Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Theoretical Physics: Second Edition
Theoretical Physics: Second Edition
Theoretical Physics: Second Edition
Ebook953 pages14 hours

Theoretical Physics: Second Edition

Rating: 4 out of 5 stars

4/5

()

Read preview

About this ebook

This authoritative volume by a renowned Russian scientist offers advanced students a thorough background in theoretical physics. The treatment's review of basic methods takes an approach that's as rigorous and systematic as it is practical.
Chiefly devoted to mechanics, electrodynamics, quantum mechanics, and statistical mechanics, this book stresses atomic, nuclear, and microscopic matters. Subjects include the quantum theories of radiation, dispersal, and scattering and the application of statistical mechanics to electromagnetic fields and crystalline bodies. Particularly strong in its coverage of statistical physics, the text examines Boltzmann statistics, Bose and Fermi distributions, Gibbs statistics, thermodynamic quantities, thermodynamic properties of ideal gases in Boltzmann statistics, fluctuations, phase equilibrium, weak solutions, chemical equilibria, and surface phenomena. Many of the 137 exercises feature complete solutions. Translated by George Yankovsky under the author's supervision.
LanguageEnglish
Release dateJun 19, 2013
ISBN9780486311227
Theoretical Physics: Second Edition

Related to Theoretical Physics

Titles in the series (100)

View More

Related ebooks

Physics For You

View More

Related articles

Reviews for Theoretical Physics

Rating: 4 out of 5 stars
4/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Theoretical Physics - A. S. Kompaneyets

    Index

    FROM THE PREFACE TO THE FIRST EDITION

    This book is intended for readers who are acquainted with the course of general physics and analysis of nonspecializing institutions of higher education. It is meant chiefly for engineer-physicists, though it may also be useful to specialists working in fields associated with physics—chemists, physical chemists, biophysicists, geophysicists, and astronomers.

    Like the natural sciences in general, physics is based primarily on experiment, and, what is more, on quantitative experiment. However, no series of experiments can constitute a theory until a rigorous logical relationship is established between them. Theory not only allows us to systematize the available experimental material, but also makes it possible to predict new facts which can be experimentally verified.

    All physical laws are expressed in the form of quantitative relationships. In order to interrelate quantitative laws, theoretical physics appeals to mathematics. The methods of theoretical physics, which are based on mathematics, can be fully mastered only by those who have acquired a very considerable volume of mathematical knowledge. Nevertheless, the basic ideas and results of theoretical physics are readily comprehensible to any reader who has an understanding of differential and integral calculus, and is acquainted with vector algebra. This is the minimum of mathematical knowledge required for an understanding of the text that follows.

    At the same time, the aim of this book is not only to give the reader an idea about what theoretical physics is, but also to furnish him with a working knowledge of the basic methods of theoretical physics. For this reason it has been necessary to adhere, as far as possible, to a rigorous exposition. The reader will more readily agree with the conclusions reached if their inevitability has been made obvious to him. In order to activize the work of the student, some of the applications of the theory have been shifted into the exercises, in which the line of reasoning is not so detailed as in the basic text.

    In compiling such a relatively small book as this one it has been necessary to cut down on the space devoted to certain important sections of theoretical physics, and omit other branches entirely. For instance, the mechanics of solid media is not included at all since to set out this branch, even in the same detail as the rest of the text, would mean doubling the size of the book. A few results from the mechanics of continuous media are included in the exercises as illustrations in thermodynamics. At the same time, the mechanics and electrodynamics of solid media are less related to the fundamental, gnosiological problems of physics than microscopic electrodynamics, quantum theory, and statistical physics. For this reason, very little space is devoted to macroscopic electrodynamics: the material has been selected in such a way as to show the reader how the transition is made from microscopic electrodynamics to the theory of quasi-stationary fields and the laws of the propagation of light in media. It is assumed that the reader is familiar with these problems from courses of physics and electricity.

    On the whole, the book is mainly intended for the reader who is interested in the physics of elementary processes. These considerations have also dictated the choice of material; as in all nonencyclopaedic manuals, this choice is inevitably somewhat subjective.

    In compiling this book, I have made considerable use of the excellent course of theoretical physics of L. D. Landau and E. M. Lifshits. This comprehensive course can be recommended to all those who wish to obtain a profound understanding of theoretical physics.

    I should like to express my deep gratitude to my friends who have made important observations: Ya. B. Zeldovich, V. G. Levich, E. L. Feinberg, V. I. Kogan and V. I. Goldansky.

    A. Kompaneyets

    PREFACE TO THE SECOND EDITION

    In this second edition I have attempted to make the presentation more systematic and rigorous without adding any difficulties. In order to do this it has been especially necessary to revise Part III, to which I have added a special section (Sec. 30) setting out the general principles of quantum mechanics; radiation is now considered only with the aid of the quantum theory of the electromagnetic field, since the results obtained from the correspondence principle do not appear sufficiently justified.

    Gibbs’ statistics are included in this edition, which has made it necessary to divide Part IV into something in the nature of two cycles: Sec. 39-44, where only the results of combinatorial analysis are set out, and Sec. 45-52, an introduction to the Gibbs’ method, which is used as background material for a discussion of thermodynamics. A phenomenological approach to thermodynamics would nowadays appear an anachronism in a course of theoretical physics.

