Prévia do material em texto
Free ebooks ==> www.Ebook777.com Emil de Souza Sánchez Filho Tensor Calculus for Engineers and Physicists www.Ebook777.com http://www.ebook777.com Free ebooks ==> www.Ebook777.com Tensor Calculus for Engineers and Physicists www.Ebook777.com http://www.ebook777.com Emil de Souza Sánchez Filho Tensor Calculus for Engineers and Physicists Free ebooks ==> www.Ebook777.com Emil de Souza Sánchez Filho Fluminense Federal University Rio de Janeiro, Rio de Janeiro Brazil ISBN 978-3-319-31519-5 ISBN 978-3-319-31520-1 (eBook) DOI 10.1007/978-3-319-31520-1 Library of Congress Control Number: 2016938417 © Springer International Publishing Switzerland 2016 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, express or implied, with respect to the material contained herein or for any errors or omissions that may have been made. Printed on acid-free paper This Springer imprint is published by Springer Nature The registered company is Springer International Publishing AG Switzerland www.Ebook777.com http://www.ebook777.com To Sandra, Yuri, Nat�alia and Lara Preface The Tensor Calculus for Engineers and Physicist provides a rigorous approach to tensor manifolds and their role in several issues of these professions. With a thorough, complete, and unified presentation, this book affords insights into several topics of tensor analysis, which covers all aspects of N-dimensional spaces. Although no emphasis is placed on special and particular problems of Engineer- ing or Physics, the text covers the fundamental and complete study of the aim of these fields of the science. The book makes a brief introduction to the basic concept of the tensorial formalism so as to allow the reader to make a quick and easy review of the essential topics that enable having a dominium over the subsequent themes, without needing to resort to other bibliographical sources on tensors. This book did not have the framework of a math book, which is a work that seeks, above all else, to organize ideas and concepts in a didactic manner so as to allow the familiarity with the tensorial approach and its application of the practical cases of Physics and the areas of Engineering. The development of the various chapters does not cling to any particular field of knowledge, and the concepts and the deductions of the equations are presented so as to permit engineers and physicists to read the text without being experts in any branch of science to which a specific topic applies. The chapters treat the various themes in a sequential manner and the deductions are performed without omission of the intermediary steps, the subjects being treated in a didactic manner and supplemented with various examples in the form of solved exercises with the exception of Chap. 3 that broaches review topics. A few problems with answers are presented at the end of each chapter, seeking to allow the reader to improve his practice in solving exercises on the themes that were broached. Chapter 1 is a brief introduction to the basic concepts of tensorial formalism so as to permit the reader to make a quick and easy review of the essential topics that make possible the knowledge of the subsequent themes that come later, without needing to resort to other bibliographic sources on tensors. vii http://dx.doi.org/10.1007/978-3-319-31520-1_1 http://dx.doi.org/10.1007/978-3-319-31520-1_3 The concepts of covariant, absolute, and contravariant derivatives, with the detailed development of all the expressions concerning these parameters, as well as the deductions of the Christoffel symbols of the first and second kind, are the essence of Chap. 2. Chapter 3 presents the Green, Stokes, and Gauss–Ostrogradsky theorems using a vectorial formulation. The expansion of the concepts of the differential operators studied in Differential Calculus is performed in Chap. 4. The scalar, vectorial, and tensorial fields are defined, and the concepts and expressions for gradient, divergence, and curl are formulated. With the definition of the nabla operator, successive applications of this linear differential operator are carried out and various fundamental relations between the differential operators are deducted, defining the Laplace operator. All the formulas are deducted by means of tensorial approach. The definition of metric spaces with several dimensions, with the introduction of Riemann curvature concept, and the Ricci tensor formulations, the scalar curvature, and the Einstein tensor are the subjects studied in detail in Chap. 5. Various particular cases of Riemann spaces are analyzed, such as the bidimensional spaces, the spaces with constant curvature, the Minkowski space, and the conformal spaces, with the definition of the Weyl tensor. Chapter 6 broaches metric spaces provided with curvature with the introduction of the concepts of the geodesics and the geodesics and Riemann coordinate systems. The geodesics deviation and the parallelism of vectors in curved spaces are studied, with the definition of the torsion tensor concept. The purpose of this book is to give a simple, correct, and comprehensive mathematical explanation of Tensor Calculus, and it is self-contained. Postgraduate and advanced undergraduate students and professionals will find clarity and insight into the subject of this textbook. The preparation of a book is a hard and long work that requires the participation of other people besides the author, which are of fundamental importance in the preparation of the originals and in the tiresome task of reviewing the typing, chiefly in a text such as the one in this book. So, our sincere thanks to all those who helped in the preparation and editing of these pages. In relation to the errors in this text which were not corrected by a more diligent review, it is stressed that they are the author’s responsibility and the author apologizes for them. Rio de Janeiro, Brazil Emil de Souza Sánchez Filho December 17, 2015 viii Preface http://dx.doi.org/10.1007/978-3-319-31520-1_6 http://dx.doi.org/10.1007/978-3-319-31520-1_5 http://dx.doi.org/10.1007/978-3-319-31520-1_4 http://dx.doi.org/10.1007/978-3-319-31520-1_3 http://dx.doi.org/10.1007/978-3-319-31520-1_2 Free ebooks ==> www.Ebook777.com Historical Introduction This brief history of Tensor Calculus broaches the development of the idea of vector and the advent of the concept of tensor in a synthetic way. The following paragraphs aim to show the history of the development of these themes in the course of time, highlighting the main stages that took place in this evolution of the mathematical knowledge. A few items of bibliographic data of the mathematicians and scientists who participated on this epic journey in a more striking manner are described. The perception of Nature under a purely philosophical focus led Plato in 360 BC to the study of geometry. This philosopher classified the geometric figures into triangles, rectangles, and circles, and with this system, he grounded the basic conceptsof geometry. Later Euclid systemized geometry in axiomatic form, starting from the fundamental concepts of points and lines. The wise men of ancient Greece also concerned themselves with the study of the movement of bodies by means of geometric concepts. The texts of Aristotle (384– 322 BC) inMechanics show that he had the notion of composition of movements. In this work, Aristotle enounced in an axiomatic form that the force that moves a body is collinear with the direction of the body’s movement. In a segment ofMechanics, he describes the velocity of two bodies in linear movement with constant pro- portions between each other, explaining that “When a body moves with a certain proportion, the body needs to move in a straight line, and this is the diameter of the figure formed with the straight lines which have known proportions.” This state- ment deals with the displacements of two bodies—the Greek sage acknowleding that the resultant of these displacements would be the diagonal of the rectangle (the text talks about the diameter) from the composition of the speeds. In the Renaissance, the prominent figure of Leonardo da Vinci (1452–1519) also stood out in the field of sciences. In his writings, he reports that “Mechanics is the paradise of mathematical science, because all the fruits of mathematics are picked here.” Da Vinci conceived concepts on the composition of forces for maintaining the balance of the simple structures, but enunciated them in an erroneous and contradictory manner in view of the present-day knowledge. ix www.Ebook777.com http://www.ebook777.com The awakening of a new manner of facing the uniform was already blossoming in the 1600s. The ideas about the conception and study of the world were no longer conceived from the scholastic point of view, for reason more than faith had become the way to new discoveries and interpretations of the outside world. In the Nether- lands, where liberal ideas were admitted and free thought could be exercised in full, the Dutch mathematician Simon Stevin (1548–1620), or Stevinus in a Latinized spelling, was the one who demonstrated in a clear manner the rule for the compo- sition of forces, when analyzing the balance of a body located in an inclined plane and supported by weights, one hanging at the end of a lever, and the other hanging from a pulley attached to the vertical cathetus of the inclined plane. This rule is a part of the writings of Galileo Galilei (1564–1642) on the balance of bodies in a tilted plane. However, it became necessary to conceive mathematical formalism that translated these