To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure no-reply@cambridge.org
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
Nature is found to conspire in just such a way that the time integral of the Lagrangian is smallest if the motion obeys Newton's Laws. Mechanics can be based on the single principle: Minimize the time integral of the Lagrangian. Three laws of motion can be condensed into one universal principle!
The mathematical language needed to provide the framework for this is called variational calculus. The variational calculus can be used as a powerful tool in solving mechanics problems with explicit constraints. It is also the most general means of solving nonholonomic problems with constraints on the velocities such as for rolling motion. This type of problem cannot be solved by choosing coordinates equal to the number of degrees of freedom but must be embedded in a higher-dimensional space.
The well-known theoretical physicist E. P. Wigner refers to the “unreasonable effectiveness of mathematics in theoretical physics.” Mathematical beauty is and should be the chief guiding principle of theorists, according to P. A. M. Dirac, one of the inventors of quantum mechanics. Although it is hard to define exactly what mathematical beauty is, the search for beauty was the guiding principle in the invention of two major advances in physics in the twentieth century: relativistic quantum mechanics and general relativity. In this chapter, we will discover an elegant formulation of classical mechanics. The mathematical techniques uncovered here are not only beautiful, but they have become the language of modern theoretical physics. […]
Joseph Louis Lagrange reformulated Newton's Laws in a way that eliminates the need to calculate forces on isolated parts of a mechanical system. Any convenient variables obeying the constraints on a system can be used to describe the motion. If Lagrangian mechanics rather than Newtonian mechanics is used, it is only necessary to consider a single function of the dynamical variables that describe the motion of the entire system. The differential equations governing the motion are obtained directly from this function without any vector force diagrams. Lagrangian mechanics is extremely efficient: There are only as many equations to solve as there are physically significant variables.
Lagrange did not introduce new physical principles to mechanics. The physical concepts are due to Newton and Galileo. But he succeeded in giving a more powerful and sophisticated way to formulate the mathematical equations of classical mechanics, an approach that has spread its influence over physics far beyond the purely mechanical problems.
We will begin by solving some examples that lead us toward this new formulation of mechanics. We plan to use the concept of virtual work to derive this. We will consider extended rigid bodies to be made up of collections of massive point particles. Summing over the constituent particles will lead to an efficient and general method for obtaining the differential equations of motion for any frictionless mechanical system. […]
By
David N. Schramm, Department of Astronomy and Astrophysics, University of Chicago, Chicago, IL 60637, USA
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
This lecture series provides an overview of modern physical cosmology with an emphasis on nuclear arguments and their role in the larger framework. In particular, the current situation on the age of the universe and the Hubble constant are reviewed and shown now to be in reasonable agreement once realistic systematic uncertainties are included in the estimates. Big bang nucleosynthesis is mentioned as one of the pillars of the big bang along with the microwave background radiation. It is shown that the big bang nucleosynthesis constraints on the cosmological baryon density, when compared with dynamical and gravitational lensing arguments, demonstrate that the bulk of the baryons are dark and also that the bulk of the matter in the universe is non–baryonic. The recent extragalactic deuterium observations as well as the other light element abundances are examined in detail. Comparison of nucleosynthesis baryonic density arguments with other baryon density arguments is made.
Introduction
Modern physical cosmology has entered a “golden period” where a multitude of observations and experiments are guiding and constraining the theory in a heretofore unimagined manner. Many of these constraints involve nuclear physics arguments, so the interface with nuclear astrophysics is extemely active. This review opens with a discussion of the three pillar of the big bang: the Hubble expansion, the cosmic microwave background, and big bang nucleosynthesis (BBN).
The Lagrangian and Hamiltonian formalisms of particle dynamics can be generalized and extended to describe continuous systems such as a vibrating rod or a fluid. In such systems each point x, influenced by both external and internal forces, moves independently. In the rod, points that start out very close together always remain close together, whereas in the fluid they may end up far apart. The displacement and velocity of the points of the system are described by functions ψ(x, t) called fields.