    In order not to increase the size of the book overmuch, it has been necessary to omit the theory of beta decay, the variational properties of eigenvalues, and certain other problems included in the first edition.

    I am greatly indebted to A. F. Nikiforov and V. B. Uvarov for pointing out several inaccuracies in the first edition of the book.

    A. Kompaneyets

    PART I

    MECHANICS

    Sec. 1. Generalized Coordinates

    Frames of reference. In order to describe the motion of a mechanical system, it is necessary to specify its position in space as a function of time. Obviously, it is only meaningful to speak of the relative position of any point. For instance, the position of a flying aircraft is given relative to some coordinate system fixed with respect to the earth; the motion of a charged particle in an accelerator is given relative to the accelerator, etc. The system, relative to which the motion is described, is called a frame of reference.

    Specification of time. As will be shown later (Sec. 20), specification of time in the general case is also connected with defining the frame of reference in which it is given. The intuitive conception of a universal, unique time, to which we are accustomed in everyday life, is, to a certain extent, an approximation that is only true when the relative speeds of all material particles are small in comparison with the velocity of light. The mechanics of such slow movements is termed Newtonian, since Isaac Newton was the first to formulate its laws.

    Newton’s laws permit a determination of the position of a mechanical system at an arbitrary instant of time, if the positions and velocities of all points of the system are known at some initial instant, and also if the forces acting in the system are known.

    Degrees of freedom of a mechanical system. The number of independent parameters defining the position of a mechanical system in space is termed the number of its degrees of freedom.

    The position of a particle in space relative to other bodies is defined with the aid of three independent parameters, for example, its Cartesian coordinates. The position of a system consisting of N particles is determined, in general, by 3N independent parameters.

    However, if the distribution of points is fixed in any way, then the number of degrees of freedom may be less than 3N. For example, if two points are constrained by some form of rigid nondeformable coupling, then, upon the six Cartesian coordinates of these points, x1, y1, z1, x2, y2, z2, imposed the condition

    where R12 is the given distance between the points. It follows that the Cartesian coordinates are no longer independent parameters: a relationship exists between them. Only five of the six values x1, ..., z2 are now independent. In other words, a system of two particles, separated by a fixed distance, has five degrees of freedom. If we consider three particles which are rigidly fixed in a triangle, then the coordinates of the third particle must satisfy the two equations:

    Thus, the nine coordinates of the vertices of the rigid triangle are defined by the three equations (1.1), (1.2) and (1.3), and hence only six of the nine quantities are independent. The triangle has six degrees of freedom.

    The position of a rigid body in space is defined by three points which do not lie on the same straight line. These three points, as we have just seen, have six degrees of freedom. It follows that any rigid body has six degrees of freedom. It should be noted that only such motions of the rigid body are considered as, for example, the rotation of a top, where no noticeable deformation occurs that can affect its motion.

    Generalized coordinates. It is not always convenient to describe the position of a system in Cartesian coordinates. As we have already seen, when rigid constraints exist, Cartesian coordinates must satisfy supplementary equations. In addition, the choice of coordinate system is arbitrary and should be determined primarily on the basis of expediency. For instance, if the forces depend only on the distances between particles, it is reasonable to introduce these distances into dynamical equations explicitly and not by means of Cartesian coordinates.

    In other words, a mechanical system can be described by coordinates whose number is equal to the number of degrees of freedom of the system. These coordinates may sometimes coincide with the Cartesian coordinates of some of the particles. For example, in a system of two rigidly connected points, these coordinates can be chosen in the following way: the position of one of the points is given in Cartesian coordinates, after which the other point will always be situated on a sphere whose centre is the first point. The position of the second point on the sphere may be given by its longitude and latitude. Together with the three Cartesian coordinates of the first point, the latitude and longitude of the second point completely define the position of such a system in space.

    For three rigidly bound points, it is necessary, in accordance with the method just described, to specify the position of one side of the triangle and the angle of rotation of the third vertex about that side.

    The independent parameters which define the position of a mechanical system in space are called its generalized coordinates. We will represent them by the symbols , where the subscript α signifies the number of the degree of freedom.

    As in the case of Cartesian coordinates, the choice of generalized coordinates is to a considerable extent arbitrary. It must be chosen so that the dynamical laws of motion of the system can be formulated as conveniently as possible.

    Sec. 2. Lagrange’s Equation

    In this section, equations of motion will be obtained in terms of arbitrary generalized coordinates. In such form they are especially convenient in theoretical physics.

    Newton’s Second Law. Motion in mechanics consists in changes in the mutual configuration of bodies in time. In other words, it is described in terms of the mutual distances, or lengths, and intervals of time. As was shown in the preceding section, all motion is relative; it can be specified only in relation to some definite frame of reference.

    In accordance with the level of knowledge of his time, Newton regarded the concepts of length and time interval as absolute, which is to say that these quantities are the same in all frames of reference. As will be shown later, Newton’s assumption was an approximation (see Sec. 20). It holds when the relative speeds of all the particles are small compared with the velocity of light; here Newtonian mechanics is based on a vast quantity of experimental facts.

    In formulating the laws of motion a very convenient concept is the material particle, that is, a body whose position is completely defined by three Cartesian coordinates. Strictly speaking, this idealization is not applicable to any body. Nevertheless, it is in every way reasonable when the motion of a body is sufficiently well defined by the displacement in space of any of its particles (for example, the centre of gravity of the body) and is independent of rotations or deformations of the body.

