Sie sind auf Seite 1von 13

Chapter 1.

1.1 Classical mechanics

Introduction and Newtonian review

What is classical mechanics? Well, the word mechanics refers to the study of the motion of matter, and it is called classical to differentiate it from quantum mechanics, which also deals with the motion of matter but on much smaller size scales. So classical mechanics is just the study of how particles and objects move. This field is typically considered to include studies of the motion of solid material and hypothetical point particles only, since liquids and gases are more easily treated by different methods usually referred to as fluid dynamics. Ok, so its the study of the motion of solid bodies. Blocks sliding down inclined planes, the swing of a pendulum, masses bobbing on springs, sounds like fun. Tremendous fun of course, but theres a little more to it than that. The study of the motion of bodies is the oldest study in the sciences, going back thousands of years. And it took thousands of years for people to finally get it right. Why, because its really hard? Yes and no. To do mechanics, you have to think like a physicist or mathematician. It took a long time for people to learn how to do that. Newton is considered the greatest scientist of all time for (partly) inventing classical mechanics. Seems like a lot for F=ma. But theres a lot more to it than that of course. In this class, you will delve into the history of mechanics as a science. You will examine some of the ideas (right and wrong) that have been proposed along the way. You will look into the lives of the giants of science who made the discoveries. Most of all, you will learn the powerful mathematical techniques that are used to calculate the motions of bodies, techniques which are the necessary foundations of and jumping-off points for more modern fields of study, like quantum mechanics and Einsteins special and general theories of relativity, both of which are just extensions of classical mechanics. By the end of this course, you too will be able to think like a physicist.

1.2 Some terminology


There are a number of specialist terms to describe mechanics, outlined briefly below. The meanings of these terms are sometimes more precise than or different from the common meanings. Mechanics: The branch of physics that is concerned with the analysis of the action of forces on matter or material systems (American Heritage Dictionary). Statics: The mechanics of stationary bodies. Dynamics: The mechanics of objects in motion. Kinetics and kinematics are two other words sometimes used to denote dynamics, however they are subtle differences. Kinetics refers more to the study of the forces involved (eg. in the human body). Kinematics refers to motions that are heavily constrained or predictable (eg. the motion of a person on a roller coaster or other thrill ride).

-1-

1.3 Chapter 1: Newtonian mechanics of a single particle


(Most of the material presented in this chapter is taken from Thornton and Marion, Chap. 2) Newtons work built on that of his predecessors, particularly Galileo (who basically knew Newtons first two laws, though he didnt state them clearly). But Newtons Laws are now the natural starting point of any study of classical mechanics. So lets do a quick review

1.4 Newtons Laws


Newtons Laws are usually simply stated as: I. A body remains at rest or in uniform motion unless acted upon by a force. II. A body acted upon by a force moves in such a manner that the time rate of change of the momentum equals the force. III. If two bodies exert forces on each other, these forces are equal in magnitude and opposite in direction. The First Law is meaningless without the concept of force, but conveys a precise meaning for the concept of zero force. The Second Law is very explicit: Force is the time rate of change of the momentum. But what is the momentum p?
p mv ,

(1.1)

with m the mass, and v the velocity of the particle. We, therefore, re-write the Second Law as
F= dp d = ( mv ) . dt dt

(1.2)

Basically, Newton defined force with this equation. We have an intuitive notion of force, but it can mislead us. For example, if I push on a heavy box, I seem to exert a force but the box may not move. Then if I push harder, the box starts to move. This seems in direct contradiction to Newtons Law, where there should be some motion regardless of the force. This kind of problem endlessly confused early researchers. Youve probably already seen the answer in earlier courses. The problem is that there are often many forces on an object (in the case of the box, the force of friction has a role to play) that makes Newtons Laws tricky to apply in real cases.