In this chapter we describe the classical (nonquantum) theory of fields. Particle dynamics will be transformed to field theory by allowing the number of particles to increase without bound while their masses and the distance between them go to zero in such a way that a meaningful limit exists. This is called passing to the continuum limit.
LAGRANGIAN FORMULATION OF CONTINUUM DYNAMICS
PASSING TO THE CONTINUUM LIMIT
THE SINE–GORDON EQUATION
In this section we present an example to show how to pass to the continuum limit. Consider the system illustrated in Fig. 9.1, consisting of many simple plane pendula of length l and mass m, all suspended from a horizontal rod with a screw thread cut into it. The planes of the pendula are perpendicular to the rod (only two of the planes are drawn in the figure), and the pendula are attached to (massless) nuts that move along the rod as the pendula swing in their planes [for more details, see Scott (1969)].
By
F.-K. Thielemann, Departement für Physik und Astronomie, Universität Basel, CH–4056 Basel, Switzerland, institute for Theoretical Physics, University of California, Santa Barbara, CA 93106–4030,
T. Rauscher, Departement für Physik und Astronomie, Universität Basel, CH–4056 Basel, Switzerland,
C. Freiburghaus, Departement für Physik und Astronomie, Universität Basel, CH–4056 Basel, Switzerland, institute for Theoretical Physics, University of California, Santa Barbara, CA 93106–4030,
K. Nomoto, Department of Astronomy and Research Center for the Early Universe, University of Tokyo, Tokyo 113, Japan, institute for Theoretical Physics, University of California, Santa Barbara, CA 93106–4030,
M. Hashimoto, Department of Physics, Faculty of Science, Kyushu University, Pukuoka 810, Japan,
B. Pfeiffer, Institut für Kernchemie, Universität Mainz, D–55128 Mainz, Germany,
K.-L. Kratz, Institut für Kernchemie, Universität Mainz, D–55128 Mainz, Germany
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
This review concentrates on nucleosynthesis processes in general and their applications to massive stars and supernovae. A brief initial introduction is given to the physics in astrophysical plasmas which governs composition changes. We present the basic equations for thermonuclear reaction rates and nuclear reaction networks. The required nuclear physics input for reaction rates is discussed, i.e. cross sections for nuclear reactions, photodisintegrations, electron and positron captures, neutrino captures, inelastic neutrino scattering, and beta–decay half–lives. We examine especially the present state of uncertainties in predicting thermonuclear reaction rates, while the status of experiments is discussed by others in this volume (see M. Wiescher). It follows a brief review of hydrostatic burning stages in stellar evolution before discussing the fate of massive stars, i.e. the nucleosynthesis in type II supernova explosions (SNe II). Except for SNe la, which are explained by exploding white dwarfs in binary stellar systems (which will not be discussed here), all other supernova types seem to be linked to the gravitational collapse of massive stars (M>8M⊙) at the end of their hydrostatic evolution. SN1987A, the first type II supernova for which the progenitor star was known, is used as an example for nucleosynthesis calculations. Finally, we discuss the production of heavy elements in the r–process up to Th and U and its possible connection to supernovae.
Most dynamical systems cannot be integrated in closed form: the equations of motion do not lend themselves to explicit analytic solution. There exist strategies, however, for approximating solutions analytically and for obtaining qualitative information about even extremely complex and difficult dynamical systems. Part of this chapter is devoted to such strategies for Hamiltonian dynamics.
The first two sections describe Hamilton–Jacobi (HJ) theory and introduce action–angle (AA) variables on T*ℚ. HJ theory presents a new and powerful way for integrating Hamilton's canonical equations, and motion on T*ℚ is particularly elegant and easy to visualize when it can be viewed in terms of AA variables. Both schemes are important in their own right, but they also set the stage for Section 6.3.
