8. That gives
Solution 6.7
From symmetry, the center of mass lies on the -axis. By dividing the
shell into very thin rings each of mass we have
Solution 6.8
By taking the midpoint between the boy and the container as the origin
(see Fig. 6.10) and by neglecting the mass of the rope, the center of
mass of the system is
Solution 6.9
By neglecting air and water resistance, the net external force on the
boat) system is zero. Therefore the center of mass of the system
must remain at rest. Suppose that the boat is a symmetrical
homogeneous object where its center of mass is at its geometrical
center. The center of mass of the boat is therefore at a distance of 2.5
from the origin. Thus, the center of mass of the system is
If the boy moves a distance of 1 , the center of mass is still at the
same position, and we have
where , or
(6.1)
(6.2)
where is the acceleration of the ith particle.
(6.3)
Solution 6.10
That gives
The total linear momentum is
at
and
(6.5)
where is the net force acting on the ith particle. If both the external
forces on the system and the internal forces between the particles in
the system are included, then may be written as
(6.6)
(6.7)
Therefore, the second term in Eq. 6.7 is equal to zero. Hence the net
force acting on the system is due only to external forces. That gives
thus
Thus
and
(6.8)
By using Newton’s third law of action and reaction, the double sum in
Eq. 6.8 has terms of the form
Now, suppose that the internal forces between the two particles lie
along the line joining the particles (i.e., the vectors and
have the same direction). This condition is known as the strong law of
action and reaction. It requires the internal forces to be central. If the
internal forces are equal and opposite but not central, then they are
said to satisfy the weak law of action and reaction. The force of gravity
is an example of a force satisfying the strong law of action and reaction.
Some forces such as the forces between two moving charges are not
central. From this, it follows that the double summation in Eq. 6.8 is
equal to zero.
Therefore, the total torque on the system about the origin is only the
torque associated with external forces
(6.9)
i.e., the net external torque about some origin exerted on a system of
particles is equal to the time rate of change of the total angular
momentum of the system.
6.3.14 Work
Since the total force acting on the ith particle is given by
it follows that
and
The factor 1/2 occurs since each term in the summation appears twice.
Now, consider the total work done by the external conservative forces
To show that energy is conserved when both the external and internal
forces are conservative, we may de ine a total potential of the system as
From the work–energy theorem, the work done by the total force
acting on the ith particle is equal to the change in the kinetic energy of
that particle
and since
or
or
Thus
6.3.17 Impulse
In Sect. 6.3.7, we have seen that the net external force on a system of
particles is equal to the rate of change of the total linear momentum of
the system
The total linear impulse on the system as the system goes from one
state to another is de ined as
That is, the total linear impulse on the system is equal to the change in
the total momentum of the system.
Fig. 6.12 The position vector of the ith particle relative to the center of mass
(6.10)
From Fig. 6.12, the position vector of the ith particle relative to the
center of mass is
or
(6.11)
Where is the position vector of the ith particle relative to the origin
O. Substituting Eq. 6.11 into Eq. 6.10 gives
therefore
That gives
(6.12)
(6.13)
or
or
That is, the total linear momentum of the system is zero when observed
from the center of mass frame.
The second and third terms are zero followed from Eqs. 6.12 and 6.13
where and , hence
(6.15)
6.4.3 The Total Kinetic Energy of a System of Particles
About the Center of Mass
The total kinetic energy of a system of particles relative to an origin in
an inertial frame of reference is given by
From Eq. 6.13, the term in brackets in the second term is equal to zero.
Hence
total angular momentum of the system; the total external torque acting
on the system; and the total kinetic energy of the system all relative to
the origin at any time; (b) repeat (a) relative to the center of mass.
Solution 6.11
(a)
that gives
From Eq. 6.15, the total angular momentum relative to the center of
mass is
Example 6.12 Two particles of equal mass m are rotating about their
center of mass with a constant speed v as in Fig. 6.13. If they are
separated by a distance 2d, ind the total angular momentum of the
system.
Solution 6.12
Fig. 6.13 Two particles rotating about their center of mass
and
That is, when viewed from the center of mass frame the two objects
approach each other with equal and opposite momenta and move away
from each other with an equal and opposite momenta. Therefore, the
center of mass frame simpli ies the analysis since it exhibits a particular
symmetry to the problem (see Fig. 6.15).
Fig. 6.14 Consider a system consisting of two bodies undergoing a one-dimensional collision
Solution 6.13
When the rocket reaches the top its velocity immediately before
explosion is zero. Since and are the velocities of the fragments
immediately after explosion, we have from the conservation of
momentum
and
That gives
and
Thus,
Example 6.14 Find the center of mass of the Earth–Moon System and
describe its motion around the sun.
Solution 6.14
As we shall see in Chap. 9, the center of mass of two bodies with
different masses moving under gravity will trace an ellipse. Since the
external forces on the sun can be neglected, we may consider it to be at
rest in an inertial frame of reference and at the origin of a coordinate
system (see Fig. 6.17). The center of mass of the Earth–Moon system is
where and are unit vectors in the direction of and
respectively. The equation of motion of the center of mass is
Since the distance between the earth and the moon is so small
compared to their distance from the sun we may write
Fig. 6.18 The center of mass of the Earth-Moon system moves as a single planet of mass
about the sun
Example 6.15 Describe the motion of a rocket in space using the law of
conservation of momentum.
Fig. 6.19 A rocket moving in space is a system with varying mass. Its motion is analyzed using the
law of conservation of momentum
Solution 6.15
A rocket moving in space is a system with varying mass. Its motion is
analyzed using the law of conservation of momentum. In order for a
rocket to move in space, its fuel is burned and gases are produced and
ejected from its rear. This will cause the mass of the rocket to decrease
continuously The ejected gases produce momentum in the backward
direction and as a result the rocket receives a forward momentum and
its velocity increases (see Fig. 6.19). Suppose at an instant t, the rocket
has a mass M and velocity v relative to a stationary frame of reference.
During a time interval t, a mass of the fuel is expelled as gas with a
velocity u relative to the rocket. The speed of the rocket increases to
and the speed of the fuel relative to the stationary frame of
reference is . The initial momentum of the rocket is
Thus
hence
That gives
Therefore, the inal speed of the rocket depends on the exhaust speed
and on the ratio of the initial and inal masses.
Fig. 6.20 A system of particles in x-y plane
Problems
1.