experimental verifications. The start of the concept of vector came about in an empirical mode with the formulation of the parallelogram rule, for Stevinus, in a paper published in 1586 on applied mechanics, set forth this principle of Classic Mechanics, formalizing by means of the balance of a force system the concept of a variety depending on the direction and orientation of its action, enabling in the future the theoretical preparation of the concept of vector. The creation of the Analytical Geometry by René Du Perron Descartes (1596– 1650) brought together Euclid’s geometry and algebra, establishing a univocal correspondence between the points of a straight line and the real numbers. The introduction of the orthogonal coordinates system, also called Cartesian coordi- nates, allowed the calculation of the distance ds between two points in the Euclid- ean space by algebraic means, given by ds2 ¼ dx2 þ dy2 þ dz2, where dx, dy, dz are the coordinates of the point. The movement of the bodies was a focus of attention of the mathematicians and scientists, and a more elaborate mathematical approach was necessary when it was studied. This was taken care of by Leonhard Paul Euler (1707–1783), who con- ceived the concept of inertia tensor. This concept is present in his book Theoria Motus Corporum Solidorum seu Rigidorum (Theory of the Movement of the Solid and Rigid Bodies) published in 1760. In this paper, Euler studies the curvature lines, initiating the study of Differential Geometry. He was the most published x Historical Introduction mathematician of the all time, 860 works are known from him, and it is known that he published 560 papers during his lifetime, among books, articles, and letters. In the early 1800s, Germany was becoming the world’s largest center in math- ematics. Among many of its brilliant minds, it counted Johann Karl Friedrich Gauß (1777–1855). On occupying himself with the studies of curves and surfaces, Gauß coined the term non-Euclidean geometries; in 1816, he’d already conceived con- cepts relative to these geometries. He prepared a theory of surfaces using curvilin- ear coordinates in the paper Disquisitones Generales circa Superfı́cies Curvas, published in 1827. Gauß argued that the space geometry has a physical aspect to be discovered by experimentation. These ideas went against the philosophical concepts of Immanuel Kant (1724–1804), who preconized that the conception of the space is a priori Euclidian. Gauß conceived a system of local coordinates system u, v,w located on a surface, which allowed him to calculate the distance between two points on this surface, given by the quadratic expression ds2 ¼ Adu2þ Bdv2 þ Cdw2 þ 2Edu � dvþ 2Fdv � dwþ 2Gdu � dw, where A,B,C,F,G are func- tions of the coordinates u, v,w. The idea of force associated with a direction could be better developed analyt- ically after the creation of the Analytical Geometry by Descartes. The representa- tion of the complex numbers by means of two orthogonal axes, one axis representing the real numbers and the other axis representing the imaginary num- ber, was developed by the Englishman John Wallis (1616–1703). This representa- tion allowed the Frenchman Jean Robert Argand (1768–1822) to develop in 1778, in a manner independent from the Dane Gaspar Wessel (1745–1818), the mathe- matical operations between the complex numbers. These operations served as a framework for the Irish mathematician William Rowan Hamilton (1805–1865) to develop a more encompassing study in three dimensions, in which the complex numbers are contained in a new variety: the Quaternions. Historical Introduction xi This development came about by means of the works of Hamilton, who had the beginning of his career marked by the discovery of an error in the bookMécanique Celeste authored by Pierre Simon-Laplace (1749–1827), which gave him prestige in the intellectual environment. In his time, there was a great discrepancy between the mathematical production from the European continent and from Great Britain, for the golden times of Isaac Newton (1642–1727) had already passed. Hamilton studied the last advances of the continental mathematics, and between 1834 and 1835, he published the books General Methods in Dynamics. In 1843, he published the Quaternions Theory, printed in two volumes, the first one in 1853 and the second one in 1866, in which a theory similar to the vector theory was outlined, stressing, however, that these two theories differ in their grounds. In the first half of the nineteenth century, the German Hermann G€unther Graßmann (1809–1877), a secondary school teacher of the city of Stettin located in the region that belongs to Pomerania and that is currently a part of Poland, published the book Die Lineale Ausdehnunsgleher ein neuer Zweig der Mathematik (Extension Theory), in which he studies a geometry of more than three dimensions, treating N dimensions, and formulating a generalization of the classic geometry. To outline this theory, he used the concepts of invariants (vectors and tensors), which later enabled other scholars to develop calculus and vector analysis. xii Historical Introduction The great mathematical contribution of the nineteenth century, which definitely marked the development of Physics, is due to Georg Friedrich Bernhard Riemann (1826–1866). Riemann studiedin G€ottingen, where he was a pupil of Gauß, and afterward in Berlin, where he was a pupil of Peter Gustav Lejeune Dirichlet (1805– 1859), and showed an exceptional capacity for mathematics when he was still young. His most striking contribution was when he submitted in December 1853 his Habilititationsschrift (thesis) to compete for the position of Privatdozent at the University of G€ottingen. This thesis titled €Uber die Hypothesen welche der Geometrie zu Grunde liegen enabled a genial revolution in the structure of Physics in the beginning of the twentieth century, providing Albert Einstein (1879–1955) with the mathematical background necessary for formulating his Theory of Rela- tivity. The exhibition of this work in a defense of thesis carried out in June 10, 1854, sought to show his capacity to teach. Gauß was a member of examination board and praised the exhibition of Riemann’s new concepts. His excitement for the new formulations was expressed in words: “. . . the depth of the ideas that were presented. . ..” This work was published 14 years later, in 1868, two years after the death of its author. Riemann generalized the geometric concepts of Gauß, conceiving a system of more general coordinates spelled as dxi, and established a fundamental relation for the space of N dimensions, where the distance between two points ds is given by the quadratic form ds2 ¼ gijdxidxj, having gij a symmet- rical function, positive and defined, which characterizes the space in a unique manner. Riemann developed a non-Euclidean, elliptical geometry, different from the geometries of János Bolyai and Nikolai Ivanovich Lobachevsky. The Riemann Geometry unified these three types of geometry and generalized the concepts of curves and surfaces for hyperspaces. The broaching of the Euclidean space in terms of generic coordinates was carried out for the first time by Gabriel Lamé (1795–1870) in his work Leçons sur les Fonctions Inverses des Transcedentes et les Surfaces Isothermes, published in Paris in 1857, and in another work Leçons sur les Coordonées Curvilignes, published in Paris in 1859. Historical Introduction xiii The new experimental discoveries in the fields of electricity and magnetism made the development of an adequate mathematical language necessary to translate them in an effective way. These practical needs led the North American Josiah Willard Gibbs (1839–1903) and the Englishman Olivier Heaviside (1850–1925), in an independent manner, to reformulate the conceptions of Graßmann and Hamilton, creating the vector calculus. Heaviside had thoughts turned toward the practical cases and sought applications for the vectors and used vector calculus in electro- magnetism problems in the industrial areas. With these practical applications, the vectorial formalism became a tool to be used in problems of engineering and physics, and Edwin Bidwell Wilson, a pupil of Gibbs, developed his master’s idea in the book Vector Analysis: A Text Book for the Use for Students of Mathematics and Physics Founded upon Lectures of Josiah Willard Gibbs, published in 1901 where he disclosed this mathematical apparatus, making it popular. This was the first book to present the modern system of vectorial analysis and became a landmark in broadcasting the concepts of calculus and vectorial analysis. xiv Historical Introduction The German mathematician and prominent professor Elwin Bruno Christoffel (1829–1900) developed researches on the Invariant Theory, writing six articles about this subject. In the article €Uber die Transformation der Homogenel Differentialausdr€ucke zweiten Grade, published in the Journal f€ur Mathematik, 70, 1869, he studied the differentiation of the symmetric tensor gij and introduced two functions formed by combinations of partial derivatives of this tensor, con- ceiving two differential operators called Christoffel symbols of the first and second kind, which are fundamental in Tensorial Analysis. With this, he contributed in a fundamental way to the arrival of Tensor Calculus later developed by Gregorio Ricci-Curbastro and Tullio Levi Civita. The metrics of the Riemann spaces and the Christoffel symbols are the fundaments of Tensor Calculus. The importance of tensors in problems of Physics is