Sei sulla pagina 1di 96

L ECTURE NOTES ON C LASSICAL M ECHANICS

Jos Thijssen

Spring 2012

Copyright 2012 by TU Delft An electronic version of these notes is available at http://blackboard.tudelft.nl/.

P REFACE
These notes have been developed over several years for use with the courses Classical and Quantum Mechanics A and B, which for several years have been part of the third year applied physics degree program at Delft University of Technology. Part of these notes stem from courses which I taught at Cardiff University of Wales, UK. Nowadays, in the Delft programme, the subjects of classical and quantum mechanics have been separated again, and I have extracted the part dealing with classical mechanics into this set of lecture notes. These notes are intended to be used alongside standard textbooks. Several texts can be used, such as the books by Morin (Introduction to Classical Mechanics, Cambridge University Press, 2008), Hand and Finch (Analytical Mechanics, Cambridge University Press, 1999) and Goldstein (Classical Mechanics, third edition, Addison Wesley, 2004), the older book by Corben and Stehle (Classical Mechanics, second edition, Dover, 1994, reprint of 1960 edition), and the textbook by Kibble and Berkshire, (Classical Mechanics, 5th edition, World Scientic, 2004). You should be aware that consultation of one or more of the texts mentioned is essential for a thorough understanding of this eld. In the appendix there is a large problem set, which is more essential than the notes themselves. There are many things in life which you can only learn by doing it yourself. Nobody would seriously believe you can master any sport or playing a musical instrument by reading books. For physics, the situation is exactly the same. You have to learn the subject by doing it yourself even by failing to solve a difcult problem you learn a lot, since in that situation you start thinking about the structure of the subject. In writing these notes I had numerous discussions with and advice from Herre van der Zant and Miriam Blaauboer. Part of the problems were chosen and developed or formulated by Miriam, Wim Bouwman and Ad Verbrugge. Wim Bouwman has contributed to the nal A edit. Jos Seldenthuis has provided the LTEX style les with which these notes were formatted. I hope the resulting set of notes and problems will help students learn and appreciate the beautiful theory of classical mechanics. Delft, January 2012

iii

C ONTENTS
Preface 1 Introduction: Newtonian mechanics and conservation laws 1.1 Newtons laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Systems of point particles symmetries and conservation laws . . . . . . . . . . 2 Lagrange and Hamilton formulations 2.1 Generalised coordinates and virtual displacements . . . . . . . . . . . . 2.2 dAlemberts principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 The pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.2 The block on the inclined plane . . . . . . . . . . . . . . . . . . . 2.3.3 Heavy bead on a rotating wire . . . . . . . . . . . . . . . . . . . . 2.4 dAlemberts principle in generalised coordinates . . . . . . . . . . . . . 2.5 Conservative systems the mechanical path . . . . . . . . . . . . . . . . 2.6 Summary and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6.1 Overview of the theory . . . . . . . . . . . . . . . . . . . . . . . . . 2.6.2 A system of pulleys . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6.3 Example: the spinning top . . . . . . . . . . . . . . . . . . . . . . 2.7 Non-conservative forces charged particle in an electromagnetic eld 2.7.1 Charged particle in an electromagnetic eld . . . . . . . . . . . . 2.8 Hamilton mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9 Applications of the Hamiltonian formalism . . . . . . . . . . . . . . . . . 2.9.1 The three-pulley system . . . . . . . . . . . . . . . . . . . . . . . . 2.9.2 The spinning top . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9.3 Charged particle in an electromagnetic eld . . . . . . . . . . . . iii 1 2 3 9 10 12 13 13 14 16 17 18 21 21 23 24 26 26 27 32 32 33 34 35 36 38 40 43 44 44 46 47 47 47 50 52 53 54 55

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

3 The two-body problem 3.1 Formulation and analysis of the two-body problem . . . . . . . . . . . . . . . . . 3.2 Solution of the Kepler problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Examples of variational calculus, constraints 4.1 Variational problems . . . . . . . . . . . . . 4.2 The brachistochrone . . . . . . . . . . . . . 4.3 Fermats principle . . . . . . . . . . . . . . . 4.4 The minimal area problem . . . . . . . . . 4.5 Constraints . . . . . . . . . . . . . . . . . . . 4.5.1 Constraint forces . . . . . . . . . . . 4.5.2 Global constraints . . . . . . . . . . 4.6 Summary . . . . . . . . . . . . . . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

5 Symmetry and conservation laws 5.1 Noethers theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Liouvilles theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

vi 6 Systems close to equilibrium 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 Analysis of a system close to equilibrium . . . . . . . . . . . . 6.2.1 Examples: a spring system and the double pendulum 6.3 Normal modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.1 Normal coordinates . . . . . . . . . . . . . . . . . . . . 6.4 Vibrational analysis in molecules . . . . . . . . . . . . . . . . . 6.5 The chain of particles . . . . . . . . . . . . . . . . . . . . . . . . 6.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A Problems and Exercises Classical Mechanics B Antwoorden

C ONTENTS 59 60 61 62 64 65 67 70 72 73 85

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

1
I NTRODUCTION : N EWTONIAN
MECHANICS AND CONSERVATION LAWS

In this lecture course, we shall introduce some mathematical techniques for studying problems in classical mechanics and apply them to several systems. In a previous course, you have already met Newtons laws and some of its applications. In this chapter, we briey review the basic theory, and consider Newtons laws in some detail. Furthermore, we consider conservation laws of classical mechanics which are connected to symmetries of forces, and derive these conservation laws from Newtons laws. 1

1. I NTRODUCTION : N EWTONIAN MECHANICS AND CONSERVATION LAWS

{ {

1.1 N EWTON S LAWS


The aim of a mechanical theory is to predict the motion of objects. It is convenient to start with point particles which have no dimensions. The trajectory of such a point particle is described by its position at each time. Denoting the spatial position vector by r, the trajectory of the particle is given as r(t ), a three-dimensional function depending on a one-dimensional coordinate: the time. The velocity is dened as the time-derivative of the vector r(t ), and by convention it is denoted as r(t ): d r(t ) = r(t ), (1.1) dt and the acceleration a is dened as the second derivative of the position vector with respect to time: a(t ) = r(t ). (1.2) The last concept we must introduce is that of momentum p: it is dened as p = m(t ), r (1.3)

where m is the mass. Although we have an intuitive idea about the meaning of mass, this is also a rather subtle physical concept, as is clear from the frequent confusion of mass with the concept of weight (see below). Now let us state Newtons laws: 1. A body not inuenced by any other matter will move at constant velocity 2. The rate of change of momentum of a body is equal to the force, F: dp = F(r, t ). dt (1.4)

3. When a particle exerts a force F on another particle, then the other particle exerts a force on the rst particle which is equal in magnitude but opposite in direction to the force F these forces are directed along the line connecting the two particles. Denoting the particle by indices 1 and 2, and the force exerted on 1 by 2 by F1,2 and the force exerted on 2 by 1 by F2,1 , we have: F1,2 = F2,1 = F 1,2 r1,2 . (1.5)

where r1,2 is a unit vector pointing from r1 to r2 . The denotes whether the force is repulsive () or attractive (+). Some remarks about these laws are in place. It is questionable whether the second law is really a statement, as a new vector quantity, called force, is introduced, which is not yet dened. Only if we know the force, we can predict how a particle will move. In that sense, a real law is only formed by combining Newtons second law together with an explicit expression for the force. In table 1.1, known forces are given for several systems. Note that the force generally depends on the position r, on the velocity r, and also explicitly on time (e.g. when an external, time-varying eld is present). An implicit dependence on time is further provided by the time dependence of the position vector r(t ). In most cases, the mass is taken to be constant, although this is not always true: you may think of a rocket burning its fuel, or disposing of its launching system, or bodies moving at a speed of the order of the speed of light, where the mass deviates from the rest mass. With constant mass, the second law reads: m(t ) = F(r, t ). r (1.6)

In fact, the second law disentangles two ingredients of the motion. One is the mass m, which is a property of the moving particle which is acted upon by the force, and the other the

1.2. S YSTEMS OF POINT PARTICLES SYMMETRIES AND CONSERVATION LAWS

TABLE 1.1: Forces for various systems. The symbol m i stand for the mass of point particle i , q i stands for electric charge of particle i , B is a magnetic, and E an electric eld. G, 0 and g are known constants. The gravitational and the electrostatic forces are directed along the line connecting the two particles i = 1, 2.

{ {

Forces in nature System Force Gravity F G = GmM r12 Gravity near earths surface Fg = mg z 1 Electrostatics F C = 4 0 q 1 q 2 r12 Particle in an electromagnetic eld FEM = q (E + r B) Air friction Ffr = r

force, which itself arises from some external origin. In the case of gravitational interaction, the force depends on the mass, which drops out of the equation of motion. Generally, mass can be described as the resistance to velocity change, as the second law states that the larger the mass, the smaller the change in velocity (for the same force). It is an experimental fact that the mass which enters the expression for the gravitational force is the same as this universal quantity mass, which occurs for any force in the equation of motion. The weight is the gravity force acting on a body. Usually, the rst law is phrased as follows: when there is no force acting on a point particle, the particle moves at constant velocity. This statement obviously follows from the second law by taking F = 0. The formulation adopted above (page 2) emphasises that force has a material origin. It is impossible to fulll the requirements of this law, as everywhere in the universe gravitational forces are present: the rst law is an idealisation. The rst law is not obvious from everyday life, where it is never possible to switch friction off completely: in everyday life, motion requires a force in order to be maintained. The third law is a statement about forces. It turns out that this statement does not hold exactly, as the forces of this statement should act simultaneously. In quantum eld theory, particles travelling between the interacting particles are held responsible for the interactions, and these particles cannot travel at a speed faster than that of light in vacuum (about 3 108 m/s). However, for everyday life mechanics, the third law holds to sufcient precision, unless the moving particles carry a charge and interact through electromagnetic interactions. In that case, the force acts no longer along the line connecting the two particles.

1.2 S YSTEMS OF POINT PARTICLES SYMMETRIES AND CONSERVATION


LAWS Real objects which we describe in mechanics are not point particles, but to very good accuracy they can be considered as large collections of interacting point particles in this section we consider systems consisting of N point particles. It is possible to disentangle the mutual forces acting between these particles from the external ones. The mutual forces satisfy Newtons third law: for every force Fi , j , which is the force exerted by particle j on particle i , the force F j ,i is equal in magnitude but opposite in direction to Fi , j . For a particle i , we consider all the mutual forces Fi , j for j = i the remaining forces on i must then be due to external sources (i.e., not depending on the other particles in our system), and we lump these forces together in one external force, FExt : i
N

Fi =
j =1; j =i

Fi , j + FExt , i

i = 1, . . . , N .

(1.7)

1. I NTRODUCTION : N EWTONIAN MECHANICS AND CONSERVATION LAWS

{ {

The equations of motion read:


N

m i ri =
j =1; j =i

Fi , j + FExt . i

(1.8)

The total momentum of the system is the sum of the momenta of all the particles:
N N

p=
i =1

pi =
i =1

m i ri .

(1.9)

We can view the total momentum of the system as the momentum of a single particle with a mass equal to the total mass M of the system, and position vector rC . This position vector is then dened through:
N N

p = M rC =
i =1

m i ri ;

M=
i =1

mi .

(1.10)

This is equivalent to rC = 1 M
N

m i ri
i =1

(1.11)

up to an integration constant which is always taken to be zero. The vector rC is called centre of mass of the system. A particle of mass M at the centre of mass (which obviously changes in time) represents the same momentum as the total momentum of the system. Let us nd an equation of motion for the centre of mass. We do this by summing Eq. (1.8) over i :
N N N

m i ri =
i =1 i , j =1,i = j

Fi , j +
i =1

FExt . i

(1.12)

In the rst term on the right hand side, for every term Fi , j , there will also be a term F j ,i , but this is equal in magnitude and opposite to Fi , j ! So the rst term vanishes, and we are left with
N N

m i ri = p =
i =1 i =1

FExt FExt . i

(1.13)

We see that the centre of mass behaves as a point particle with mass M subject to the total external force acting on the system. Conservation of physical quantities, such as energy, momentum etcetera, is always the result of some symmetry. This deep relation is borne out in a beautiful theorem, formulated by E. Noether, which we shall consider in the next semester. In this section we shall derive three conservation properties from Newtons laws and the appropriate symmetries. The rst symmetry we consider is that of a system of particles experiencing only mutual forces, and no external ones. We then see immediately from Eq. (1.13) with FExt = 0 that p = 0, in other words, the total momentum is conserved. Conservation of momentum In a system consisting of interacting particles, not subject to an external force, the total momentum is always conserved. Next, let us consider the angular momentum L. This is a vector quantity, which for particle i is dened as Li = ri pi . The total angular momentum L is the sum of the vectors Li :
N

L=
i =1

Li .

(1.14)

To see how L varies in time, we calculate the time derivative of Li : Li = ri pi + ri pi . (1.15)

1.2. S YSTEMS OF POINT PARTICLES SYMMETRIES AND CONSERVATION LAWS

r2 dr

{ {

r1

F IGURE 1.1: Path from r1 to r2 . The force at some point along the path is shown, together with the contribution to the work of a small segment d r along the path.

The rst term of the right hand side vanishes because pi is parallel to ri , so we are left with Li = ri pi = Ni ; (1.16)

Ni is the torque acting on particle i . Now we calculate the torque on the total system by summing over i and replacing pi by the force (according to the second law): L=
N N N

ri F i =
i =1 i =1

ri
j =1, j =i

Fi , j + FExt . i

(1.17)

The rst term in right hand side vanishes, again as a result of the third law: ri Fi , j + r j F j ,i = Fi , j ri r j = 0, (1.18)

where the last equality is a result of the direction of Fi , j coinciding with that of the line connecting ri and r j (which excludes electromagnetic interactions between moving particles from the discussion). We therefore have L=
N i =1

ri FExt . i

(1.19)

We see that if the external forces vanish, the angular momentum does not change: Conservation of angular momentum In a system consisting of interacting particles (not electromagnetic), not subject to an external force, the angular momentum is always conserved. Finally, we consider the energy. Let us evaluate the work W done by moving a single particle from r1 to r2 along some path (see gure 1.1). This is by denition the inner product of the force and the innitesimal displacements, summed over the path:
t2

W=

F d r(t ) =

t1

Fr dt.

(1.20)

Using Newtons second law, we can write:


t2 t2

W=

t1

mrd t = r

t1

m 2 2 m d 2 r dt = r r , 2 dt 2 2 1

(1.21)

where r1 is the velocity at time t 1 and similar for r2 . We see that from Newtons second law it follows that the work done along the path is equal to the change in the kinetic energy T = m2 /2. r

1. I NTRODUCTION : N EWTONIAN MECHANICS AND CONSERVATION LAWS

{ {

A conservation law can be derived for the case where F is a conservative and time-independent force. This means that F can be written as the negative gradient of some scalar function, called the potential:1 F(r) = V (r). (1.22) In that case we can write the work in a different way:
t2

W =

V (r)d r(t ) =

t1

dV (r) d t = V (r1 ) V (r2 ). dt

(1.23)

From this and from Eq. (1.21) it follows that T 1 + V1 = T 2 + V2 , (1.24)

where T1 is the kinetic energy in the point r1 (or at the time t 1 ) etcetera. Thus T + V is a conserved quantity, which we call the energy E . Of course, now that we know the expression for the energy, we can verify that it is a conserved quantity by calculating its time derivative, using Newtons second law: E = m r + V (r) r = F r F r = 0. r (1.25)

For a many-particle system, the derivation is similar the condition on the force is then that there exists a potential function V (r1 , r2 , . . . , rN ), such that the force Fi on particle i is given by Fi = i V (r1 , r2 , . . . , rN ). (1.26) Note that V depends on 3N coordinates the gradient i acting on V gives a 3-dimensional vector V V V i V = , , . (1.27) x i y i z i The kinetic energy is the sum of the one-particle kinetic energies. Now the energy conservation is derived as follows: E=
i

m i ri r i +
i

i V ri =
i

(Fi ri Fi ri ) = 0.

(1.28)

The function V above depends on time only through the time-dependence of the arguments ri . If we consider a charged particle in a time-dependent electric eld, this is no longer the case: then t occurs as an additional, explicit argument in V . If V would depend explicitly on time, the energy would change at a rate E= V (r1 , r2 , . . . , rN , t ), t (1.29)

where the arguments ri also depend on time (but do not take part in the differentiation with respect to t ). If V does not depend explicitly on time, we can dene the zero of time (i.e. the time when we set our clock to zero) arbitrarily. This time translation invariance is essential for having conservation of energy. Similarly, the conservation of momentum is related to space translation invariance of the potential, i.e. this potential should not change when we translate all particles all over the same vector. Finally, angular momentum is related to rotational symmetry of the potential. In quantum mechanics, all these symmetries lead to the same conserved quantities (or rather their quantum mechanical analogues).
1 From vector calculus it is known that a necessary and sufcient condition for this to be possible is that the force

is curl-free, i.e. F = 0.

1.2. S YSTEMS OF POINT PARTICLES SYMMETRIES AND CONSERVATION LAWS There exists a correspondence between symmetries and conservation laws. We have considered the following examples of this correspondence: Spatial translation symmetry implies conservation of the total momentum P. Rational symmetry implies conservation of angular momentum L. If the forces in a system are conservative, time translation symmetry implies conservation of total energy E .

{ {

A nal remark concerns the evaluation of the kinetic energy of a many-particle system. As we have seen above, the motion of the centre of mass can be split off from the analysis in a suitable way. This procedure also works for the kinetic energy. Let us decompose the position vector ri of particle i into two parts: the centre of mass position vector rC and the position relative to the centre of mass, which we call ri : ri = rC + ri . As, by denition, rC =
i

(1.30)

m i ri /M , we have m i ri = m i ri M rC = 0.
i

(1.31)

We can use this decomposition to rewrite the kinetic energy: T=


i

mi rC + ri 2

M 2 r + rC 2 C

m i ri +

mi 2 r . 2 i

(1.32)

The second term vanishes as a result of (1.31) and therefore we have succeeded in writing the kinetic energy of the many-particle system as the kinetic energy of the centre of mass plus the kinetic energy of the relative coordinates: T = TCM +
i

mi 2 r . 2 i

(1.33)

This formula is a convenient device for calculating the kinetic energy in many applications. In a many-particle system, it is often useful to split the velocities into two components: the centre of mass velocity, and the velocities relative to the centre of mass. The kinetic energy can be written as the sum of the centre of mass kinetic energy and the internal kinetic energy, i.e. the kinetic energy associated with the relative velocities.

2
L AGRANGE AND H AMILTON
FORMULATIONS OF CLASSICAL MECHANICS

10

2. L AGRANGE AND H AMILTON FORMULATIONS


z y x

2{

F IGURE 2.1: The pendulum in three dimensions. The position of the mass is described by the two angles and .

The laws of classical mechanics, formulated by Newton, and the various laws for the forces (see table 1.1) supply sufcient ingredients for predicting the motion of mechanical systems in the classical limit. Working out the solution for particular cases is not always easy, however. In this chapter we shall develop an alternative formulation of the laws of classical mechanics, which renders the analysis of many systems easier than the traditional Newtonian formulation, in particular when the moving particles are subject to constraints. The new formulation will not only enable us to analyse new applications more easily than using Newtons laws, but it also leads to an important example of a variational formulation of a physical theory. Broadly speaking, in a variational formulation, a physical solution is found by minimising a mathematical expression involving a function by varying that function. Many physical theories can be formulated in a variational way, in particular quantum mechanics and electrodynamics.