Find the coordinate of the center of mass of the system shown in
Fig. 6.20.
2.
Find the center of mass of a uniform plate bounded by
and the -axis from to
3.
Find the center of mass of the homogeneous sheet shown in
Fig. 6.21.
4.
Find the center of mass of the homogeneous sheet shown in
Fig. 6.22.
5.
Find the center of mass of a uniform solid circular cone of radius a
and height h.
6.
Find the center of mass of a uniform solid hemisphere of radius R.
7.
Two masses initially at rest are located at the points shown in
Fig. 6.23. If external forces act on the particles as in Fig. 6.23, ind
the acceleration of the center of mass.
A projectile of mass 15 kg is ired from the ground with an initial
p j g g
8. velocity of 12 at an angle of to the horizontal. 1 second
later, the projectile explodes into two fragments A and B. If
immediately after explosion, fragment A has a mass of 5 kg and a
Fig. 6.24 By neglecting friction between the boat and water, the center of mass can be used to
ind the distance moved by the boat
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_7
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
Note that if the particle completes one revolution, will not become zero
again, it is then equal to . Thus for example for three revolutions the
angular position is given by
Fig. 7.1 A rigid body of an arbitrary shape is in pure rotational motion about the -axis
Fig. 7.2 The motion of a particle that lies in a slice of the body in the x-y plane
Fig. 7.3 The particle is at point at and at at , where it changes its angular position from
to
Example 7.1 Convert each of the following into the other angular units:
, 0.25 , 3
Solution 7.1
Example 7.2 A rotating rigid object has an angular position given by
rad. Determine: (a) the angular displacement of
the object and the average angular velocity during the time interval from
to . (b) the instantaneous angular velocity and the
instantaneous angular acceleration at .
Solution 7.2
(a)
and
(b)
at
at
Solution 7.3
(a)
Since
we have
(7.1)
Furthermore
Hence
(7.2)
or
(7.3)
Finally solving for t from Eq. 7.1 and substituting into Eq. 7.2 gives
or
(7.4)
Note that as mentioned earlier, if a rigid object is in pure rotational motion,
all particles in the object have the same angular velocity and angular
acceleration. Different particles move in different circles but the center of
these circles lies at the axis of rotation. As the rigid body rotates, a particle
in the body will move through a distance s along its circular path (see
Fig. 7.6). The angular displacement of the particle is related to s by
where r is the radius of the circle in which the particle is moving along.
Differentiating the above equation with respect to t gives
(7.6)
or
The total linear acceleration of the particle (see Fig. 7.7) is given by
Fig. 7.6 As the rigid body rotates, a particle in the body will move through a distance s along its circular
path
Fig. 7.7 The total acceleration of the particle
Rotational motion about a ixed axis with constant Linear motion with constant a
Solution 7.4
(a) We have and rad. By
choosing the reference position we have
(b)
at
(c)
(d)
Example 7.5 Two sprockets are attached to each other as in Fig. 7.8. There
radii are cm and cm. If the angular velocity of the smaller
sprocket is 2 ind the angular velocity of the other.
Solution 7.5
A point at the rim of one sprocket has the same linear speed as a point at
the rim of the other sprocket since they are attached to each other, i.e.,
hence
Fig. 7.8 Two sprockets connected at the rim
Example 7.6 Find the angular speed of the moon in its orbit about the
earth in rev/day.
Solution 7.6
Assuming that the moon’s orbit is circular, the linear speed of the moon is
given by , where r is the mean distance from the earth to the
moon and T is its period. Thus, the angular velocity of the moon is
or
where is the position vector of the particle from the origin (see Fig. 7.9)
and is the angle between the position vector and the -axis. As shown in
Fig. 7.9, the direction of is perpendicular to the plane formed by and
where it can be veri ied using the right-hand rule. Therefore, by using the
de inition of vector product we may write
(7.7)
The total linear acceleration is
(7.9)
Equations 7.7–7.9 are the vector relationship between angular and linear
quantities.
Fig. 7.9 A rigid body in pure rotational motion about a ixed axis (here the -axis)
velocity of the ith particle respectively (see Fig. 7.10). From Eq. 7.5, we have
In solving problems , and (see Sect. 6.3.4) are often used to express
dm in terms of its position coordinates.
where is the moment of inertia about an axis passing through the center
of mass, M is the total mass of the system, and D is the perpendicular
distance between the two parallel axes.
Proof
Consider an axis that is perpendicular to the page and passing through the
center of mass of the object. Figure 7.11 shows a thin slice of the object that
lies in the x-y plane. Because the origin is taken at the center of mass we
have
The moment of inertia of the object about the center of mass axis is
where x and y are the coordinates of the mass element dm from the center
of mass (the origin). Now consider another axis that is parallel to the irst
axis and that passes through a point as shown in Fig. 7.11. Suppose that
the and coordinates of from the center of mass are and . The
moment of inertia about an axis passing through is
and
it follows that the second and third terms are zero. Thus
where
Fig. 7.12 The rotational inertia of various rigid bodies of uniform density
7.7 Angular Momentum of a Rigid Body Rotating
about a Fixed Axis
Consider a rigid body rotating about a ixed axis (the -axis) with an
angular speed as shown in Fig. 7.13. The angular momentum of the ith
particle with respect to the origin is given by
Fig. 7.13 A rigid body rotating about a ixed axis (the -axis) with an angular speed
where I is the moment of inertia of the rigid body about the rotational axis
(z-axis). This equation can also be written in component form since is
parallel to , that is,
(7.10)
Therefore, if a rigid body is rotating about a ixed axis (say the -axis), the
component of the angular momentum along that axis is given by Eq. 7.10.
Now suppose that the rigid body is symmetric and homogeneous and that it
is rotating about its symmetrical axis (see Fig. 7.14). For any two particles
(1 and 2) opposing each other with an equal angular momenta and ,
the perpendicular components, and , of the angular momenta
cancel each other out since they are in opposite directions. That leaves the
parallel components and which add up since they have the same
direction. For all particles in the object the total angular momentum is,
therefore, given by
In the case of any rigid object symmetrical or not, the net external torque
acting on the object about the axis of rotation (say the -axis) is equal to the
rate of change of the component of angular momentum that is along that
axis
Fig. 7.14 A homogenous symmetrical rigid body rotating about its symmetrical axis
Example 7.7 A 5 kg wheel of radius of 0.1 decelerates from an angular
speed of 5 to rest after going through an angular displacement of 10
rev If a frictional force causes the wheel to decelerate, ind the torque due to
this force.