-2-

At this point, we have used the concept of mass, but what exactly are we talking about here? We have (again) an intuitive notion of mass, closely connected to weight. However to be precise, we really need some definition of what mass is, or at least a mathematical definition. The Third Law, sometimes called the Law of Action and Reaction really incorporates a definition of what mass means in Newtonian terms. It is rewritten as follows to incorporate the appropriate definition of mass: III'. If two bodies constitute an ideal, isolated system, then the accelerations of these bodies are always in opposite direction, and the ratio of the magnitudes of the accelerations is constant. This constant ratio is the inverse ratio of the masses of the bodies. If we have two isolated bodies, 1 and 2, then the Third Law states that F1 = F2 , and from the Second Law, we have
dp1 dp = 2 , dt dt

(1.3)

(1.4)

or, if the masses are constant (not always the case: consider a rocket losing mass through its exhaust) dv1 dv 2 m1 = m2 , (1.5) dt dt and using the acceleration a

m1a1 = m2 a 2 m1 a = 2. m2 a1 (1.6)

If one chooses m1 as the reference or unit mass, m2, or the mass of any other object, can be measured by comparison (if it is allowed to interact with m1). Incidentally, we can use equation (1.4) to provide a different interpretation of Newtons Second Law
d ( p1 + p 2 ) = 0 dt

(1.7)

or p1 + p 2 = constant. (1.8)

-3-

The momentum is conserved in the interaction of two isolated particles. This is a special case of the conservation of linear momentum. One should note that the Third Law is not a general law of nature. It applies when dealing with central forces (e.g., gravitation (in the non-relativistic limit), electrostatic, etc.), but not necessarily to other types of forces (e.g., velocity-dependent forces such as between to moving electric charges).

1.5 Frames of Reference


It turns out that Newtons Laws implicitly define a frame of reference. This is because when we define a velocity, we have to define it with respect to something. Two cars moving down the highway may be stationary with respect to each other, but moving quickly with respect to the road. This principle that motion is relative is sometimes referred to as Newtonian or Galilean relativity (nothing to do with Einsteins relativity). It turns out that Newtons Laws are NOT valid in all frames of reference. A frame of reference is just something we measure velocity with respect to, and we could define it in any arbitrary way we want (eg. oscillating rapidly, spinning, etc), so we cant expect Newtons Laws to work for all possible reference frames. Defining a suitable reference frame is so hard that we actually approach the problem backwards. A reference frame is called a good frame (technically an inertial frame) if Newtons laws are valid in that frame. More precisely, If a body subject to no external forces moves in a straight line with constant velocity, or remains at rest in a reference frame, then this frame is inertial. If Newtons laws are valid in one reference frame, they are also valid in any other reference frame in uniform motion (or not accelerated) with respect to the first one.

One is tempted to define an inertial reference frame as one which is at rest with respect to space but we cannot measure absolute positions in space without objects as markers and references. Newtonian relativity is full of its own thorny problems. However, the good news is that most frames we encounter are inertial or nearly-inertial frames and so we can use Newtons Laws most of the time. The last bullet point above can be expressed mathematically like this. If the position of a free particle of mass m is represented by r in a first inertial frame, and that a second frame is moving at a constant velocity V2 relative to the first frame, then we can write the position r ' of the particle in the second frame by r ' = r + V2t . The particles velocity v ' in that same frame is (1.9)

-4-

v' =

d dr ( r + V2t ) = + V2 dt dt = v + V2 ,

(1.10)

where v is the velocity of the particle in the first frame. Similarly, we can calculate the particles acceleration in the second frame ( a ' ) as a function of its acceleration (a) in the first one
a' = dv ' d = ( v + V2 ) dt dt dv = = a. dt

(1.11)

We conclude that the second reference frame is inertial since Newtons laws are still valid in it ( F ' = ma ' = ma ). This result is called Galilean invariance, or the principle of Newtonian relativity. Newtons equations do not describe the motion of bodies in non-inertial reference frame (e.g., rotating frames). That is to say, in such frames Newtons Second Law, or the equation of motion, does not have the simple form F = ma .

1.6 Conservation Theorems


We now derive three conservation theorems that are consequences of Newtons Laws of dynamics. 1.6.1 Conservation of linear momentum This theorem was derived in section 1.4 for the case of two interacting isolated particles (see equations (1.7) and (1.8)). We now re-write it more generally from Newtons Second Law (equation (1.2)) for cases where no forces are acting on a (free) particle

F =0p=0

(1.12)

where p is the time derivative of p , the linear momentum. Note that equation (1.12) is a vector equation, and, therefore, applies for each component of the linear momentum. In cases where a force is applied in a well-defined direction, a component of the linear momentum vector may be conserved while another is not. For example, if we consider a constant vector s such that F s = 0 (the force F is in a direction perpendicular to s), then
 

p s = F s = 0. If we integrate with respect to time, we find

(1.13)

-5-

p s = constant.