due to the fact that physical phenomena are analyzed by means of models which include these varieties, which are described in terms of reference systems. However, the coordinates which are described in terms of the reference systems are not a part of the phenomena, only a tool used to represent them mathematically. As no privileged reference systems exist, it becomes necessary to establish relations which transform the coordinates from one referential system to another, so as to relate the tensors’ components. These components in a coordinate system are linear and homogeneous functions of the components in another reference system. The technological development at the end of the nineteenth century and the great advances in the theory of electromagnetism and in theoretical physics made the conception of a new mathematical tool which enabled expressing new concepts and Historical Introduction xv laws imperious. The vectorial formalism did not fulfill the broad field and the variety of new knowledge that needed to be studied more and interpreted better. This tool began to be created by the Italian mathematician Gregorio Ricci- Curbastro (1853–1925), who initiated the conception of Absolute Differential Calculus in 1884. Ricci-Curbastro was a mathematical physicist par excellence. He was a pupil of the imminent Italian professors Enrico Betti (1823–1892) and Eugenio Beltrami (1835–1900). He occupied himself mainly with the Riemann geometry and the study of the quadratic differential form and was influenced by Christoffel’s idea of covariant differentiation which allowed achieving great advances in geometry. He created a research group in which Tullio Levi-Civita participated and worked for 10 years (1887–1896) in the exploration of the new concepts and of an elegant and synthetic notation easily applicable to a variety of problems of mathematical analysis, geometry, and physics. In his article,Méthodes de Calcul Differéntiel Absolu et leurs Applications, published in 1900 in vol. 54 of the Mathematische Annalen, in conjunction with his pupil Levi-Civita, the appli- cations of the differential invariants were broached, subject of the Elasticity Theory, of the Classic Mechanics and the Differential Geometry. This article is considered as the beginning of the creation of Tensor Calculus. He published the first explanation of his method in the Volume XVI of the Bulletin des Sciences Mathématiques (1892), applying it to problems from Differential Geometry to Mathematical Physics. The transformation law of a function system is due to Ricci-Curbastro, who published it in an article in 1887, and which is also present in another article published 1889, in which he introduces the use of upper and lower indexes, showing the differences between the contravariant and covariant transfor- mation laws. In these papers, he exhibits the framework of Tensor Calculus. The pupil and collaborator of Ricci-Curbastro, Tullio Levi-Civita (1873–1941) published in 1917 in the Rediconti del Circolo Matem�atico di Palermo, XLII (pp. 173–215) the article Nozione di Parallelismo in una Variet�a Qualunque e Conseguente Specificazione Geometrica della Curvatura Riemanniana, contribut- ing in a considerable way to the development of Tensor Calculus. In this work, he describes the parallelism in curved spaces. This study was presented in lectures addressed in two courses given at the University of Rome in the period of xvi Historical Introduction 1920–1921 and 1922–1923. He corresponded with Einstein, who showed great interest in the new mathematical tool. In 1925, he published the book Lezione di Calcolo Differenziale Absoluto which is a classic in the mathematicalliterature. It was the German Albert Einstein in 1916 who called the Absolute Differential Calculus of Ricci-Curbastro and Levi-Civita Tensor Calculus, but the term tensor, such as it is understood today, had been introduced in the literature in 1908 by the physicist and crystallographer G€ottingen, Waldemar Voigt (1850–1919). The development of the theoretical works of Einstein was only possible after he became aware of by means of his colleague from Zurich, Marcel Grossmann (1878–1936), head professor of descriptive geometry at the Eidgen€ossische Technische Hochschule, the article Méthodes de Calcul Differéntiel Absolut, which provided him the mathematical tool necessary to conceive his theory, publishing in 1916 in the Annalen der Physik the article Die Grundlagen der algemeinnen Relativitatstheorie. His contribution Tensor Calculus also came about with the conception of the summation rule incorporated to the index notation. The term tensor became popular mainly due to the Theory of Relativity, in which Einstein used this denomination. His researches on the gravitational field also had the help of Grossmann, Tulio Levi-Civita, and Gregorio Ricci-Curbastro, conceiving the Gen- eral Relativity Theory. On the use of the Tensor Calculus in his Gravitation Theory, Einstein wrote: “Sie bedeutet einen wahren Triumph der durch Gauss, Riemann, Christoffel, Ricci . . . begr€undeten Methoden des allgemeinen Differentialkalculus.” Historical Introduction xvii Other notable mathematics contributed to the development of the study of tensors. The Dutch Jan Arnoldus Schouten (1873–1941), professor of the T. U. Delft, discovered independently of Levi-Civita the parallelism and systematized the Tensor Calculus. Schouten published in 1924 the book Ricci-Kalk€ulwhich became a reference work on the subject, where he innovates the tensorial notation, placing the tensor indexes in brackets to indicate that it was an antisymmetric tensor. The Englishman Arthur Stanley Eddington (1882–1944) conceived new in Tensor Calculus and was major promoter of the Theory of Relativity to the lay public. The German Hermann Klaus Hugo Weyl (1885–1955) published in 1913 Die Idee der Riemannschen Fl€ache, which gave a unified treatment of Riemann xviii Historical Introduction Free ebooks ==> www.Ebook777.com surfaces. He contributed to the development and disclosure of Tensor Calculus, publishing in 1918 the book Raum-Zeit-Materie a classic on the Theory of Rela- tivity. Weyl was one of the greatest and most influential mathematicians of the twentieth century, with broad dominium of themes with knowledge nearing the “universalism.” The American Luther Pfahler Eisenhart (1876–1965) who contributed greatly to semi-Riemannian geometry wrote several fundamental books with tensorial approach. The work of French mathematician Élie Joseph Cartan (1869–1951) in differ- ential forms, one of the basic kinds of tensors used in mathematics, is principal reference in this theme. He published the famous book Leçons sur la Géométrie des Espaces de Riemann (first edition in 1928 and second edition in 1946). Historical Introduction xix www.Ebook777.com http://www.ebook777.com Contents 1 Review of Fundamental Topics About Tensors . . . . . . . . . . . . . . . . . 1 1.1 Preview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1.1 Index Notation and Transformation of Coordinates . . . . . 1 1.2 Space of N Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3.2 Kronecker Delta and Permutation Symbol . . . . . . . . . . . . 3 1.3.3 Dual (or Reciprocal) Basis . . . . . . . . . . . . . . . . . . . . . . . 3 1.3.4 Multilinear Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 1.4 Homogeneous Spaces and Isotropic Spaces . . . . . . . . . . . . . . . . . 16 1.5 Metric Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 1.5.1 Conjugated Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 1.5.2 Dot Product in Metric Spaces . . . . . . . . . . . . . . . . . . . . . 30 1.6 Angle Between Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 1.6.1 Symmetrical and Antisymmetrical Tensors . . . . . . . . . . . 43 1.7 Relative Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 1.7.1 Multiplication by a Scalar . . . . . . . . . . . . . . . . . . . . . . . 54 1.8 Physical Components of a Tensor . . . . . . . . . . . . . . . . . . . . . . . . 62 1.8.1 Physical Components of a Vector . . . . . . . . . . . . . . . . . . 62 1.9 Tests of the Tensorial Characteristics of a Variety . . . . . . . . . . . . 66 2 Covariant, Absolute, and Contravariant Derivatives . . . . . . . . . . . . 73 2.1 Initial Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 2.2 Cartesian Tensor Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 2.2.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 2.2.2 Cartesian Tensor of the Second Order . . . . . . . . . . . . . . . 77 2.3 Derivatives of the Basis Vectors . . . . . . . . . . . . . . . . . . . . . . . . . 78 2.3.1 Christoffel Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 2.3.2 Relation Between the Christoffel Symbols . . . . . . . . . . . 83 2.3.3 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 2.3.4 Cartesian Coordinate System . . . . . . . . . . . . . . . . . . . . . 84 xxi 2.3.5 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 2.3.6 Number of Different Terms . . . . . . . . . . . . . . . . . . . . . . 85 2.3.7 Transformation of the Christoffel Symbol of First Kind . . 86 2.3.8 Transformation of the Christoffel Symbol of Second Kind 87 2.3.9 Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . 88 2.3.10 Orthogonal Coordinate Systems . . . . . . . . . . . . . . . . . . . 88 2.3.11 Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 2.3.12 Christoffel Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 2.3.13 Ricci Identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 2.3.14 Fundamental Relations . . . . . . . . . . . . . . . . . . . . . . . . . . 