THE HAMILTON–JACOBI METHOD
It was mentioned at the end of Worked Example 5.5 that there would be some advantage to finding local canonical coordinates on T*ℚ in which the new Hamiltonian function vanishes identically, or more generally is identically equal to a constant. In this section we show how to do this by obtaining a Type 1 generating function for a CT from the initial local coordinates on T*ℚ to new ones, all of which are constants of the motion. The desired generating function is the solution of a nonlinear partial differential equation known as the Hamilton–Jacobi (or HJ) equation. Unfortunately, the HJ equation is in general not easy to solve.
This chapter is devoted almost entirely to nonlinear dynamical systems, whose equations of motion involve nonlinear functions of positions and velocities. We concentrate on some general topics, like stability of solutions, behavior near fixed points, and the extent to which perturbative methods converge. Because nonlinear systems involve complicated calculations, the chapter stresses mathematical detail. It also presents results of some numerical calculations. The importance of numerical methods can not be overemphasized: because nonlinear systems are inherently more complicated than linear ones, it is often impossible to handle them by purely analytic methods. Some results depend also on topics from number theory, and these are discussed in an appendix at the end of the chapter.
We have already dealt with nonlinear systems perturbatively (e.g., the quartic oscillator of Section 6.3.1), but now we will go into more detail. It will be shown, among other things, that in the nonperturbative regime nonlinear systems often exhibit the kind of complicated behavior that is called chaos.
The first four sections of the chapter do not use the Lagrangian or Hamiltonian description to deal with dynamical systems. In them we discuss various kinds of systems, largely oscillators, which are often dissipative and/or driven. We go into detail concerning some matters that have been touched on earlier (e.g., stability, the Poincaré map), introducing ideas and terminology that are important in understanding the behavior of nonlinear systems of all kinds. In particular, considerable space is devoted to discrete maps.
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
Among the first courses taken by graduate students in physics in North America is Classical Mechanics. This book is a contemporary text for such a course, containing material traditionally found in the classical textbooks written through the early 1970s as well as recent developments that have transformed classical mechanics to a subject of significant contemporary research. It is an attempt to merge the traditional and the modern in one coherent presentation.
When we started writing the book we planned merely to update the classical book by Saletan and Cromer (1971) (SC) by adding more modern topics, mostly by emphasizing differential geometric and nonlinear dynamical methods. But that book was written when the frontier was largely quantum field theory, and the frontier has changed and is now moving in many different directions. Moreover, classical mechanics occupies a different position in contemporary physics than it did when SC was written. Thus this book is not merely an update of SC. Every page has been written anew and the book now includes many new topics that were not even in existence when SC was written. (Nevertheless, traces of SC remain and are evident in the frequent references to it.)
From the late seventeenth century well into the nineteenth, classical mechanics was one of the main driving forces in the development of physics, interacting strongly with developments in mathematics, both by borrowing and lending. The topics developed by its main protagonists, Newton, Lagrange, Euler, Hamilton, and Jacobi among others, form the basis of the traditional material.
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
By
Thomas K. Gaisser, Bartol Research Institute, University of Delaware, Newark, DE 19716, USA
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
This chapter is a review of the background and status of several current problems of interest concerning cosmic rays of very high energy and related signals of photons and neutrinos.
Introduction
The steeply falling spectrum of cosmic rays extends over many orders of magnitude with only three notable features:
(a) The flattened portion below 10 GeV that varies in inverse correlation with solar activity,
(b) The “knee” of the spectrum between 1015 and 1016 eV, and
(c) the “ankle” around 1019 eV.
For my discussion here I will divide the spectrum into three energy regions that are related to the two high–energy features, the knee and the ankle: I: E < 1014 eV, II: 1014 < E < 1018 eV and III: > 1018 eV.
In Region I (VHE) there are detailed measurements of primary cosmic rays made from detectors carried in balloons and on spacecraft. These observations, and related theoretical work on space plasma physics, form the basis of what might be called the standard model of origin of cosmic rays. Cosmic rays are accelerated by the first order Fermi mechanism at strong shocks driven by supernova remnants (SNR) in the disk of the galaxy. The ionized, accelerated nuclei then diffuse in the turbulent, magnetized plasma of the interstellar medium, eventually escaping into intergalactic space at a rate that depends on their energy.