2.1 G ENERALISED COORDINATES AND VIRTUAL DISPLACEMENTS


When observing motion in everyday life, we often encounter systems in which the moving particles are subject to constraints. For example, when a car moves on the road, the road surface withholds the car from moving downward, as is the case with the balls on a billiard table. Another example is a particle suspended on a rigid rod (i.e. the pendulum), which can only move on the the sphere around the suspension point with radius equal to the rod length. The constraints are realised by forces, which we call the forces of constraint. The forces of constraint guarantee that the constraints are met they often do not inuence the motion within the subspace.1 The main object of the next few sections is to show that it is possible to eliminate these constraint forces from the description of the mechanical problem. As the presence of constraints reduces the actual degrees of freedom of the system, it is useful to use a smaller set of degrees of freedom to describe the system. As an example, consider a ball on a billiard table. In that case, the z-coordinate drops out of the description, and we are left with the x and y coordinates only. This is obviously a very simple example, in which one of the Cartesian coordinates is simply left out of the description of the system. More interesting is a ball suspended on a rod. In that case we can use the angular coordinates and to describe the system that is, we replace the coordinates x, y and z by the angles and see gure 2.1. In this case, we see that the coordinates no longer represent distances, that is, they do not have the dimension of length, but rather they are values of angles, and therefore dimensionless. This is the reason why we speak of generalised coordinates. These coordinates form a reduced representation of a system subject to constraints. In chapter 2 of the Schaum book you nd many examples of constraints and generalised coordinates. Generalised constraints are denoted by q j , where j is an index which runs over the degrees of freedom of the constrained system.
1 The subspace on which the particle is allowed to move is not necessarily a linear subspace, e.g. the spherical

subspace in the case of a pendulum. Mathematicians would use the term submanifold rather than subspace.

2.1. G ENERALISED COORDINATES AND VIRTUAL DISPLACEMENTS

11

We now shall look at constraints and generalised coordinates from a more formal viewpoint. Let us consider a system consisting of N particles in 3 dimensions, so that the total number of coordinates is 3N . The system is subject to a number of constraints, which are of the form g (k) (r1 , . . . , rN , t ) = 0, k = 1, . . . , K . (2.1) Constraints of this form (i.e. independent of the velocities) are called holonomic. Usually, it is then possible to transform the 3N degrees of freedom to a reduced set of 3N K generalised coordinates {q} = q j , j = 1, . . . , 3N K . It is now possible to express the position vectors in terms of these new coordinates: ri = ri ({q}, t ). (2.2) As an example, consider the particle suspended on a rod; see gure 2.2. The Cartesian coordinates are x, y and z and they can be written in terms of the generalised coordinates and as: x = l sin cos ; y = l sin sin ; z = l cos , (2.3) (2.4) (2.5)

{ {

where l is the length of the rod (and therefore xed). These equations are a particular example of Eqs. (2.2). The velocity can be expressed in terms of the q j using the chain rule:
3N K

ri =
j =1

ri ri qj + . q j t

(2.6)

If we now take the partical derivative with respect to q j in the left and right hand side, we nd: i r ri = , (2.7) q j q j a result which will be very useful further on. Newtons laws predict the evolution of a mechanical system without ambiguity from a given initial state (if that state is not on an unstable point, such as zero velocity at the top of a hill). However, we are sometimes interested in a variation of the path of a system, i.e. a displacement of one or more particles in some direction. Such displacements are called virtual displacements in order to distinguish them from the actual displacement, which is always governed by the Newton equations of motion. If we now generalise the denition of work, Eq. (1.20) to include virtual displacements ri rather than the mechanical displacements which actually take place, then the work done due to this displacement is dened as W =
N i =1

F ri .

(2.8)

The notion of virtual work is very important in the following section. Summarizing this section, we have A system consisting of N point particles has 3N coordinates, usually given as ri , i = 1, . . . , N . If the system is subject to S independen constraints, it can be described by K = 3N K new variables, called generalised coordinates. We may consider instantaneous displacements ri of the particles in the system and assign a virtual work W to these displacements as in Eq. (2.8).

12

2. L AGRANGE AND H AMILTON FORMULATIONS

2.2 D A LEMBERT S PRINCIPLE


We start from Newtons law of motion for an N -particle system. pi = mi = Fi , r i = 1, . . . , N . (2.9)

2{

It is always possible to decompose the total force on a particle into a force of constraint FC and the remaining force, which we call the applied force FA : F = FC + FA . (2.10)

If you consider any system consisting of a single particle (or nonrotating rigid body), subject to constraints, you will nd that the work forces of constraint are always perpendicular to the space in which the particle is allowed to move. For example, if a particle is attached to a rigid rod which is suspended such that it can rotate freely, the particle can only rotate on a spherical surface. The force of constraint, which is the tension in the rod, is always normal to that surface. Similarly, the force of the billiard table on the balls is always vertical, i.e. perpendicular to the plane of motion. This notion provides a way to eliminate these forces from the description. Consider an arbitrary but small virtual displacement r within the subspace allowed by the constraint. Because the force of constraint is perpendicular to this subspace, we have: p r = FC + FA r = FA r. (2.11) We see that the force of constraint drops out of the system, and we are left with a motion determined by the applied force only. Because (2.11) holds for every small r, we have p = FA (2.12)

if we restrict both p and FA to be tangential to the constraint subspace. The principle we have formulated in Eq. (2.11) is called dAlemberts principle. For systems consisting of a single rigid body, it expresses the fact that the forces of constraint are perpendicular to the subspace of the constraint. This is a nontrivial fact: it can be veried for particular systems and can be used as a general rule for handling constraints. The expression F r is the virtual work done as a result of the virtual displacement. It is important to note that the virtual displacements are always considered to be spatial the time is not changed. This is particularly important in cases where the constraints are time-dependent. In the next section we shall consider an example of this. For more than one object, the contributions to the virtual work must be added, so that we obtain:
N i

pi ri =
i

FA ri . i

(2.13)

In this form, the contributions of the constraint forces to the virtual work do not all vanish for each individual object, but the total virtual work due to the constraint forces vanishes:
i

FC ri = 0. i

(2.14)

In summary, we can formulate dAlemberts principle in the following, concise form: According to dAlemberts principlem the virtual work due to the forces of constraint is always zero for virtual displacements which do not violate the constraint. The use of dAlemberts principle can simplify the analysis of systems subject to constraints, although we often use this principle tacitly in tackling problems in the Newtonian approach. In that approach we usually demand that the forces of constraint balance the components of the applied force perpendicular to the constraint subspace. Nevertheless, it is convenient to skip this step, using dAlemberts principle, especially in complicated problems (many applied forces and constraints).

2.3. E XAMPLES

13

l FT m F g

{ {

F IGURE 2.2: The pendulum moving in a plane. The rod of length is rigid, massless, and is suspended frictionless.

2.3 E XAMPLES
2.3.1 T HE PENDULUM
As a simple example, let us consider a pendulum moving in a plane. This system is shown in gure 2.2. Using Newtons mechanics, we say that the ball of mass m is kept on the circle by the tension in the suspension rod. This tension is directed along the rod, and it precisely compensates the component of the gravitational force along the same line. The component of the gravitational force tangential to the circle of motion determines the motion. The distance traveled by the pendulum is given by r (t ) = l (t ), where is the angle shown in the gure. The speed is therefore given by r (t ) = l (t ), and the equation of motion is l (t ) = g sin (t ). (2.15)

Using dAlemberts principle simplies the rst part of this analysis. We can simply say that the motion is determined by the component of the applied force (i.e. gravity) lying in the subspace of the motion (i.e. the circle) and this leads to the same equation of motion. Although in this simple case the difference between the approaches with and without dAlemberts principle is minute, in more complicated systems, the possibility to avoid analysing the forces of constraint is a real gain. Let us redo this calculation in terms of the Cartesian components x, y and z. These are given by x(t ) = l sin (t ) y(t ) = l cos (t ) z(t ) = 0 (2.16a) (2.16b) (2.16c)

where the last equation denes the plane of the pendulum motion as the x y-plane. These equations clearly show how the r(t ) as 3 Cartesian coordinates can be written in terms of a single generalised coordinate , as a result of the two constraints: motion in one plane (z = 0) and the length l = x 2 + y 2 xed. We can now calculate the velocities: x(t ) = l (t ) cos (t ) y(t ) = l (t ) sin (t ) z(t ) = 0. From this we immediately see that v= x 2 + y 2 + z 2 = l . (2.18) (2.17a) (2.17b) (2.17c)

14

2. L AGRANGE AND H AMILTON FORMULATIONS

F 2

F1

2{
-F 1 F z

SB

^ n

^ d

IP

^ y ^ x

FZ

F IGURE 2.3: Small block on an inclined plane.

Similarly, we nd for the acceleration a = l , a result we used above.

2.3.2 T HE BLOCK ON THE INCLINED PLANE


Now we consider a more complicated example: that of a block sliding on a wedge. We shall denote the block by SB (small block) and the wedge by IP (Inclined plane). The setup is shown in gure 2.3. It consists of the wedge (inclined plane) of mass M which can move freely (i.e. without friction) over a horizontal table, and the small block off mass m, which can slide over the inclined plane (also frictionless). The aim is to nd expressions for the accelerations of IP and SB. The Cartesian unit vectors are x and y, and the unit vector along the inclined plane, pointing to the right, is d, and the upward normal unit vector to the plane is called n. Let us solve this problem using the standard approach. The acceleration of IP is called A, and that of the small block is A + a, i.e., a is the acceleration of the small block with respect to the inclined plane. Newtons second law for the two bodies reads: M A = M g y + F 2 y F 1 n, m(A + a) = mg y + F 1 n. (2.19a) (2.19b)

As we know that the motion of IP is horizontal, we know that all y components of the forces acting on it will cancel, and A is directed along x. Similarly, we know that a is zero along n. This allows us to simplify the equations: M A = F 1 sin ; = mg y + F 1 n m(A x + a d) (2.20a) (2.20b)

where a is the component of a directed along d. The rst of these equations is a scalar equation. The second equation represents in fact two equations, one for the x and one for the y component. We have three unknowns: A, a and F 1 . Translating d and n into the x- and ycomponents is straightforward, and (2.20b) becomes: m(A + a cos ) = F 1 sin , ma sin = F 1 cos mg . (2.21a) (2.21b)

2.3. E XAMPLES Now we can solve for the accelerations by eliminating F 1 from our equations, and we nd: a =g (M + m) sin M + m sin2 m sin cos M + m sin2 ; .

15

(2.22a) (2.22b)
{ {

A = g

The solution of this problem involved one nontrivial step: the fact that we have split the acceleration of SB into the acceleration of IP plus the acceleration of the SB with respect to the IP has enabled us to remove the latters component along n. This is not so easy a step when a different representation is used (e.g. when the acceleration is not split into these parts). Now we turn to the solution using dAlemberts principle: pSB rSB + pIP rIP = FA rSB + FA rIP . SB IP (2.23)

We identify two natural candidates to be used as generalized coordinates: the coordinate X of the IP along the horizontal direction, and the distance d from the top of the IP to the SB. The total virtual work done as a result of displacements X and d is the sum of the virtual work done by each body: rSB = d d + X x and rIP = X x. (2.24)

The applied forces are the gravity forces we do not care about constraint forces any longer and we nd FA rIP = 0, (2.25) IP as the displacement is perpendicular to the applied (gravity) force. Furthermore FA rSB = mg sin d . SB On the other hand: pSB = m( X x + d d) and pIP = M X x, (2.28) so that pIP = M A x and pSB = m(A x + a d). Taking time derivatives of (2.27) and (2.28) and using dAlemberts equations (2.23) for this problem, together with (2.25) and (2.26), we obtain m A X + ma d + m A cos d + ma cos X + M A X = mg sin d . (2.29) (2.27) (2.26)

This equation should hold for any pair of virtual displacements X and d . We may therefore take X = 0 or d = 0. From this we see that the coefcients of both X and d should each vanish individually, giving the equations: (m + M )A + ma cos = 0. m(a + A cos ) = mg sin . (2.30a) (2.30b)

Not surprisingly, these equations lead to the same result (2.22) as obtained before. Although the second approach using dAlemberts principle does not seem simpler, it is less prone to errors since the constraint forces do not have to be taken into account explicitly. This manifests itself explicitly in the fact that we do not have to eliminate the constraint force F 1 as in the direct approach. An important remark concerns this neglect of the constraint force: the inclined plane moves along the x-axis therefore, the constraint force F1 is not perpendicular to the displacement of the inclined plane. This seems to be in contradiction with what we said above about all constrained forces being perpendicular to the allowed motion. However, when the

16

2. L AGRANGE AND H AMILTON FORMULATIONS

^ q

{ {

^ t

y x
F IGURE 2.4: Bead on a rotating wire.

inclined plane moves over a distance x in the horizontal direction, the displacement of the block can be decomposed into (i) the same displacement x in the horizontal direction plus (ii) a distance d over which the block slides over the wedges surface. The second displacement is perpendicular to the constraint force F1 , but as the constraint force on the block is opposite to the constraint force F1 on the inclined plane, the rst displacement gives a contribution to the virtual work which is opposite to that on the inclined plane. Therefore the total work done by the constraint forces is zero: note every displacement in the system is perpendicular to the constraint force, but the sum of all contributions to the virtual work vanishes.

2.3.3 H EAVY BEAD ON A ROTATING WIRE


In this section, we consider a system with a time-dependent constraint. A bead slides without friction along a straight wire which rotates along a vertical axis, under an angle (see gure 2.4). The position of the bead along the wire is denoted by q, which is the distance of the bead from the origin. The momentum of the bead is given by p = mq sin + m q q t (2.31)

It should however be noted that the unit vectors and q rotate themselves, and hence their t time derivatives occur in p. The latter occurs in dAlemberts equation, in which gravity enters as the applied force FA . Instead of working out p explicitly, we can use the following trick: p r = d (p r) p . r dt (2.32)

At rst sight, you might think that the second term on the right hand side is zero as r = q q and q does not involve any time dependence: virtual displacements are always assumed to be instantaneous and do not involve any time dependence. However, even with a timeindependent q, the displacement r is time-dependent as the displacement is carried out in a rotating frame. This can also be seen from the fact that q is time-dependent. In fact, in our system the displacement along the wire will cause a change in the rotational velocity, and it is this velocity change which gives . If the bead is moved upward, for example, the bead r will move along a circle which has a larger radius, but still at the same angular velocity, so that the orbital speed increases. The orbital speed is given as q sin , so that we have: = sin q . r t As r is given by q q, we nd p r = m q q m2 sin2 q q = Fa r = mg cos q (2.34) (2.33)

2.4.

D A LEMBERT S PRINCIPLE IN GENERALISED COORDINATES

17

and we nd the equation of motion: q 2 sin2 q = g cos . The solution to this equation can be found straightforwardly: q(t ) = q 0 + Ae
t

(2.35)
{ {

+Be

(2.36)

with q 0 = g cot /(2 sin ), A and B arbitrary constants and = sin . This is quite a peculiar behaviour: as soon as we release the bead a bit away from the position where the centrifugal and gravity forces are in balance, it will run away exponentially fast from this position as the rst term in the solution grows with time. This is reasonable if you think of it: displacing the bead to a higher position increases the centrifugal force; therefore the bead will be pushed even further outward. The same holds for a downward displacement, which causes the gravity force to become dominant and pull the bead further downward. Later we shall encounter more powerful techniques which enable us to solve such a problem more easily.

2.4 D A LEMBERT S PRINCIPLE IN GENERALISED COORDINATES


In the previous section we have encountered a few examples of systems subject to constraints, and analysed them using dAlemberts principle. In this section we shall do the same for an unspecied system and derive the equations of motion for a general constrained system using dAlemberts principle. We start from dAlemberts equation for N objects:
N i =1

pi ri =

N i =1

FiA ri .

(2.37)

Now let us consider a small displacement. This can be described in terms of a small change in the generalised coordinates q j q j + q j . (2.38) The Cartesian coordinates ri are functions of the q j . Therefore we can express the changes in these ri via a rst order Taylo expansion in the q j : ri =
3N K j =1

ri q j . q j

(2.39)

Now we realise ourselves that the q j can be chosen independently (provided they are small). Choosing all but one of them zero, we see that the coefcients of each q j in dAlemberts principle must be identical on the left and right hand side in equation (2.37). This leads to
N

pi
i =1

N ri ri = FA . q j i =1 i q j

(2.40)

This is just a new formulation of dAlemberts principle. In order to make progress, we use the sum rule for differentiation, similar to the one we applied already to the bead sliding along the wire. Using the product rule for differentiation:
N

pi
i =1

ri d = q j d t

pi
i =1

N ri d ri pi . q j d t q j i =1

(2.41)

We note furthermore that in time derivative in the second term on the right hand side, we can swap the derivative with respect to time with the derivative with respect to q j to obtain: i r d ri = . d t q j q j (2.42)

18

2. L AGRANGE AND H AMILTON FORMULATIONS

We see that we can formulate Eq. (2.40) in the form d dt


N

pi
i =1

N N i r ri ri pi FA = . q j q j i =1 i q j i =1

(2.43)

2{

In section 1.2 we have seen that the work done equals the change in kinetic energy. This suggests that the kinetic energy might be a convenient device for expressing dAlemberts equation in generalised coordinates. To see that this is indeed the case, we rst calculate the derivative of the kinetic energy with respect to q j :
N T i r = m i ri . q j i =1 q j

(2.44)

Then we calculate its derivative with respect to q j :


N N T i r ri mi r pi = = , q j i =1 q j i =1 q j

(2.45)

where we have used (2.7). If we stare long enough at these derivatives and at (2.43), we see that the left hand side of dAlemberts equation as in (2.43) can be written as d T T . d t q j q j Lets now tackle the right hand side of Eq. (2.43). Dening
N i =1

(2.46)

FA i

ri = Fj , q j

(2.47)

where F j is the generalised force, we have the following Formulation for dAlemberts principle in generalised coordinates: d T T = Fj . d t q j q j (2.48)

There is no sum over j in this equation because the variations q j are arbitrary and independent. It is then possible to obtain the form (2.48) from dAlemberts principle by taking only one particular q j to be nonzero.

2.5 C ONSERVATIVE SYSTEMS THE MECHANICAL PATH


Consider now a particle which moves in a constrained subspace under the inuence of a potential. As an example you can imagine a non-at surface on which a ball is moving from r1 to r2 . If the ball is not forced to obey the laws of mechanics, it can move from r1 at time t 1 to r2 at time t 2 along many different paths. Instead of approaching the problem of nding the motion of the ball from a time-evolution point of view, where we update the position and the velocity of a particle at each innitesimal time step, we consider the path allowed for by the laws of mechanics2 as a special one among all the available paths from r1 at t 1 to r2 at t 2 . We thus try to nd a condition on the path as a whole rather than for each of its innitesimal segments. To this end, we start from dAlemberts principle, and apply it to two paths, ra (t ) and rb (t ), which are close together for all times. The difference between the two paths at any time t between t 1 and t 2 is r(t ) = rb (t ) ra (t ), and we write down dAlemberts principle at time t using this r(t ): m(t ) r(t ) = F r(t ), r (2.49)
2 This path is not always, but nearly always, unique.

2.5. C ONSERVATIVE SYSTEMS THE MECHANICAL PATH

19

where it is understood that F is the applied force only, as r lies in the constrained subspace.3 This equation holds for every t between t 1 and t 2 , and we can formulate a global condition on the path by integrating it over time from t 1 to t 2 :
t2 t1

m(t ) r(t )d t = r

t2 t1

F r(t )d t .

(2.50)

{ {

The analysis which follows resembles that of the previous chapter when we derived the conservation property of the energy. Indeed, the right hand side looks like an expression for the work, but it should be kept in mind that r is not a real displacement of the particle, but a difference between two possible paths. Via partial integration, and using the fact that the begin and end point of the path are xed, we can transform the left hand side of (2.50), using partial integration:
t2 t1

m(t ) r(t )d t = r

t2 t1

m(t ) (t )d t = r r

t2 t1 t2 t1

m 2 r d t = r 2 r t2 T r r d t r [T (b ) T (a )] d t , (2.51) r t1

where the approximation holds to rst order in r as we have performed a rst-order Taylor expansion in . The resulting expression is the difference in kinetic energy between the two r paths, integrated over time. If we are dealing with a conservative force eld, the right hand side of (2.50) can also be transformed to a difference between two global quantities:
t2 t1

F r(t )d t =

t2 t1

V r(t )d t

t2 t1

[V (rb ) V (ra )] d t .