93 2.4 Covariant Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 2.4.1 Contravariant Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 2.4.2 Contravariant Tensor of the Second-Order . . . . . . . . . . . 104 2.4.3 Covariant Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 2.4.4 Mixed Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 2.4.5 Covariant Derivative of the Addition, Subtraction, and Product of Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 2.4.6 Covariant Derivative of Tensors gij, g ij, δij . . . . . . . . . . . . 117 2.4.7 Particularities of the Covariant Derivative . . . . . . . . . . . . 121 2.5 Covariant Derivative of Relative Tensors . . . . . . . . . . . . . . . . . . . 123 2.5.1 Covariant Derivative of the Ricci Pseudotensor . . . . . . . . 125 2.6 Intrinsic or Absolute Derivative . . . . . . . . . . . . . . . . . . . . . . . . . 128 2.6.1 Uniqueness of the Absolute Derivative . . . . . . . . . . . . . . 131 2.7 Contravariant Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 3 Integral Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 3.1 Basic Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 3.1.1 Smooth Surface . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . 137 3.1.2 Simply Connected Domain . . . . . . . . . . . . . . . . . . . . . . . 137 3.1.3 Multiply Connected Domain . . . . . . . . . . . . . . . . . . . . . 138 3.1.4 Oriented Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 3.1.5 Surface Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 3.1.6 Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 3.2 Oriented Surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 3.2.1 Volume Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 3.3 Green’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 3.4 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 3.5 Gauß–Ostrogradsky Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 4 Differential Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 4.1 Scalar, Vectorial, and Tensorial Fields . . . . . . . . . . . . . . . . . . . . . 155 4.1.1 Initial Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 4.1.2 Scalar Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 xxii Contents 4.1.3 Pseudoscalar Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 4.1.4 Vectorial Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 4.1.5 Tensorial Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 4.1.6 Circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 4.2 Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 4.2.1 Norm of the Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 4.2.2 Orthogonal Coordinate Systems . . . . . . . . . . . . . . . . . . . 165 4.2.3 Directional Derivative of the Gradient . . . . . . . . . . . . . . 166 4.2.4 Dyadic Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 4.2.5 Gradient of a Second-Order Tensor . . . . . . . . . . . . . . . . 169 4.2.6 Gradient Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 4.3 Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 4.3.1 Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 4.3.2 Contravariant and Covariant Components . . . . . . . . . . . . 179 4.3.3 Orthogonal Coordinate Systems . . . . . . . . . . . . . . . . . . . 181 4.3.4 Physical Components . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 4.3.5 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 4.3.6 Divergence of a Second-Order Tensor . . . . . . . . . . . . . . 183 4.4 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194 4.4.1 Stokes Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196 4.4.2 Orthogonal Curvilinear Coordinate Systems . . . . . . . . . . 201 4.4.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 4.4.4 Curl of a Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 4.5 Successive Applications of the Nabla Operator . . . . . . . . . . . . . . 207 4.5.1 Basic Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 4.5.2 Laplace Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214 4.5.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216 4.5.4 Orthogonal Coordinate Systems . . . . . . . . . . . . . . . . . . . 218 4.5.5 Laplacian of a Vector . . . . . . . . . . . . . . . . . . . . . . . . . . 218 4.5.6 Curl of the Laplacian of a Vector . . . . . . . . . . . . . . . . . . 219 4.5.7 Laplacian of a Second-Order Tensor . . . . . . . . . . . . . . . . 220 4.6 Other Differential Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224 4.6.1 Hesse Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224 4.6.2 D’Alembert Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 5 Riemann Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 5.1 Preview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 5.2 The Curvature Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 5.2.1 Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228 5.2.2 Differentiation Commutativity . . . . . . . . . . . . . . . . . . . . 231 5.2.3 Antisymmetry of Tensor Ri‘jk . . . . . . . . . . . . . . . . . . . . . 233 5.2.4 Notations for Tensor Ri‘jk . . . . . . . . . . . . . . . . . . . . . . . . 233 5.2.5 Uniqueness of Tensor R‘ijk . . . . . . . . . . . . . . . . . . . . . . . 234 5.2.6 First Bianchi Identity . . . . . . . . . . . . . . . . . . . . . . . . . . . 234 Contents xxiii 5.2.7 Second Bianchi Identity . . . . . . . . . . . . . . . . . . . . . . . . . 235 5.2.8 Curvature Tensor of Variance (0, 4) . . . . . . . . . . . . . . . . 238 5.2.9 Properties of Tensor Rpijk . . . . . . . . . . . . . . . . . . . . . . . . 240 5.2.10 Distinct Algebraic Components of Tensor Rpijk . . . . . . . . 241 5.2.11 Classification of Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 245 5.3 Riemann Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246 5.3.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246 5.3.2 Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247 5.3.3 Normalized Form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248 5.4 Ricci Tensor and Scalar Curvature . . . . . . . . . . . . . . . . . . . . . . . 250 5.4.1 Ricci Tensor with Variance (0, 2) . . . . . . . . . . . . . . . . . . 251 5.4.2 Divergence of the Ricci Tensor with Variance Ricci (0, 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253 5.4.3 Bianchi Identity for the Ricci Tensor with Variance (0, 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253 5.4.4 Scalar Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254 5.4.5 Geometric Interpretation of the Ricci Tensor with Variance (0, 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254 5.4.6 Eigenvectors of the Ricci Tensor with Variance (0, 2) . . . 256 5.4.7 Ricci Tensor with Variance (1, 1) . . . . . . . . . . . . . . . . . . 257 5.4.8 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259 5.5 Einstein Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262 5.6 Particular Cases of Riemann Spaces . . . . . . . . . . . . . . . . . . . . . . 264 5.6.1 Riemann Space E2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265 5.6.2 Gauß Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267 5.6.3 Component R1212 in Orthogonal Coordinate Systems . . . . 269 5.6.4 Einstein Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271 5.6.5 Riemann Space with Constant Curvature . . . . . . . . . . . . 273 5.6.6 Isotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274 5.6.7 Minkowski Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 5.6.8 Conformal Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281 5.7 Dimensional Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291 6 Geodesics and Parallelism of Vectors . . . . . . . . . . . . . . . . . . . . . . . . 295 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295 6.2 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295 6.2.1 Representation byMeans of Curves in the Surfaces . . . . . 299 6.2.2 Constant Direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299 6.2.3 Representation by Means of the Unit Tangent Vector . . . 301 6.2.4 Representation by Means of an Arbitrary Parameter . . . . 302 6.3 Geodesics with Null Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307 6.4 Coordinate Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309 6.4.1 Geodesic Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . 309 6.4.2 Riemann Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . 311 6.5 Geodesic Deviation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313 xxiv Contents 6.6 Parallelism of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319 6.6.1 Initial Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319 6.6.2 Parallel Transport of Vectors . . . . . . . . . . . . . . . . . . . . . 