Solution 7.7
The angular displacement is
Example 7.8 Three masses are connected by massless rods as in Fig. 7.15.
If ind the moment of inertia of the system and the
corresponding kinetic energy if it rotates with an angular speed of 5
about: (a) the -axis; (b) the -axis and; (c) the -axis .
Solution 7.8
(a)
(b)
(c)
Example 7.9 Fig. 7.16 shows a uniform thin rod of mass M and length L.
Find the moment of inertia of the rod about an axis that is perpendicular to
it and passing through: (a) the center of mass; (b) at one end; (c) at a
distance of L / 6 from one end.
Solution 7.9
(a) The mass dm of an element in the rod is
(b)
(c)
Example 7.10 Fig. 7.17 shows a uniform thin plate of mass M and surface
density . Find the moment of inertia of the plate about an axis passing
through its center of mass if its length is b and its width is a (the -axis).
Solution 7.10
A mass element dm has an area dxdy and is at a distance from
Solution 7.11
Method 1: Using a single integration by dividing the cylinder into thin
cylindrical shells each of radius r, length L and thickness dr as in Fig. 7.18,
then each volume element is given by
and
Since
then
Method 2: Using double integration: dividing the cylinder into thin rods
each of mass
Since
We have
Method 3: Using triple integration Dividing the cylinder into small cubes
each of mass given by
Since
Therefore,
Example 7.12 Three rods of length L and mass M are connected together
as in Fig. 7.19. Determine the moment of inertia of the system about an axis
passing through and perpendicular to the page (the rods lie in the same
plane).
Solution 7.12
The moment of inertia of a thin rod about an axis that is perpendicular to it
and passing through one end is . The total moment of inertia at
Solution 7.13
Let us divide the spherical shell into thin rings each of area (see Fig. 7.20)
given by
since , we have
Fig. 7.20 A spherical shell divided into thin rings
then
That is, the angular momentum is not necessarily conserved in all
directions. It is conserved in the direction where the net external torque is
equal to zero.
Since and are parallel, (the force lies in the x-y plane therefore the total
torque is parallel to the -axis) we have
(7.12)
Table 7.2 Analogous Equations in linear Motion and Rotational Motion about a Fixed Axis
7.10 Power
The instantaneous power delivered to rotate an object about a ixed axis is
found from
Solution 7.14
(a) Since the rotational axis is the axis of symmetry of the disc, then the
moment of inertia is
and
and
(c) The acceleration of a point in the unwinding rope is the same as the
acceleration of a point at the rim of the cylinder, i.e.,
(d)
Since
or
(e) If the rope has moved a distance of lm, the angular displacement of the
cylinder is
Solution 7.16
(a) Since the normal force exerted by the pin on the rod passes through
then the only force that contributes to the torque is the force of gravity This
force acts at the center of gravity which is at the center of mass (see
Sect. 8.4). Therefore the net external torque is
and hence
(c) When the rod reaches its lowest position, the potential energy of its
center of mass is transformed into rotational kinetic energy of the rod. From
conservation of energy we have . Taking the potential
energy to be zero at the lowest position, gives
That gives
Fig. 7.24 A cylinder with a core section is free to rotate about its center. Ropes wrapped around the
inner and outer sections exert different forces
Fig. 7.25 A block of mass m is attached to a light string that is wrapped around the rim of a uniform
solid disk of radius R and mass M
Example 7.17 Find the net torque on the system shown in Fig. 7.24 where
cm, cm, and . Neglect the
mass and friction of the ropes and pulleys.
Solution 7.17
Since all forces lie in the same plane the net torque is
Solution 7.18
The free-body diagrams of the disc and the block are shown in Fig. 7.25.
Applying Newton’s second law to the block gives
or
(7.13)
or
(7.14)
Equating Eqs. 7.13 and 7.14 gives
that gives
gives
Finally
acceleration; (b) the work done on the sphere after 7 revolutions; (c) the
work done after 7 revolutions using the work–energy theorem.
Solution 7.19
(a)
(b)
and
assuming
(c) After seven revolutions the angular velocity is
Since , we have
Example 7.20 Fig. 7.26 shows Atwood’s machine when the mass of the
pulley is considered. If the system is released from rest (and assuming that
the string does not stretch or slip) and that the friction of the pulley is
negligible, ind linear acceleration of the blocks and the angular acceleration
of the pulley.
Fig. 7.26 AtwoodOs machine
Solution 7.20
Fig. 7.26 shows the free-body diagram for each block and for the pulley
Applying Newton’s second law gives
and
The torque is negative because the pulley rotates in the clockwise direction.
Therefore we have
and
That gives
and
Solution 7.21
(a) The moment of inertia of the cylinder is
at that instant
(c)
(d)
the angular momentum of the sphere and the applied torque as a function
of time.
Solution 7.22
and
Fig. 7.27 A uniform solid sphere rotating about an axis tangent to the sphere
Example 7.23 In Example 7.8 ind the angular momentum in each case.
Solution 7.23
(a)
(b)
(c)
Example 7.24 A uniform solid sphere of radius of 0.2 is rotating about
its central axis with an angular speed of 5 . If an impulsive force that
has an average value of 100 acts at the rim of the sphere at the center
level for a short time of 2 : ind the angular impulse of the force; (b)
the inal angular speed of the sphere.
(b)
That gives
Solution 7.25
Because the resultant external torque on the system is zero, it follows that
the total angular momentum of the system is conserved. That is
hence
(b)
This increase in the kinetic energy is because the man does work when he
moves the dumbbells inwards.
Fig. 7.29 A uniform disc rotating without friction. Another disc that is initially at rest is dropped on the
irst, the two will eventually rotate with the same angular speed due to friction between them
Solution 7.26
(a) Since the net external torque acting on the system is zero, it follows that
the total angular momentum of the system is conserved, i.e.,
or
hence
(b)
Problems
1.
A wheel is initially rotating at 60 in the clockwise direction. If a
counterclockwise torque acts on the wheel producing a
counterclockwise angular acceleration , ind the time
the point at
3. A wheel of radius of 0.5 rotates from rest at a constant angular
acceleration of 2.5 . At Find (a) the angular speed of the
wheel (b) the angle in radians through which the wheel rotates (c) the
wheel (b) the angle in radians through which the wheel rotates (c) the
tangential and radial acceleration of a point at the rim of the wheel.