(2.52)

Combining (2.51) and (2.52) we obtain:


t2 t1

(T V )d t = 0,

(2.53)

where the notation with the means that we calculate the integral for two paths that differ only slightly from each other and then calculate the difference between these two integrals. We conclude that dAlemberts principle for a conservative force can be transformed to the condition that the linear variation (2.53) vanishes. This global condition distinguishes the mechanical path from all other ones. The quantity T V is called the Lagrangian, L. The integral over time of this quantity t2 t 1 L d t is called the action, denoted by S:
t2 t2

S= We have derived a new principle:

t1

d t (T V ) =

d t L.
t1

(2.54)

The mechanical path of a particle moving in a conservative potential eld from a position r1 at time t 1 to a position r2 at t 2 is a stationary solution of the action, i.e. the linear variation of the action with respect to an allowed variation of the path around the mechanical path, vanishes. This principle is called Hamiltons principle. Note that the variations of the path are restricted to lie within the constrained subspace. The advantage of this new formulation of mechanics with conservative force elds over the Newtonian formulation is that it holds for
3 We suppose that the constrained subspace is smooth and that r (t ) is close to r (t ) for all t . a b

20

2. L AGRANGE AND H AMILTON FORMULATIONS

2{

any system subject to constraints, and that it holds independently of the coordinates which were chosen to represent the motion. This is clear from the fact that we search for the minimum of the action within the subspace allowed for by the constraint, and this subspace is properly described by the generalised coordinates q j . When solving the motion of some particular mechanical system our task is therefore to properly express T and V in terms of these generalised coordinates, plug the Lagrangian L = T V into the action, and minimise the latter with respect to the generalised coordinates (which are functions of time). Although this might seem a complicated way of solving a simple problem, it should be realised that the transformation of forces and accelerations to generalised coordinates is usually more complicated than writing the kinetic energy and the potential in terms of these new coordinates. Furthermore we shall see below that the problem of nding the stationary solution for a given action leads straightforwardly to a second-order differential equation, which is the correct form of the Newtonian equation of motion in terms of the chosen generalised coordinates. As an example, consider the simple pendulum (see gure 2.1). The position of the mass m is given by the 2 coordinates x and y (we neglect the third coordinate z). The constraint obeyed by these coordinates is x 2 + y 2 = l 2 . This constraint allows us to use only a single generalised coordinate : x = l sin and y = l cos . The velocity is given by v = l . This example shows that the generalised coordinate q = does not necessarily have to have the dimension of length ( is an angle and is therefore dimensionless), and likewise q = does not necessarily have the dimension of velocity (with dimension 1/time rather than displace ment/time). The kinetic energy is now given as T = ml 2 2 /2, and the potential energy by V = mg l cos . The Lagrangian of the pendulum is therefore L = T V = m l 2 2 + g l cos . 2 (2.55)

We now turn to the problem of determining the stationary solution for an action with such a Lagrangian. The Lagrangian can have many different forms, depending on the particular set of generalised coordinates chosen; therefore we shall now work out a general prescription for determining the stationary solution of the action without making any assumptions concerning the form of the Lagrangian, except that it may depend on the q j and on their time derivatives qj:
t2

S[q] =

L(q, q, t )d t .
t1

(2.56)

Here q(t ) is any vector-valued function, q(t ) = q 1 (t ), . . . , q N (t ) . We now consider an arbitrary, but small variation q(t ) of the path q(t ), and calculate the change S in S as a result of this variation: S[q] = S[q + q] S[q] =
t2 t1

L q + q, q + , t d t q
t2 t1

t2 t1

L q, q, t d t q + L q, q, t q d t . (2.57) q

L q, q, t q

In the last line of this equation, we have performed a Taylor expansion in q and in . Note q that both q and q depend on time. Note further that /q is a vector the derivative must be interpreted as a gradient with respect to all the components of q. The use of the partial derivative and not the full derivatives d in the derivatives indicates that when calculating the gradient with respect to q, q is considered as a constant, and vice-versa. Of course, q and are not independent: if we know q(t ) for all t in the interval under q consideration, we also know the time derivative q. We can remove by partial integration: q
t2 t1

L q, q, t q

q +

L q, q, t q

d t = q

t2 t1

L d L qd t . q d t q

(2.58)

2.6. S UMMARY AND EXAMPLES

21

Because q is small but arbitrary, this variation can only vanish when the term in brackets on the right hand side vanishes. Consider for example a q which is zero except for a very small range of t -values around some t 0 in the interval between t 1 and t 2 . Then the term between the square brackets must vanish in that small range, which is means that it should vanish at t 0 . We can do this for any small interval on the time axis, and we conclude that the term in brackets vanishes for all t in the integration interval. So our conclusion reads The action S[q] is stationary, that is, its variation with respect to q vanishes to rst order, if the following equations are satised: d L L = , q j d t q j for j = 1, . . . , N . (2.59)

{ {

The equations (2.59) are called Euler equations. In the case where L is the Lagrangian of classical mechanics, L = T V , the equations are called EulerLagrange equations (note that in the above derivation, no assumption has been made with respect to the form of L nor what it means the only assumption is that L depends at most on q, q and t ). The Euler equations have many applications outside mechanics. Often the following notation is used: L = and
N j =1

L d L q j q j d t q j

(2.60)

L L d L = , q q d t q L L d L = . q j q j d t q j

(2.61)

or, written in another way:

(2.62)

Note that (2.61) is an equality between (N -dimensional) vector quantities. Solving problems using this formalism usually proceeds along a number of xed steps. We will outline those at the beginning of the next subsection. Turning again to our simple example of a pendulum, we use the Lagrangian found in (2.55) and write down the EulerLagrange equation for this: L d L = mg l sin = = ml 2 . d t (2.63)

The solution to this equation can be found through numerical integration. In the next section we shall encounter some more complicated examples which show the advantages of the new approach more clearly.

2.6 S UMMARY AND EXAMPLES


We start by summarising the general results we have obtained so far in this chapter. Then we shall consider several nontrivial examples where these results are used to nd the equations of motion for several nontrivial mechanical systems.

2.6.1 OVERVIEW OF THE THEORY


When dealing with a system of M degrees of freedom subject to K constraints, there are constraint forces present in order for these constraints to be realized. The constraints leave M K degrees of freedom. These degrees of freedom can be represented using M K generalised coordinates q j . These coordinates may be distances, or angles etcetera.

22

2. L AGRANGE AND H AMILTON FORMULATIONS In addition to the constraint forces, we usually have applied forces, FA . In order to get rid of the constraint forces, we use dAlemberts principle: pi ri =
i i

FA ri , i

(2.64)

2{

where FA is the applied force on particle i . i dAlemberts principle can be reformulated in terms of the generalised coordinates q j : T d T = Fj , d t q j q j where T is the total kinetic energy, and F j is the generalised force: Fj =
N i =1

(2.65)

FA i

ri . q j

(2.66)

If the forces are all conservative, we can reformulate dAlemberts principle as Hamiltons principle, which says that the mechanical path q(t ) running from a xed initial position at t 1 to a xed nal position at t 2 , is given by the condition that the action
t2

S=

L(q, q, t )d t
t1

(2.67)

is stationary, i.e. S varies to second order in the deviation from the mechanical path. L is the Lagrangian which has the form L = T V where T is the kinetic, and V is the potential energy. The path q j (t ) for which S is stationary obeys the Euler-Lagrange equations: d L L . = d t q j q j (2.69) (2.68)

The art of solving problems in analytical mechanics usually involves the following steps. Find the number of degrees of freedom without constraints. For N point particles, this is 3N . Identify the constraints (by carefully reading the description of the system). Find a convenient set of generalised coordinates. This is the step where experience and intuition are convenient. Find the kinetic energy. This is sometimes difcult. If so, you write the particle coordinates ri as functions of the generalised coordinates q j and, from this, the velocities ri as functions of the q j and q j . Then the kinetic energy T = i mi r2 can be written as 2 i j . The resulting expression must be quadratic in the q j ! It is a function of the q j and q always recommended to check this. Is the force is non-conservative, solve dAlemberts equation in generalised coordinates. If it is conservative, write V in terms of the q j and write L = T V . You can now write down the Euler-Lagrange equation and these may be solvable or not. Often, the rst few steps are not necessary as it is usually possible to directly identify the number of degrees of freedom, together with a convenient set of q j . If also the kinetic energy can be written easily in terms of the q j and q j , there is no need to involve the full set of coordinates ri .

2.6. S UMMARY AND EXAMPLES

23

2.6.2 A SYSTEM OF PULLEYS


We consider a system of massless pulleys as in the gure below.

{ {

l l
1

ma m
b

mc

The string is also massless and furthermore inextensible. It is quite complicated to nd out what the forces on the system are when taking all the forces on the pulleys and the tension of the string into account. However, it turns out that using Hamiltons principle makes it an easy problem. The total string length is l = l 1 +l 2 +l 3 +l 4 and this is xed. This is the rst constraint. Note that we have omitted the motion in the horizontal direction and perpendicular to the plane, which eliminates 6 degrees of freedom from the problem. From the 9 original degrees of freedom, there are therefore only 9 1 6 = 2 left. From the coordinates l 1 , l 2 , l 3 and l 4 , one is redundant as l 2 = l 3 . The constraint on the total string length xes l 2 = l 3 when l 1 and l 4 are given: 1 l 2 = l 3 = (l l 1 l 4 ). (2.70) 2 Therefore we use only l 1 and l 4 as generalised coordinates. We can calculate the total potential energy now entirely in terms of l 1 and l 4 . This energy is composed of three contributions, corresponding to the masses m a , m b and m c . The heights of the outer two are directly given by l 1 and l 4 , and the height of the central mass is given by that of the central pulley, which is given by l 2 (or l 3 ), and the total potential energy is therefore given as: mb V = g m a l 1 + (l l 1 l 4 ) + m c l 4 . (2.71) 2 The speed of the left mass m a is given by l 1 , and that of the right one, m c , by l 4 . Using (2.70) 1 1 l 4 ). The Lagrangian is therefore we nd that the speed of the central pulley is given by 2 (l given as 1 mb 2 1 2 1 2 2 L = m a l 1 + m c l 4 + m b (l 1 + l 4 + 2 l 1 l 4 ) + g m a l 1 + (l l 1 l 4 ) + m c l 4 . 2 2 8 2 The Euler-Lagrange equations can be derived straightforwardly: d L 1 1 = (m a + m b )l 1 + m b l 4 = d t l 1 4 4 d L 1 1 = (m c + m b )l 4 + m b l 1 = 4 d t l 4 4 L 1 = m a mb g ; l 1 2 L 1 = mc mb g . l 4 2 (2.73a) (2.73b) (2.72)

24

2. L AGRANGE AND H AMILTON FORMULATIONS

The two equations can be solved for l 1 and l 4 and the result is 4m a m c + m a m b 3m c m b l1 = g; m c m b + 4m a m c + m a m b 4m a m c + m c m b 3m a m b g. l4 = m a m b + 4m a m c + m b m c (2.74a) (2.74b)

2{

To check whether the answer is reasonable we verify that a stationary motion (i.e. a motion with constant velocity) is possible if m b = 2m a = 2m c (convince yourself of this). The solution is now trivial, since the right hand sides of (2.74) vanish as should indeed be the case. We see that the Lagrange equations provide a framework which enables us to nd the equations of motion quite easily.

2.6.3 E XAMPLE : THE SPINNING TOP


?? Consider a top with cylindrical symmetry. In this section, we shall formulate and analyze the equations of motion of the top. This is material that you have already covered in your rst year classical mechanics. If you have forgotten most of this, refer back to that course! The position of the top is dened by its two polar angles and and a third angle, , denes the rotation of the top around its symmetry axis. The angular velocity is given in terms of these three polar angles as: z e = + + d (2.75) where z is a unit vector along the z-axis; e is a unit vector in the x y plane which is perpendic ular to the axis of the top, and d is a unit vector along the axis of the top. The axis of the top is shown in the gure:

d e

x f

From this gure, it is clear that e = ( sin , cos , 0) and d = (cos sin , sin sin , cos ). And it follows that = e d = (cos cos , sin cos , sin ). f The rotational kinetic energy of the top is given by 1 T = T I 2 (2.78) (2.77) (2.76a) (2.76b)

2.6. S UMMARY AND EXAMPLES

25

(the superscript T denotes transpose in a linear algebra sense: it turns the column vector into a row vector). Diagonalizing the 3 3 matrix I, it is always possible to nd some axes with respect to which this moment of inertia tensor is diagonal, and as a result of the axial symmetry of the top one diagonal element, which we shall denote by I 3 , corresponds to the symmetry axis d, and two other diagonal elements correspond to axes in the plane perpendicular to the body axis, such as e and f we call these elements I 1 (the axial symmetry forces them to be equal). The kinetic energy is then given by 1 1 1 1 I1 1 f) T = I 1 ( e)2 + I 1 ( 2 + I 3 ( d)2 = I 1 2 sin2 + 2 + I 3 ( + cos )2 . 2 2 2 2 2 2 (2.79)

{ {

The gravitational force results in a potential V = M g R cos , where M is the tops mass and R the distance from the point where it rests on the ground to the centre of mass clearly, the height of the centre of mass is then given as R cos . The Lagrangian therefore reads: 1 I1 1 L = I 1 2 sin2 + 2 + I 3 ( + cos )2 M g R cos . 2 2 2 The Lagrange equations for , and are then given by: d L L = I 1 2 sin cos I 3 ( + cos ) sin + M g R sin ; = I1 = d t d L d L = I 1 sin2 + I 3 ( + cos ) cos = = 0; d t d t L d L d = I 3 ( + cos ) = = 0. d t dt (2.81a) (2.81b) (2.81c) (2.80)

We immediately see from the last equation that + cos is a constant of the motion we shall call this 3 : 3 = + cos = Constant. (2.82) 3 denotes the component of angular velocity along the spining axis. Let us search for solutions of constant precession: = constant, or = 0. We furthermore set = . The rst Euler-Lagrange equation then gives: I 1 2 cos I 3 3 + M g R = 0. If 3 is large, we nd the two solutions = MgR I 3 3 (2.84) (2.83)

for which is inversely proportional to 3 and = I 3 3 I 1 cos (2.85)

i.e. is proportional to 3 . The rst solution corresponds to slow precession and fast spinning around the spinning axis; the second solution corresponds to rapid precession in which the gravitational force is negligible. For general 3 , the quadratic equation (2.83) with = 0 has two real solutions for if
2 I 3 2 > 4I 1 cos M g R. 3

(2.86)

For smaller values of 3 , a wobbling motion sets in (nutation).

26

2. L AGRANGE AND H AMILTON FORMULATIONS

2.7 N ON - CONSERVATIVE FORCES CHARGED PARTICLE IN AN ELEC TROMAGNETIC FIELD


{

2{

In this section we consider one particular type of force which is not conservative, but which can still be analysed fully within the Lagrangian approach. This is the very important example of a charged particle in an electromagnetic eld. Before we turn to this specic example of a particle in an electromagnetic eld, we discuss how the Euler-Lagrange technique can be extended to a particular type of nonconservative forces. Suppose we have a collection of N particles which experience a non-conservative force which can be derived from a generalised potential W (ri , ri ) in the following way: F= W d W + . ri d t ri (2.87)

Analogous to the previous section we can then derive a variational condition, starting from dAlemberts principle:
t2 t1

mi ri d t = r

t2 t1

T d t =

t2 t1

W d W + ri d t . ri d t ri

(2.88)

The left hand side has been transformed as in (2.51). To rewrite the the right hand side, we need a special partial integration step for the second term:
t2

t1

T d t =

t2 t1

W W ri i d t = r ri ri

t2 t1

W d t .

(2.89)

So we see that the variation of the action


t2

S[q] =

t1

[T W ] d t

(2.90)

vanishes. We can also work out the Euler-Lagrange equations for the Lagrangian L = T W , which, for
N

T=
i =1

i m i r2 ,

(2.91)

directly leads to the classical equation of motion: d dL d dW dW = m i ri + = , i i dt dr dt dr d ri which is equivalent to m i ri = F i . (2.93) (2.92)

2.7.1 C HARGED PARTICLE IN AN ELECTROMAGNETIC FIELD


A point particle with charge q moving in an electromagnetic eld experiences a force F = q (E + v B) . (2.94)

The charge q of the particle should not be confused with the generalised coordinates q i introduced before. E is the electric eld, B is the magnetic eld. Note that the force depends on the velocity, which shows that it cannot be conservative. Therefore we try to cast it into the form (2.87). The elds are not independent, but they are related through the Maxwell equations. We use the following two Maxwell equations B = 0 E+ and (2.95a) (2.95b)

B = 0. t

2.8. H AMILTON MECHANICS

27

We know from vector calculus that a vector eld whose divergence is zero, can always be written as the curl of a vector function depending on space (and, in our case, time); applying this to (2.95a) we see that we can write B in the form B = A, where A is a vector function, called the vector potential, depending on space and time. Substituting this expression for B in Eq. (2.95b) leads to A E+ = 0. (2.96) t Now we use another result from vector calculus, which says that any function whose curl is zero can be written as the gradient of a scalar function, which in this case we call the potential, (r, t ). This results in the following representations of the electromagnetic eld: E(r, t ) = (r, t ) B(r, t ) = A(r, t ). A (r, t ); t (2.97a) (2.97b)

{ {

In fact, by using two Maxwell equations, we have reduced the set of 6 eld values (3 for E and 3 for B) to 4 (3 for A and 1 for ). We are after a function W (r, r) which, when used in an action of the usual form, yields the correct equation of motion with the force (2.94). The potential which does the job is W (r, r) = q(r, t ) q r A(r, t ) = q q(x A x + y A y + z A z ). (2.98)

Note that A x denotes the x-component and not the partial derivative with respect to x. The Lagrangian occurring in the action is therefore: 1 L = m2 + q r A(r, t ) q(r, t ). r 2 (2.99)

To see that this Lagrangian is indeed correct we work out the force component in the x-direction. First we calculate the derivative of the potential W with respect to x: Furthermore d W d Ax A x A x A x A x = q = q + x+ y+ z . d t x dt t x y z (2.101) A y A x A z W = q +q x +y +z . x x x x x (2.100)

The Euler-Lagrange equations for the action contain the two contributions resulting from the potential. We have mx = A x W d W = q + + + x d t x x t A y A x A z A x q y +z x y x z

= qE x + q( yB z zB y ) (2.102)

i.e. precisely the equation of motion (for the x-component) with the force given in (2.94)! The other components follow in a similar way.

2.8 H AMILTON MECHANICS


In the previous sections of this chapter, we have seen that Hamilton principle, formulted as the Lagrangian equations of motion, enable us to solve complicated systems relatively quickly. It is possible to formulate Lagrangian mechanics in a different way. At rst sight this does not add anything new to the formalism which was constructed in the previous sections, but we shall see that this new formalism provides us with a conserved quantity which is the

28

2. L AGRANGE AND H AMILTON FORMULATIONS

energy or some analogous object. More importantly, this formalism is essential for setting up quantum mechanics in a structured way, as will be shown in a later course. Let us again consider a system described by a Lagrangian formulated in terms of generalised coordinates, with the equations of motion given by:

2{

L d L = . d t q j q j

(2.103)

This is a second order differential equation, which we shall transform into two rst order ones. We dene the conjugate momentum p j as pj = L . q j (2.104)

The conjugate momentum should not be confused with the mechanical momentum, which is simply i mi , although the two coincide when the generalised coordinates are simply the r ri . Using the conjugate momenta, the equations of motion can be formulated as: pj = L . q j (2.105)

In the particular example of a conservative system formulated in terms of the position coordinates ri : N m i 2 r V (r1 , . . . , rN ), (2.106) L= 2 i i =1 the momenta are given as pi = mi r and the equations of motion are pi = V . ri (2.108) (2.107)

We see that in the case of a particle moving in a conservative force eld, the conjugate momentum corresponds to the mechanical momentum. We have reformulated the Euler-Lagrange equation as two rst-order differential equations. The Euler-Lagrange equations were derived from a variational principle, the Hamilton principle, which requires the action to be stationary for the mechanical path. This led to a relation between the derivatives of the Lagrangia, L/q j and L/q j . We may ask ourselves if it is possible to dene our two new equations also as relations of a derivative of some other function. This means that the Lagrangian, which is formulated as a function of q j and q j (and perhaps the time t ) is traded in for a new function which depends on the coordinates q j and p j . We call this new function the Hamiltonian, H (p j , q j , t ). We shall for now not discuss how this function is constructed (see the box below), but just give its form:
N

H (p j , q j , t ) =
j =1

p j q j L q j , q j (q k , p k ), t .