321 6.6.3 Torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341 Contents xxv Notations ℜ Set of the real numbers Z Set of the complex numbers � � �j j Determinant � � �k k Modulus, absolute value · Dot product, scalar product, inner product � Cross product, vectorial product � Tensorial product � Two contractions of the tensorial product δij, δ ij, δji Kronecker delta δij . . .m, δ ij . . .m, δj:::: ni ...m Generalized Kronecker delta eijk, e ijk Permutation symbol eijk . . .m, e ijk . . .m Generalized permutation symbol εijk, ε ijk Ricci pseudotensor εi1i2i3���in , ε i1i2i3���in Ricci pseudotensor for the space EN E3 Euclidian space J Jacobian EN Vectorial space or tensorial space with N dimension ‘n� � � Natural logarithm εijk . . .m, ε ijk . . .m Ricci pseudotensor for the space EN d . . . dxk Differentiation with respect to variable xk ϕ, i Comma notation for differentiation _x: Differentiation with respect to time ∂ . . . ∂xk Partial differentiation with respect to variable xk ∂k� � � Covariant derivative δ� � � δ t Intrinsic or absolute derivative ∇� � � Nabla operator xxvii ∇2� � � Laplace operator, Laplacian H� � � Hesse operator, Hessian □. . . D’Alembert operator, D’alembertian div� � � Divergent grad� � � Gradient lap� � � Laplacian rot� � � Rotational, curl gij, g ij, gji Metric tensor Γij,k Christoffel symbol of first kind Γmip Christoffel symbol of the second kind Gij, G ij Einstein tensor Gkm Einstein tensor with variance (1,1) K Riemann curvature R Scalar curvature Rij Ricci tensor of the variance (0,2) Rij Ricci tensor of the variance (1,1) R‘ijk Riemann–Christoffel curvature tensor, Riemann–Christoffel mixed tensor, Riemann–Christoffel tensor of the second kind, curvature tensor Rpijk Curvature tensor of variance (0, 4) tr� � � Trace of the matrix Wijk‘ Weyl curvature tensor Greek Alphabets Sound Letter Alpha α, Α Beta β, Β Gamma γ, Γ Delta δ, Δ Epsilon ε, Ε Zeta ζ, Ζ Eta η, Η Theta θ, Θ Iota ι, I Kappa κ, K Lambda λ, Λ M€u μ, M N€u ν, Ν Ksi ξ, Ξ Omicron o, Ο Pi π, Π Rho ρ, Ρ xxviii Notations Sigma σ, Σ Tau τ, Τ Üpsı́lon υ, Υ Phi φ,ϕ, Φ Khi χ, Χ Psi ψ , Ψ Omega ω, Ω Notations xxix Chapter 1 Review of Fundamental Topics About Tensors 1.1 Preview This chapter presents a brief review of the fundamental concepts required for the consistent development of the later chapters. Various subjects are admitted as being previously known, which allows avoiding demonstrations that overload the text. It is assumed that the reader has full knowledge of Differential and Integral Calculus, Vectorial Calculus, Linear Algebra, and the fundamental concepts about tensors and dominium of the tensorial formalism. However, are presented succinctly the essential topics for understanding the themes that are developed in this book. 1.1.1 Index Notation and Transformation of Coordinates On the course of the text, when dealing with the tensorial formulations, the index notation will be preferably used, and with the summation rule, for instance, yj ¼ X3 i¼1 X3 j¼1 aijxi ¼ aijxi, where i is a free index and j is a dummy in the sense that the sum is independent of the letter used, this expression takes the forms y1 ¼ a11x1 þ a12x2 þ a13x3 y2 ¼ a21x1 þ a22x2 þ a23x3 y3 ¼ a31x1 þ a32x2 þ a33x3 8><>: ) y1 y2 y3 8><>: 9>=>; ¼ a11 a12 a13 a21 a22 a23 a31 a32 a33 264 375 x1x2 x3 8><>: 9>=>; The transformation of the coordinates from a point in the coordinate system Xi to the coordinate system X i given by xi ¼ aijxj þ ai0 where the terms aij, ai0 are constants is called affine transformation (linear). In this kind of transformation, the points of the space E3 are transformed into points, the straight lines in straight © Springer International Publishing Switzerland 2016 E. de Souza Sánchez Filho, Tensor Calculus for Engineers and Physicists, DOI 10.1007/978-3-319-31520-1_1 1 lines, and the planes in planes. When ai0 ¼ 0, this transformation is called linear and homogeneous. The term ai0 represents only a translation of the origin of the referential. 1.2 Space of N Dimensions The generalization of the Euclidian space at three dimensions for a number N of dimensions is prompt, defining a space EN. This expansion of concepts requires establishing a group of N variables xi, i ¼ 1, 2, 3, . . .N, relative to a point P xið Þ2EN , related to a coordinate system Xi, which are called coordinates of the point in this reference system. The set of points associated in a biunivocal way to the coordinates of the reference system Xi defines the N-dimensional space EN. In an analogous way a subspace EM � EN is defined, with M < N, in which the group of pointsP xið Þ2EM is related biunivocally with the coordinates defined in the coordinate system Xi. To make a few specific studies easier, at times the space is divided into subspaces. The space EN is called affine space, and if it is linked to the notion of distance between two points, then it is a metric space. 1.3 Tensors 1.3.1 Vectors The structure of a vectorial space is defined by two algebraic operations: (a) the sum of the vectors and (b) the multiplication of vector by scalar. The conception of vectors u, v,w as geometric varieties is extended to a broad range of functions, as long as the set of these functions forms a vectorial space (linear space) on a set of scalars (numbers). The functions f, g, h, . . .with continuous derivatives that fulfill certain axioms are assumed as vectors, and all the formula- tions and concepts developed for the geometric vectors apply to these formulations. A vectorial space is defined by the following axioms: 1. uþ v ¼ vþ u or f þ g ¼ gþ f . 2. uþ vð Þ þ w ¼ uþ vþ wð Þ or f þ gð Þ þ h ¼ f þ gþ hð Þ. 3. The null vector is such that uþ 0 ¼ u or 0þ f ¼ f . 4. To every vector u there is a corresponding unique vector �u, such that �uð Þ þ u ¼ 0 or �fð Þ þ f ¼ 0. 5. 1 � uk k ¼ u or 1 � fk k ¼ fk k. 6. m nuð Þ ¼ mn uð Þ or m nfð Þ ¼ mn fð Þ, where m, n are scalars. 7. mþ nð Þu ¼ muþ nu or mþ nð Þf ¼ mf þ nf . 8. m uþ vð Þ ¼ muþ mv or m f þ gð Þ ¼ mf þ mg. 2 1 Review of Fundamental Topics About Tensors 1.3.2 Kronecker Delta and Permutation Symbol The Kronecker delta is defined by δ ¼ δij ¼ δ ij ¼ 1, i ¼ j 0, i 6¼ j ( ð1:3:1Þ that is symmetrical, i.e., δij ¼ δji,8i, j. The Kronecker delta is the identity tensor. This tensor is used as a linear operator in algebraic developments, such as ∂xi ∂xj δki ¼ ∂x k ∂xj ∂xj ∂xi δki ¼ ∂x j ∂xk Tijδ ikuk ¼ Tkjuk ¼ Tj The permutation symbol is defined by eijk ¼ eijk ¼ 1 is an even permutation of the indexes �1 is an odd permutation of the indexes 0 when there are repeated indexes 8><>: ð1:3:2Þ and the generalized permutation symbol is given by ei1i2i3���in ¼ ei1i2i3���in¼ 1 is an even permutation of the indexes �1 is an odd permutation of the indexes 0 when there are repeated indexes 8><>: ð1:3:3Þ Figure 1.1 shows an illustration how to obtain the values of this symbol. 1.3.3 Dual (or Reciprocal) Basis The vector u expresses itself in the Euclidean space E3 by means of the linear combination of three linearly independent unit vectors, which form the basis of this space. For the case of oblique coordinate systems, there are two kinds of basis 2 2 1 1 -1+1 3 3 Fig. 1.1 Values of the permutation symbol 1.3 Tensors 3 called reciprocal or dual basis. Let vector u expressed by means of their compo- nents relative to a coordinate system with orthonormal covariant basis ej: u ¼ ujej and with ei � ek ¼ δjk, the dot product takes the form u � ek ¼ ujej � ek ¼ ujδjk ¼ uk, which are the components’ covariant of the vector u. These components are the projections of this vector on the coordinate axes. In the case of oblique coordinate system, the basis ej, ek is called reciprocal basis, which fulfills the condition ej � ek ¼ δ kj . In Fig. 1.2 the axes OXi and OXk are perpendicular, as are also the axes OXk and OXi. This definition shows that the dot product of two reciprocal basis fulfills eik k ei �� �� cos 90o � αð Þ ¼ 1 > 0 ) ei�� �� ¼ 1 eik k sin α and with eik k ¼ 1 results in ei �� �� > 1, then ei and ek have different scales. Let the representation of the vector u in a coordinate system with covariant basis ei, ej, ek, where the indexes of the vectors of the basis indicate a cyclic permutation of i, j, k; thus, u ¼ uiei. These vectors do not have to be coplanar ei � ek � ek 6¼ 0; thus, the volume of the parallelepiped is given by the mixed product ei � ek � ek ¼ V and with the relation between the two reciprocal basis ei � ej ¼ δij follows iX i X k X k X O ie k e ie ke Fig. 1.2 Reciprocal basis 4 1 Review of Fundamental Topics About Tensors Free ebooks ==> www.Ebook777.com 1 ei ¼ e i ¼ ej � ek V Then vector u in terms of reciprocal basis is defined by u ¼ ujej where uj is the components of this vector in the new basis (contravariant), having these new components expressed in terms of the original components. Consider the representation u in terms of the two basis u ¼ uiei ¼ ujej and with the dot product of both sides of this expression by ej, and applying the definition of reciprocal basis e j � ei ¼ 1 provides uj ¼ uiej � ei In an analogous way V ¼ e i � ej � ek whereV is the volume of the parallelepiped defined by the mixed product of the unit vectors of the reciprocal basis. The height of the parallelepiped defined by the mixed product of the unit vectors of a base is collinear with one of the unit vectors of the reciprocal basis (Fig. 1.3). The volume of the parallelepiped is determined by means of the mixed product of three vectors and allows assessing the relations between the same by means of the reciprocal basis in the levorotatory and dextrorotatory coordinates systems. Consider the mixed product of the vectors of the basis of a levorotatory coordi- nate system V ¼ ei � ej � ek ¼ ei � e123 e2e3eið Þ which will cancel itself only if i ¼ 1, whereby V ¼ e123 e1ð Þ2e2e3 ¼ e1ð Þ2e2e3 ) e1ð Þ2 ¼ V e2e3 h 1 e 2 e 3e 1e 2e 3e Fig. 1.3 Parallelepiped defined by means of the reciprocal basis 1.3 Tensors 5 www.Ebook777.com http://www.ebook777.com and for the reciprocal basis e1 � � 2 ¼ V e1ð Þ 2 ¼ e2e3 V V ¼ ei � ej � ek ¼ ei � e123e2e3ei � � ¼ e1� � 2e2e3 e1 ¼ 1 e1 V ¼ 1 e1 � � e2e3 ¼ e2e3 V � � e2e3 ¼ 1 V ) VV ¼ 1e1 ¼ 1 e1 For a dextrorotatory coordinate system V ¼ e2 � ej � e3 ¼ e2 � e123 eie3e2ð Þe1 ¼ 1 e1 which cancels itself for i ¼ 1, so V ¼ e2 � �eie3e2ð Þ ¼ e2ð Þ2e1e3e1 ¼ 1 e1 1 e2ð Þ2 ¼ � e1e3 V e1 ¼ 1 e1 and for the case of reciprocal basis V ¼ e2 � ei � e3 ¼ e2 e123 eie3e2 � � ¼ � e2� �2e1e3 e2 ¼ � 1 e2 In an analogous way VV ¼ 1 If e1, e3, e2 are the unit vectors of an orthogonal coordinate system, then the reciprocal basis e1, e2, e3 also defines this coordinate system. 