4.
Find the angular speed in radians per second of the earth about (a) its
axis (b) the sun.
5.
An -shaped bar rotates counterclockwise with an angular
acceleration of (see Fig. 7.30). Find (in vector form) the linear
velocity and acceleration of the point on the bar.
6.
Four masses are connected by light rigid rods as in Fig. 7.31. Calculate
the moment of inertia of the system about (a) the -axis (b) the -
axis (c) the -axis.
7.
Find the moment of inertia of a uniform solid sphere of radius R and
mass M about an axis passing through its center of mass.
8.
Find the moment of inertia of an elliptical quadrant about the -axis
(see Fig. 7.32).
9.
A 5 kg uniform solid cylinder of radius 0.2 rotate about its center of
mass axis with an angular speed of 10 rev/min. Find (a) its rotational
kinetic energy (b) its angular momentum.
10.
A wheel of mass of 20 kg and radius of 0.75 is initially rotating at
120 rev/min. If its angular speed is increased to 300 rev/min in 20 ,
ind (a) the work done on the wheel (b) the average power delivered
to the wheel.
11.
A wheel of mass 10 kg and radius 0.4 accelerates uniformly from
rest to an angular speed of 800 rev/min in 20 . Find (a) the torque
applied to the wheel (b) the work done on the wheel (c) the work
done using the work–energy theorem.
12. A uniform rod of length L and mass M is pivoted at (see Fig. 7.33). If
a projectile of mass m moving at velocity v collide with the rod and
stick to it, ind the angular momentum of the system immediately
before and immediately after the collision.
13.
A disc of radius 2.2 and mass of 120 kg rotate about a frictionless
vertical axle that passes through its center. A man of mass 65 kg walks
slowly from the rim of the disc towards the center. Find the angular
speed of the disc when the man is at a distance of 0.7 from the
center if its angular speed when the man starts walking is 1.6
Fig. 7.33 A uniform rod of length L and mass M is pivoted at . A projectile of mass m moving at
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not included in
the chapter's Creative Commons license and your intended use is not permitted by statutory regulation
or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_8
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
(8.1)
Thus, the total kinetic energy of a rolling object is the sum of the
translational kinetic energy of its center of mass and the rotational
kinetic energy about its center of mass.
Fig. 8.2 As the wheel rotates through an angle , its center of mass moves through a distance
Fig. 8.4 Another way to view rolling without slipping is to consider the wheel to be in pure
rotational motion about an instantaneous axis that passes through the point of contact
Fig. 8.5 A statistical frictional force acts on it at the instantaneous point of contact producing a
torque about the center
In most situations, the body and the surface are not perfectly rigid.
As a result, the normal force would not be a single force; rather it would
be a number of forces that are distributed over the area of contact (see
Fig. 8.6). Therefore, each normal force will exert an opposing torque
since its line of action will not pass through the center of mass.
Furthermore, as the object rolls over the surface, both the object and
the surface undergo deformation resulting in a loss in the mechanical
energy.
Fig. 8.6 If the body and the surface are not perfectly, the normal force would not be a single
force; rather it would be a number of forces that are distributed over the area of contact
Solution 8.1
(a) the total energy is given by
If the hoop slides without rolling its total kinetic energy is , that
is, its value is half of that if the hoop were to roll without slipping.
(b)
Example 8.2 A uniform solid cylinder, sphere, and hoop roll without
slipping from rest at the top of an incline (see Fig. 8.7). Find out which
object would reach the bottom irst.
Fig. 8.7 A uniform solid cylinder, sphere and hoop roll without slipping from rest at the top of an
incline
Solution 8.2
For each object, we have
Hence, the speed of the center of mass of any object at the bottom of the
incline does not depend on its mass or size; it depends only on its
shape. Therefore, all objects of the same shape such as spheres (of any
mass or size) have the same speed at the bottom. That is, the smaller
the ratio the faster the object moves since less of its energy
Fig. 8.8 A marble ball of radius R and mass M rolls without slipping down the incline
Solution 8.3
(a) Applying Newton’s second law in both linear and angular form (see
Fig. 8.7) we have
(8.3)
and
that gives
(8.4)
Substituting Eq. 8.4 into Eq. 8.3 gives
hence
and
Solution 8.4
(a) Applying Newton’s second law in both the linear and angular form
gives
(8.5)
hence
(8.6)
that gives
Solution 8.5
That gives
Fig. 8.10 A block of mass m is attached to a light string that passes over a light pulley connected
to a uniform solid sphere of radius R and mass M
Solution 8.6
From conservation of energy, we have
Since the block and the sphere are connected, they have the same
speed, therefore
Therefore, the speed of the system when the block is at the bottom of
the incline is
or
that gives
(8.7)
(8.8)
(8.11)
This force can be replaced by a single force that is equal to the weight of
the object (Mg) and that acts at a single point called the center of
gravity Now consider an object that is near the earth’s surface where
the force of gravity is assumed to be constant over that range.
Equation 8.11 becomes
To locate the center of gravity, let us calculate the net torque acting on
an object about an origin due to gravity This torque is the vector sum of
the individual torques acting on different mass elements. That is,
Therefore, we conclude that if the gravitational ield (g) is constant over
the body, the center of gravity of the object coincides with its center of
mass.
Fig. 8.13 The resultant gravitational force acting on an object is the resultant of the individual
gravitational forces acting on different mass elements of the object
Solution 8.7
(a) The free-body diagram of the system in shown in Fig. 8.14 where
, and . Applying Newton’s second
law to the beam gives
and
(b) The net external torque about an axis passing through the center of
the beam and perpendicular to the page is
Solution 8.8
(a) Figure 8.15 shows the free-body diagram of the ladder. Applying
Newton’s second law to the ladder gives
and
Applying Newton’s second law in angular form about (the point
must be chosen to give minimum unknowns) we have
(8.12)
hence
(b) The resultant of the contact forces on the ladder at the base is
the direction of is
(c) The free-body diagram is shown in Fig. 8.15. From the equilibrium
condition, we have
and
or
and
thus
Hence
Fig. 8.16 A uniform beam of weight w and length L balanced by two supports
Solution 8.9
The free-body diagram of the system is shown in Fig. 8.16. Because the
beam has a uniform density its center of mass and gravity are located at
its geometrical center. Applying Newton’s second law gives
(8.13)
Taking the torque about an axis passing through one end (at ) gives
(8.14)
and
Solution 8.10
(a) The free-body diagram is shown in Fig. 8.17. Applying Newton’s
second law to the beam gives
(8.15)
and
(8.16)
Dividing Eq. 8.16 by Eq. 8.15 gives
and
Solution 8.11
The free-body diagram is shown in Fig. 8.18. Applying Newton’s second
law to the beam gives
or
(8.17)
Also
or
(8.18)
Taking the resultant torque on the beam about one end (at ) gives
or
and
Hence
or
Also we have
or
Problems
1. A uniform cylinder of mass 3 kg and radius of 0.05 rolls without
slipping along a horizontal surface. Find the total energy of the
cylinder at the instant its speed is 2
2.