(1.14)

Equation (1.14) states that the component of linear momentum in a direction in which the forces vanishes is constant in time. 1.6.2 Conservation of angular momentum The angular momentum L of a particle with respect to an origin (in an inertial frame) from which its position vector r is being measured is given by

L rp .
The torque or moment of force N with respect to the same origin is given by
N rF ,

(1.15)

(1.16)

where the force F is being applied at the position r. Because the force is the time derivative of the linear momentum, we can write
 

L
  

dL d = (r p) = ( r p ) + (r p ) , dt dt

(1.17)

but r p = 0 , since r = v and p = mv . We, therefore, find that

L = r p = N .
  

(1.18)

It follows that the angular momentum vector L will be constant in time ( L = 0 ) if no torques are applied to the particle ( N = 0 ). That is, the angular momentum of a particle subject to no torque is conserved. Note that if you measure r with respect to a different origin, you get a different value for L but the conservation still holds. There is usually an obvious origin to choose but you can use whichever you prefer. Note that equation (1.18) is an equation of (angular) motion that can be written in a form similar to F = ma if we substitute N for F , the moment of inertia tensor {I} for m, and the angular acceleration for a . So, if we introduce the angular velocity , which is related to v through v = r , we can write
L = r p = mr v = mr ( r )


(1.19)
!

= m r 2 r ( r )
 

-6-

tensor {1} , we can write = {1} and


#

where we used the vector relation A ( B A ) = A2 B A ( A B ) . Introducing the unit


"

L = m[r 2 {1} r (r {1})] = {I}


$ $ $

(1.20)

with the inertia tensor given by


) ' % ( & A

{I} = m

r 2 {1} r (r {1}) .
0

(1.21)

If we now use equation (1.18), and = , we finally get


N = L = {I} .
2 1

(1.22)

1.6.3 Conservation of energy

If we consider the resultant of all forces (i.e., the total force) applied F to a particle of mass m between two points 1 and 2, we define the work done by this force on the particle by
4 3 6 E

W12

2 1

F dr .

(1.23)

We can rewrite the integrand as


A

F dr = m
A

dv dr dv dt = m vdt dt dt dt m d md 2 = ( v v ) dt = ( v ) dt 2 dt 2 dt 1 2 =d mv . 2
8 @ A 5 7 9

(1.24)

Since equation (1.24) is an exact differential, we can integrate equation (1.23) and find the work done on the particle by the total force

= T2 T1 ,

1 2 W12 = mv 2
D F

=
1

1 2 m ( v2 v12 ) 2

(1.25)

-7-

where T

1 2 mv is the kinetic energy of the particle. The particle has done work 2 when W12 < 0 .

Similarly, we can also define the potential energy of a particle as the work required, from the force F, to transport the particle from point 1 to point 2 when there is no change in its kinetic energy. We call this type of forces conservative (e.g., gravity). That is
2 1
I H

F dr U1 U 2 ,

(1.26)

Where U i is the potential energy at point i . The work done in moving the particle is simply the difference in the potential energy at the two end points. Equation (1.26) can be expressed differently if we consider F as being the gradient of the scalar function U (i.e., the potential energy)
F = U .

(1.27)

The potential energy is, therefore, defined only within an additive constant. Furthermore, in most systems of interest the potential energy is a function of position and time, i.e., U = U ( r , t ) , not the velocity r . We now define the total energy E of a particle as the sum of its kinetic and potential energies (1.28) E T +U. The total derivative of E is
dE dT dU = + . dt dt dt
P

(1.29)

Since we know from equation (1.24) that dT = F dr , then


dT dr =F = F r. dt dt
R S Q

(1.30)

On the other hand


T

dU = dt

U xi U + xi t t
U V

U = ( U ) r + . t

-8-

(1.31)

Inserting equations (1.30) and (1.31) in equation (1.29), we find dE U = F r + ( U ) r + dt t U = ( F + U ) r + t U = . t


W X W X W X

(1.32)

The last step is justified because we are assuming that F is a conservative force (i.e., F = U ). Furthermore, for cases where U is not an explicit function of time, we have
dE = 0. dt U = 0 and t

(1.33)

We can now state the energy conservation theorem as: the total energy of a particle in a conservative field is a constant in time. Finally, we group the three conservation theorems that we derived from Newtons equations: I. The total linear momentum p of a particle is conserved when the total force on it is zero. II. The angular momentum of a particle subject to no torque is constant. III. The total energy of a particle in a conservative field is a constant in time.