1.3.3.1 Orthonormal Basis If the basis is orthonormal ei � ej � ek ¼ V ¼ V ¼ 1 ei ¼ ej � ek ¼ ei uj ¼ ui 6 1 Review of Fundamental Topics About Tensors This shows that for the Cartesian vectors, it is indifferent, covariant, or contravariant, of which the basis is adopted. The vector components in terms of this basis are equal, and the orthonormal basis is defined by their unit vectors ei ¼ ui uik k The linear transformations 8m, u, v2E3: (a) F muð Þ ¼ mF uð Þ; (b)F u � vð Þ ¼ F mð Þ defined in the Euclidean space E3 are also defined in the vectorial space E � 3, for there is an intrinsic correspondence between these two spaces. The rules of calculus in E�3 are analogous to those of E3, so these parameters are isomorphous. The existence of this duality is extended to the case of a vectorial space of finite dimension E�N , having E � N �ℜ or E*N � Z, for this space is dual to the Euclidean space EN. 1.3.3.2 Transformation Law of Vectors The transformation of the coordinates from one point in the coordinate system Xi to the coordinate system X i is given by xi ¼ ∂xi∂xj xj, where ∂x i ∂xj ¼ cos αij are the matrix rotation elements, and its terms are the director cosines of the angles between the coordinate axes. In this linear and homogeneous transformation, the points of the space E3 are transformed into points expressed in terms of the new coordinates. Thus, the unit vectors ofX i and ofX i transform according to the law ei ¼ ∂xj∂xi ej, where the values of ∂xi ∂xj ¼ cos xixj are the components of the unit vectors ēi in the coordinate system Xi. For the position vector, u provides ui ¼ ∂xj∂xi uj. In the case of the inverse transfor- mation, i.e., of X i to Xi, provides analogously ej ¼ ∂xi∂xj ei, following for the components ∂x j ∂xi ¼ cos xjxi of the unit vectors ej in the coordinate system X i . The determinant of the rotation matrix ∂x i ∂xj assumes the value þ1 in the case of the transformation taking place between coordinate systems of the same direction, which is then called proper transformation (rotation). Otherwise ∂x i ∂xj ¼ �1, and the transformation is called improper transformation (reflection). 1.3.3.3 Covariant and Contravariant Vectors The representation of the vectors in oblique coordinate systems highlights various characteristics which are more general than the Cartesian representation. In these systems the vectors are expressed by means of two kinds of components. Let the representation of vector u in the plane coordinate system of oblique axes OXiXj that 1.3 Tensors 7 form an angle α, with basis vectors ei, ej (Fig. 1.4). The contravariant components are obtained by means of straight lines parallel to the axes OXi and OXj and graphed, respectively, as ui, uj (indicated with upper indexes). The covariant componentsare obtained by means of projection on the axes OXi and OXj given, respectively, by ui, uj (indicated with lower indexes). The projection of vector u on an axis provides its component on this axis, and by means of the dot product of u ¼ uiei and ej: u � ej ¼ uiei � ej ¼ ui ei � ej � � ¼ uiδij ¼ ui that is the contravariant component of vector and in the same way by the covariant component u � ej ¼ uiei � ej ¼ ui ei � ej � � ¼ uiδij ¼ ui Thus, the vector is defined by its components u ¼ uiei ¼ uiei These components are not, in general, equal, and in the case of α ¼ 90o (Cartesian coordinate systems), the equality of these components is verified. j j u e O O j X j X i X i X j X j e j u i X j e i e i u ie j e i e i e i u a b Fig. 1.4 Vector components: (a) contravariant, (b) covariant 8 1 Review of Fundamental Topics About Tensors 1.3.3.4 Transformation Law of Covariant Vectors The transformation law of base ei of an axis OX i for a new axis OX j , with base ēj (Fig. 1.5a), is given by ej ¼ projei ej �� ��� �ei ¼ 1 cos αð Þei cos α ¼ ∂x i ∂xj ej ¼ ∂x i ∂xj ei thatis the transformation law of the covariant basis. For the vector u the transfor- mation of its covariant components is given by uj ¼ ∂xi∂xj ui, where the variables relative to the original axis in relation to which the transformation performed are found in the numerator of the equation. 1.3.3.5 Transformation Law of Contravariant Vectors The projection of the vector ei on the axis OX j (Fig. 1.5b) provides ej ¼ projej eik k � � ei ¼ 1 cos αð Þei cos α ¼ ∂x j ∂xi ej ¼ ∂x j ∂xi ei that is the transformation law of the contravariant basis. For the vector u follows the transformation law of its contravariant components uj ¼ ∂xj∂xi ui, where the variables relative to the new axis, for which the transformation is carried out, are found in the numerator of the expression. ix∂ ix∂ O O i e i e j e j e jx∂jx∂ ie e j pro j a b Fig. 1.5 Transformation of coordinates: (a) covariant, (b) contravariant 1.3 Tensors 9 Free ebooks ==> www.Ebook777.com 1.3.4 Multilinear Forms The tensors of the order p are multilinear forms, which are vectorial functions, and linear in each variable considered separately. The concept of tensor is conceived by means of the following approaches: (a) the tensor is a variety that obeys a trans- formation law when changing the coordinate system; (b) this variety is invariant for any coordinate system; and (c) there is an equivalence between these definitions (equivalence law). A tensor of the order p is defined by a multilinear function with Np components in the space EN, where R 1� N2 � � ¼ 0 represents its order, which is maintained invariant if a change of the coordinate system occurs, and on the rotation of the reference axes (linear and homogeneous transformation) its coordi- nates modify according to a certain law. Consider the space EN and the coordinate system X i, i ¼ 1, 2, 3, . . .N, defined in this space, where there are N equations that relate the coordinates of the points in EN, given by continuously differentiable functions xi ¼ xi xj� � i, j ¼ 1, 2, 3, . . .N ð1:3:4Þ that transform these functions to a new coordinate systemX i . These transformations of coordinates require only that N functions xi(xj) be independent. The necessary and sufficient condition for this transformation to be possible is that J ¼ ∂xi ∂xj 6¼ 0. The inverse function has an inverse Jacobian J ¼ ∂xj∂xi and implies that JJ ¼ 1. 1.3.4.1 Transformation Law of the Second-Order Tensors Let the position vector ui(x i) expressed in the coordinate system Xi of base ei and a new coordinate system X i , with same origin, in which the vector is expressed by ui x ið Þ. Consider the elements ∂xk ∂xi of the rotation matrix that relates the coordinates of these two systems, then follow by means of the transformation law of covariant vectors ui ¼ ∂x k ∂xi uk i, k ¼ 1, 2, 3 ð1:3:5Þ vj ¼ ∂x ‘ ∂xj v‘ j, ‘ ¼ 1, 2, 3 ð1:3:6Þ The vectors �ui(x i) and vi x ið Þ define the transformation of the second-order tensor in terms of its covariant components 10 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com Tij ¼ uivj ¼ ∂x k ∂xi ∂x‘ ∂xj ukv‘ ¼ ∂x k ∂xi ∂x‘ ∂xj Tk‘ ð1:3:7Þ and for the contravariant components provides an analogous manner T ij ¼ uivj ¼ ∂x i ∂xk ∂xj ∂x‘ ukv‘ ¼ ∂x i ∂xk ∂xj ∂x‘ Tk‘ ð1:3:8Þ In a same way, it follows for the transformation law in terms of the mixed components T i j ¼ uivj ¼ ∂xi ∂xk ∂x‘ ∂xj ukv‘ ¼ ∂x i ∂xk ∂x‘ ∂xj T k‘ ð1:3:9Þ 1.3.4.2 Transformation Law of the Third-Order Tensors The transformations of the vectors u, v, w in terms of their covariant components are given by u‘ ¼ ∂x i ∂x‘ ui vm ¼ ∂x j ∂xm vj wn ¼ ∂x k ∂xn wk following by substitution T‘mn ¼ u‘ vm wn ¼ ∂x i ∂x‘ ∂xj ∂xm ∂xk ∂xn uivjwk that leads to the following transformation law for the covariant components of the third-order tensors T‘mn ¼ ∂x i ∂x‘ ∂xj ∂xm ∂xk ∂xn Tijk and for the contravariant components T ‘mn ¼ ∂x ‘ ∂xi ∂xm ∂xj ∂xn ∂xk Tijk and in an analogous way, for the mixed components T mn ‘ ¼ ∂x‘ ∂xi ∂xm ∂xj ∂xn ∂xk Tjki T n ‘m ¼ ∂x‘ ∂xi ∂xm ∂xj ∂xn ∂xk T kij T m ‘n ¼ ∂x‘ ∂xi ∂xm ∂xj ∂xn ∂xk T jik T ‘m n ¼ ∂xi ∂x‘ ∂xm ∂xj ∂xn ∂xk Tijk 1.3 Tensors 11 1.3.4.3 Inverse Transformation Let the inverse transformation of the vectors u and v of the coordinate system X i for the coordinate system Xi, given by the covariant components of the vectors ui ¼ ∂x k ∂xi uk vj ¼ ∂x ‘ ∂xj v‘ ð1:3:10Þ It follows that Tij ¼ uivj ¼ ∂x k ∂xi ∂x‘ ∂xj ukv‘ ¼ ∂x k ∂xi ∂x‘ ∂xj Tk‘ ð1:3:11Þ Expression (1.3.11) allows concluding that a second-order tensor can be interpreted as a transformation in the linear space E3, which associates the vector u to the vector v by means of the tensorial product and that this linear and homogeneous transformation has an inverse transformation. The inverse transfor- mations are defined for the contravariant and mixed components in an analogous way. Expressions (1.3.7) and (1.3.11) show that if the components of a second- order tensor are null in a coordinates system, they will cancel each other in any other coordinate system. For the definition of the transformation law of second- order tensor to be valid, it is necessary that the transitive property apply to the linear operators (Fig. 1.6). 