A uniform solid cylinder of mass 10 kg and radius of 0.2 rolls up
the incline of angle with an initial velocity of 15 . Find the
height in which the cylinder will stop.
3.
A wheel of mass 2 kg and radius of 0.05 rolls without slipping
with an angular speed of 3 on a horizontal surface. How
much work is required to accelerate the wheel to an angular speed
of 15
4.
A block weighing 1000 is held by a cable that is attached to a
uniform rod of weight 500 (see Fig. 8.20). Find (a) the tension in
the cable, (b) the horizontal and vertical components of the force
exerted on the base of the rod.
5.
A uniform sphere of radius r and mass m is held by a light string
and leans on a frictionless wall as in Fig. 8.21. If the string is
attached a distance d above the center of the sphere, ind (a) the
tension in the string, (b) the reaction force exerted by the wall on
the sphere.
6.
Find the minimum force applied at the top of a wheel of mass M
and radius R to raise it over a step of height h as in Fig. 8.22.
Assume that the wheel does not slip on the step.
7.
Three identical uniform blocks each of length L are on top of each
other as in Fig. 8.23. Find the maximum value of h in order for the
stack to be in equilibrium.
Fig. 8.20 A block suspended by a cable attached to a uniform rod
Fig. 8.21 A uniform sphere suspended by a light string and leaning on a frictionless wall
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_9
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
Example 9.1 Which of the following forces are repulsive and which are
attractive?
Solution 9.1 (a) Attractive, (b) repulsive, and (c) attractive if
and repulsive if
Hence,
or
Thus,
(9.2)
or
or
Thus,
Fig. 9.1 The central force
Fig. 9.2 During an in initesimally small time interval dt, the radius vector sweeps out an area
equal to dA
(9.4)
In Sect. 2. 6, we’ve also seen that the velocity of a particle in polar
coordinates is given by
Therefore, we have
Hence,
(9.5)
and Eq. 9.2 can be written as
or
(9.6)
Let , then . Since , we have . Thus
(9.7)
And
(9.8)
or
(9.9)
origin acts on it
Fig. 9.4 The central force is always acting in the direction of the radial segments and is
perpendicular to the displacement along any of the curved segments. Therefore, the total work
done in moving the particle along any path is equal to the work done along a radial line from to
From this, we conclude that the central force is a conservative force.
You may also prove that . Hence, there exists a potential
energy and the work done by the gravitational force may be written as
Since , we have
or
(9.10)
Since
we have
(9.11)
(9.12)
The gravitational force is effective when one or both the masses are
very large. This is because G is a very small number. Note that, the
gravitational force is not a contact force; it is a ield force that can act
through any medium. The direction of the gravitational force is along
the line joining the two particles.
Therefore, the gravitational force is a central force since its
magnitude is proportional only to the distance between the two
particles (where one of the particles can be considered as the center of
force), and its direction is along the line joining them (toward the
center of force).
Fig. 9.5 Two particles of masses and . Each particle exerts a gravitational force on the
other
That is, the two forces form an action and reaction pair. In terms of unit
vectors, we may write
and
where is a unit vector that is directed along the line joining the two
particles (directed from to ) and is a unit vector directed
from to . The negative sign indicates that the force is attractive.
That is, the force exerted on by will move in the direction
opposite of , i.e., toward . Where the force exerted on by
will move opposite to (toward ). If particle of mass of
interacts with a system of particles, the resultant gravitational force
exerted on particle due to all particles in the system is the vector sum
of the individual forces that each particle in the system exerts on
particle :
where is a unit vector directed from the ith particle in the system
toward the particle and is the force exerted on particle by the
ith particle. If particle of mass m interacts with an extended body of
mass M, the resultant gravitational force exerted on particle is
the vector sum of the individual forces exerted on particle due to
each mass element dM in the object, but in this case, the sum is
replaced by an integral
Solution 9.2
(a)
(b)
If the net force on is zero, we have
or
that gives
From the symmetry of the ring, if a particle (1) on the upper side
exerts a gravitational force on m, there is always another particle
(2) at the opposite side of the ring exerting another force ( ) on the
particle. Because and are equal in magnitude, then their
components cancel each other out and their components add up (see
Fig. 9.7). Thus, the resultant force exerted on m due to all particles of
the sphere is the sum of the components of their forces. Therefore
the resultant force on m is along the direction (toward the center of
the shell). The gravitational force exerted on m by a thin ring of mass
dM is
(9.13)
(9.14)
(9.16)
Thus
(9.17)
Since , it follows that
That is, the spherical shell behaves as a particle of mass M located at its
center.
Case II: A Particle inside the Shell If a particle is inside a uniform
spherical shell, the derivation of the gravitational force exerted on the
particle by the spherical shell is the same as if the particle were outside
the shell, except that the lower integration limit is different. At
since . Thus, we have
where . That is, if the particle is inside the shell, the gravitational
force exerted on it by the shell is zero. However, objects outside the
shell may still exerts forces on the particle. In summary, we have
volume density of the sphere and a is the distance from the shell to the
center of the sphere and da is the thickness of the shell, Hence,
Fig. 9.9 If a particle of mass m is located inside a uniform solid sphere of mass M, then the
gravitational force exerted on the particle is due only to the part of the sphere of radius and
of mass of
(9.19)
or
or
(9.20)
Fig. 9.11 The force exerted on a particle of mass m that is at a distance of a from a thin rod of
mass M and length L
Solution 9.3
(a)
In vector form,
(b) if , then
Solution 9.4
Let us divide the disk into thin concentric rings of radius r and
thickness dr. By symmetry, the resultant force on the particle is directed
along the axis of the ring, since the -components of the forces exerted
by all particles of the ring will cancel out, where their -components
will add up. That is,
or
Solution 9.5
(a)
(b)
(c)
(d)
Solution 9.6
(a)
(b)
(c)
(d)
Example 9.7 A spaceship of mass is moving along a straight line
path between the earth and the sun. At what distance from the center of
the earth will the gravitational force of the sun balances that of the
earth?