    If we start with the concept of a particle as the fundamental entity of mechanics, then the law of motion (Newton’s Second Law) is formulated thus:

    Here, F is the vector acceleration, the Cartesian components of which are

    The quantity m involved in equation (2.1) characterizes the particle and is called its mass.

    Force and mass. Equality (2.1) is the definition of force. However, it should not be regarded as a simple identity or designation, because (2.1) establishes the form of the interaction between bodies in mechanics and thereby actually describes a certain law of nature. The interaction is expressed in the form of a differential equation that includes only the second derivatives of the coordinates with respect to time (and not derivatives, say, of the fourth order).

    In addition, certain limiting assumptions are usually made in relation to the force. In Newtonian mechanics it is assumed that forces depend only on the mutual arrangement of the bodies at the instant to which the equality refers and do not depend on the configuration of the bodies at previous times. As we shall see later (see Part II), this supposition about the character of interaction forces is valid only when the speeds of the bodies are small compared with the velocity of light.

    The quantity m in equality (2.1) is a characteristic of the body, its mass. Mass may be determined by comparing the accelerations which the same force imparts to different bodies; the greater the acceleration, the less the mass. In order to measure mass, some body must be regarded as a standard. The choice of a standard body is completely independent of the choice of standards of length and time. This is what makes the dimension (or unit of measurement) of mass a special dimension, not related to the dimensions of length and time.

    The properties of mass are established experimentally. Firstly, it can be shown that the mass of two equal quantities of the same substance is equal to twice the mass of each quantity. For example, one can take two identical scale weights and note that a stretched spring gives them equal accelerations. If we join two such weights and subject them to the action of the same spring, which has been stretched by the same amount as for each weight separately, the acceleration will be found to be one half what it was. It follows that the overall mass of the weights is twice as great, since the force depends only on the tension of the spring and could not have changed. Thus, mass is an additive quantity, that is, one in which the whole is equal to the sum of the quantities of each part taken separately. Experiment shows that the principle of additivity of mass also applies to bodies consisting of different substances.

    In addition, in Newtonian mechanics, the mass of a body is a constant quantity which does not change with motion.

    It must not be forgotten that the additivity and constancy of masses are properties that follow only from experimental facts which relate to very specific forms of motion. For example, a very important law, that of the conservation of mass in chemical transformations involving rearrangement of the molecules and atoms of a body, was established by M. V. Lomonosov experimentally.

    Like all laws deduced from experiment, the principle of additivity of mass has a definite degree of precision. For such strong interactions as take place in the atomic nucleus, the breakdown of the additivity of mass is apparent (for more detail see Sec. 21).

    We may note that if instead of subjecting a body to the force of a stretched spring it were subjected to the action of gravity, then the acceleration of a body of double mass would be equal to the acceleration of each body separately. From this we conclude that the force of gravity is itself proportional to the mass of a body. Hence, in a vacuum, in the absence of air resistance, all bodies fall with the same acceleration.

    Inertial frames of reference. In equation (2.1) we have to do with the acceleration of a particle. There is no sense in talking about acceleration without stating to which frame of reference it is referred. For this reason there arises a difficulty in stating the cause of the acceleration. This cause may be either interaction between bodies or it may be due to some distinctive properties of the reference frame itself. For example, the jolt which a passenger experiences when a carriage suddenly stops is evidence that the carriage is in nonuniform motion relative to the earth.

    Let us consider a set of bodies not affected by any other bodies, that is, one that is sufficiently far away from them. We can suppose that a frame of reference exists such that all accelerations of the set of bodies considered arise only as a result of the interaction between the bodies. This can be verified if the forces satisfy Newton’s Third Law, i.e., if they are equal and opposite in sign for any pair of particles (it is assumed that the forces occur instantaneously, and this is true only when the speeds of the particles are small compared with the speed of transmission of the interaction).

    A frame of reference for which the acceleration of a certain set of particles depends only on the interaction between these particles is called an inertial frame (or inertial coordinate system). A free particle, not subject to the action of any other body, moves, relative to such a reference frame, uniformly in a straight line or, in everyday language, by its own momentum. If in a given frame of reference Newton’s Third Law is not satisfied we can conclude that this is not an inertial system.

    Thus, a stone thrown directly downwards from a tall tower is deflected towards the east from the direction of the force of gravity. This direction can be independently established with the aid of a suspended weight. It follows that the stone has a component of acceleration which is not caused by the force of the earth’s attraction. From this we conclude that the frame of reference fixed in the earth is noninertial. The noninertiality is, in this case, due to the diurnal rotation of the earth.

    On the forces of friction. In everyday life we constantly observe the action of forces that arise from direct contacts between bodies. The sliding and rolling of rigid bodies give rise to forces of friction. The action of these forces causes a transition of the macroscopic motion of the body as a whole into the microscopic motion of the constituent atoms and molecules. This is perceived as the generation of heat. Actually, when a body slides an extraordinarily complex process of interaction occurs between the atoms in the surface layer. A description of this interaction in the simple terms of frictional forces is a very convenient idealization for the mechanics of macroscopic motion, but, naturally, does not give us a full picture of the process. The concept of frictional force arises as a result of a certain averaging of all the elementary interactions which occur between bodies in contact.