1.3.4.4 Transitive Property Let a second-order tensor Tk‘ defined in the coordinate system X i, that is expressed in the coordinate system X i by means of the expression (1.3.7), and with the transformation of X i for eXi iX iX iX�k T � � kji k ij Tx x x x T ¶ ¶ ¶ ¶ = ijq j p i pq Tx~ x x~ x T ~ ¶ ¶ ¶ ¶ = � � kjp k pq Tx x x~ x T ~ ¶ ¶ ¶ ¶ = Fig. 1.6 Transitive property of the second-order tensors 12 1 Review of Fundamental Topics About Tensors eTpq ¼ ∂xi∂exp ∂xj∂exq Tij ð1:3:12Þ However, the tensor eTpq can be expressed in terms of tensor Tk‘, thereby avoiding the intermediary transformation, so substituting expression (1.3.7) in expression (1.3.12), it follows that eTpq ¼ ∂xk ∂xi ∂x‘ ∂xj ∂xi ∂exp ∂xj∂exq Tk‘ ð1:3:13Þ and simplifying ∂xk ∂xi ∂xi ∂exp ¼ ∂xk∂exp ∂x‘∂xj ∂xj∂exq ¼ ∂x‘∂xj ð1:3:14Þ Then eTpq ¼ ∂xk∂exp ∂x‘∂xj Tk‘ ð1:3:15Þ Expression (1.3.15) is the transformation law of the second-order tensor of the coordinate system Xi for the coordinate system eXi, which proves that the transitive property applies to these tensors. This property is also valid when using the contravariant and mixed components. The tensors studied in this book belong to metric spaces. If a variety is a tensor with respect to the linear transformations, it will be a tensor with respect to all the orthogonal linear transformations, but the inverse usually does not occur. The tensors are produced in spaces more general than the vectorial space. Table 1.1 shows the covariant, contravariant, and mixed tensors and their transformation laws for the space EN. 1.3.4.5 Multiplication of a Tensor by a Scalar It is the multiplication that provides a new tensor as a result, which components are the components of the original tensor multiplied by the scalar. Let the tensor Tijk Table 1.1 Kinds of tensors Tensor Expression Transformation law Covariant Tij���k Trs���t ¼ ∂x i ∂xr ∂xj ∂xs � � �∂x k ∂xt Tij���k Contravariant Tij���k T rs���t ¼ ∂x r ∂xi ∂xs ∂xj � � � ∂x t ∂xp Tij���k Mixed Tk‘���hij���f Tmn���hrs���t ¼ ∂xi ∂xr ∂xj ∂xs ∂xm ∂xk ∂xn ∂x‘ � � �∂x f ∂xt ∂xh ∂xh Tk‘���hij���f 1.3 Tensors 13 and the scalar m which product Pijk is given by Pijk ¼ mTijk. For demonstrating this expression represents a tensor, all that is needed is to apply the tensor transforma- tion law to the same. 1.3.4.6 Addition and Subtraction of Tensors The addition of tensors of the same order and the same type is given by T kij ¼ Akij þ Bkij The addition of the mixed tensors given by the previous expression provides as a result a mixed tensor of the third order, which is twice covariant and oncecontravariant. To demonstrate this expression represents a tensor, all that is needed is to apply the tensor transformation law to the same. The subtraction is defined in the same way as the addition, however, admitting that a tensor is multiplied by the scalar�1. As an exampleT kij ¼ Akij þ �1ð ÞBkij; thus, this expression provides as a result a mixed tensor of the third order, which is twice covariant and once contravariant. To demonstrate that previous expression represents a tensor all that is needed is to carry out the analysis developed for the addition considering the negative sign. 1.3.4.7 Contraction of Tensors The contraction of a tensor is carried out when two of its indexes are made equal, a covariant index and a contravariant index, and thus reducing the order of this tensor in two. For instance, the tensor Tk‘ij contracted in the indexes ‘ and j results as Tkji‘ ¼ Tkjij ¼ T ki . 1.3.4.8 Outer Product of Tensors The outer product is the product of two tensors that provide as a new tensor, which order is the sum of the order of these two tensors. Let, for example, the tensor Ak ...ij ... with variance index number ( p, q) and the tensor B... ‘m... rs with variance index number (u, v), which if multiplied provides a tensor Tk...‘mij...rs ¼ Ak...ij...B...‘m...rs with variance index number pþ u, qþ vð Þ. The order of the tensor is given by the sum of these two indexes. To demonstrate that the previous expression is a tensor, all that is needed is to apply the tensor transformation law to the same. 14 1 Review of Fundamental Topics About Tensors Free ebooks ==> www.Ebook777.com 1.3.4.9 Inner Product of Tensors The inner product of two tensors is defined as the tensor obtained after the contracting of the outer product of these tensors. Let, for example, tensors Aij and B‘k which the outer product is P ‘ ijk ¼ AijB ‘k that provides as a result a tensor of the fourth order, which contracted in the indexes ‘ and k provide the inner product P ‘ij‘ ¼ AijB ‘‘ ¼ Pij. This shows that the resulting tensor is of the second order. To demonstrate that this expression represents a tensor, all that is needed is to apply the tensor transformation law to the same. 1.3.4.10 Quotient Law This law allows verifying if a group of Np functions of the coordinates of the referential system Xi has tensorial characteristics. Its application serves to test if a variety is a tensor. The systematic for applying this law is to make the dot product of the variety that is to be tested by a vector, for the outer product of two tensors generates a tensor, and then carry out the contraction of this product and afterward, by means of applying the tensor transformation law, verify if the variety fulfills this law. Let, for example, the contravariant tensor of the first order Tk and the variety A(i, j, k) composed of 27 functions defined in the space EN, for which it is desired to verify if it is tensor. The fundamental premise is that the vector Tk is independent of A(i, j, k). If the inner product A i, j, kð ÞTk ¼ Bij originates a contravariant tensor of the second order, then A(i, j, k) has the characteristics of a tensor. Applying the transformation law of tensors to the tensor Bij B pq ¼ ∂x p ∂xi ∂xq ∂xj Bij ¼ ∂x p ∂xi ∂xq ∂xj A i; j; kð ÞTk and for the vector Tk, it follows that Tk ¼ ∂x k ∂xr T r By substitution B pq ¼ ∂x p ∂xi ∂xq ∂xj ∂xk ∂xr A i, j, kð ÞTr and in a new coordinate system, the tensor Bij is given by B pq ¼ A p, q, rð ÞTr 1.3 Tensors 15 www.Ebook777.com http://www.ebook777.com following by substitution A p, q, rð Þ � ∂x p ∂xi ∂xq ∂xj ∂xk ∂xr A i, j, kð Þ � T r ¼ 0 As T r is an arbitrary vector the result is A p; q; rð Þ ¼ ∂x p ∂xi ∂xq ∂xj ∂xk ∂xr A i; j; kð Þ that represents the transformation law of third-order tensors. This shows that the variety A(i, j, k) has tensorial characteristics. 1.4 Homogeneous Spaces and Isotropic Spaces The isotropic space has properties which do not depend on the orientation being considered, and the components of isotropic tensors do not change on an orthogonal linear transformation. The sum of isotropic tensors results in an isotropic tensor, and the product of isotropic tensors is also an isotropic tensor. There is no isotropic tensor of the first order. The isotropic tensor of the fourth order is given by Tijk‘ ¼ λδijδk‘ þ μδikδj‘ þ νδi‘δjk ð1:4:1Þ where λ, μ, ν are scalars. The Kronecker delta δij is the only isotropic tensor of the second order. The homogeneous space has properties which are independent of the position of the point. The homogeneous tensors have constant components when the coordinate system is changed. A homogeneous tensor of the fourth order is given by Tij‘k ¼ λδijδk‘ þ μ δikδj‘ þ δi‘δjk � � ð1:4:2Þ where λ, μ are scalars. 1.5 Metric Tensor The study of tensors carried out in affine spaces applies to another type, called metric space, in which the length of the curves is determined by means of a variety that defines this space, in which the basic magnitudes are the length of a curve and the vector’s norm, just as the angle between vectors and the angle between two curves. The distinction between these two types of spaces is of fundamental importance in the study of tensors. 16 1 Review of Fundamental Topics About Tensors The metric space is determined by the definition of its fundamental tensor which is related with its intrinsic properties. The conception of this metric tensor, which gives an arithmetic form to the space, considers the invariance of distance between two points, the concept of distance being acquired from the space E3. The geometry grounded in the concept of metric tensor is called Riemann geometry. The angle between two curves is calculated by means of the dot product between vectors using the metric tensor, which awards a generalization to this tensor’s formulation. Let the arc element length of a curve ds defined in the Cartesian coordinate system Xi with unit vectors g1, g2, g3 by means of its coordinates x i (Fig. 1.7), with two neighboring pointsP xið Þ , Q xi þ dxið Þ, which define the position vectors r and rþ dr, respectively. The coordinates of increment of the position vector dr are given byQ� P ¼ dxi; thus, lim Q!P Q� Pð Þ ¼ ds, and the dot product of this vector by himself takes the form ds2 ¼ dr � dr ¼ dxidxi ð1:5:1Þ Consider a transformation of the coordinates xi ¼ xi xið Þ for a new coordinate system X i dxi ¼ ∂x i ∂xk dxk ð1:5:2Þ becomes ds2 ¼ ∂x i ∂xk ∂xi ∂x‘ dxkdx‘ ð1:5:3Þ Putting gk‘ ¼ ∂xi ∂xk ∂xi ∂x‘ ð1:5:4Þ 2 X 1 X 3 X ii d xxQ + i xP d s O 1 g 2 g 3 g Fig. 1.7 Elementary arc of a curve 1.5 Metric Tensor 17 thus the metric takes the form ds2 ¼ gk‘dxkdx‘ ð1:5:5Þ The symmetry of the variety given by expression (1.5.4) is obvious, because gk‘ ¼ g‘k, then gij ¼ gji ¼ g1g1 g1g2 g1g3 g2g1 g2g2 g2g3 g3g1 g3g2 g3g3 24 35 ¼ g11 g12 g13g21 g22 g23 g31 g32 g33 24 35 ð1:5:6Þ The analysis of expression (1.5.4) shows that gk‘ relates with the Jacobian J½ � ¼ ∂xi ∂xk h i of a linear transformation by means of the following expression gk‘½ � ¼ ∂xi ∂xk �T ∂xi ∂x‘ � ¼ J½ �T J½ � ð1:5:7Þ For the coordinate system Xi, the variety gij is defined by his unit vectors gi, gj. Consider a new coordinate system X i , with respect to which these unit vectors are expressed by gk ¼ ∂xi ∂xk gi g‘ ¼ ∂xj ∂x‘ gj ð1:5:8Þ Thus gk‘ ¼ ∂xi ∂xk gi � � ∂xj ∂x‘ gj � � ¼ ∂x i ∂xk ∂xj ∂x‘ gigj � � ¼ ∂xi ∂xk ∂xj ∂x‘ gij then gij is a symmetric tensor of the second order. The arc length is invariable when changing the coordinate system. The coeffi- cients of gk‘(x i) are class C2, and the N equations xi ¼ xi xið Þ must satisfy the 1 2 N N þ 1ð Þ partial differential equations given by expression (1.5.4). However, if gk‘(x i) is specified arbitrarily, this system of 1 2 N N þ 1ð Þ partial differential equa- tions, in general, has no solution. The fundamental tensor gk‘ related toa coordinate system Xi, in a region of the space EN, must fulfill the following conditions: (a) gk‘(x i) is a class C2 function, i.e., its second-order derivatives exist and are continuous. (b) Be symmetrical, i.e., gk‘ ¼ g‘k. (c) detgk‘ ¼ g 6¼ 0, i.e., gk‘ is not singular. (d) ds2 ¼ gk‘dxkdx‘ is an invariant after a change of coordinate system. 