Solution 9.7
At that point, we have
or
Solution 9.8
where r is the distance between the center of the earth and the satellite.
That is,
Hence,
exact form of the gravitational force between any two objects was given
earlier in this chapter by Newton’s law of gravity In the case of an
earth–particle system, the gravitational force that each one exerts on
the other is
where is the mass of the earth and m is the mass of the particle
that is at a distance r from the center of the earth. Note that, it is
assumed that the earth is a perfect sphere of uniform mass distribution,
and therefore behaves as a particle. In reality, the earth is not a perfect
sphere but rather an ellipsoid. Furthermore, the earth’s density is not
uniform since it varies with the radius of earth.
The earth’s density also varies at the earth’s surface from one region
to another. In addition, if the earth’s rotation is included, then the
resultant force on an object will be its weight plus the centripetal force
exerted on the object due to the rotation. However, these variations are
often neglected. From the de inition of weight, we have
therefore
(9.21)
As you can see the free-falling acceleration does not depend on the
mass of the object as was predicted before. If the object is falling near
the earth’s surface, then distance r in Eq. 9.21 can be replaced by
which is the radius of the earth and we have
Altitude h (km)
1000 7.34
6000 2.6
10000 1.49
30000 0.3
Altitude h (km)
60000 0.09
Example 9.9 A man can jump vertically upward from the earth’s
surface and reach an altitude of 0.2 . Find the altitude the man can
reach if he jumps with the same initial velocity on the surface of the
moon.
Solution 9.9
Using the formula and by taking at the earth’s
Since the initial velocity of the man is the same on earth and on moon,
we have
That is, the maximum height reached by the man on the moon is six
times the height reached on earth.
Solution 9.10
The gravitational acceleration of a particle near the surface of the star is
Solution 9.11
Solution 9.12
Since all masses are equal, the net gravitational force at is due to the
sum of the -components of and . That is,
Fig. 9.15 Finding the magnitude and direction of the gravitational ield at P
Hence,
(9.22)
and
Since is the length of the major axis which is equal to 2a, (a is the
length of the semimajor axis) we have
(9.23)
or
Hence,
Or
(9.24)
and
(9.25)
The distance C between the center of the ellipse and the focus is
Since from Fig. 9.17, we have , i.e., the distance between the foci is
less than that between the vertices, then . Furthermore, you can
prove that or where b is the length of the
or
(9.26)
This equation is a nonhomogeneous linear differential equation. Its
solution may be given by
(9.27)
or
Thus, the path of the particle under the in luence of the gravitational
force ield is a conic with and and
That is, as the particle of mass m moves toward or away from M, the
potential energy of the system decreases and increases respectively
Note that, the lighter particle (m) gains most of the kinetic energy as
the potential energy changes. By choosing the reference point at in inity
then and taking gives
Solution 9.13
Solution 9.14
(a)
J.
(9.29)
That gives
or
(9.30)
Proof
The gravitational force between the sun and a planet is
where and are the masses of the sun and the planet,
respectively The acceleration of the planet is
Using
Also we have
or
we have
or
This equation is of a conic section and since the only closed conic
section is an ellipse the law is proved.
Proof
This was proved in Sect. 9.1 as a property of a central force, where
we’ve seen that for any central force, the position vector of the
particle from the center of force sweeps out equal areas in equal
times. That is,
or
Here, the center of force is the sun and the particle is the planet, hence
we have
9.4.3 Kepler’s Third Law
The square of the period of revolution of any planet about the sun is
proportional to the cube of the semimajor axis of its orbit.
Proof
The area of an ellipse is given by , where a and b are the
semimajor and semiminor axis of the ellipse, respectively. From
Kepler’s second law, the areal velocity is a constant given by
Also, we’ve seen that the eccentricity for the gravitational force is given
by or in the case of the planet–sun
or
Thus,
or
This proves Kepler’s third law. Note that, Kepler’s laws apply also for
satellites. In such cases, the mass of the sun in the previous equations is
replaced by the earth or any other planet about which the satellite
revolves.
or
(9.32)
(9.33)
, we have
and
Hence the orbit is circular. The potential, kinetic, and total energy as
functions of r of an object in circular orbit are shown in Fig. 9.23.
Fig. 9.23 The potential, kinetic and total energy as functions of r of an object in a circular orbit
Solution 9.15
(9.34)
Body Mass (kg) Radius (m) Semimajor axis a (m) Escape speed (km/s)
Body Mass (kg) Radius (m) Semimajor axis a (m) Escape speed (km/s)
Mercury 4.3
Venus 10.3
Earth 11.2
Mars 5
Jupiter 60
Saturn 36
Uranus 22
Neptune 24
Pluto 1.1
Moon 2.3
Sun 618
Hence
where R is the radius of the planet. If the object’s initial speed is greater
than the escape speed from that planet, then the object will still have
some kinetic energy at in inity. Table.9.2 shows planetary data escape
speeds
Example 9.16 What is the escape speed from the surface of: (a) Earth;
(b) Mars; (c) Pluto.
Solution 9.16
(a)
(b)
(c)
Solution 9.17
The minimum speed is that in which the spacecraft has zero total
mechanical energy at in inity,
Example 9.18 Given that the period of Mars in its orbit about the sun
is 1.88 years and its semimajor axis of the orbit is , ind
Solution 9.18
The period in seconds is
Solution 9.19
From Kepler’s third law,
or
Example 9.20 If Pluto’s distance from the sun at perihelion is
, ind (a) the ratio of its speed at perihelion to its speed at
Solution 9.20
From Table. 9.2, we have , therefore
hence,
(c)
Example 9.21 Two stars of equal mass M revolve about their center of
mass with a speed v as shown in Fig. 9.24. Find the period of motion of
each star.
Fig. 9.24 Two stars of equal mass M revolve about their center of mass with a speed v
Solution 9.21
The gravitational force that one star exerts on the other is
and
Solution 9.22
That gives
Problems
1.