(2.109)

Note that we can indeed express q j in terms of the p k and q k , as indicated in the second argument of L by inversion of Eq. (2.104).4 This function can be used to formulate the two equations (2.104) and (2.105) in an elegant way in terms of the Hamiltonian.
4 For this inversion to be possible, the Lagrangian should be convex, but we shall not go into details concerning

this point.

2.8. H AMILTON MECHANICS To see this, let us calculate the derivatives of H with respect to q j and p j :

29

H = qj + p j

pk
k

q k p j

L q k . q k p j

{ {

(2.110)

Note that it follows from (2.104) that the second and third terms on the right hand side cancel, so that we are left with

H = qj. p j

(2.111)

Now let us calculate the derivative with respect to q j :

H L = + q j q j

pk
k

q k q j

L q k . q k q j

(2.112)

Again using (2.104) we see that the second and third term on the right hand side cancel furthermore the rst term on the right hand side is equal to p i and we are left with:

H = p j . q j

(2.113)

Eqs. (2.111) and (2.113), together with the denition of the Hamiltonian (2.109) and of the momentum (2.104) are equivalent to the equations of motion. Eqs (2.111) and (2.113) are called Hamiltons equations. Note that we must consider the generalised coordinates and the canonical momenta as independent coordinates.

30

2. L AGRANGE AND H AMILTON FORMULATIONS


** Legendre transformations It seems that the Hamiltonian has been found by some magic. Now we shall explain how this magic works. Let us assume that we have a function f (x, y) depending on two variables x and y. A variation of these variables induces a variation of f :

2{
Calling

df =

f f dx + d y. x y

(2.114)

u = f /x and v = f /y, we have d f = ud x + vd y.

(2.115) (2.116)

In mathematical terms, we say that f is a exact differential. Most well-behaved functions are exact differentials, in particular the Lagrangian is an exact differential of the q j and q j . Now we search for a function g (u, y) which depends on u and y such that dg = g g du + d y. u y (2.117)

It turns out that such a function is given by g = xu f . Let us check this. We have d g = xd u + ud x d f . Using (2.116), we can work this out as d g = xd u + ud x ud x vd y = xd u vd y. (2.120) (2.119) (2.118)

We see that indeed the new function g does the job! The procedure outlined here is known as Legendre transformation. It occurs in many area of physics. Now let us turn back to L and assume that it is an exact differential of q j and q j (we forget the explicit time dependence for simplicity): dL = L L L dqj + dqj = dqj + q j q j q j pj dqj ,
j

(2.121)

where we have used the denition of p j , (2.104). Now we nd the exact differential H in terms of q j and p j analogous to our recipe for nding g from f (applied to each pair of coordinates q j , q j ): (2.122) H (p j , q j ) = p j q j L(q j , q j ),
j

which is the form we have given above.

If the system does not depend explicitly on time, the Hamiltonian is the analogue of the energy. The simplest case is a conservative system with the positions ri as coordinates. In that case it is easy to see that
N

H=

2 pi

i =1 2m

+ V (r1 , . . . , rN ).

(2.123)

More generally, let us consider a conservative system formulated in terms of generalised coordinates q 1 , . . . , q s . Note the difference with Eq. (2.2), where ri may contain an explicit time-dependence in the present case we assume that the constraints have no explicit timedependence. In that case it is possible to express the position coordinates ri in terms of the s generalised coordinates q j , j = 1, . . . , s: ri = ri (q 1 , . . . , q s ) (2.124)

2.8. H AMILTON MECHANICS and therefore the velocities can be calculated as ri qj. q j j =1
s 1 r2 i 2 mi

31

ri =

(2.125)
{ {

in terms of the generalised coordinates, Therefore, if we formulate the kinetic energy we obtain an expression which is quadratic in the q j :
s

T=
k, j =1

M k j (q 1 , . . . , q s )q j q k

(2.126)

where M j k = Mk j = m i ri ri . i =1 2 q j q k
N

(2.127)

If we calculate the contribution to the momenta arising from the kinetic energy, we nd that they depend linearly on the q j :
s s T = M j k + Mk j qk = 2 Mk j qk . q j k=1 k=1

pj =

(2.128)

Hence
s

q j p j = 2T
j =1

(2.129)

and H = 2T (T V ) = T + V = Energy. (2.130)

For a general system Hamiltons equations of motion can be used to derive the time derivative of the Hamiltonian:
s H s H dH H = qj + pj + . dt t j =1 q j j =1 p j

(2.131)

Using Hamiltons equation of motion (2.111) and (2.113) we see that the rst two terms on the right hand side cancel and we are left with: d H H = . dt t

(2.132)

We see therefore that if H (or L) does not depend explicitly on time, then H is a conserved quantity. If the potential does not contain a q j dependence, this implies conservation of energy. If the potential on the other hand does contain such a dependence, then (2.132) implies conservation of some quantity which plays a role more or less equivalent to energy.

32

2. L AGRANGE AND H AMILTON FORMULATIONS Let us summarize the formalism which we have developed in this section. We started with the Euler Lagrange equation L d L = . j q j d t q Dening pj = L , q j (2.133)

{ {

(2.134)

we could write the Euler Lagrange equation in the form pj = L . q j (2.135)

It is possible to formulate these equations using a new functional depending on q j and p j , the Hamiltonian: H (p j , q j ) =
j

p j q j L.

(2.136)

The new equations are called Hamilton equations: qj = H , p j pj = H . q j (2.137)

The Hamiltonian is preserved in time. In the case where the Lagrangian (and therefore the Hamiltonian) have no explicit time dependence, the Hamiltonian coincides with the total mechanical energy.

2.9 A PPLICATIONS OF THE H AMILTONIAN FORMALISM


In this section we shall reconsider the systems studied before in the Lagrange framework and point out which features are different when these systems are considered within the Hamiltonian framework. From the derivation of the Hamiltonian and Hamiltons equations, it is seen that the latter can be viewed as a different way of writing Lagranges equations. The reason for introducing the Hamiltonian and Hamiltons equations is that they are often used in quantum mechanics and because the Hamilton formalism is more convenient for discovering some conserved quantities.

2.9.1 T HE THREE - PULLEY SYSTEM


From the Lagrangian (2.72): 1 mb 2 1 2 1 2 2 L = m a l 1 + m c l 4 + m b (l 1 + l 4 + 2 l 1 l 4 ) + g m a l 1 + (l l 1 l 4 ) + m c l 4 , 2 2 8 2 (2.138)

the momenta p 1 and p 4 associated with the degrees of freedom l 1 and l 4 are found as: L mb mb l1 + l4; = ma + 1 4 4 dl L mb mb l4 + l1; p4 = = mc + 4 4 4 dl p1 = We should express l 1 and l 4 in terms of the momenta: 1 mb mb l1 = mc + p1 p4 , 4 4 1 mb mb l4 = p1 + ma + p4 , 4 4 (2.140a) (2.140b) (2.139a) (2.139b)

2.9. A PPLICATIONS OF THE H AMILTONIAN FORMALISM with = (m a + m c )m b /4 + m a m c . After some calculation, we nd for the Hamiltonian: H= mb mb 1 2 2 mc p 1 + m a p 4 + (p 1 p 4 )2 g m a l 1 + (l l 1 l 4 ) + m c l 4 . 2 4 2

33

(2.141)

(2.142)

{ {

Note that this Hamiltonian only depends on the p j and q j (which are in this particular prob lem the l j s) there is no dependence of this Hamiltonian on the q j = l j here). The Hamilton equations read: mb )g ; 2 mb )g . p 4 = (m c 2 p 1 = (m a The solution is simple since the right hand sides of these equations are constants: mb )g t ; 2 mb p 4 = (m c )g t , 2 p 1 = (m a (2.144a) (2.144b) (2.143a) (2.143b)

where the initial conditions are that the system is standing still at t = 0. Together with Eqs. (2.139), we obtain the same solution as in the Lagrangian case. We see that the difference between the two approaches are not very dramatic in this case. Note that it is now easy to see that for m a = m c = 2m b the system is in equilibrium.

2.9.2 T HE SPINNING TOP


1 1 I1 L = I 1 2 sin2 + 2 + I 3 ( + cos )2 M g R cos . (2.145) 2 2 2 From this Lagrangian, we can derive the momenta associated with the three degrees of freedom , and : p = I 1 sin2 + I 3 ( + cos ) cos ; p = I 1 ; p = I 3 ( + cos ). (2.146a) (2.146b) (2.146c)

If we want to express the kinetic energy in terms of these momenta, we need to solve for the time derivatives of the angular coordinates , and in terms of these momenta: = p p cos I 1 sin2 ; (2.147a) (2.147b) (2.147c)

p ; = I1 p p p cos = cos . I3 I 1 sin2 After some calculation, the Hamiltonian is then found to be H= (p p cos )2 2I 1 sin2 +
2 p

2I 1

2 p

2I 3

+ M g r cos .

(2.148)

As the Hamiltonian does not depend on and , we see immediately that p and p must be constant. Coordinates of which only the momentum does appear in the Hamiltonian are called ignorable: these momenta are constant in time they represent constants of the motion. We have seen that both p and p are constants of motion.

34

2. L AGRANGE AND H AMILTON FORMULATIONS The Hamiltonian now reduces to a simple form: H=
2 p

2I 1

+U (),

(2.149)

2{

where U () =

(p p cos )2 2I 1 sin2

2 p

2I 3

+ M g r cos .

(2.150)

The Hamilton equations yield dU I 1 = p = . d (2.151)

This equation is difcult to solve analytically. Note that apart from the ignorable coordinates, we have an additional constant of the motion, the energy:
2 p

I1

+U () = E = constant.

(2.152)

2.9.3 C HARGED PARTICLE IN AN ELECTROMAGNETIC FIELD


Finally we consider again the charged particle in an electromagnetic eld. The momentum can be found as usual from the Lagrangian we obtain p = m + qA. r The Hamiltonian is H = p r m 2 m (p qA)2 r q r A + q = r2 + q(r) = + q(r). 2 2 2m (2.154) (2.153)

You might already know that this Hamiltonian is used in quantum mechanics for a particle in an electromagnetic eld.

3
T HE TWO - BODY PROBLEM

35

36

3. T HE TWO - BODY PROBLEM

In this chapter we consider the two-body problem within the framework of Lagrangian mechanics. One of the most impressive results of classical mechanics is the correct description of the planetary motion around the sun, which is equivalent to electric charges moving in each others eld. With the analytic solution of this problem, we shall recover the famous Kepler laws. The problem is also important in quantum mechanics: the hydrogen atom is a quantum version of the Kepler problem.
{ {

3.1 F ORMULATION AND ANALYSIS OF THE TWO - BODY PROBLEM


The two-body problem describes two point particles with masses m 1 and m 2 . We denote their positions by r1 and r2 respectively, and their relative position, r2 r1 , by r. Finding the Lagrangian is quite simple. The kinetic energy is the sum of the kinetic energies of the two particles, and the potential energy is the interaction, which depends only on the separation r = |r| of the two particles, and is directed along the line connecting them (note that this last restriction excludes magnetic interactions). We therefore have: L= m1 2 m2 2 r + r V (r ). 2 1 2 2 (3.1)

Before deriving the equations of motion, we note that instead of writing the kinetic energy as the sum of the kinetic energies of the two particles, it can also be separated into the kinetic energy of the centre of mass and that of the relative motion, as in Eq. (1.33): T = TCM + where ri = ri rCM ; m 1 r1 + m 2 r2 rCM = , M and TCM = M 2 r , 2 CM (3.3) (3.4) mi 2 ri , i =1 2
2

(3.2)

M = m1 + m2 .

(3.5)

As there are only two particles, we can work out the coordinates ri relative to the centre of mass rC , and we nd, using Eq. (3.4): r1 = r1 rCM = and r2 = r2 rCM = m2 (r1 r2 ), M m1 (r2 r1 ). M (3.6)

(3.7)

We can take time derivatives by simply putting a dot on each r in these equations and then, after some calculation, we nd for the kinetic energy: T= The Lagrangian is therefore L = T V = M 2 m1 m2 2 rCM + r V (r ). 2 2M (3.9) M 2 m1 m2 2 r + r . 2 CM 2M (3.8)

We see that the kinetic energy of the relative motion has the form of the kinetic energy of a single particle of mass m 1 m 2 /(m 1 +m 2 ) and position vector r(t ). The mass term = m 1 m 2 /(m 1 + m 2 ) is called reduced mass.

3.1. F ORMULATION AND ANALYSIS OF THE TWO - BODY PROBLEM

37

Of course we could write down the EulerLagrange equations for this Lagrangian as before, but it is convenient to perform a further separation: that of the kinetic energy of the relative coordinate into a radial and a tangential part. First we must realise that the plane through the origin and the initial velocity vector r of the relative position will always remain the plane of the motion, as the force acts only within that plane. In this plane, we choose an x and a y axis. Then we can conveniently introduce polar coordinates r and , in which the x and y coordinate can be expressed as follows: x = r cos ; y = r sin . (3.10) (3.11)
{ {

It then immediately follows that the kinetic energy of the relative motion can be rewritten as 2 2 x + y2 = r + r 2 2 . 2 2 The Lagrange equations given in Eq. (2.59) then take the form: M rCM = 0; d r 2 = 0; dt r r 2 = dV (r ) . dr (3.13a) (3.13b) (3.13c) (3.12)

The rst equation tells us that the centre of mass moves at constant speed: it does not feel a net force. This follows from the fact that it does not appear in the potential, and is in accordance with the conservation of total momentum in the absence of external forces. Coordinates such as rCM with constant conjugate momentum, are called ignorable see section 2.9.2. The second and third equation do not depend on rCM therefore we see that the relative motion can entirely be understood in terms of a single particle with mass and moving in a plane under the inuence of a potential V (r ). We now use the second equation to eliminate . First note that the term in brackets occurring in this equation must be a constant, we call this constant ( = r 2 is precisely the angular momentum, and we see that it is conserved); the third equation then transforms into 2 dV (r ) r 3 = , (3.14) r dr Note that this equation can be viewed as that of a one-dimensional particle subject to a force 2 (r F = r 3 dV r ) . Such a force can in turn be derived from a conservative potential: d F (r ) = d d VEff (r ) = dr dr
2

2r 2

+ V (r ) .

(3.15)

The subscript Eff is used to distinguish between this effective potential and the original, bare attraction potential V (r ). The potential VEff is represented in gure 3.1 for the case V (r ) = 1/r . From Eqs. (3.13b) and (3.13c) and from gure 3.1, we can infer the qualitative behaviour of the motion. We have seen [Eq. (3.13b)] that the angular momentum is constant. This implies that the motion will always keep the same orientation (i.e. clockwise or anti-clockwise). If the particles move apart, the speed at which they orbit around each other will be slower (since increasing r implies decreasing r ). The motion in the radial direction can be understood qualitatively as follows. Note that we can interpret Eq. (3.13c) as the motion of a particle in one dimension. This particle has a mechanical energy which is the sum of its kinetic energy and the effective potential, and this energy should remain constant. Furthermore, the energy cannot be lower than the lowest

38

3. T HE TWO - BODY PROBLEM

3 2

VEff (r )

1
rmin

E> 0

3{

0
rmin rmax

E< 0

-1 0 1 2 3 4

r
F IGURE 3.1: Effective potential for a two-particle system.

value of the effective potential shown in gure 3.1. If it lies between this value and 0, then then r will vary between some minimum and maximum value as shown in this gure. If E is on the other hand positive, r will vary between some minimum value and innity. We have seen that the r -component of the two-body motion can be described in terms of a single particle in one dimension. The energy of this particle is the sum of its kinetic and potential energy the latter is the effective potential [see Eq. (3.15)]. It turns out that this energy is equal to the total energy of the two-particle system (neglecting the contribution of the centre of mass motion to the latter). As we have already worked out the kinetic and potential energy of the two-body problem above, we immediately see that E = T +V = 2 r + r 2 2 + V (r ), 2 (3.16)

which can easily be identied as the kinetic energy r 2 /2 of the one-dimensional particle plus the effective potential.

3.2 S OLUTION OF THE K EPLER PROBLEM


The special case V (r ) = A/r is very important as it describes the gravitational and the Coulomb attraction. Also, in this special case, the motion can be studied further by analytical means. Finding the solution in the form r (t ), (t ) is not convenient rather, we search for r (), which contains explicit information about the shape of the orbit. We use the fact that the angular momentum = r 2 is constant and combine this with the fact that the energy is constant and given by (3.16): = r2 = r 2 ; (3.17) (3.18)

2 2 (E V ) 2 2 . r

Eliminating the d t of the time derivatives by dividing (3.17) by the square root of (3.18) leads to d = . (3.19) 2 2(E V (r )) 2 /r 2 1/2 dr r With V (r ) = A/r this can directly be integrated to give C = arcsin Ar Ar
2

(3.20)

3.2. S OLUTION OF THE K EPLER PROBLEM

39

In addition to the integration constant C on the left hand side, we see a constant , called the eccentricity, which is given in terms of the problem parameters as = 1+ 2E 2 . A 2
2

(3.21)

Inverting Eq. (3.20) to nd r as a function of the polar angle gives: r= A 1 sin( C ) . (3.22)
{ {

We have some freedom in choosing C it changes the denition of the angle . If we take = 0 as the angle for which the two particles are closest (perihelion), we see that C = /2. The motion can now be classied according to the value of . We take positive changing the sign of does not change the shape of the orbit (putting + compensates this sign change). For = 0, r does not depend on . This corresponds to a circle. If 0 < < 1, we have an ellipse (r varies between some maximum and minimum value). For = 1, we have a parabola (r for = ), and for > 1 we have an hyperbola (r for cos = 1/ ). Dening
2

, (3.23) A We see that the radius of the orbit varies between r 0 /(1 + ) and r 0 /(1 ). In gure 3.2, the focal point F 1 must be viewed as the origin where r = 0. We see that the shortest and longest distance lie in this picture along the horizontal axis. If we add the two, we get r0 r0 r0 . (3.24) + =2 1+ 1 1 2 From the gure, we see that this sum must be equal to twice the long axis, 2a. Therefore r0 . (3.25) a= 1 2 Usually, the notation 2 1 = (3.26) A 1 + is used for the perihelion distance, so that the equation relating the two polar coordinates on the curve of the motion reads: (1 + ) . (3.27) r= 1 + cos In gure 3.2, we indicate the semi-major and semi-minor axis a and b respectively and the focal points. From the denitions for and the value of the semi-major axis, we immediately see that a= . (3.28) 1 The area of an ellipse in terms of its semi-major axis is a 2 1 2 . This can be related to the angular momentum by realising that the innitesimal area swept by a line from the origin to the point of the motion is given by r 2 d /2. This tells us that the rate at which this area changes is given as 2 /(2), so that the total area, which is swept in one revolution of period T is equal to T /(2), so that we have: T 2 = a 2 1
2.

r0 =

(3.29) related

The quantities a and are not independent remember a = /(1 ); furthermore to and [see Eq. (3.26)]. Using this to eliminate nally leads to 42 3 a . A We have now recovered all three laws of Kepler: T2 =

(3.30)

40

3. T HE TWO - BODY PROBLEM

b a a F2

{ {

F 1

F IGURE 3.2: Ellipse with various parameters indicated.

All planets move around the sun in elliptical paths. In fact, most planets have eccentricities very close to zero. A line drawn from the sun to a planet sweeps out equal areas in equal times. The rate at which this area increases is given by 2 /(2) as we have seen above. The squares of the periods of revolution of the planets about the sun are proportional to the cubes of the semimajor axes of the ellipses. See Eq. (3.30).

3.3 S UMMARY
In this chapter, we have analysed the two-body problem, with a focus on the Kepler case, where the interaction potential of the two bodies is V (r ) = A/r . Very generally, we have seen that we can decompose any two-particle system with an ineraction potential between the two particles (and no external eld), into the centre of mass motion with mass M and coordinate m 1 r1 + m 2 r 2 R= . (3.31) M This coordinate experiences a uniform motion. The relative motion is described by the coordinate r = r2 r1 . Its motion is described by the Lagrangian L = r2 V (r ), (3.32) 2 where is the reduced mass m1 m2 = . (3.33) m1 + m2 This motion takes place in a plane spanned by the velocity and the force at any time and is therefore essentially two-dimensional. For the particular case of the Kepler problem, where V (r ) = A/r , with positive A, we expect the particles to either orbit around each other (this happens when their total energy is negative) or they scatter off each other (positive energy). We have solved this problem analytically for the negative energy solutions. We have seen that the shape of the orbit was determined by r0 r= , (3.34) 1 + cos where r, are polar coordinates.