    In this part, which is concerned only with elementary laws, we shall not consider averaged interactions where motion is transferred to the internal, microscopic, degrees of freedom of atoms and molecules. Here, we will study only those interactions which can be completely expressed with the aid of elementary laws of mechanics and which do not require an appeal to any statistical concepts connected with internal, thermal, motion.

    Ideal rigid constraints. Bodies in contact also give rise to forces of infraction which can be reduced to the kinematic properties of rigid constraints. If rigid constraints act in a system they force the particles to move on definite surfaces. Thus, in Sec. 1 we considered the motion of a single particle on a sphere, at the centre of which was another particle.

    This kind of interaction between particles does not cause a transition of the motion to the internal, microscopic, degrees of freedom of bodies. In other words, motion which is limited by rigid constraints is completely described by its own macroscopic generalized coordinates .

    If the limitations imposed by the constraints distort the motion, they thereby cause accelerations (curvilinear motion is always accelerated motion since velocity is a vector quantity). This acceleration can be formally attributed to forces which are called reaction forces of rigid constraints.

    Reaction forces change only the direction of velocity of a particle but not its magnitude. If they were to alter the magnitude of the velocity, this would produce a change also in the kinetic energy of the particle. According to the law of conservation of energy, heat would then be generated. But this was excluded from consideration from the very start.

    To summarize, the reaction forces of ideally rigid constraints do not change the kinetic energy of a system. In other words, they do not perform any work on it, since work performed on a system is equivalent to changing its kinetic energy (if heat is not generated).

    In order that a force should not perform work, it must be perpendicular to the displacement. For this reason the reaction forces of constraints are perpendicular to the direction of particle velocity at each given instant of time.

    However, in problems of mechanics, the reaction forces are not initially given, as are the functions of particle position. They are determined by integrating equations (2.1), with account taken of constraint conditions. Therefore, it is best to formulate the equations of mechanics so as to exclude constraint reactions entirely. It turns out that if we go over to generalized coordinates, the number of which is equal to the number of degrees of freedom of the system, then the constraint reactions disappear from the equations. In this section we shall make such a transition and will obtain the equations of mechanics in terms of the generalized coordinates of the system.

    The transformation from rectangular to generalized coordinates. We take a system with a total of 3N n Cartesian coordinates of which ν are independent. We will always denote Cartesian coordinates by the same letter xi, understanding by this symbol all the coordinates x, y, z; this means that i varies from 1 to 3N, that is, from 1 to n. Since the generalized coordinates completely specify the position of their system, xi are their unique functions:

    From this it is easy to obtain an expression for the Cartesian components of velocity. Differentiating the function of many variables xt (... ) with respect to time, we have

    In the subsequent derivation we shall often have to perform summations with respect to all the generalized coordinates , and double and triple sums will be encountered. In order to save space we introduce the following summation convention.

    If a Greek symbol is met twice on one side of an equation, it will be understood as denoting a summation from unity to ν, that is, over all the generalized coordinatescan be rewritten thus:

    Here the summation sign is omitted.

    The total derivative with respect to time is usually denoted by a dot over the corresponding variable:

    In this notation, (2.3) is written in an even more abbreviated form:

    Differentiating (2.5) with respect to time again, we obtain an expression for the Cartesian components of acceleration:

    The total derivative in the first term is written as usual:

    The Greek symbol over which the summation is performed is denoted by the letter β to avoid confusion with the symbol α, which denotes the summation in the expression for velocity :

    The first term on the right-hand side contains a double summation with respect to α and β.

    Potential of a force. We now consider components of force. In many cases, the three components of the vector of a force acting on a particle can be expressed in terms of one scalar function U according to the formula:

    Such a function can always be chosen for the force of Newtonian attraction, and for electrostatic and elastic forces. The function U is called the potential of the force.

    It is clear that by far not every system of forces can, in the general case, be represented by a set of partial derivatives (2.7), since if

    then we must have the equality

    for all i, k, which is not, in advance, obvious for the arbitrary functions Fi, Fk. The definite form of the potential, in those cases where it exists, will be given below for various forces.

    Expression (2.7) defines the potential function U to the accuracy of an arbitrary constant term. U is also called the potential energy of the system. For example, the gravitational force F = —mg, while the potential energy of an elevated body is equal to mgz, where g~980 cm/sec² is the acceleration of a freely falling body and z is the height to which it has been raised. It can be calculated from any level, which in the given case corresponds to a determination of U to the accuracy of a constant term. A more precise expression for the force of gravity than F = —mg (with allowance made for its dependence on height also admits of a potential, which we shall derive a little later [see (3.4)].

    We denote the component reaction forces of rigid constraints by Fi. We now note that

    if the displacements are compatible with the constraints. Indeed, (2.8) expresses precisely the work performed by the reaction forces for a certain possible displacement of the system; but this work has been shown to be equal to zero.

    Lagrange’s equations.* We will now write down the equations of motion with the aid of (2.7) and (2.8) as

    Here, of course, m1 = m2 = m3 is equal to the mass of the first particle, m4 = m5 = mand sum from 1 to n over i.