Calculate the gravitational force between the earth and (a) the
sun, (b) the moon.
2.
Calculate the gravitational acceleration at the surface of Mars.
3.
Three particles of masses kg, kg, and kg
are located at the points (0, 0), (0, 5), and (5, 0), respectively. Find
magnitude and direction of the resultant gravitational force
exerted on
4.
The Geosynchronous satellites move in a circular orbit in the
equatorial plane of the earth. They move in such a way that they
always remain over the same point on the earth. Find the height
and velocity of this satellite.
5. If the eccentricity of the orbit of Mercury about the sun is
and its semimajor axis is AU, ind (a) the
distance of its farthest and closest approach to the sun (the
aphelion and perihelion), (b) its period, (c) its total energy, (d) its
angular momentum. .
6.
A body is released at a distance r from the center of the earth.
Find its velocity just as it hits the surface of the earth.
7.
Show that the speed of a satellite in an elliptical orbit about the
earth at apogee and perigee are given by
and
8.
An arti icial satellite moves in an elliptical orbit about the earth.
Its perigee and apogee altitudes are 1100 km and 4100 km
respectively Find (a) the velocity of the satellite at perigee and
apogee, (b) its semimajor axis, (c) its eccentricity, (d) the equation
of its orbit, (e) its period, (f) its speed when it is at a distance of
3000 km above the earth’s surface.
9.
A satellite is at a distance of from the center of the earth.
Find the speed required for the satellite at this altitude (where it
represents the orbit perigee) to be in (a) circular orbit, (b)
parabolic orbit, (c) elliptical orbit of eccentricity of
10.
Suppose the earth suddenly stops moving about the sun, ind the
time it would take the earth to fall to the sun.
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
© The Author(s) 2019
Salma Alrasheed, Principles of Mechanics, Advances in Science, Technology & Innovation: IEREK
Interdisciplinary Series for Sustainable Development
https://doi.org/10.1007/978-3-030-15195-9_10
Salma Alrasheed
Email: salma.alrasheed@kaust.edu.sa
or
(10.1)
where and are arbitrary constants that can be found from the
initial conditions. Therefore, there are many possible motions with the
same angular frequency . By multiplying and dividing Eq. 10.2 by
, you can show that the solution may be written as
(10.3)
Fig. 10.1 A block of mass m attached to a light spring of spring constant k that is ixed at the other
end
or
The frequency is de ined as the number of complete cycles per unit
time
This frequency is called the natural frequency of the motion. The unit of
the frequency is cycles/s or hertz (Hz).
but differing in amplitude. b Two simple harmonic motions of the same frequency and amplitude
but differing in phase by
(10.4)
(10.5)
(10.6)
or
(10.7)
Hence, the velocity and acceleration also vary harmonically with time
with amplitudes and , respectively, but they all have the same
angular frequency From Eqs. 10.5 and 10.7 you can see that the velocity
leads the displacement by or 90. The acceleration on the other
hand leads the velocity by and the displacement by or 180.
Figure 10.4 shows the displacement, velocity, and acceleration versus
time.
Fig. 10.4 The displacement, velocity and acceleration versus time
or
Solution 10.1
(a)
and
(b) At
At
at
(c) at
or
that gives
Example 10.2 A 9 kg object is moving along the -axis under the
in luence of a force given by N. Find (a) the equation of
motion; (b) the displacement of the mass at any time if at
and
Solution 10.2
(a)
hence,
also we have at , or
Solution 10.3
Solution 10.4
Assuming that the earth is a perfect sphere of uniform density and
since the particle is inside the earth, then from Sect. 9. 2, the
gravitational force exerted on the particle by the earth is
Fig. 10.5 A particle of mass m is dropped in a straight tunnel that is drilled through the earth and
which passes through the center of earth
Solution 10.5
The force that each spring exerts on the block acts in the opposite
direction of the displacement, therefore we have
Solution 10.6
The maximum acceleration of the lower block is . In order
for the upper block not to slip, the force of static friction between the
two blocks must produce the same acceleration as the lower block. The
maximum statistical frictional force that can be exerted on the upper
block is and hence, the maximum acceleration that the force of
static friction can produce is . Therefore, . Since
we have
Hence,
and
Thus,
or
(10.10)
or
Hence
where . As the mass moves, its kinetic energy is
transformed into potential energy and vice versa. Figure 10.11 shows
the kinetic energy and potential energy of the system as a function of
time and as a function of the displacement respectively Note that the
variation of U and K with time is at twice the angular frequency of the
variation of x, v, and a with time. This is because the potential energy is
converted to kinetic energy twice in each cycle. The velocity of the
simple harmonic oscillator can be obtained from the total energy of the
system. From Eq. 10.10, we have
Solution 10.7
hence
Solution 10.8
(a)
(b)
(c)
(d)
(e)
(f)
Solution 10.9
therefore
Solution 10.10
At any instant the total mechanical energy is
or
this equation is of a simple harmonic motion with
The minus sign indicates that the torque is a restoring torque, since it
always tends to decrease . The moment of inertia of the bob about an
axis passing through is
From Newton’s second law in angular form, we have
Hence,
or
(10.11)
or
(10.12)
10.3.4.1 Energy
The kinetic energy of the simple pendulum is
Taking the reference point of potential energy of the system to be zero
when the bob is at the bottom, we have
thus
Since
we have
or
or
Solution 10.11
At Mars’s surface, the gravitational acceleration is
Solution 10.12
(a) The amplitude of motion is
Solution 10.13
Taking the potential energy to be zero at the bottom, we have
and
where M is the mass of the body and d is the moment arm of the
tangential component of the weight . From Newton’s second
law, we have
or
Thus,
Therefore, the moment of inertia of a body can be found by measuring
its period when it is in simple harmonic motion as a physical pendulum.