3.3. S UMMARY

41

From the solution, the Kepler laws follow. They are summarized at the end of the previous section.

{ {

4
E XAMPLES OF VARIATIONAL CALCULUS ,
CONSTRAINTS

Suspended chain; this shape is called catenary, from the latin word catena.

43

44

4. E XAMPLES OF VARIATIONAL CALCULUS , CONSTRAINTS

4.1 VARIATIONAL PROBLEMS


In the previous chapters, we have considered a reformulation of classical mechanics in terms of a variational principle. This will lead the way to formulating quantum mechanics this is the subject of the next chapter. In this chapter we make an excursion which is still in the eld of classical problems, though not classical dynamics as in the previous chapters. In fact, variational calculus is not only useful for mechanics. Many physical problems which occur in every day life can be formulated as variational problems. In the next sections we shall consider a few examples. We shall rst introduce some further analysis concerning the problems we are about to treat in this chapter. Consider an expression of the form

4{

J=

d x F (y, y , x).

(4.1)

We have used a notation which differs from that used in previous chapter in order to emphasise that J is not always the action and F not always the Lagrangian. J assigns a real value to every function y(x) it is called a functional. There is a whole branch of mathematics, called functional analysis, dedicated to such objects. Here we shall only consider nding the stationary solutions (minima, maxima or saddle points) of J ; they are given as the solutions to the Euler equations F d F =0 (4.2) y d x y In the case where F does not depend explicitly on x, i.e. F = F (y, y ), (4.3)

we can directly integrate the Euler equation(s) once: by multiplying the Euler equation by y we nd F (y, y ) d F (y, y ) y = 0. (4.4) dx y From this it follows that the solution must obey F (y, y ) y F (y, y ) = Constant. y (4.5)

This is a rst order differential equation: we have integrated the second order Euler equations once. This is a quick way of solving the Euler equations. Note that it only works for functionals which do not depend explicitly on x!

4.2 T HE BRACHISTOCHRONE
Near the end of the 18-th century, Jean Bernouilli was studying a problem, which we formulate as follows. Suppose you are to design a monorail in an amusement park. There is a track in your monorail where the trolleys, which arrive at some high point A with low (approximately zero) speed should move to another place B under the inuence of gravity (no motor is used and friction is neglected) in the shortest possible time. The problem is to design the shape of the track in order to achieve this goal. It will be clear that the track lies in a plane. Let us rst consider the possible solutions heuristically. One could argue that a straight line would be the best solution because it is the shortest path between A and B . On the other hand, it would seem favourable to increase the particles velocity as much as possible in the beginning of the motion. This would call for a steep slope near the starting point A, followed by a more or less horizontal path to B , but the resulting curve is considerably longer than the straight line, which is the shortest path between A and B . We must therefore nd the optimum between the shortness of the path and the earliest increase of the velocity by a steeper slope.

4.2. T HE BRACHISTOCHRONE

45

We can solve this problem using the techniques of the previous section. We must minimise the time for a curve which can be parametrised as x(s), y(s). Obviously, there are many ways to parametrise a curve we shall use for s the distance along the curve, measured from the point A. The innitesimal segment d s is given by d x2 + d y 2 = d x dy dx
2 2

ds =

1+

= dx

1+ y

(4.6)

s can be expressed as a function of t the relation between the two is given by d s = vd t where v =
2 v x + v 2 is the particle speed. The time needed to go from A to B is given by y B

(4.7)
{ {

t=

ds . v

(4.8)

We need an equation for v in terms of the path length. As the gravitational force is responsible for the change in velocity it is useful to consider the x- and y-components of the path. In fact, we have the following relation between v and y as a result of conservation of energy: 1 2 v = g y, 2 (4.9)

where the height y is measured from the point A. This means that we have put in the boundary condition that when y = 0, then v = 0, which is correct since the particle is released from A with zero velocity. Therefore, using (4.9), we arrive at
x0

t [y] =

dx
0

1+ y 2g y

(4.10)

where x 0 is the horizontal distance between A and B . We have to nd the stationary function y(x) for the functional t [y]. The Euler-Lagrange equations have the solution [see Eq. (4.5)]: 1+ y 2 y 2g y
2

1 1+ y
2

2g y

= Constant.

(4.11)

Rewriting the rst term on the left hand side as 1+ y 2 = 2g y this can be simplied to y(1 + y ) = C = Constant. In order to solve this equation we substitute y = tan so that we have: y = C cos2 = C And 1 1 + cos(2) . 2 2 (4.14)
2

1+ y

(4.12)

( 1 + y 2 )2g y

(4.13)

(4.15)

dx 1 d y C sin(2) = = = 2C cos2 . d y d tan

(4.16)

46 The solution is therefore

4. E XAMPLES OF VARIATIONAL CALCULUS , CONSTRAINTS

x = C +

1 sin(2) + D 2 1 1 y =C + cos(2) . 2 2

(4.17a) (4.17b)

4{

D and C are integration constants if we identify the point A with (0, 0), the curve starts at = /2, and D/C = /2. The two coordinates of B are used to x the value of at point B and the constants D and C . Note that the boundary condition y = 0 at point A was already realised in Eq. 4.9). The resulting curve is called the cycloid it is the curve described by a point on a horizontally moving wheel.

4.3 F ERMAT S PRINCIPLE


The path traversed by a ray of light in a medium with a varying refractive index is not a straight line. According to Fermats principle this path is determined by the requirement that the light ray follows the path which allows it to go from one point to another in the shortest possible time. The time needed to traverse a path is determined by the speed of light along that path, and this quantity is given as c c(n) = (4.18) n where c is the speed of light in vacuum and n is the refractive index. The latter might vary with position. If the path lies in the x y plane, the path length d l of a segment corresponding to a distance d x along the x-axis is given by dy dx
2

dl = dx

1+

(4.19)

The time d t needed to traverse the path d l is given as: dt = dl , c/n (4.20)

so that the total time can now be given as an integral over d x:


L

t=

n(y) dx c

dy 1+ dx

(4.21)

Here we have assumed that n depends on the coordinate y only. Now take n(y) = then we must minimise L d y 2 ct = d x 1 + y2 1 + . dx 0 For this case, the Euler-Lagrange equations reduce to the equation [see Eq. (4.5)] dy 1 = dx A The solution is given as y(x) = 1 A 2 sinh The possible range of A-values is |A| 1. x +B . A (1 A 2 ) + y 2 .

1 + y 2,

(4.22)

(4.23)

(4.24)

4.4. T HE MINIMAL AREA PROBLEM

47

y(x) x

{ {

4.4 T HE MINIMAL AREA PROBLEM


Consider a soap lm which is suspended between two parallel hoops (see gure). The soap lm has a nite surface tension, which means that its energy scales linearly with its area. As the lm tends to minimise its energy, it minimises its area. The minimal area for a surface of revolution described by a function y(x) is given by:
L

2
0

dx y

1 + y 2.

(4.25)

Minimising this functional of y leads to the standard Euler-Lagrange solution Eq. (4.5) for functionals with no explicit time dependence: y 1+ y The solution to this equation is given by y(x) = C cosh x+A C (4.27)
2

=C

(4.26)

We now assume that the hoops have the same diameter, and that they are placed at a distance L from each other. Let us furthermore choose the x-axis such that the origin is in the middle between the two hoops. Using the fact that cosh is an even function, we have: R = C cosh L , 2C (4.28)

where R is the radius of the hoops. Consider now the graph of C cosh[L/(2C )] as a function of C for xed L: It is clear that for R lying in the gap of the graph, no solution can be found. What happens is that if the hoops are not too far apart, the soap lm will form a nice coshformed shape. However, when we pull the hoops apart, there will be moment at which the lm can no longer be sustained and it collapses. It can be seen from the graph that usually there are two different solutions. The one with the smallest surface is to be selected. The surface area is found as A(y) = LC + 2R C 2 + R2 . (4.29)

4.5 C ONSTRAINTS
4.5.1 C ONSTRAINT FORCES
In dAlemberts approach, the forces of constraint are neglected as they are usually of limited physical importance. In some cases, however, it might be useful to know what these forces

48

4. E XAMPLES OF VARIATIONAL CALCULUS , CONSTRAINTS

x*cosh(0.5/x)

R/L

-2

-4

4{

-4

-2

C /L

are. For example, a designer of a monorail would like to know the force which is exerted on that rail by the train in order to certify that the monorail is robust enough. In fact, it is possible to work out within a Lagrangian analysis what the forces of constraint are. Let us rst recall the solution to the following problem Find the minimum of the function f (x), where x = (x 1 , x 2 , . . . , x N ), under the condition g (k) (x) = C k , where C k are constants; k = 1, . . . , K . Consider a small variation x such that g (k) (x+x) = C k still holds for all k. Then it holds that g (k) (x) + x g (k) (x) = g (k) (x) = C k hence x g (k) (x) = 0 (4.31) for all k. On the other hand, for variations x satisfying (4.31), f should not change to rst order along x, so we have x f (x) = 0. (4.32) Now we can show that f (x) must lie in the span of the set g (k) (x). If it would lie outside the span, we can write it as the sum of a vector lying in the span of g (k) (x) plus a vector perpendicular to this space. If take x to be proportional to the latter, then (4.31) is satised, but (4.32) is not. Therefore we conclude that f can be written as a linear combination of the gradients g (k) :
K

(4.30)

f (x) =
k=1

k g (k) (x).

(4.33)

This is the well-known Lagrange multiplier theorem. Let us consider a simple example: nding the minimum or maximum of the function f (x, y) = x y on the unit circle: g (x, y) = x 2 + y 2 1 = 0. There is only one Lagrange parameter , and Eq. (4.33) for this case reads (y, x) = (2x, 2y), (4.34)

whose solution is x = y, = 1/2. The constraint x 2 + y 2 = 1 then xes the solution to x = y = 1/ 2 and x = y = 1/ 2. Indeed, the symmetry of the problem allows only the axes x = y and the x = 0 or y = 0 as the possible solutions, and it is easy to identify the stationary points (minima, maxima, saddle points) among these.

4.5. C ONSTRAINTS

49

Now suppose we have a mechanical N -particle system without constraints. For such a system we know the Lagrangian L. The combined coordinates of the system are represented as a vector R = (r1 , . . . , rN ). Then we have for any displacement R that the corresponding change in Lagrangian vanishes:
M i =1

ri

L(R, R, t ) L(R, R, t ) = 0 ri

(4.35)

for all t , where we have used the notation of (2.60). Now suppose that there are constraints present of the form g (k) (R) = 0. (4.36) The argument used above for ordinary functions of a single variable can be generalised to show that we should have K L(R, R, t ) k R g (k) (R); (4.37) = R k=1 the reader is invited to verify this. As L is the Lagrangian of a mechanical system without constraints, we know that the left hand side of this equation can be written as p F A . The right hand side has the dimension of a force and must therefore coincide with the constraint force. Let us analyse the simple example of the pendulum once again. Without constraints we have m 2 L= x + y 2 mg y. (4.38) 2 The constraint is given by (4.39) l 2 = x 2 + y 2. So the pendulum equations of motion become m x = 2x m y = mg + 2y. (4.40a) (4.40b)
{ {

These equations cannot be solved analytically, as they describe the full pendulum, and not the small angle limit. In the small angle limit the force in the x-direction dominates, and therefore mg /(2) should be approximately equal to l . We then see that is negative, so that the solution is oscillatory and the frequency is given by = 2/m = g /l . The dependent terms in the equation of motion represent indeed a force in the +y direction of magnitude mg : this is the tension in the string or rod on which the weight is suspended. When using polar coordinates, we have L=m r 2 (r )2 + + mg cos , 2 2 (4.41)

with the constraint r = l . which leads to the Lagrange equations: m r = mg cos + mr 2 ; m r 2 + r r = mg r sin . Filling the constraint is particularly easy. The constraint force is given by = mg cos + mr 2 . (4.44) (4.42) (4.43)

The constraint force consists of a term which compensates for the gravity force (rst term) and an extra term which is necessary for keeping the circular motion going (a centripetal force, the second term). The equation for reduces to the usual pendulum equation when the constraint is used: = g /l sin . (4.45) In practice, constraints are seldom used explicitly in the solution of mechanical problems.

50

4. E XAMPLES OF VARIATIONAL CALCULUS , CONSTRAINTS

4.5.2 G LOBAL CONSTRAINTS


In (4.5.1) we have analysed constraints of the form: g (k) (r1 , r2 , . . . , rN ; t ) = 0. (4.46)

4{

This type of constraint is called holonomic, and it frequently allows us to represent the system using generalised coordinates. This type of constraint imposes conditions on the system which should hold at any moment in time. We may therefore consider this type of constraints as an innite set (one constraint for each time). Such constraints are called local, where this term refers to the fact that the constraint is local in time. Sometimes however, we must deal with constraints of a different form. Consider for example the problem of nding the shape of a chain of homogeneous density (= mass/unit length) suspended at its two end points. We represent this shape by a function y(x) where x is the coordinate along the line connecting the two end points and y(x) is the height of the chain for coordinate x. The shape is determined by the condition that it minimises the (gravitational) potential energy, and it is readily seen that this energy is given by the functional V = g
X

dx y
0

1+

dy dx

(4.47)

We leave out the constants g and in the following as they do not affect the shape. If we would minimise the potential energy (4.47), we would get divergences, as we have not yet restricted the total length L of the wire to have a xed length. This requirement can be formulated as X dy 2 dx 1+ L= . (4.48) dx 0 This is a constraint which is not holonomic, and there is no way to reduce the number of degrees of freedom. This type of constraint is called global as it is formulated as a condition on an integral of the same type as the functional to be minimised, and is not to be satised for all values of the integration variable. Therefore, we must generalise the derivation of the Euler equations to cases in which a functional constraint is present. Let us consider two functionals, J and K :
b

J= K=

d x F (y, y , x)
a b

(4.49a) (4.49b)

d x G(y, y , x),
a

and suppose we want to minimise J under the condition that K has a given value, i.e., for each variation y which satises: G(y, y , x) d x = 0; we require that F (y, y , x) d x = 0. (4.51) (4.50)

Consider now a particular variation which is nonzero only in a small neighbourhood of two values x 1 and x 2 (see gure 4.1). If the areas under these two humps are A 1 and A 2 respectively, we have
b a

G(y, y , x) d x = A 1

G[y(x 1 ), y (x 1 ), x 1 ] G[y(x 2 ), y (x 2 ), x 2 ] + A2 , y y G 1 /y A2 = G 2 /y A1

(4.52)

and therefore

(4.53)

4.5. C ONSTRAINTS

51

x1
F IGURE 4.1: Example of a function y(x) which is nonzero only near x 1 and x 2 .

x2

(0,0)

(X,0)

{ {

x
F IGURE 4.2: The cosh solution to the suspended chain problem.

with an obvious shorthand notation. Applying this argument once again we see that for functions y(x) satisfying requirement (4.53), we should have F 1 /y A2 = (4.54) F 2 /y A1 But this can only be true for arbitrary x 1 and x 2 when F /y and G/y are proportional: F G = . y y Therefore, we must solve the Euler equations for the combined functional J (y) K (y) (4.56) (4.55)

where is xed by putting the solution of this minimisation back into the constraint. This is the Lagrange multiplier theorem for functionals. We shall now apply this to the suspended chain problem. We have F = y G= 1 + y 2 . Therefore, the Euler equations read:
(y + )

1+ y

and

1+ y 2

2 2

= Constant,

(4.57)

1+ y

which leads to y + =C The solution is given by y(x) = A cosh [(x x 0 )] + B (4.59) 1 + y 2. (4.58)

52 with

4. E XAMPLES OF VARIATIONAL CALCULUS , CONSTRAINTS

A = C = 1/ B = .

(4.60a) (4.60b)

Boundary conditions are y(0) = y(X ) = 0 and the length of the wire must be equal to L. These conditions x x 0 and : x 0 = X /2, = C cosh [X /(2C )] and C = L/ {sinh [X /(2C )]}. In gure 4.2 the solution is shown.

4.6 S UMMARY

4{

In this chapter we have visited a few applications of functional minimization to several everydaylife problems. An expression
b

J=

F (y, y , x)d x
a

(4.61)

where y(x) is a function dened between a and b, and y its derivative with respect to x, assumes its extremal values when d F (y, y , x) F (y, y , x) = . d x y y (4.62)

This is the general form of the Euler equation. We have seen that, for the particular case where F does not depend explicitly on x, the Euler equation can be integrated once, leading to the new equation F (y, y ) F (y, y ) y = C, (4.63) y where C is an integration constant. We also considered constraints of two kinds: local and global. A local constraint is a condition on the function y which should hold for any x. If a set of K such constraints of the form g (k) (r1 , . . . , rN ) = 0, k = 1, . . . , K (4.64)

holds for an N -particle problem, then the Euler Lagrange equations take the form
K d L L = + j k g (k) (r1 , . . . , rN ). d t j r j k=1 r

(4.65)

The solutions for r j obviously depend on the parameters k these parameters are xed at the end by requiring that the solutions r j satisfy the constraints. Now we turn to the global constraints: they are conditions on the function y as a whole, usually formulated in terms of one or more integrals. In that case, if we want to minimize the functional J = F (y, y , x)d x where y is subject to the constraint K (y) = G(y)d x = constant, (4.66)

This problem is solved by the solutions of the Euler equation of the functional J K = F (y, y , x) G(y, y , x) d x. (4.67)

The resulting solution depends on the Lagrange parameter , which is solved by substituting the solution into the constraint equation.

5
S YMMETRY AND CONSERVATION LAWS

53

54

5. S YMMETRY AND CONSERVATION LAWS

In this chapter, we return to classical mechanics and shall explore the relation between the symmetry of a physical system and the conservation of physical quantities. In the rst chapter, we have already seen that translational symmetry implies momentum conservation, that time translation symmetry implies energy conservation and that rotational symmetry implies conservation of angular momentum. There exists a fundamental theorem, called Noethers theorem, which shows that, indeed, for every spatial continuous symmetry of a system which can be described by a Lagrangian, some physical quantity is conserved, and the theorem also allows us to nd that quantity. The special form of the equations of motion for a system described by a Lagrangian (or Hamiltonian) leads already to a large number of conserved quantities, called Poincar invariants. We shall consider only one Poincar invariant here: phase space volume. The associated conservation law is called Liouvilles theorem.

5.1 N OETHER S THEOREM

{ {

Suppose a mechanical system is invariant under symmetry transformations which can be parametrised using some real, continuous parameter. Examples include those mentioned already above: rotations (parametrised by the rotation angles) or translations in space or time. The fact that the system is invariant under these transformations is reected by the Lagrangian being invariant under these symmetries. For simplicity we shall restrict ourselves to a single continuous parameter, s. In the case of rotations one could imagine s to be the rotation angle about an axis xed in space, such as the z-axis. The mechanical path for some system, i.e. the solution of the Euler-Lagrange equations of motion, is called q(t ). Now we perform a symmetry transformation. This gives rise to a different path, which we call Q(s, t ), with Q(0, t ) = q(t ). The path Q(s, t ) should have the same value of the Lagrangian L as the path q(t ), in other words, L should not depend on s: d L(Q(s, t ), Q(s, t )) = 0. ds This leads to
N j =1

(5.1)

L Q j L Q j + Q j s Q j s

= 0.