    Let us first consider the right-hand side. We obviously have

    in accordance with the law for differentiating composite functions. For the forces of reaction we obtain

    since this equality is a special case of (2.8), in which the displacements dxi are taken for all constants q except for q. It is clear that in such a special displacement the work done by the reaction forces of the constraints is equal to zero, as in the case of a general displacement.

    and summation, the left-hand side of equality (2.9) can be written in a more compact form, without resorting to explicit Cartesian coordinates. It is precisely the purpose of this section to give such an improved notation. To do this we express the kinetic energy in terms of generalized coordinates:

    Substituting the generalized velocities by using (2.5), we obtain

    The summation indices for q must, of course, be denoted by different letters, since they independently take all values from 1 to ν inclusive. Changing the order of summation for Cartesian and generalized coordinates we have

    Henceforth, T will have to be differentiated both with respect to generalized coordinates . The coordinate and its corresponding generalized velocity are independent of each other since, in the given position in which the coordinate has a given value permitted by the constraints. Naturally, and all the coordinates, including , should be regarded as constant.

    . In the double summation (2.13), the quantity γ can be taken as the index α and also the index β, so that we obtain

    Both these sums are the same except that in the first the index is denoted by β and in the second by α. They can be combined, replacing β by α in the first summation; naturally the value of the sum does not change due to renaming of the summation sign. Then we obtain

    Let us calculate the total derivative of this quantity with respect to time:

    Here we have had to write down the derivatives of each of the three factors of all the terms in the summation (2.14) separately.

    consists of two terms which may be amalgamated into one. Differentiating (2.13), we obtain

    Subtracting (2.16) from (2.15), we see that (2.16) and the last term of (2.15) cancel. As a result we obtain

    However, the expression on the right-hand side of and sum over i. Thus we find

    In mechanics it is usual to consider interaction forces that are independent of particle velocities. In this case U , so that (2.18) may be rewritten in the following form:

    The difference between the kinetic and potential energy is called the Lagrangian function (or, simply, Lagrangian) and is denoted by the letter L:

    Thus we have arrived at a system of ν equations with ν independent quantities , the number of which is equal to the number of degrees of freedom of the system:

    These equations are called Lagrange’s equations. Naturally, in (2.21) L is considered to be expressed solely in terms of , the Cartesian coordinates being excluded. It turns out that this type of equation holds also in cases when the forces depend on the velocities (see Sec. 21).*

    The rules for forming Lagrange’s equations. Since the derivation of equations (2.21) from Newton’s Second Law is not readily evident we will give the order of operations which, for this given system, lead to the Lagrange equations.

    1) The Cartesian coordinates are expressed in terms of generalized coordinates:

    2) The Cartesian velocity components are expressed in terms of generalized velocities:

    3) The coordinates are substituted in the expression for potential energy so that it is defined in relation to generalized coordinates:

    4) The velocities are substituted in the expression for kinetic energy

    which is now a function of q. It is essential that in generalized coordinates, T is a function both of q.

    are found.

    6) Lagrange’s equations (2.21) are formed according to the number of degrees of freedom.

    In the next section we will consider some examples in forming Lagrange’s equations.

    Exercises

    1) Write down Lagrange’s equation, where the Lagrangian function has the form:

    2) A point moves in a vertical plane along a given curve in a gravitational field. The equation of the curve in parametric form is x = x (s), z = z (s). Write down Lagrange’s equations.

    The velocities are

    The Lagrangian has the form:

    Lagrange’s equation is

    Sec. 3. Examples of Lagrange’s Equations

    Central forces. Central forces is the name given to those whose directions are along the lines joining the particles and which depend only on the distances between them. Corresponding to such forces, there is always a potential energy, U, dependent on these distances. As an example, we consider the motion of a particle relative to a fixed centre and attracting it according to Newton’s law. We shall show how to find the potential energy in this case by proceeding from the expression for a gravitational force.

    Gravitational force is known to be inversely proportional to the square of the distance between the particles and is directed along the line joining them:

    Here a is the factor of proportionality which we will not define more precisely at this point, r is a unit vector. The minus sign signifies that the particles attract each other, so that the force is in the opposite direction to the radius vector r, According to (3.1), the attractive-force component along x is equal to

    since x is a component of rso that

    and similarly for the two other component forces. Comparing (3.3) and (2.7), we see that in the given case

    We note that the potential energy U is chosen here in such a way that U (∞) = 0 when the particles are separated by an infinite distance. The choice of the arbitrary constant in the potential energy is called its gauge. In this case it is convenient to choose this constant so that the potential energy tends to zero at infinity.

    It is obvious that an expression similar to (3.4) is obtained for two electrically charged particles interacting in accordance with Coulomb’s law.

    Spherical coordinates. Formula (3.4) suggests that in this instance it is best to choose precisely r as a generalized coordinate. In other words, we must transform from Cartesian to spherical coordinates. The relationship between Cartesian and spherical coordinates is shown in Fig. 1. The z-axis is called the polar axis of the spherical coordinate system. The angle ϑ between the radius vector and the polar axis is called the polar angle; it is complementary (to 90°) to the latitude. Finally, the angle φ is analogous to the longitude and is called the azimuth. It measures the dihedral angle between the plane zOx and the plane passing through the polar axis and the given point.

    Let us find the formulae for the transformation from Cartesian to spherical coordinates. From Fig. 1 it is clear that

    The projection ρ of the radius vector onto the plane xOy is

    Whence,

    We will now find an expression for the kinetic energy in spherical coordinates. This can be done either by a simple geometrical construction or by calculation according to the method of Sec. 2.