Note that, the simple pendulum is a special case of the physical
pendulum since for a simple pendulum of mass m, the moment of
inertia is
Solution 10.14
Fig. 10.17 A uniform rod suspended at one end oscillated with a small amplitude
Fig. 10.18 A uniform square plate pivoted at one of its corners and oscillates in a vertical plane
Solution 10.15
The moment of inertia of a uniform rectangular plate about its center of
mass is
Thus for a uniform square plate, we have
From the parallel axis theorem, the moment of inertia of the plate about
an axis that is parallel to the center of mass axis and passing through
one corner is
and hence
or
That gives
or
Solution 10.16
for a uniform solid sphere
hence,
Fig. 10.20 A uniform solid sphere suspended at its midpoint by a light string
where b is a positive constant called the damping coef icient. Its SI units
is . The negative sign shows that the direction of the
hence
or
(10.13)
where and . The units of is . This equation
and
size of the damping force. The roots and are either distinct real
roots, equal real roots or a conjugate complex roots. Therefore, there
are three possible motions of the system.
and
where
Since we have
(10.15)
(10.16)
Fig. 10.22 In A lightly damped oscillator, the system oscillates in a decreasing harmonic motion
where the amplitude of motion decreases exponentially with time until eventually the oscillation
dies out
Solution 10.17
(a)
and
or
since at , then
(10.17)
(10.18)
Solving Eqs. 10.17 and 10.18 for A and gives rad and
Therefore,
(b)
In that case, the motion decays without oscillation (see Fig. 10.23) and
the general solution of Eq. 10.13 is
and
or
That gives
Fig. 10.23 In a critically damped motion, the motion decays without oscillation
and
or
where
Fig. 10.24 As critical damping, the resulting motion here is non-periodic but the system returns
to its equilibrium position at large values of t unlike critical damping
critically damped if
Hence,
Thus, the energy decreases with time in any damped motion and the
rate in which it decreases is not uniform.
then, we have
and
That gives
and
and
Hence,
i.e.,
and
where . Hence,
(10.22)
If the driving force is applied for a long time compared with the time
that the damped vibration dies out, then the system will eventually
vibrate at the same frequency of the deriving force. Therefore, the
general solution of Eq. 10.13 is called the transient solution since it
approaches zero in a relativity short time whereas Eq. 10.21 is called
the steady-state solution where the system oscillates with the same
frequency as the deriving force. Therefore, the amplitude of a steady-
state vibration is
When the deriving frequency approaches the natural frequency of
the system , the amplitude of the resulting forced oscillation will
increase. This is known as resonance. If the damping is very light, the
amplitude reaches its peak when the deriving frequency is nearly equal
to the natural frequency . As the damping becomes heavier, the
maximum amplitude shifts to lower frequencies (see Fig. 10.25). In the
case where there is no damping at all , the amplitude of
resonance is in inite at
Fig. 10.25 When the deriving frequency approaches the natural frequency of the system ,
the amplitude of the resulting forced oscillation will increase. This is known as resonance. If the
damping is very light the amplitude reaches its peak when the deriving frequency is nearly equal
to the natural frequency . As the damping becomes heavier, the maximum amplitude shifts to
lower frequencies
Solution 10.19
Hence,
Therefore, the forced vibration has the same frequency as the deriving
force but lag in phase by
Solution 10.20
The amplitude of the forced oscillation when the angular frequency
of the deriving force is varied.
Problems
1. A 2 kg block is fastened to a spring of force constant 98 on a
horizontal frictionless surface. If the block is released a distance
of 6 cm from its equilibrium position, ind (a) the angular
frequency, the frequency and the period of the resulting motion,
(b) the time it takes the block to irst reach cm and its
velocity at that time, (c) the maximum speed and maximum
acceleration of the oscillating block, (d) the total mechanical
energy of the oscillator.
2.
A 10 kg block is attached to a light spring of force constant 200
on a smooth horizontal surface. Find the amplitude of
motion if at the velocity of the block is
3.
A particle rotate counterclockwise in a circle of radius 0.2 with
a constant angular speed of 2 . If at the -coordinate
of the particle is 0.14 , ind the displacement, velocity and
acceleration of the particle at any time.
4.
If a simple pendulum has a period of 2 , ind its period when its
length is increased by .
5.
A simple pendulum of length lm and mass of 0.4 kg oscillates in a
region where . If the amplitude of oscillation is ,
Fig. 10.26 A uniform solid cylinder of radius R and mass M rolls without slipping on a track of
radius 4R
Open Access This chapter is licensed under the terms of the Creative Commons
Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons
license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter's Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder.
References
Alonso, M., Finn, E.J.: Fundamental University Physics: Volume 1 Mechanics. Addison-Wesley
Publishing Company (1967)
Anton, H., Rorres, C.: Elementary Linear Algebra Application Version. Wiley (1994)
Ar ken, G.B., Weber, H.J.: Mathematical Methods for Physicists. Academic Press (1995)
Bueche, F.J., Hecht, E.: Schaum’s Outline of Theory and Problems of College Physics. The McGraw-
Hill Companies Inc. (1997)
Giancoli, D.C.: Physics Principles and Applications. Prentice-Hall International Inc. (1991)
Goldstein, H., Poole, C., Sa ko, J.: Classical Mechanics. Addison Wesley (2002)
Halliday, D., Resnick, R., Walker, J.: Fundamentals of Physics Extended. Wiley (1997)
Halliday, D., Resnick, R., Krane, K.S.: Physics: Volume 1. Wiley (2003)
King, A.R., Regev, O.: Physics with Answers. Cambridge University Press (1997)
Kittel, C., Knight, W.D., Ruderman, M.A., Helmholz, C.A., Moyer, B.J.: Mechanics. McGraw-Hill Inc.
(1973)
Kleppner, D., Kolenkow, R.J.: An Introduction to Mechanics. McGraw-Hill Book Company (1973)
Main, I.G.: Vibrations and Waves in Physics. Cambridge University Press (1994)
Meriam, J.L., Kraige, L.G.: Engineering Mechanics: Volume 2 Dynamics. Wiley (1998)
Narula, G.K.: Physics for Class XI. Vikas Publishing House Pvt, Ltd (1996)
Serway, R.A.: Physics for Scientists and Engineers with Modern Physics. Saunders College
Publishing (1996)
Seto, W.W.: Theory and Problems of Mechanical Vibrations. McGraw-Hill Book Company (1964)
Spiegel, M.R.: Theory and Problems of Theoretical Mechanics. McGraw-Hill Book Company (1982)
Swokowski, E.W., Olinick, M., Pence, D., Cole, J.A.: Calculus. PWS Publishing Company, Boston
(1994)
Young, H.D., Freedman, R.A.: University Physics with Modern Physics. Addison Wesley Longman
Inc. (2000)
Zill, D.G.: A First Course in Differential Equations. PWS-KENT Publishing Company (1993)