These conservations theorems are part of our physicists toolbox: you will use them over and over again in solving various mechanics problems. Note that these tools are inherently more mathematical than physical in some sense. The fact that physics can be explained by mathematics has caused no small amount of wonder in the scientific community. The reason for it remains unclear, though it is at the heart of the power of science. Along with the mathematical nature of these theorems, we should note that there is no physical thing called momentum that you can pull out of a body: it is a mathematical quantity defined relative to some frame of reference. The same applies to angular momentum and even to energy. This starts to get a bit hazy as we now know that energy can be changed into matter and other strange forms

-9-

Problems
(The numbers refer to the problems at the end of Chapter 2 in Thornton and Marion.) 2-2. A particle of mass m is constrained to move on the surface of a sphere of radius R by an applied force F ( , ) . Write the equation of motion. Solution Using spherical coordinates, we can write the force applied to the particle as
F = Fr e r + F e + F e .

(1.34)

But since the particle is constrained to move on the surface of a sphere, there must exist a reaction force Fr er that acts on the particle. Therefore, the total force acting on the particle is
Ftotal = F e + F e = mr.
YY

(1.35)

The position vector of the particle is r = Re r , (1.36)

where R is the radius of the sphere and is constant. The acceleration of the particle is a = r = Re r .
`` `` aa

(1.37)

We must now express e r in terms of e r , e , and e . Because the unit vectors in rectangular coordinates e1 , e 2 , and e3 , do not change with time, it is convenient to make the calculation in terms of these quantities. Using Figure 1.1 for the definition of a spherical coordinate system we get
e r = e1 sin cos + e 2 sin sin + e3 cos e = e1 cos cos + e 2 cos sin e3 sin e = e1 sin + e2 cos .

(1.38)

Then
e r = e1 sin sin + cos cos
2

= e sin + e .
b b

- 10 -

( ) +e ( cos sin + sin cos ) e


b b b b

sin

(1.39)

Figure 1.1 Spherical coordinate system. Similarly, e = e r + e cos e = e r sin e cos . And, further,
c c c c c c

(1.40)

e r = e r 2 sin 2 + 2 + e 2 sin cos + e 2 cos + sin , (1.41)


which is the only second time derivative needed. The total force acting on the particle is Ftotal = mr = mRe r , and the components are
ee ee dd d d d dd d d dd

(1.42)

F = mR 2 sin cos F
ff f f f ff

( ) = mR ( 2 cos + sin )

(1.43)

2-32. A string connects two blocks of unequal mass over a smooth pulley. If the coefficient of friction is k , what angle of the incline allows the masses to move at constant speed?

- 11 -

Figure 1.2 - What inclination angle will make the masses move at constant speed? Solution

The forces on the hanging mass are easily determined from the following figure

The equation of motion is (calling downward positive)

mg T = ma, or
T = m ( g a).

(1.44)

(1.45)

The forces on the other mass can be derived from the following figure

- 12 -

The y equation of motion gives N 2mg cos = my = 0, or N = 2mg cos . The x equation of motion gives ( with Ff = k N = 2k mg cos )
T 2mg sin 2k mg cos = ma. Substituting from (1.45) into (1.48) mg 2mg sin 2k mg cos = 2ma. When = 0 and a = 0 we get g 2 g sin 0 2 k g cos 0 = 0, and
1 = sin 0 + k cos 0 2
gg

(1.46)

(1.47)

(1.48)

(1.49)

(1.50)

= sin 0 + k (1 sin 0 ) .
2 12

(1.51)

Isolating the square root, squaring both sides, and rearranging gives
q p r

Using the quadratic formula gives sin 0 = 1 k 3 + 4 k2 2 (1 + k2 )

- 13 -

(1 + ) sin
2 k

0 sin 0 +

1 k2 = 0. 4

(1.52)

(1.53)

Das könnte Ihnen auch gefallen