    Fig. 1

    Fig. 2

    Although the construction is simpler, let us first follow the computation procedure in order to illustrate the general method. We have:

    Squaring these equations and adding, we obtain, after very simple manipulations, the following:

    The same is clear from the construction shown in Fig. 2. An arbitrary displacement of the point can be resolved into three mutually perpendicular displacements: dr, rdϑ and ρdφ = r sin ϑdφ. Whence

    , (3.9) is obtained from (3.10) simply by dividing by (dt.

    Hence, in spherical coordinates, the Lagrangian function is expressed as

    Now in order to write down Lagrange’s equations it is sufficient to calculate the partial derivatives. We have:

    These derivatives must be substituted into (2.21), which, however, we will not now do since the motion we are considering actually reduces to the plane case (see beginning of Sec. 5).

    Two-particle system. So far we have considered the centre of attraction as stationary, which corresponds to the assumption of an infinitely large mass. In the motion of the earth around the sun, or of an electron in a nuclear field, the mass of the centre of attraction is indeed large compared with the mass of the attracted particle. But it may happen that both masses are similar or equal to each other (a binary star, a neutron-proton system, and the like). We shall show that the problem of the motion of two masses interacting only with one another can always be easily reduced to a problem of the motion of a single mass.

    Let the mass of the first particle be m1 and of the second m2. We call the radius vectors of these particles, drawn from an arbitrary origin, r1 and r2, respectively. The components of r1 are x1, y1, z1; the components of r2 are x2, y2, z2. We now define the radius vector of the centre of mass of these particles R by the following formula:

    Synonymous terms for the centre of mass are the centre of gravity and the centre of inertia.

    In addition, let us introduce the radius vector of the relative position of the particles

    . From (3.12) and (3.13) we have

    The kinetic energy is equal to

    Differentiating (3.14) and (3.15) with respect to time and substituting in (3.16), we obtain, after a simple rearrangement,

    If we introduce the Cartesian components of the vectors R(X, Y, Z) and r(x, y, z), then we obtain an expression for the kinetic energy in terms of Cartesian components of velocity.

    Since no external forces act on the particles, the potential energy can be a function only of their relative positions: U = U (x, y, z). Thus, the Lagrangian is

    Transition to the centre-of-mass system. Let us write down Lagrange’s equations for the coordinates of the centre of mass. We have

    Hence, in accordance with (2.21)

    These equations can be easily integrated:

    where the letters with the index 0 signify the corresponding values at the initial time.

    Combining the coordinate equations into one vector equation, we obtain

    Thus, the centre of mass moves uniformly in a straight line quite independently of the relative motion of the particles.

    Reduced mass. If we now write down Lagrange’s equations for relative motion in accordance with (2.21) the coordinates of the centre of mass do not appear. It follows that the relative motion occurs as if it were in accordance with the Lagrangian

    ), formed in exactly the same way as the Lagrangian for a single mass m equal to

    This mass is called the reduced mass.

    The motion of the centre of mass does not affect the relative motion of the masses. In particular, we can consider, simply, that the centre of mass lies at the coordinate origin R = 0.

    In the case of central forces (for example, Newtonian forces of attraction) acting between the particles, the potential energy is simply equal to U(r) [this is taken into consideration in . Then, if we describe the relative motion in spherical coordinates, the equations of motion will have the same form as for a single particle moving relative to a fixed centre of attraction.

    The centre of mass can now be considered as fixed, assuming R = 0. From this, in accordance with (3.14) and (3.15), we obtain the distance of both masses from the centre of mass:

    , i.e., the centre of mass is close to the larger mass. This is the case for a sun-planet system. At the same time the reduced mass can also be written thus:

    From here it can be seen that it is close to the smaller mass. That is why the motion of the earth around the sun can be approximately described as if the sun were stationary and the earth revolved about it with its own value of mass, independent of the mass of the sun.

    Simple and compound pendulums. In concluding this section we shall derive the Lagrangian for simple and compound pendulums.

    The simple plane pendulum is a mass suspended on a flat hinge at a certain point of a weightless rod of length l. The hinge restricts the swing of the pendulum. Let us assume that swinging occurs in the plane of the paper (Fig. 3). It is clear that such a pendulum has only one degree of freedom. We can take the angle of deflection of the pendulum from the vertical φ as a generalized coordinate. Obviously the velocity of particle m , so that the kinetic energy is

    The potential energy is determined by the height of the mass above the mean position z = l (1 — cos φ). Whence, the Lagrangian for the pendulum is

    Fig. 3

    A double pendulum can be described in the following way: in mass m there is another hinge from which another pendulum, which is forced to oscillate in the same plane (Fig. 4), is suspended. Let the mass and length of the second pendulum be m1 and l1, respectively, and its angle of deflection from the vertical, ψ. The coordinates of the second particle are

    Whence we obtain its velocity components:

    Fig. 4

    :

    The potential energy of the second particle is determined in terms of z1. Finally, we get an expression of the Lagrangian for a double pendulum in the following form:

    All the formulae for the Lagrangian functions (3.11), (3.22), and (3.23) will be required in the sections that follow.

    Exercises

    , where l0 is the equivalent length of the unstretched rod and k is a constant, characteristic of its elasticity.