(5.2)

Now we use the Euler-Lagrange equations: L d L = Q j d t Q j in order to write


N dL L Q L Q j = + d s j =1 Q j s Q j s N

(5.3)

=
j =1

d L Q j L Q j + d t Q j s Q j s

d dt

L dQ j j =1 Q j d s
N

= 0,

(5.4) and we see that the term within brackets in the last expression must be a constant of the motion: N L dQ N dQ j j = pj = Constant in time. (5.5) j d s ds j =1 Q j =1 We see that any continuous symmetry of the Lagrangian leads to a constant of the motion, given by (5.5). This analysis is obviously rather abstract, so let us now consider an example. Suppose a one-particle system in three dimensional space is invariant under rotations around the z-axis. The rotation angle is called . In order to be able to evaluate the derivatives of the coordinates with respect to , we use cylindrical coordinates (r, , z) with x = r cos and y = r sin . A rotation about the z axis over an angle then corresponds to + . (5.6)

5.2. L IOUVILLE S THEOREM so that we have dx = p x r sin( + ) = p x y; d dy py = p y r cos( + ) = p y x; d dz =0 pz d px so that the conserved quantity, from (5.5) is xp y y p x = L z ,

55

(5.7) (5.8) (5.9)

(5.10)

the z-component of the angular momentum. Similarly, we would nd L x and L y for the conserved quantities associated with rotations around the x- and y- axes respectively. Also, it is easy to verify that for more than one particle, the total angular momentum is conserved. The reader is invited to check that space translation symmetry results in momentum conservation.

{ {

5.2 L IOUVILLE S THEOREM


A special conservation law is due to the fact that the equations of motion can be derived from a Hamiltonian (or from a Lagrangian). Such equations of motion are called canonical. The fact that the equations of motion are canonical reects a symmetry which is called symplecticity (or symplecticness), a discussion of which is outside the scope of these notes. The important notion is that this type of symmetry leads to a number of conserved quantities, called Poincar invariants, of which we shall consider only one, the volume of phase space. The proof of Liouvilles theorem hinges upon the fact that whenever in a volume integral like V = dnx (5.11)

we perform a variable transformation x y, we must put a correction factor det(J) in the integral, where J is the Jacobian matrix, given by Ji j = We thus have V=

y i . x j d n y.

(5.12)

d n x det(J) =

(5.13)

where is the volume transformed to y-space. The state of a mechanical system consisting of N degrees of freedom is represented by a point in 2N -dimensional phase space (q i , p j ). In the course of time, this point moves in phase space, and forms a trajectory. We now consider not a single mechanical system in phase space, but a set of systems which are initially homogeneously distributed over some region 0 , with volume V0 . In the course of time, every point in 0 will move in phase space, 0 will therefore transform into some new region (t ). The volume of this new space is given as V (t ) = dnq dnp (5.14)
(t )

We want to show that V (t ) = V0 , hence the volume does not change in time. To this end, we consider a transformation from time t = 0 to t : H (p, q) t + O (t 2 ); p i H (p, q) p i p i (t ) = p i (0) t + O (t 2 ), q i q i q i (t ) = q i (0) + (5.15) (5.16)

56

5. S YMMETRY AND CONSERVATION LAWS

where we have used a rst order Taylor expansion and replaced time derivatives of q i and p i using Hamiltons equations. Now we can evaluate the original volume V0 as follows: V0 = dnq dnp = d n q d n p det [J(t )] . (5.17)

(t )

The Jacobi determinant can be written in block-form as follows: det [J(t )] =


H 1 + t qi p j
2

H t p i p j
2

H 1 t q j p i
2

H t qi qi
2

(5.18)

Careful consideration of this expression should convince you that det [J(t )] = 1+O (t 2 ). We see therefore that V (t ) = V0 + O (t 2 ), from which it follows that dV (0) = 0. dt (5.19)

5{

This argument can be extended for arbitrary times, so that we have proven that V is a constant of the motion. We have found Liouvilles theorem in the form: The volume of a region phase occupied by a set of systems does not change in time. Of course, the region can change in shape, but its total volume will remain constant in time. We could have put any density distribution of points in phase space in the integrals, which does not change the derivation. Liouvilles theorem is important in equilibrium statistical mechanics. So-called ergodic systems are assumed to move to a time-independent distribution in phase space, that is, any large set of systems setting off at time t = 0 from different points in phase space and moving according to the Hamiltonian equations of motion will assume the same, invariant distribution after long times. Liouvilles theorem, moreover tells us that the systems will not all end up in the same point in phase space, but spread over a region with a volume equal to the initial volume. There exists a more specic theorem concerning the behaviour of systems in time. This is Poincars theorem, which says that a system which is to evolve under the mechanical equations of motion will always return arbitrarily close to its starting point within a nite time. Consider for example a box partitioned into two sub-volumes (gure 5.1). There is a small hole in the middle, and there are N particles in the right hand volume. Obviously, a fraction of these particles will move to the left hand volume, but the Poincar recurrence theorem tells us that after a nite time, all particles will reassemble in the right hand volume! This seems to be in contradiction with the second law of thermodynamics. This law states that the entropy will not decrease in the course of time. Here we see an increase of entropy when the particles distribute themselves over the two volume halves rather than a single one, but come back in a more ordered (less entropic) state after some time. This is only an apparent contradiction, as the Poincar theorem holds for a nite number of particles (nite-dimensional phase space). What we see here is an example of the inequivalence of interchanging the order in which limits are taken: if we rst take the system size to innity (the approach of statistical mechanics and thermodynamics), the recurrence time will become innite. If, on the other hand, we consider a nite system over innitely large times (the mechanics approach), we see that it returns arbitrarily close to its initial state innitely often. Taking then the system size to innity does not alter this conclusion.

5.2. L IOUVILLE S THEOREM

57

F IGURE 5.1: A box divided into two halves by a wall with a hole. Initially, particles will be in the right hand volume, and will move to the left. After large times, they will all come back to the right hand volume.

'' '( '

 

 

)0 )   

 

    % %& ! !" #$

{ {

6
S YSTEMS CLOSE TO EQUILIBRIUM

Vibrational modes of carbon molecules and methane play a crucial role in the greenhouse effect see section 6.4. The right gure shows the vibrational modes of methane.

59

60

6. S YSTEMS CLOSE TO EQUILIBRIUM

V M S
F IGURE 6.1: System with a metastable (M) and a stable (S) state. A strong perturbation may kick the ball out of its metastable state, and under the inuence of dissipation it will then move to the stable state.

6.1 I NTRODUCTION
When we prepare a conservative system in a state with a certain energy, it will conserve this energy ad innitum. In practice, such is never the case, as it is impossible to decouple a system from its environment or from its internal degrees of freedom. This requires some explanation. We usually describe macroscopic objects in terms of the coordinates of their centre of mass and their Euler angles. These are the macroscopic degrees of freedom. As these objects consist of particles (atoms, molecules), they have very many additional, internal, or microsopic degrees of freedom. In fact the heat which is generated during friction is nothing but a transfer of mechanical energy associated with the macroscopic degrees of freedom to a mechanical energy of the internal (microscopic) degrees of freedom: if we throw a ball at the wall, the molecules of the ball will vibrate more as a result of the collision, resulting in a loss of the mechanical energy associate with the centre of mass motion of the ball. So heat in the end is a form of mechanical energy. As a result of friction, macroscopic objects will, when subject to a conservative, time-independent force (apart from friction), always end up at rest in a point where there potential energy is minimal. Any system at a point where its potential energy is minimal is said to be in a stable state. A system which looses its kinetic energy via friction is called a dissipative system. All the macroscopic systems we know are dissipative, although some can approach conservative systems very well. If the interactions are not all harmonic (harmonic means that the potential energy is a quadratic functions of the coordinates) then there may be more than one minimum. Local minima correspond to metastable states. A system in a metastable state will return to that state under a small perturbation, but, when it is strongly perturbed, it might move to another metastable with lower potential energy, or to the stable state. An example is shown in gure 6.1, for a particle in a one-dimensional potential. Molecular systems, in which we take all degrees of freedom explicitly into account, are believed to be non-dissipative. We know from statistical physics that every degree of freedom in a system in thermal equilibrium carries a kinetic energy equal (on average) to k B T , where k B is Boltzmanns constant. At low temperatures, the energies of the particles are small as can be seen from the Boltzmann distribution, which gives the probability of nding a system with energy E as exp[E /(k B T )]. Therefore, for low temperatures, the kinetic and the potential energy of a system are small. It can therefore be inferred that at low temperatures, a system is close to a (meta-)stable state. In this section, we analyse systems close to mechanical equilibrium. The beautiful result of this analysis is a description in terms of a set of uncoupled harmonic oscillators, which are themselves trivial to analyse. Moreover we obtain a straightforward recipe for nding the resonance frequencies (related to the coupling strengths) of those oscillators.

6{

6.2. A NALYSIS OF A SYSTEM CLOSE TO EQUILIBRIUM

61

6.2 A NALYSIS OF A SYSTEM CLOSE TO EQUILIBRIUM


Consider a conservative system characterised by generalised coordinates q j , j = 1, . . . , N . The system is in equilibrium, dened by q 1 , . . . , q N , if its potential energy is minimal. In that case we have V (q 1 , . . . , q N ) = 0; j = 1, . . . , N . (6.1) q j Now suppose that we perturb the system slightly, i.e. we change the values of the q j slightly with respect to their equilibrium values. As the rst derivative of the potential with respect to each of the q j vanishes, a Taylor expansion of the potential only contains second and higher order terms: V (q 1 , . . . , q N ) = V (q 1 , . . . , q N ) + 1 2
N

(q j q j )(q k q k )
j ,k=1

2 V (q 1 , . . . , q N ) + . . . . q j q k

(6.2)

The terms of order higher than two will be neglected, as we are interested in systems close to equilibrium (i.e. q j q j small). We can represent the resulting expansion using matrix notation. Introduce the matrix 2 V 2 V 2 V q1 q N q1 q1 q1 q2 2 2 V 2 V V q2 q1 q2 q2 q2 q N K= (6.3) . . . . . . . . . . . . . 2 2 V V 2 V q q N q N q 1 q N q 2 N The matrix K is obviously symmetric. We can write 1 V (q 1 , . . . , q N ) = V (q 1 , . . . , q N ) + qT K q 2 (6.4)

{ {

where q is a column vector with components q j q j , j = 1, . . . , N ; the superscript T denotes the transpose of the vector. For clarity, we write the last term in this equation explicitly: K 11 K 12 K 1N q1 K 21 K 22 K 2N q 2 1 1 T . . . q K q = q 1 , . . . , q N (6.5) . . . . . . 2 2 . . . . KN1 KN2 KN N qN Now we write down the kinetic energy in terms of the generalised coordinates. We assume that the constraints only depend on the generalised coordinates q j and not on their derivatives or on time. In that case, the kinetic energy can be written in the form (see page 30): 1 N T= M j k q j qk , (6.6) 2 j ,k=1 where the matrix M is symmetric: M j k = M k j . Note that M j k may depend on the q j . In terms of q j q j , and using vector notation, we can rewrite the kinetic energy as 1 T = T M . q q 2 The equations of motion now read:
N k=1

(6.7)

(M j k q k + q k M k j ) =

N V T + = (K k j q k + q k K k j ). q j q j k=1

(6.8)

We have omitted the dependence of M i j on q j this dependence generates terms on the left and right hand side, which are both of order q 2 and can therefore be neglected. The easiest

62

6. S YSTEMS CLOSE TO EQUILIBRIUM

way of nding the matrix M in practical applications is by writing up the kinetic energy and then identifying each M j k as the coefcient of the term containing q j q k . Formally, we can also formulate M j k as the second derivative matrix of the kinetic energy with respect to the velocity coordinates: 2 T Mjk = . q j q k Using the symmetry of the matrices M and K , (6.8) reduces to
N k=1

M j k q k =

N k=1

K j k q k .

(6.9)

Let us consider the two-dimensional case to clarify the procedure. We consider two generalised coordinates q 1 and q 2 with the matrix M j k equal to the identity. Then, the kinetic energy has the form: 1 2 1 2 T = q1 + q2 . (6.10) 2 2 The potential energy depends on the two coordinates q 1 and q 2 : V = V (q 1 , q 2 ). (6.11)
{

6{

The equations of motion read: V q 1 V q 2 = q 2 = q 2 q 1 = q 1 = Expanding about the point q 1 , q 2 , where V is supposed to be minimal, we have 2 1 V (q 1 , q 2 ) = V (q 1 , q 2 ) + (q 1 q 1 )2 2 V (q 1 , q 2 )+ 2 q 1 (q 1 q 1 )(q 2 q 2 ) This can be written in the form:
1 V (q 1 , q 2 ) = V (q 1 , q 2 ) + (q 1 q 1 , q 2 q 2 ) 2
2 V 2 q 1 2 V q 1 q 2 2 V q 1 q 2 2 V 2 q 1

(6.12) (6.13)

1 2 2 V (q 1 , q 2 ) + (q 2 q 2 )2 2 V (q 1 , q 2 ). (6.14) q 1 q 2 2 q 2

q1 q1 q2 q2

(6.15)

Dening q 1 = q 1 q 1 and similarly for q 2 , this equation reads: 2 V 2 V 2 1 q 1 q 2 q 1 V (q 1 , q 2 ) = V (q 1 , q 2 ) + (q 1 , q 2 ) 2 V 2 V 2 2 q q


1 2

q 1

q 1 q 2

(6.16)

The 2 2 matrix occurring in this expression is our matrix K.

6.2.1 E XAMPLES : A SPRING SYSTEM AND THE DOUBLE PENDULUM


As a trivial example, we consider two masses m 1 and m 2 and three springs with spring constants 1 and 2 and 3 in the arrangement shown in the gure.

x1
1

x2
3

m1

m2

6.2. A NALYSIS OF A SYSTEM CLOSE TO EQUILIBRIUM

63

In terms of the displacements x 1 and x 2 of the masses with respect to the their equilibrium positions when the system is at rest, the Lagrangian is L= m 1 2 m 2 2 1 2 2 2 3 x + x x x (x 2 + x 1 )2 . 2 1 2 2 2 1 2 2 2 (6.17)

We do not have to linearise as the Lagrangian is already in quadratic form. The matrices M and K are 1 + 3 3 m1 0 . (6.18) ; K= M= 3 2 + 3 0 m2 Note the off-diagonal elements which have been chosen such as to make K symmetric. The second, somewhat more interesting example is the double pendulum, consisting of two rigid massless rods of length l and L, with masses M and m:

L
{ {

M l m

The velocity of the upper mass is L , that of the lower one is a vector sum of the velocity of the upper one and that of the lower one with respect to the upper one. For very small angles and both velocities will be approximately in the horizontal direction so that they can simply be added: v m = L + l . The kinetic energy therefore reads: T= M 2 m L + l 2 . L + 2 2 (6.19)

Let us perform a transformation to more convenient variables x = L y = L + l . (6.20a) (6.20b)

Note that x and y do not denote cartesian coordinates. In that case the kinetic energy can simply be written as M m T = x 2 + y 2. (6.21) 2 2 The potential energy of the upper mass is M g L(1 cos ) M g L2 /2, and that of the lower mass is given as mg L(1 cos ) + l (1 cos ) , which, in the small angle approximation becomes 1 VLower (, ) = mg L2 + l 2 . (6.22) 2 The total potential energy, written in terms of x and y, therefore reads: V= (M + m)g 2 mg x + (y x)2 . 2L 2l (6.23)

64 The equations of motion can therefore be written as M 0 0 m x y = M +m g L


mg l mg l

6. S YSTEMS CLOSE TO EQUILIBRIUM

mg l mg l

x y

(6.24)

This is the form given in (6.9). We shall return to this example in the next section.

6.3 N ORMAL MODES


The equations of motion of the double pendulum would be easy to solve if both matrices M and K would be diagonal. In that case, we would end up with a set of uncoupled harmonic oscillators. It is possible indeed to nd a linear transformation of the generalised coordinates q j such that these matrices are diagonal. The key for nding the diagonalised representation resides in nding solutions to Eq. (6.16) of the form q j = A j e i t (6.25)

6{

where does not depend on j all the degrees of freedom oscillate at the same frequency. Such a motion is called a normal mode. In the following we shall use q j instead of q j : q j is the generalised coordinate measured with respect to its equilibrium value. We have q j = 2 q j , so (6.9) reduces to
N k=1

M j k 2 A k =

K j k Ak .
k=1

(6.26)

If the mass tensor M j k would be the identity, Eq. (6.26) would be an eigenvalue equation. For general mass tensors, the equation is a generalised eigenvalue equation. We can reduce this equation to an ordinary eigenvalue equation by multiplying the left and right and side by the inverse M1 of the mass matrix. We then have: 2 A j =
N k,l =1

M 1 K kl A l . jk

(6.27)

In algebraic terms, the solutions to these equations are the eigenvectors A (with components A j ) and the corresponding eigenvalues 2 . In physical terms, the components A j are the amplitudes of the oscillatory motions of the generalised coordinates, and is the frequency of the oscillation. In order for the normal modes to exist, the eigenvalues should be real. That the eigenvalues are indeed real follows from the fact that both M and K are real, symmetric matrices. It is a well-known result of linear algebra that the eigenvalues of the generalised eigenvalue problem are then real. Another question is whether the eigenvalues are positive or negative. Assuming that we are expanding the potential around a minimum, the matrix K can be shown to be positive denite. A positive denite matrix has only positive eigenvalues1 . Moreover, the mass matrix can be shown to be positive. Then its inverse M1 is also positive. Multiplying two positive matrices yields a product matrix which is positive. Therefore M1 K is positive, and the 2 are positive. Hence the frequencies of the oscillations are always real we do not nd expontial growth or decay. In physical terms one could say that perturbing the system from equilibrium always pushes it back to this equilibrium therefore the spring force experienced by the coordinates is always opposite to the perturbation, and therefore an oscillation arises, and not a drift away from equilibrium, or some exponential decay. Such decay may however be found near a local maximum or near a saddle point of the potential.
1 In fact, we shall occasionally allow for zero eigenvalues; in that case, the matrix is called positive semidenite.

6.3. N ORMAL MODES

65

6.3.1 N ORMAL COORDINATES


We will now show that it is possible to linearly transform the generalised coordinates such that both M and K assume a diagonal form. First note that the matrix M is real and symmetric therefore its eigenvalues are real. Furthermore they are positive, as we have just noted. So, M can by put in diagonal form by an orthogonal transformation O (as you should know from your linear algebra course): OT MO = d, (6.28) where the superscript T denotes the transpose as usual, and d is a diagonal matrix with the eigenvalues d j of M on the diagonal. It is easy to see that we can bring d to unit form by multiplying on the left and right by the matrix d1/2 , which is a diagnonal matrix with the values 1/ d j . The positivity guarantees that the 1/ d j are real. So we can write (6.29)

d1/2 OT MOd1/2 = I (I is the unit matrix). Using the fact that OOT = OT O = I, Starting now from the generalized eigenvalue equation 2 MA = KA,

(6.30)
{ {

(6.31)

where A is a vector and M and K are matrices, and multiply on both sides from the left with d1/2 OT we get 2 d1/2 OT MA = d1/2 OT KA. (6.32) Now we insert on the left and right hand side the unit matrix in the form I = Od1/2 d1/2 OT , to obtain 2 d1/2 OT MOd1/2 d1/2 OT A = d1/2 OT KOd1/2 d1/2 OT A. Using (6.29), and dening K = ST KS where S = Od1/2 , we then have 2 d1/2 OT A = Kd1/2 OT A. If we now dene A = ST A, the result can concisely be written in the form 2 A = K A , (6.39) (6.38) (6.37) (6.36) (6.35) (6.34) (6.33)

in other words, we must diagonalise the matrix K . Having done so, we have found a set of vectors A which, after transforming them back to A: A = ST
1

A,

(6.40)

give the eigenvectors of the original generalised eigenvalue problem (6.27). It is easy to show that K is symmetric (use (AB)T = BT AT and the fact that K is symmetric): K
T

= d1/2 OT KOd1/2

= d1/2 OT KOd1/2 .