By
Madappa Prakash, Department of Physics and Astronomy, SUNY at Stony Brook, Stony Brook, NY 11794, USA
Edited by
Jorge G. Hirsch, Center of Research and Advanced Studies, National Polytechnic Institute, Mexico City,Danny Page, Universidad Nacional Autónoma de México
The structure of neutron stars is discussed with a view to explore (1) the extent to which stringent constraints may be placed on the equation of state of dense matter by a comparison of calculations with the available data on some basic neutron star properties; and (2) some astrophysical consequences of the possible presence of strangeness, in the form of baryons, notably the Λ and Σ−, or as a Bose condensate, such as a K− condensate, or in the form of strange quarks.
Introduction
Almost every physical aspect of a neutron star tends to the extreme when compared to similar traits of other commonly observed objects in the universe. Stable matter containing A ∼ 1057 baryons and with a mass in the range of (1 − 2) M⊙ {M⊙ ≅ 2 × 1033 g) confined to a sphere of radius R ∼ 10 km (recall that R⊙ = 6.96 × 105 km) represents one of the densest forms of matter in the observable universe. Depending on the equation of state (EOS) of matter at the core of a neutron star, the central density could reach as high as (5 − 10)p0, where p0 ≅ 2.65 × 1014 g cm−3 (corresponding to a number density of n0 ≅ 0.16 fm−3) is the central mass density of heavy laboratory nuclei (compare this to P⊙= 1.4 g cm−3).
In its treatment of dynamical systems, this book has emphasized several themes, among them nonlinearity, Hamiltonian dynamics, symplectic geometry, and field equations. The last section on the Hamiltonian treatment of nonlinear field equations unites these themes in bringing the book to an end. To arrive at this point we have traveled along the roads laid out by the Newtonian, Lagrangian, Hamiltonian, and Hamilton–Jacobi formalisms, eventually replacing the traditional view by a more modern one. On the way we have roamed through many byways, old and new, achieving a broad understanding of the terrain.
The end of the book, however, is not the end of the journey. The road goes on: there is much more terrain to cover. In particular, statistical physics and quantum mechanics, for both of which the Hamiltonian formalism is crucial, lie close to the road we have taken. These fields contain still unsolved problems bordering on the matters we have discussed, in particular questions about ergodic theory in statistical physics and about chaos in quantum mechanics. As usual in physics, the progress being made in such fields builds on previous work. In this case it depends on subject matter of the kind treated in this book.
Although they are deterministic, classical dynamical systems are generically so sensitive to initial conditions that their motion is all but unpredictable. Even inherently, therefore, classical mechanics still offers a rich potential for future discoveries and applications.
This chapter contains material that grows out of and sometimes away from Lagrangian mechanics. Some of it makes use of the Lagrangian formulation (e.g., scattering by a magnetic dipole, linear oscillations), and some does not (e.g., scattering by a central potential, chaotic scattering). Much of this material is intended to prepare the reader for topics to be discussed later: chaotic scattering off hard disks is a particularly understandable example of chaos, and linear oscillators are the starting point for the discussion of nonlinear ones, of perturbation theory, and field theory.
SCATTERING
SCATTERING BY CENTRAL FORCES
So far we have mostly concentrated on bounded orbits of central-force systems. Yet unbounded orbits are very common for both attractive and repulsive central forces. Indeed, for repulsive ones all orbits are unbounded. For attractive ones, the outer turning point of bounded orbits – the maximum distance of the particle from the attracting center – increases with increasing energy. If the force is weak enough at large distances, the particle may escape from the force center even at finite energies: the orbit may break open and the outer turning point may move off to infinity. Then the orbit becomes unbounded, as in the hyperbolic orbits in the Coulomb potential. The result is what is called scattering.
GENERAL CONSIDERATIONS
Consider a central-force dynamical system that possesses unbounded orbits. Let the initial conditions be that a particle is approaching the force center from far away, so far away that it may be thought of as coming from infinity, which means that it is moving on one of the unbounded orbits.