    2) Write down the kinetic energy for a system of three particles with masses m1, m2, and m3 in the form of a sum of the kinetic energy of the centre of mass and the kinetic energy of relative motion, using the following relative coordinates:

    Sec. 4. Conservation Laws

    The problem of mechanics. If a mechanical system has ν degrees of freedom, then its motion is described by ν [see (2.17)]. From general theorems of analysis we conclude that after integration of this system we obtain 2 ν arbitrary constants. The solution can be represented in the following form:

    Differentiating these equations with respect to time, we obtain expressions for the velocities:

    Let us assume that equations (4.1) and (4.2) are solved with respect to the constants C1, . . ., C2ν, so that these values are expressed in terms of t .

    Then

    From the equations (4.3) we see that in any mechanical system described by 2ν second-order equations there must be 2 ν functions of generalized coordinates, velocities, and time, which remain constant in the motion. These functions are called integrals of motion.

    It is the main aim of mechanics to determine the integrals of motion.

    If the form of the function (4.3) is known for a given mechanical system, then its numerical value can be determined from the initial conditions, that is, according to the given values of generalized coordinates and velocities at the initial instant.

    In the preceding section we obtained the so-called centre-of-mass integrals R(3.18).

    Naturally, Lagrange’s equations cannot be integrated in general form for an arbitrary mechanical system. Therefore the problem of determining the integrals of motion is usually very complicated. But there are certain important integrals of motion which are given directly by the form of the Lagrangian. We shall consider these integrals in the present section.

    Energy. The quantity

    *

    is called the total energy of a system. Let us calculate its total derivative with respect to time.

    We have

    The last three terms on the right-hand side are the derivatives of the Lagrangian L, which, in the general form, depend on qand t. The result is, therefore,

    , the energy is an integral of motion. Let us find the conditions for which time does not appear explicitly in the Lagrangian.

    If the formulae expressing the generalized coordinates q in terms of Cartesian coordinates x do not contain time explicitly (which corresponds to constant, time-independent constraints) then the transformation from x to q cannot introduce time into the Lagrangian.

    , the external forces must also be independent of time. When these two conditions—constant constraints and constant external forces—are fulfilled, the energy is an integral of motion. To take a particular case, when no external forces act on the system its energy is conserved. Such a system is called closed.

    When frictional forces act inside a closed system, the energy of macroscopic motion is transformed into the energy of molecular microscopic motion. The total energy is conserved in this case, too, though the Lagrangian, which involves only the generalized coordinates of macroscopic motion of the system, no longer gives a complete description of the motion of the system. The mechanical energy of only macroscopic motion, determined by means of such a Lagrangian, is not an integral. We will not consider such a system in this section.

    Let us now consider the case when our definition of energy . Let the kinetic energy be a homogeneous quadratic function of generalized velocities, as expressed in equation (2.13). For this it is necessary that the constraints should not involve time explicitly, otherwise equation (2.5) would have the form

    where the partial derivative of the function (2.2) with respect to time is taken for all constants in the first degree would appear in the expression for T.

    Since we assume that the potential energy U does not depend on velocity [see (2.18) and (2.19)], then

    and the energy is

    But according to Euler’s theorem, the sum of partial derivatives of a homogeneous quadratic function, multiplied by the corresponding variables, is equal to twice the value of the function (this can easily be verified from the function of two variables ax²+2 bxy+cy²). Thus,

    that is, the total energy is equal to the sum of the potential energy and the kinetic energy, in agreement with the elementary definition.

    We note that the definition (4.4) is more useful and general also in the case when the Lagrangian is not represented as the difference L = TU. Thus, in electrodynamics (Sec. 15) L (if, of course, there are no frictional forces).

    The application of the energy integral to systems with one degree of freedom. The energy integral allows us, straightway, to reduce problems of the motion of systems with one degree of freedom to those of quadrature. Thus, in the pendulum problem considered in the previous section we can, with the aid of (4.7), write down the energy integral directly:

    . Whence

    Substituting this in (4.8), we have

    From this, the relationship between the deflection angle and time is determined by the quadrature

    It is essential that the law governing the oscillation of a pendulum depend only on the value of the ratio l/g and is independent of the mass. The integral in (4.11) cannot be evaluated in terms of elementary functions.

    A system in which mechanical energy is conserved is sometimes called a conservative system. Thus, the energy integral permits reducing to quadrature the problem of the motion of a conservative system with one degree of freedom.

    In a system with several degrees of freedom the energy integral allows us to reduce the order of the system of differential equations and, in this way, to simplify the problem of integration.

    Generalized momentum. We shall now consider other integrals of motion which can be found directly with the aid of the Lagrangian. To do this we shall take advantage of the following, quite obvious, consequence of Lagrange’s equations. If some coordinate , then in accordance with Lagrange’s equations

    But then

    i.e., it is an integral of motion. The quantity is called the generalized momentum . Summarizing, if a certain generalized coordinate does not appear explicitly in the Lagrangian, the generalized momentum corresponding to it is an integral of motion, i.e., it remains constant for the motion.

    In the preceding section we saw that the coordinates X, Y, Z of the centre of gravity of a system of two particles, not subject to the action of external forces, do not appear in the Lagrangian. From this it is evident that

    are constants of motion.