(6.41)

66

6. S YSTEMS CLOSE TO EQUILIBRIUM

so the set of vectors An is, or can be chosen orthogonal. It may seem strange that we can always nd simultaneous eigenvectors of the transformed M (i.e. the unit matrix) and K , while these two matrices need not commute. It should however be noticed that we do not diagonalise as usual through a similarity transform (i.e. by an orthogonal transformation) but allow for a more general, linear transformation (a so-called congruence transform). It turns out that the eigenvectors are not orthogonal as would be the case with a similarity transform (as is done in quantum mechanics). We can now rewrite the Lagrangian as follows: 1 1 L = qT M qT Kq = q 2 2 1 T T q S 2

1 T

1 S MSS1 q qT ST 2

1 T

S KSS1 q = 1 T 1 T q q q K q , (6.42) 2 2

6{

where in the rst step, we have inserted unit matrices of the form SS1 etcetera, and in the second step we have used q = S1 q. Let us now write q = C n An , (6.43)
n

then, using the fact that the An can be taken orthonormal, we see that the Lagrangian takes the form 1 C n C n C n 2 C n . L= (6.44) n 2 n This represents a system consisting of a set of uncoupled harmonic oscillators. The new coordinates C n are called normal coordinates. Note that the vectors An = SAn satisfy the equations (6.26). These vectors need not be orthonormal, but the vectors An are. As a rst example of normal modes, we consider the system consisting of two masses and three springs. For simplicity, we take the masses and the spring constants equal m and respectively). We then obtain the following equation: m2 2 . m2 2 This yields a quadratic equation in 2 with solutions 2 = The eigenvectors are 1 1 and 1 1 (6.47) and 2 = 3 . m m (6.46) (6.45)

respectively. The rst mode corresponds to the two masses moving parallel and feeling only the springs with which they are attched to the walls. The middle spring does not inuence the motion and the frequency is given by a single mass and single spring (2 = /m). In the second mode, the masses move in opposite directions. If we introduce the coordinates u= v= we see that the Lagrangian takes the form L= m 2 u + v 2 v 2 3 u2, 2 2 2 (6.49) x1 + x2 2 x1 x2 2 , (6.48a) (6.48b)

6.4. V IBRATIONAL ANALYSIS IN MOLECULES

67

so that we have indeed written the Lagrangian as a sum of two Lagrangians of single harmonic oscillators: this is the normal coordinate representation discussed above. As a second example, we consider nding the normal modes for the coupled pendulums. Note that this problem is relatively simple as a result of the fact that the mass matrix is diagonal and therefore trivial to invert. After multiplying both sides of Eq. (6.24) by M 1 , we have a standard diagonalisation problem for the matrix:
M +m ML g g

mg Ml

Ml
g l

mg

(6.50)

The eigenvalues are the solutions of the so-called secular equation which has the from
M +m ML g

+ M l 2 g l

mg

Ml g 2 l

mg

= 0.

(6.51)

This reduces to the following quadratic equation in 2 : 4 M +m g g 2 M +m g2 + + = 0. M L l M Ll (6.52)


{ {

This equation has two solutions for 2 . We will examine some special cases. If M m, then, provided that l is not too close to L, the two roots with corresponding eigenvectors (A x , A y ) are given by and g ; L Ax L l . Ay L (6.54) g ; l Ax m L Ay M l L (6.53)

The rst solution describes an almost stationary motion of the upper pendulum with the lower one oscillating at its natural frequency. In the second case, the motion of the upper and lower are of the same order of magnitude with the natural frequency of the upper pendulum. If M m, the solutions are 2 and 2 g L +l Ax L = Ay L + l Ax m L +l . Ay M L (6.55)

m g g + , M L l

(6.56)

The rst case describes a motion in which the two rods are aligned so that we have essentially a single pendulum of length l + L and mass m. The second case corresponds to a very high frequency of the upper mass with an almost stationary lower mass.

6.4 V IBRATIONAL ANALYSIS IN MOLECULES


The way in which atoms are bound together in molecules is described by quantum mechanics. There is a long standing tradition in the quantum mechanical calculation of stationary states of molecules. In the last fteen years or so it has become possible to perform dynamical computations of molecules to very good accuracy using fully quantum mechanical calculations. These calculations are quite demanding on computer resources and they do not always give a very good insight into the dynamics of interest. Therefore, a semi-classical approach is often adopted in order to calculate vibration spectra for example. First, the total energy of the molecule is calculated as a function of the nuclear positions Ri , i = 1, . . . , N for an N -atomic molecule. There is however a problem in doing this. Suppose

68

6. S YSTEMS CLOSE TO EQUILIBRIUM

stretch

bend

torsion
F IGURE 6.2: Interactions in a molecule.

2
6{
{

1 m
F IGURE 6.3: Triatomic molecule.

. we want to calculate this energy for 10 values of all the coordinates of a 10-atom molecule. As there are 30 coordinates, we need to perform 1030 stationary quantum calculations, which would require the age of the universe. Therefore the potential is parametrised in a sensible way, which we now describe. All the chemically bonded atoms are described by harmonic or goniometric interactions. The degrees of freedom chosen for this parametrisation are the bond length, bond angle and dihedral angle. The forces associated with these degrees of freedom are called stretch, bend, and torsion respectively. These degrees of freedom are shown in gure 6.2 The form of the potentials associated with bond stretching is given as VStretch = (l l 0 )2 2 (6.57)

where l is the bond length and l 0 is the equilibrium bond length. The spring constant determines how difcult it is to stretch the bond. The bending potential is given in terms of the bond angle : (6.58) VBend = ( 0 )2 2 A similar expression exists for the torsional energy. The constants and can be determined from stationary quantum mechanical calculations. Assuming that these parameters are known, we shall now use the given form of the potential to calculate the vibration spectrum of a triatomic, linear molecule, such as CS2 or CO2 (see gure 6.3). We neglect bending here, so only bond stretching is taken into account. If the initial conguration is linear, the motion takes place along a straight line, which we take as our X -axis. The coordinates of the three atoms are x 1 , x 2 and x 3 . The kinetic energy

6.4. V IBRATIONAL ANALYSIS IN MOLECULES can therefore be written down immediately: T= The potential energy is given by V= (x 2 x 1 l )2 + (x 3 x 2 l )2 . 2 2 2 m 2 2 x + x3 + x2 . 2 1 2

69

(6.59)

(6.60)

Here, l is the equilibrium bond length. The centre of mass of the system will move uniformly as there are no external forces acting, and we take this centre as the origin. The equilibrium coordinates are then x 1 = l , x 2 = 0 and x 3 = l . The deviations from these values are x 1 = x 1 + l ; In this representation, we have T= and m 2 2 2 x 1 + x 3 + x 2 . 2 2 (6.62) x 2 = x 2 and x 3 = x 3 l . (6.61)

(x 2 x 1 )2 + (x 3 x 2 )2 . 2 2 We can nd the matrices K and M directly from these expressions: V= 0 0 m M= 0


0 0 ,

(6.63)

{ {

(6.64)

and
1 1 0 2 1 . K = 1 0 1 1

(6.65)

The normal modes can now be found by solving (6.26) with these matrices. The eigenvectors can be found by solving the secular equation: 2 0 This leads to: ( 2 )2 (m2 m 2) = 0, from which we nd: 1 = 0; 2 = ; 3 = 1 2 + . m (6.68) (6.67) 2 m2 0 2

= 0.

(6.66)

The corresponding eigenvectors can be found after some algebra:


1 A1 = 1 ; 1 1 A2 = 0 ; 1 1 A3 = 2/m . 1

(6.69)

The rst of these, corresponding to 1 = 0, is a mode in which the atoms all slide in the same direction with the same speed. This is a manifestation of the translational symmetry of the problem, which has been recovered by our procedure. The second one represents a mode in which the middle atom stands still and the two outer atoms vibrate oppositely. Obviously,

70

6. S YSTEMS CLOSE TO EQUILIBRIUM

2 1 m 3

Mode 1 Mode 2 Mode 3


F IGURE 6.4: The three modes of the triatomic molecule.

6{

the frequency of this mode is 2 = /m, corresponding to the two springs. Finally, the last mode is one in which the two outer atoms move in one direction, and the central atom in the opposite direction. The motion can be understood by replacing the two outer masses by a single one with mass 2 at their midpoint, coupled by a spring with spring constant 2 to the central mass. The reduced mass of this system (1/(2)+1/m)1 then occurs in the expression for the resonance frequency. The three modes are depicted in gure 6.4. The vibrations of the CO2 molecule are very important: CO2 is one of the notorious greenhouse gases which are held responsible for part of the global warming. This is because infrared radiation emanating from the earth to cool down its surface, are partly absorbed by the vibrational modes of the greenhouse gases in the atmosphere. After some time, a vibrationally excited molecule relaxes by emitting radiation, part of which reaches the earth again. In addition to the modes we have studied above, there are bending modes and rotations being excited in the molecule. The vibrational spectrum of a molecule can be measured by studying the absorption of infrared radiation or Raman scattering. Obviously, the frequencies depend on the value of the spring constant, which is difcult to nd. But the ratio between the two frequencies turns out independent of : we nd 3 = 2 1+2 1.9 m (6.70)

where we have used the masses for the carbon (12 amu) and oxygen (16 amu). The frequencies have been measured at 2349 cm1 and 1388 cm1 which have a ratio of about 1.7. The difference with our result is explained by the interaction between the two oxygen atoms which we have neglected in our simple analysis.

6.5 T HE CHAIN OF PARTICLES


In the previous section we have analysed a triatomic molecule. Now we shall analyse a larger system: a chain of N particles. We assume that all particles have the same mass, and that they are connected by a string with tension . The particles are assumed to move only in the vertical (y) direction, and the x-components of adjacent particles differ by a separation d . The rst and last spring are connected to points at y = 0. The chain is depicted in gure 6.5. The chain is a model for a continuous string, which is obtained by letting N and d 0 while keeping the string length N d xed. Let us consider particle number k. The springs connecting this particle to its neighbours are stretched, and this may result in a net force acting on particle k. The spring between

6.5. T HE CHAIN OF PARTICLES

71

k1
d

yk

k k+1

F IGURE 6.5: The harmonic chain of particles.

particle k and k + 1 has a length l= d 2 + (y k+1 y k )2 d + (y k+1 y k )2 2d (6.71)

where a rst order Taylor expansion is used to obtain the second expression. The potential energy for this link is equal to the tension times the extension of the string, and therefore we nd for the total potential energy: V= with y 0 = y N +1 = 0. The kinetic energy is given by T= We now nd the matrices M kl and K kl as: M kl = mkl (6.75) m 2 yk . k=1 2
N

2d

N k=0

(y k+1 y k )2

(6.72)

{ {

(6.73)

(6.74)

where kl is the Dirac delta function, in other words, M kl is m times the unit matrix. For K kl we nd: 2d d 0 0 0 0 0 0 0 d 2d d K = 0 2 0 0 . (6.76) d d d . . . . . . .. . . . . . . . . . . . . . Note that the indices of this matrix run from 1 to N : the end points are not degrees of freedom. The normal mode equation (6.26) can be solved analytically for arbitrary N by substituting for the eigenvector A k = exp(i k), where is some constant. This trial solution does not satisfy the boundary equations (6.73), but we do not bother about this for the moment. Then for 2 k N 1 we nd m2 e i k = e i (k1) + 2e i k e i (k+1) d (6.77)

Dividing left and right hand side by exp(i k), we nd m2 = 2 (1 cos ). d (6.78)

For each , there is also a solution for for the same . This can be used to construct a solution A k = (e i k e i k ) = 2i sin k. (6.79)

72

6. S YSTEMS CLOSE TO EQUILIBRIUM

This solution always vanishes at k = 0 and it vanishes also at k = N +1 when (N +1) = n, for integer n. So the conclusion is that for each n = 0, . . . , N , we have a solution which vanishes at the two ends of the string. For values of n higher than N , or lower than 0, the solutions obtained are identical to the solutions with 0 n N . For each solution, all particles move up and down with the same frequency, given by (6.78). The wavelength is given by kd such that k = 2, so = 2d /, and the wavevector q = /d . It is possible to formulate the Lagrangian directly in a continuum form, and derive the wave equation from this. Note that in the continuum limit, and d small, we obtain from (6.78) for the frequency: 2 d 2 = = q 2. (6.80) md m Comparison with the well known dispersion equation = c q, we learn that the sound speed c is given as d /m. Dening the density = m/d , we have c= . (6.81)

6.6 S UMMARY

{ {

In this chapter, we have considered normal modes: vibrational excitations of a system close to equilibrium where all the degrees of freedom vibrate at the same frequency. Our analysis started by Taylor-expanding the potential leading to 1 V = V + qT Kq, 2 (6.82)

where q = q q is the vector of displacements with respect to the equilibrium position (for which V /q j = 0 for all j ). In this expression, K is the matrix containing the second derivatives of the potential: 2V K jk = . (6.83) q j q k We have also written the kinetic energy as a quadratic expression in the velocities: 1 T = T M , q q 2 where M is the mass matrix: Mjk = 2 M . q j q k (6.84)

(6.85)

Note that q j = q j . The Euler Lagrange equations can be formulated concisely using the matrices M and K:
N k=1

M j k q k =

N k=1

K j k q k .

(6.86)

Requiring that all degrees of freedom q j vibrate at the same frequency yields the matrix equation: 2 MA = KA, (6.87) where the vector A should satisfy this generalised eigenvalue problem dened by the matrices M and K. This equation can sometimes be solved by educated guessing of the eigenvectors. More generally, the frequencies are found as the zeroes of the polynomial 2 M K = 0. (6.88)

For each frequency found, putting it into the generalised eigenvalue equations allows us to nd the corresponding eigenvectors. It turns out that the Lagrangian can be written as a sum of Lagrangians, each of which describes an independent harmonic oscillator.

A
P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS

73

74 * ** ***

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS = Problems which test you elementary knowledge, you should be able to do these correctly! = Problems on the level of the exam = Challenging (difcult) problems

{ {

Answers can be found at the end.

N EWTON S LAWS AND CONSERVATION LAWS


* 1. A chain of length l is lying on a frictionless table and then released. A length y 0 of the chain hangs freely over the edge, so that the chain is pulled downward. Give the length of the chain as a function of time until the moment where the chain is falling completely freely. * 2. (a) What is the moment of inertia with respect to an exis through the center of a (nonmassive) wheel with radius r and mass m? (b) The same question for an axis through the rim of the wheel. (c) Give the acceleration with which such a wheel rolls down a slope which makes an angle with the horizon. (d) What should be the minimal friction coefcient to prevent the wheel from sliding? * 3. A comet approaches the solar system with velocity v and would, if the sun would not attract the comet, pass the sun at a distance d . What is in reality the shortest distance between comet and sun. (Hint: use conservation of angular momentum and energy).

C ONVENIENT COORDINATES
* 4. Two rigid bodies, attached with a small ball bearing, can move freely through space. How many degrees of freedom does this system possess? Explain your answer. If one of the two bodies is attached to a second ball bearing which is itself xed, how many degrees of freedom are then left? ** 5. A uniform rod of mass M and length l is resting with one end against a wall and starts sliding. There is no friction with the soil and the wall. How many degrees of freedom does this system have if you consider it to be two-dimensional? Express the kinetic energy in terms of these degrees of freedom. ** 6. A bead of mass m slides along a smooth conic spiral. Use cylinder coordinates in this problem. Assuming that = az and = bz, with a and b constants, show, using dAlemberts principle, that the equation of motion reads: z (a 2 + 1 + a 2 b 2 z 2 ) + a 2 b 2 z z 2 = g .
..

(A.1)

L AGRANGE -H AMILTON FORMALISM


** 7. (a) Construct the Lagrangian for a mass m attached to a pendulum of xed length l and moving in a plane. (b) Construct the Lagrangian for two masses m 1 and m 2 attached to both ends of a rope hanging over a pulley. The mass m 1 is connected to one end of the rope through a spring with spring constant k. The radius of the pulley is R and its moment of inertia is I .

75 (c) A beam of length L and mass M is connected on one end through a pivot on the ceiling such that it can move in a plane. At the other end, a mass m is attached to the beam. The size of this mass is small enough for it to be considered as a point mass. The moment of inertia I of the beam with respect to an axis through its end is M L 2 /3. How many degrees of freedom does this system have? Choose suitable coordinates and give the Lagrangian. (d) A trolly on four wheels can move frictionless on a table in the x-direction. The mass of the trolley (including its wheels) is M and the wheels each have a radius a and moment of inertia I . On top of the trolley a second mass m is placed, which is attached to the trolly through a spring with spring constant . This second mass is also allowed to move in the x-direction only. Give the Lagrangian for this system.

{ {

** 8. A mass m is attached to a pendulum with a xed length l which is suspended at a point p. The point p itself can move freely along a horizontal axis, but it is attached to two springs along the x-direction and both with spring constant k.

(a) Choose suitable generalised coodinates. (b) Construct the Lagrangian for this system and derive the equations of motion from it. *** (c) Assume that the mass is displaced a little from its equilibrium state and then released. Neglecting the mass of the springs, show that the pendulum starts oscillating with period mg + 2kl . 2kg

T = 2

(A.2)

Hints: (1) Use the small-angle approximation, (2) neglect all terms in the equation of motion of second order in the generalised coordinates and (3) assume that the pendulum and the springs oscillate with the same frequency (this implies that the acceleration of the angle of the pendulum woth the vertical axis is 2 , and similar for the motion of the springs).

** 9. Assume that the block B , to which is attached a spring, is driven in the vertical direction according to s = A sin(t ). Figure (a) shows the position of the spring if there is no external force (so also no gravitational force) acting on B and m is in equilibrium. Figure (b) shows an arbitrary position of the system. The spring has length l in equilibrium. Show that the height of the mass m with respect to its equilibrium position is given by

q = C sin( k/m t + ) +

Ak sin t . m(k/m 2 )

76

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS

{ {

0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 111 0 000 1 111 0 000 1 1 0 11111111111 00000000000 1 0 111 000 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0

l = constant

11 00 11 00 11 00

1 0 1 0 1 0 1 0 1 0 1 0 1 0 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1

(a)

(b)

** 10. A uniform rod AB with mass M and length l is suspended to a frictionless hinge A. The end B is attached to a spring as shown in the gure. For = 0, the spring is stretched. Neglect the mass of the spring and show that V for this system is given by:

1 1 V = M g l cos + k[(s 2 + l 2 2sl cos )1/2 l 0 ]2 . 2 2 Assume to be small and show that the oscillation frequency is given by:

3g 3ks(l s + l 0 ) + . 2l M l (l s)

A l M s k B

** 11. Consider the following:

77

r k
m1

{ {

1111111111111111 1111111111111111 0000000000000000 0000000000000000 1111111111111111 1111111111111111 0000000000000000 0000000000000000 1111111111111111 1111111111111111 0000000000000000 0000000000000000 1111111111111111 1111111111111111 0000000000000000 0000000000000000 1111111111111111 1111111111111111 0000000000000000 0000000000000000 1111111111111111 1111111111111111 0000000000000000 0000000000000000

l
m2
The mass m 1 can move freely on the table; the mass m 2 can only move along the vertical direction. (a) Show that the Lagrangian has the form: 1 1 1 L = m 1 r 2 + r 2 2 + m 2 l 2 + m 2 g l k(l + r b)2 , 2 2 2 where b is the length of the string plus the length of the spring in equilibrium. (b) Show that the quantity J z = m 1 r 2 is constant in time. What does this quantity represent? *** (c) Solve the equation of motion for J z = 0. ** 12. Deze opgave is bedoeld voor de genteresseerden niet verplicht! Beschouw een functie f (p k , q k ) die afhangt van alle impulsen p k en coordinaten q k . We kijken naar de tijdsafhankelijkheid van een dergelijke functie. De tijdsafhankelijkheid van f wordt gegeven als df = { f , H }. dt Werk de uitdrukking { f , H } in het rechterlid uit. De algemene denitie van { f , g } is: {f ,g} =
j

f g f g . q j p j p j q j

Neem nu f = p j , respectievelijk f = q j en laat zien dat je zo de Hamilton vergelijkingen terugvindt. * 13. From the denition of the Lagrangian in terms of the Hamiltonian, L = p q H (p, q) via d L derive the Hamilton equations: H p H p = q q= af. Hint: use q and q as independent coordinates. * 14. See problem ** 11

78

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS (a) Find the momenta corresponding to the gegeneralised coordinates used in that problem. (b) Formulate the Hamiltonian for this problem. ** (c) Consider the same problem, but now without the spring. When the motion starts, the mass m 2 is at rest and m 1 is at a distance r 0 ; r is then 0. The value of J z is then J 0 . J 0 and r 0 are therefore assumed to be known. Calculate r at the moment where m 2 has descended over a distance d . Hint: Use the conservation of energy.