    The momentum of a system of particles. The same thing is readily shown also for a system of N particles. Indeed, for N particles we can introduce the concept of the centre of mass and the velocity of the centre of mass by means of the equations:

    The velocity of the ith mass relative to the centre of mass is

    (by the theorem of the addition of velocities). The kinetic energy of the system of particles is

    However, from and R. Therefore, the kinetic energy of a system of particles can be divided into a sum of two terms: the kinetic energy of motion of the centre of mass

    and the kinetic energy of motion of the mass relative to the centre of mass

    . Consequently, they can be expressed in terms of an N-1 independent quantity by determining the relative positions of the ith and first masses. For this reason the kinetic energy of N . By definition, no external forces can act on the masses in a closed system, and the interaction forces inside the system can be determined only by the relative positions r1—ri.

    appears in the Lagrangian, and R does not. Therefore, the overall momentum is conserved:

    Equality (4.18), which contains a derivative with respect to a vector, should be understood as an abbreviated form of three equations:

    For more detail about vector derivatives see Sec. 11.

    We have seen that the overall momentum of a mechanical system not subject to any external force is an integral of motion. It is important that it is what is known as an additive integral of motion, i.e., it is obtained by adding the momenta of separate particles.

    It may be noted that the momentum integral exists for any system in which only internal forces are operative, even though they may be frictional forces causing a conversion of mechanical energy into heat.

    If we integrate (4.18) with respect to time once again, the result will be the centre-of-mass integral similar to (3.18). This will be the so-called second integral (for it contains two constants); it contains only coordinates but not velocities. (3.18) and (4.11) are also second integrals.

    Properties of the vector product. The angular momentum of a particle is defined as

    Here the brackets denote the vector product of the radius vector of the particle and its momentum. We know that (4.19) takes the place of three equations,

    for the Cartesian components of the vector M.

    Recall the geometrical definition for a vector product. We construct a parallelogram on the vectors r and p. Then [rp] denotes a vector numerically equal to the area of the parallelogram with direction perpendicular to its plane. In order to specify the direction of [rp] uniquely, we must agree on the way of tracing the parallelogram contour. We shall agree always to traverse the contour beginning with the first factor (in this case beginning with r). Then that side of the plane will be considered positive for which the direction is anticlockwise. The vector [rp] is along the normal to the positive side of the plane. In still another way, if we rotate a corkscrew from r to p, then it will be displaced in the direction of [rp]. The direction of traverse changes if we interchange the positions of r and p. Therefore, unlike a conventional product, the sign of the vector product changes if we interchange the factors. This can also be seen from the definition of Cartesian components of angular momentum.

    The area of the parallelogram is rp sin α, where α is the angle between r and p. The product r sin α is the length of a perpendicular drawn from the origin of the coordinate system to the tangent to the trajectory whose direction is the same as p. This length is sometimes called the arm of the moment.

    The vector product possesses a distributive property, i.e.,

    Hence, a binomial product is calculated in the usual way, but the order of the factors is taken into account.

    The angular momentum of a particle system. The angular momentum of a system of particles is defined as the sum of the angular momenta of all the particles taken separately. In doing so we must, of course, take the radius vectors related to a coordinate origin common to all the particles:

    We shall show that the angular momentum of a system can be separated into the angular momentum relative to the motion of the particles and the angular momentum of the system as a whole, similar to the way that it was done for the kinetic energy. To do this we must represent the radius vector of each particle as the sum of the radius vector of its position relative to the centre of mass and the radius vector of the centre of mass; we must expand the expression for the particle velocities in the same way. Then, the angular momentum can be written in the form

    In the second and third sums, we can make use of the distributive property of a vector product and introduce the summation sign inside the product sign. However, both these sums are equal to zero, by definition of the centre of mass. This was used in (4.17) for velocities. Thus, the angular momentum is indeed equal to the sum of the angular momenta of the centre of mass (M0) and the relative motion (M′):

    Let us perform these transformations for the special case of a system of two particles. We substitute r1 and r2 expressed [from (3.14) and (3.15)] in terms of r and R. This gives

    Further, we replace p1 by m, p2 by mand p1 + p1 by P, after which the angular momentum reduces to the required form:

    is the momentum of relative motion.

    .

    Accordingly, angular momentum for relative motion will be

    because

    Thus, the determination of angular momentum for relative motion does not depend on the choice of the origin of the coordinate system.

    Conservation of angular momentum. We shall now show that the angular momentum of a system of particles not acted upon by any external forces is an integral of motion.

    Let us begin with the angular momentum of the system as a whole. Its time derivative is equal to zero:

    is in the direction of P.

    We shall now prove that angular momentum is conserved for relative motion. The total derivative with respect to time is

    . We consider the second term. Let us choose the origin to be coincident with any particle, for example, the first. As a result, M′, as we have already seen, does not change. The potential energy can only depend on the differences r1—r2, r1—r3, ..., r1—rk, . . . The other differences are expressed in terms of these, for example, rkrt = (r1—rt)—(r1—rk). We introduce the abbreviations.

    will be expressed in terms of the variables ρ1 ..., ρk – 1, . . . as follows:

    Substituting this in (4.23) we obtain

    In this expression, only the relative coordinates ρ1, ...,ρN , totally irrespective of whether the initial expression was a function only of the absolute values |r1— rk|, or whether it also involved scalar products of the form (rirk, rtrn). An essential point is that the system is closed (in accordance with the definition, see page 32), and the forces in it are completely

    Enjoying the preview?
    Page 1 of 1