{ {

** 15. A particle experiences a conservative force F c in addition to a frictional force of the form: F w = r . Is it possible to derive this force from a (generalised) potential W (r, r) as is done in the lecture notes for the electromagnetic force? If the answer is yes, give W ((r, r), otherwise, explain why this is not possible. What would be the best way to analyse a problem with constraints and friction? *** 16. A coin with radius a is dropped on a stone table and starts to spin. After some time, the coin will come to a rest. The precise mechanism leading to this damping are still subject to debate. Without damping, the motion will persist forever; the total energy will then remain constant. Considering the axis d perpendicular through the centre of the coin, we assume that the angle between this axis and the vertical remains constant and that the angular velocity with which this axis precesses around the vertical, is also constant in time. We us the notation of subsection 2.9.2. (a) If the coin is rolling on its edge, the contact point with the table describes a circle with radius a cos . If the coin does not slip, we can obtain a relation between the precision speed and the angular spin velocity around d. Find this relation. (b) Give an expression for the angular velocity vector of the coin. (c) Because of the symmetry of the coin, it is convenient to use instead of the three axes z, e, d, the perpendicular axes e, d. Give an expression for in the e, d f, f, system. (d) Write down the Lagrangian for the rolling coin. (e) Show that is given by: =2 g . a sin

(f) Show that the component of along the vertical axis z is given by z = 2 g sin3 . a

** 17. Two particles with mass m 1 and m 2 are connected by a string and a spring as in the gure. The particles can move frictionless through the horizontal tubes. The tubes have a moment of inertia I around the vertical axis. Use , r 1 and r 2 as coordinates and assume that no torque is acting on the vertical axis. Show that in that case:
2 2 (I + m 1 r 1 + m 2 r 2 ) = P = constant, .. m 1 r 1 m 1 r 1 2 = k(r 1 + r 2 c),

(A.3) (A.4) (A.5)

m 2 r 2 m 2 r 2 2 = k(r 1 + r 2 c).

..

79

{ {

r1 m1

m2

r2

T WO - BODY PROBLEM
** 18. In the two-body problem, we use polar coordinates r and . (a) What are the expressions for the angular momentum rp and for the velocity r in polar coordinates? (b) The equations of motion for a particle moving in the x y plane under the inuence of a central force f r are: m r mr 2 = f r ; p = mr 2 = constant. We recognise in p the angular momentum. Use p d d d d = = dt d t d mr 2 d

and the substitution u = 1/r to show that the equation of motion for u has the following form: d 2u m fr +u = 2 2. d 2 pu (c) Assume that the path is a conic section with r = 1 A , where is the eccentricity cos and A a constant which depends on the system and the starting conditions of the 2 motion. Show that f r = P /(m Ar 2 ). Note: whether the actual path is an ellipse, parabola or hyperbola, depends on the starting conditions; in all three cases, f r is proportional to 1/r 2 . (d) Show that for an orthogonal hyperbola with x y = C , it holds that f r = (e) Show that for a cardioid r = a(1 + cos ), we must have f r = orbit of m for this case.
2 3p a mr 4 2 P r . 4mC 2

** 19. A rocket orbits around the earth in an ellipsoidal trajectory with eccentricity and semi-major axis a. By letting the rocket engine work for some short time, the rocket is given a change in velocity v which may be considered to be instantaneous. The mass m r of the rocket may be considered to remain constant.

and sketch the

80

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS (a) At which point of the orbit and in which direction one should do this in order to let the rocket escape from the gravity eld of the earth? (b) Express |v| of the rocket in terms of a, , G and M (with M m r ). Does |v|min increase or decrease with increasing when a is kept constant?

{ {

VARIATIEREKENING
* 20. De hyperbolische functies cosh(x) en sinh(x) worden gegeven door cosh(x) (e x + e x )/2 sinh(x) (e x e x )/2 (a) Laat zien dat cosh2 (x) sinh2 (x) = 1. (b) Laat zien dat cosh(x) een oplossing is van de vergelijking y = 1 + y 2.

** 21. We proberen de vorm van een ketting die symmetrisch aan twee uiteinden is opgehangen, zie de guur, af te leiden met behulp van een krachtenbalans. In de ketting heerst een spankracht. De ketting heeft een massadichtheid (massa per lengteeenheid). Omdat het gewicht van de ketting niet verwaarloosd kan worden is deze spankracht niet constant. Wel kunnen we zeggen dat de spankracht altijd gericht is langs de raaklijn aan de ketting. We deniren de x-coordinaat zo dat x = 0 het midden van de ketting vormt, en dat de ophangpunten liggen op x = L en x = L. De vorm van de ketting wordt aangeduid met y(x). Beschouw het segment van de ketting tussen x = 0 en x = a L. Op dit segment werken drie krachten: de spankracht F s in de ketting op x = 0, de spankracht T in x = a en de zwaartekracht F zw .

a 0
(a) Laat zien dat de krachtenbalans voor dit segment leidt tot g
a

dx
0

1 + y 2 = F s y (a)

(A.6)

81 (b) Differentieer het linker- en rechterlid naar a en laat zien dat de resulterende tweede orde differentiaalvergelijking de kettinglijn A cosh(x/A) + B als oplossing heeft. Wat wordt F s ? a d Hint: d a 0 d x f (x) = f (a). (c) Geef een uitdrukking voor de lengte van de ketting. ** 22. In deze opgave gebruiken we bolcordinaten en variatierekening om de geodeet (de curve die de kortste afstand tussen twee punten geeft) op een bol te vinden. Piloten noemen dit de great circle route. (a) Leid eerst m.b.v. variatierekening de differentiaalvergelijking voor deze kortste route af. *** (b) Gebruik nu de substitutie w = cot , dus sin2 = jking en leid af dat de
1 1+w 2

{ {

in deze differentiaalvergeli-

oplossing gegeven wordt door = arccot[ A sin(0 )], met A en 0 constanten.

** 23. Bereken het pad van een lichtstraal in een medium waarin de index van refractie gegeven wordt door n(x, y) = y.

** 24. Beschouw de kegel x 2 + y 2 = z 2 . Gebruik cylindercoordinaten, en beschouw een pad van de vorm (z, (z)). (a) Laat zien dat de lengte d s van een klein lijnsegment op de kegel gegeven wordt door ds = 2d z 2 + z 2 d 2 .

(b) Schrijf de Euler-Lagrange vergelijking op en laat zien dat deze leidt tot de differentiaal vergelijking: d 2C = . d z z z 2 C 2 (c) Los deze vergelijking op door gebruik te maken van dx C x x 2 C 2 = arccos C . x

S YSTEMEN NABIJ EVENWICHT


** 25. Beschouw een potentiaal van een systeem met gegeneraliseerde coordinaten 1 en 2 . De potentiaal van dit systeem is gegeven door 1 V = m 1 g l cos 1 m 2 g l cos 2 + l 2 (sin 1 sin 2 )2 , 2 waarbij g de valversnelling is en een veerconstante; m 1 en m 2 zijn massas. Wat zijn de waarden van de gegeneraliseerde cordinaten in de evenwichtsstand? Vind de matrix K (zie dictaat). Met welk fysisch systeem correspondeert deze potentiaal?

82

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS

{ {

** 26. Beschouw een systeem bestaande uit twee bollen met massas m 1 en m 2 . De bovenste bol (m 1 ) is met een veer met veerkonstante 1 aan het plafond bevestigd en bovendien is zij met een veer (veerconstante 2 ) aan m 2 bevestigd. De onderste massa m 2 is op haar beurt met een veer (3 ) aan de vloer bevestigd. We beschouwen uitsluitend verticale bewegingen van de bollen. Vind de normal modes van het systeem (houd er rekening mee dat de veren niet in hun ruststand zijn als het systeem in evenwicht is). ** 27. Beschouw onderstaand systeem:

M, I

l
m1 m2

Vind de normal modes. ** 28. Beschouw twee slingers met gelijke lengte l . Aan beide slingers hangt een massa m. De massas zijn met elkaar verbonden door een veer met veerconstante (zie de guur hieronder).

l/ 2 l
1 2

m
(a) Geef de Lagrangiaan voor de slingers zonder veer.

(b) Idem, voor de gekoppelde slingers. Leid ook de bewegingsvergelijkingen af. (c) Stel de matrix voor de kinetische en potentile energie op en vind de normal modes. Beschijf de beweging behorend bij elk van de gevonden modes.

83 ** 29. Beschouw twee massas m die met wrijvingloos draaibare, onbuigzame en massaloze staven van gelijke lengte l zijn opgehangen aan het plafond. De slingers zijn gekopppeld door een veer met veerconstante en met een massa M , die bevestigd is aan de middens van de beide staven:

{ {

l/ 2 l
1 2

(a) Geef de Lagrangiaan voor de slingers zonder veer. Geef de Lagrangiaan ook voor kleine uitwijkingen ten opzichte van de evenwichtstoestand. (b) Druk de positie van het zwaartepunt van de veer uit in termen van 1 , 2 en l , alles voor kleine uitwijkingen. Geef de kinetische en potentiele energie van het gehele systeem in de limiet voor kleine uitwijkingen. (c) Geef de matrices voor de kinetische en potentile energie en vind (m.b.v. Maple) de normal modes. ** 30. Een dunne, homogene staaf van lengte l en massa m rust op twee veren met veerconstante die verticaal kunnen trillen.

y1

y2

Bij de volgende vragen gaan we ervan uit dat de veren dicht in de buurt van hun evenwichtsstand zijn. (a) Wat is de positie van het massamiddelpunt van de staaf, uitgedrukt in termen van de hoogten y 1 en y 2 van de veren? Idem voor de hoek die de staaf maakt met de horizontale as. (b) Schrijf de Lagrangiaan op. Gebruik: het traagheidsmoment van de staaf is
1 2 12 ml .

(c) Vind de normal modes van de staaf. Leg uit met welke bewegingen deze normal modes corresponderen. ** 31. Beschouw het onderstaande systeem, bestaande uit drie karretjes, waarvan de bovenste twee verbonden met een veer. Het linker onder karretje heeft massa m 1 en het rechter m 3 . Het karretje dat op het linker karretje staat heeft massa m 2 . Verder hebben de twee wielen van m 2 elk een traagheidsmoment I en wordt het traagheidsmoment van de overige wielen verwaarloosd.

84

A. P ROBLEMS AND E XERCISES C LASSICAL M ECHANICS

{ {

x2

I
x1

2 1

x3
Laat zien dat T= en 1 2 2 2 (m 1 + m 2 ) x 1 + m 2 + 2I /r 2 q 2 + m 3 x 3 2m 2 x 1 q 2 ; 2

1 2 2 2 V = x 1 + q 2 + x 3 2x 1 x 3 + 2x 3 q 2 2x 1 q 2 , 2 waarbij q 2 = const. x 2 en waarbij r de straal van de wielen van m 2 is. Laat zien dat er twee normal modes zijn met frequentie nul. Beschrijf de bijbehorende beweging. Vind de frequenties van de derde normal mode.

B
A NTWOORDEN

85

86

B. A NTWOORDEN 1. De massa van de ketting is m. De zwaartekracht die werkt op de los hangende lengte y g is g m(l /y). Hieruit volgt voor de versnelling m y = g m(y/l ) y = y . Een oplossing van deze differtiaalvergelijking is y(t ) = Ae t + B e t met = l . Uit de startvoorwaarde y(0) = 0 rekenen we uit dat A = B en uit y(0) = y 0 dat A = B = y 0 /2. De oplossing is y(t ) = 2.
y0 2 g

B{

e t + e t .

(a) ma 2 (b) 2ma 2 (c) (d)


1 2 g sin 1 2 tan

3. 4.

2 d 2 + d 0 d 0 , met d 0 = G M /v 2 , M de massa van de zon

(a) 9 (b) 6

5.

1 2 2 6 Ml

6. 7. (a) (b) (c) (d) 8.


1 2 2 2 ml + mg l cos 1 1 1 2 2 2 1 2 2 m 1 (r 1 + u) + 2 m 2 r 1 + m 1 g (r 1 + u) + m 2 g (l r 1 ) + 2 I (r 1 /R) 2 ku 1 M M 2 2 2 ( 3 + m)L + g (m + 2 )L sin , met de hoek tussen balk en plafond 1 4I L = 2 (M + a 2 )x 2 + 1 m(x + q)2 1 q 2 2 2

(a) 2 vrijheidsgraden. Kies u (de uitwijking van de veren uit de evenwichtsstand langs de horizontale as) en (de hoek van de slinger met de verticaal). (b) L = 1 m(u 2 + 2l cos u + l 2 2 ) + mg l cos ku 2 .
2

De bewegingsvergelijkingen worden: cos u + l + g sin = 0 2 sin + l cos ) = 2ku m(u l (c) 9. 10. 11. (a) (b) (c) l (t ) = r (t ) = 1 m 1 A cos( m1 + m2 1 m 2 A cos( m1 + m2 k t ) + m 1 B sin( k t ) + m 2 B sin( k m1 1 t)+ (g + kb) + m 2 g t 2 C t D k 2 k m2 1 t)+ (g + kb) m 2 g t 2 +C t + D k 2

met m 1 m 2 /(m 1 + m 2 ), en A, B , C en D constanten. 12. 13. 14. (a) p r = m 1 r , p = m 1 r 2 , p l = m2 l

87
2 pr 2m 1 l + 2m r 2 + 2m2 m 2 g l + 1 k(l + r b)2 2 1

(b) H = (c)

p2

p2

1 2 2 (m 1 + m 2 )r

(2r 0 d 0 = 2m1 (r d )2 r)2 + m 2 g d


0 0

J 2d

15. Nee - probeer maar een potentiaal te vinden die voldoet. Gebruik wetten van Newton. 16. (a) = (b) cos z (c) = ( cos d) (d) = sin f
1 (e) L = 2 I 1 2 sin2 M g a sin

{ {

(f) (g) 17. 18. (a) Impulsmoment mr 2 z. Snelheid x = r cos r sin , y = r sin + r cos . (b) (c) (d) (e) 19. Beschouw de totale energie voor en na de snelheidsverandering. (a) E na E voor = 1 m 2v (v) + (v)2 , dus moeten v en v in dezelfde richting gericht 2 zijn. Het verschil is maximaal in het perihelion. (b) v mi n = af. 20. 21. (a) (b) F s = g A (c) lengte = 2A sinh(L/A) 22.
4 (a) 2 = sin 2 sin2 , met de polaire hoek, d.w.z. de hoek met de z-as en C een C constante.

2G M (1 )a

1+ 2

, dus neemt v mi n bij toenemende en constante a

(b) 23. Differentiaalvergelijking y = 24. (a) Gebruik dat r 2 = z 2 . (b) (c) (z) = 2 arccos(C /z).
y A2

1 met A een constante, en oplossing y =

x2 4A 2

+ A2.

25. In evenwicht geldt: 1 = 2 = 0. K= m 1 g l + l 2 l 2 l 2 m 2 g l + l 2

Fysisch systeem: twee slingers met elk een massa eraan die verbonden zijn door een veer.

88

B. A NTWOORDEN

26. Zie voorbeeld in Hoofdstuk 6.2.1 van het diktaat. Oplossen van de eigenwaardenvergelijking (met 2 en 3 verwisseld) geeft de normal modes 2 = B B 2 4AC , 2A

B{

met A m 1 m 2 , B m 1 (2 + 3 ) + m 2 (1 + 2 ) en C (1 + 2 )(2 + 3 ) 2 . 2 (Zie voor een speciaal geval (alle s hetzelfde en alle massas hetzelfde) ook (6.33) in het diktaat). 27. Kies gegeneraliseerde coordinaten y (de hoogte van het wiel) en (de rotatiehoek). We nemen aan dat het touw niet slipt. De positie van m 1 en m 2 varieert dan met R (voor m 1 ) en -R (voor m 2 ). De Lagrangiaan wordt nu gegeven door 1 1 1 1 L = m 1 ( y + R )2 + m 2 ( y R )2 + M y 2 + I 2 + 2 2 2 2 1 2 y + termen lineair in y (doen niet mee voor ). 2 De normal modes vind je uit de eigenwaardenvergelijking (m 1 + m 2 + M )2 (m 1 m 2 )R2 2 (m 1 m 2 )R ((m 1 + m 2 )R 2 + I )2 Met A (m 1 + m 2 + M ), B (m 1 m 2 )R en C (m 1 + m 2 )R 2 + I . geeft dit: C 2 = AC B 2 28. (a) Lagrangiaan zonder veer: 1 2 2 L = ml 2 (1 + 2 ) + mg l (cos 1 + cos 2 ). 2 (b) Lagrangiaan met veer: L= ml 2 2 1 + 2 + mg l (cos 1 + cos 2 ) 2 2 l2 l l (cos 1 cos 2 )2 + sin 1 sin 2 L 0 2 4 2 2 = 0.

L0

De bewegingsvergelijkingen worden (tot op eerste orde, dus kwadratische termen in de snelheden verwaarlozend, en in de kleine-hoek benadering (cos 1, sin )): l 2 (1 2 ) = 0 4 l 2 ml 2 2 + mg l 2 (1 2 ) = 0 4 ml 2 1 + mg l 1 + (c) K= mg l + l 4 2 l 4 M=
2

l 4 2 mg l + l 4
2

ml 2 0

0 ml 2

89 Oplossen van de eigenwaardenvergelijking leidt dan tot normal modes met frequentie 2 = g /l g 2 = + l 2m De eerste beschrijft de frequentie van de gekoppelde slingers wanneer zij met gelijke fase bewegen (de slingers merken niets van de koppeling), en de tweede de frequentie van de beweging wanneer de slingers in tegenfase bewegen. 29. (a) Lagrangiaan zonder veer: 1 1 2 1 2 1 2 2 2 2 L = m 1 + 2 + mg l [cos 1 + cos 2 ] m 1 + 2 + mg l 2 1 2 2 2 2 2 (b) Het zwaartepunt van de veer is gegeven door: xm ym = = 1 a l a l (x 1 + x 2 ) = + [sin 1 + sin 2 ] + [1 + 2 ] 2 2 4 2 4 1 l l (y 1 + y 2 ) = [cos 1 + cos 2 ] 2 4 2 (B.1) (B.2)

{ {

a is de afstand tussen de ophangpunten. De kinetische energie van het gehele systeem is gegeven door T = 1 1 l2 1 2 2 2 2 2 2 ml 2 [1 + 2 ] + M [x m + y m ] ml 2 [1 + 2 ] + M 1 + 2 2 2 2 32
2

De potentiele energie is gegeven door: l V = mg l cos 1 mg l cos 2 M g (cos 1 + cos 2 ) + Vveer , 4 met 1 Vveer = 2

l2 4

(cos 1 cos 2 )2 +

l (sin 1 sin 2 ) + a 2

L (B.3)

Dus, we vinden voor de potentiele energie in de limiet voor kleine uitwijkingen (constante termen weggelaten): V = mg l 1 2 2 + 2 2 1 + 1 2 2 M g l 1 + 2 8 + 2 1 2 1 2 l 1 1 2 + 2 4 2 2

(c) De matrices voor de kinetische en potentiele energie zijn M= ml 2 + M l 16


Ml 16
2 2

Ml2 16 2 ml 2 + M l 16

K=

g l m + M + l 4 4 2 l 4

l 4 2 g l m + M + l 4 4
2

en dus zijn de normal modes (gebruik Maple) 2 =


M g (m + 4 ) , l (m + M ) 8

2 =

g m + M + l 4 2 ml

90 30.

B. A NTWOORDEN (a) Voor kleine veranderingen ten opzichte van de evenwichtstoestand, is het zwaartepunt gegeven door: l y1 + y2 xM = , yM = 2 2 en dus is de hoek : y2 y1 sin = . l (b) De kinetische energie is gegeven door: 1 1 T = M y M + M l 2 2 . 2 24 De potentiele energie is: V = M g ym + (y 1 y 0 )2 + (y 2 y 0 )2 2

B{

Druk y 1 en y 2 uit in y m en sin . Maak de kleine hoek benadering dan L = T V met 1 1 2 T = m y m + I 2 2 2 1 1 1 1 V = mg y m + (y m l y 0 )2 + (y m + l y 0 )2 2 2 2 2 (c) De matrices voor de kinetische en potentiele energie zijn M= m 0 0 , I K= 2 0 0
1 2 2 l

Het oplossen van de eigenwaarden vergelijking geeft de normal modes: 2 = 1 6 , m 2 = 2 2 . m

1 correspondeert met de rotatie- en 2 met de verticale vibratie beweging. 31. -

Potrebbero piacerti anche