18 1 Review of Fundamental Topics About Tensors Expression (1.5.5) is put under parametric form with the coordinates xi ¼ xi tð Þ and i ¼ 1, 2, 3 . . .N, and the parameter a t b provides s ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi gk‘ dxk dt dx‘ dt ���� ���� s dt ð1:5:9Þ Admit a functional parameter hi ¼ 1, so as to allow the conditions gk‘ dx k dt dx‘ dt > 0 and gk‘ dxk dt dx‘ dt < 0 to be be used instead of the absolute value shown in expression (1.5.9), because the use of hi is more adequate to the algebraic manipulations; thus, s ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higk‘ dxk dt dx‘ dt r dt ð1:5:10Þ The quadratic and homogeneous form Φ ¼ gk‘dxkdx‘ is called metric or funda- mental form of the space, being invariant after a change of coordinate system. In space E3 with Φ > 0, which provides g > 0, and when Φ ¼ 0, the initial and final points of the arc coincide. If Φ ¼ 0 and dxi are not all null, the displacement between the two points is null. The possibility of Φ being undefined is admitted, for instance, in the case Φ ¼ dx1ð Þ 2 � dx2ð Þ 2, for which dx1 ¼ dx2 results in Φ ¼ 0. This case is interpreted as having a null displacement of the point. If dxi 6¼ 0, i.e., the displacements are not null, hi is adopted so that hiΦ > 0. The spaces EN (hyperspaces) are analyzed in an analogous way to the analysis of the space E3 by means of defining a metric, formalizing the Riemann geometry. The geometries not grounded on the concept of metric are called non-Riemann geometries. To demonstrate that expression (1.5.10) is invariant through a change in its parametric representation, let a curve of class C2 represented by means of the coordinates xi ¼ xi tð Þ and a t b. Consider a transformation for the new coordinates xi ¼ xi tð Þ and a t b, where t ¼ f tð Þ with f 0 tð Þ > 0, and in the new limits a ¼ f að Þ, b ¼ f bð Þ. Applying the chain rule to the function t ¼ f tð Þ: dt dt ¼ f 0 tð Þ ) dt ¼ dt f 0 tð Þ ð1:5:11Þ and with expression (1.5.11) in expression (1.5.10) 1.5 Metric Tensor 19 Free ebooks ==> www.Ebook777.com L¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higij dxi dt dxj dt r dt ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higij dxi dt dxj dt f 0 tð Þ h i2r dt ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higij dxi dt dxj dt r f 0 tð Þdt ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higij dxi dt dxj dt r dt ¼ L then the value of this expression does not vary with the change of the curve’s parameterization. The metric can be written in matrix form so as to make the usual calculations easier ds dt � �2 ¼ dx k dt �T gij h i dx‘ dt � ð1:5:12Þ In the space E3, the metric is defined by dt dt ¼ f 0 tð Þ ) dt ¼ dt f 0 tð Þ ð1:5:13Þ ds2 ¼ g11dx1dx1 þ g12dx1dx2 þ g13dx1dx3 þ g21dx2dx1 þ g22dx2dx2 þ g23dx2dx3 þ g31dx3dx1 þ g32dx3dx2 þ g33dx3dx3 ð1:5:14Þ or ds2 ¼ gii dxi � �2 þ gkk dxk� �2 þ 2gikdxidxk ð1:5:15Þ For the particular case in which the coordinate systems are orthogonal (Fig. 1.8), the segments on the coordinate axes Xi are defined by the unit vectors gi of these axes ds ið Þ ¼ gidxi ð1:5:16Þ which provide the metric ds2 ¼ h1g1dx1 � �2 þ h2g2dx2� �2 þ h3g3dx3� �2 ¼ h1dx1 � �2 þ h2dx2� �2 þ h3dx3� �2 ð1:5:17Þ then the metric tensor is defined by the elements of the diagonal of the matrix 20 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com gij ¼ h21 0 0 0 h22 0 0 0 h23 24 35 ð1:5:18Þ where h1 ¼ ffiffiffiffiffiffig11p , h2 ¼ ffiffiffiffiffiffig22p , h3 ¼ ffiffiffiffiffiffig33p , and detgij ¼ g ¼ g11g22g33. Exercise 1.1 Let gijx ixj ¼ 0, 8xi, xj show that gk‘ þ g‘k ¼ 0. Putting Φ ¼ gijxixj ¼ 0 and differentiating with respect to xk ∂Φ ∂xk ¼ gij ∂xi ∂xk xj þ gijxi ∂xj ∂xk ¼ gijδ ikxj þ gijδ jkxi ¼ gkjxj þ gikxi ¼ 0 Differentiating with respect to x‘ ∂2Φ ∂xk∂x‘ ¼ gkj ∂xj ∂x‘ þ gik ∂xi ∂x‘ ¼ gkjδ j‘ þ gikδ i‘ ¼ gk‘ þ g‘k ¼ 0 Q:E:D: Exercise 1.2 Calculate the length of the curve of class C2 given by the parametric equations x1¼ 3� t, x2 ¼ 6tþ 3, and x3 ¼ ‘n t, in the space defined by the metric tensor 1 g 2 g 3 g 1X 2X 3X O Fig. 1.8 Orthogonal coordinate systems 1.5 Metric Tensor 21 gij ¼ 12 4 0 4 1 1 0 1 x1ð Þ2 24 35 The metric of the space in matrix form stays ds dt � �2 ¼ dx k dt �T gij h i dx‘ dt � and with the derivatives dx1 dt ¼ �1 dx 2 dt ¼ 6 dx 3 dt ¼ 1 t it follows ds dt � �2 ¼ �1; 6; 1 t � 12 4 0 4 1 1 0 1 3� tð Þ2 24 35 �16 1 t 8><>: 9>=>; ¼ tþ 3ð Þ 2 t2 Making h1 ¼ 1 in expression s ¼ ðb a ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higk‘ dxk dt dx‘ dt r dt ) L ¼ ðe 1 tþ 3 t � � dt ¼ eþ 2 1.5.1 Conjugated Tensor Let the increment of the position vector expressed by means of their covariant components dr ¼ rigi in any coordinate system, where gj is the basis vector of this referential and with the dot product dr � dr ¼ dxidxj gigj � � ð1:5:19Þ and gigj � � ¼ gij ¼ gji ð1:5:20Þ whose symmetry comes from the commutative property of the dot product. 22 1 Review of Fundamental Topics About Tensors This variety with properties analogous to the properties of the metric tensor is represented by nine components of a symmetrical matrix 3� 3, which form a second-order contravariant tensor. It is called conjugated metric tensor; thus, gij ¼ g11 g12 g13 g21 g22 g23 g31 g32 g33 24 35 ð1:5:21Þ The definition of the conjugated of the metric tensor is given by gij ¼ gigj ð1:5:22Þ and with the relations between the reciprocal basis gi ¼ gk � g‘ V gj ¼ gm � gn V ð1:5:23Þ results for the conjugated metric tensor gij ¼ 1 V2 gk � g‘ð Þ gm � gnð Þ ð1:5:24Þ but with the fundamental formula of the vectorial algebra gk � g‘ð Þ � gm � gnð Þ ¼ gk � g‘ð Þ � gm½ � � gn ð1:5:25Þ and developing the double-cross product in brackets gk � g‘ð Þ � gm½ � � gn ¼ gk � gmð Þg‘ � g‘ � gmð Þgk ð1:5:26Þ So gij ¼ 1 V2 gk � gmð Þ g‘ � gnð Þ � g‘ � gmð Þ gk � gnð Þ½ � ð1:5:27Þ The term in brackets in expression (1.5.27) is the development of the determinant Gij ¼ gk � gm gk � gn g‘ � gm g‘ � gn ¼ gkm gkng‘m g‘n ð1:5:28Þ Then gij ¼ Gij V2 ð1:5:29Þ 1.5 Metric Tensor 23 Summarizing these analyses by means of the transcription of the following expressions gij ¼ Gij V 2 ¼ Gij g gij ¼ G ij V2 ¼ G ij g ð1:5:30Þ Thus V ¼ ffiffiffigp ¼ ffiffiffiffiffiffiffiffiffiffiffidetgijp V ¼ ffiffiffigp ¼ ffiffiffiffiffiffiffiffiffiffiffiffidetgijq ð1:5:31Þ The sign þð Þ in expressions (1.5.31) corresponds to a levorotatory coordinates, and the sign �ð Þcorresponds to a dextrorotatory coordinates. Knowing thatVV ¼ 1, it follows that gg ¼ 1. Exercise 1.3 Let detgij x nð Þ ¼ g xnð Þ. Calculate the derivative ∂g∂xn, n ¼ 1, 2, . . .. The matrix linked to the determinant g is a function of the variables xn: gij ¼ gij xi � � and this determinant being a function of the matrix elements g ¼ g gij � � by the chain rule ∂g ∂xn ¼ ∂g ∂gij ∂gij ∂xn As det g is expressed by its cofactors g ¼ g1kGk1 ¼ g11G11 þ g12G21 þ g13G31 þ � � � and the terms Gk1 do not contain the terms g1k, so ∂g ∂g11 ¼ G11 ∂g∂g12 ¼ G21 ∂g∂g13 ¼ G31 � � � Generalizing provides ∂g ∂gij ¼ Gji 24 1 Review of Fundamental Topics About Tensors By substitution ∂g ∂xn ¼ Gji ∂gij ∂xn Exercise 1.4 Calculate the derivative of detg ¼ x 1x2 x1ð Þ2 x1ðÞ2 2x1 with respect to the variable x1. From Exercise 1.3 ∂g ∂xi ¼ Gji ∂gij ∂xi This expression is the sum of n determinants. Each of these determinants differs from the determinant g only in the lines and columnswhich are being differentiated, so ∂g ∂x1 ¼ ∂g11 ∂x1 ∂g12 ∂x1 g21 g22 þ g11 g12 ∂g21 ∂x1 ∂g22 ∂x1 ¼ x 2 2x1 x1ð Þ 2 2x1 þ x1x2 x1ð Þ 2 2x1 2 Exercise 1.5 Let g ¼ detgij the determinant of the metric tensor gij and xk an arbitrary variable. Calculate (a) ∂ ‘n gð Þ∂gij and (b) ∂ ‘n gð Þ ∂xk . (a) From Exercise 1.3 ∂g ∂gij ¼ Gji but as gij ¼ gji it follows that ∂g ∂gji ¼ Gij Expression (1.5.30) provides gij ¼ G ij g gij ¼ Gij g ) Gij ¼ ggij By substitution ∂g ∂gji ¼ ggij ¼ ggij 1.5 Metric Tensor 25 whereby ∂ ‘ngð Þ ∂gij ¼ 1 g ∂g ∂gij ) ∂ ‘ngð Þ ∂gij ¼ gij (b) By the chain rule ∂ ‘ngð Þ ∂xk ¼ ∂ ‘ngð Þ ∂gij ∂gij ∂xk and substituting the result obtained in the previous item in this expression ∂ ‘ngð Þ ∂xk ¼ gij ∂gij ∂xk Exercise 1.6 Calculate the metric tensor, its conjugated tensor, and the metric for the Cartesian coordinate system. Let the Cartesian coordinates (x1, x2, x3), and by the definition of the distance between two points ds2 ¼ dx1� �2 þ dx2� �2 þ dx3� �2 which is the square of the metric, thus ds2 ¼ δijdxidxj By the definition of the metric tensor and the conjugated metric tensor, then gij ¼ δij ¼ 1 0 0 0 1 0 0 0 1 24 35 gij ¼ 1 gij ¼ 1 0 0 0 1 0 0 0 1 24 35 Exercise 1.7 Calculate the metric tensor, its conjugated tensor, and the metric for the cylindrical coordinate system given by r � x1, θ � x2; and z � x3 where �1 r 1, 0 θ 2π, and �1 z 1, which relations with the Cartesian coordinates are x1 ¼ x1 cos x2, x2 ¼ x1 sin x2, and x3 � x3. 26 1 Review of Fundamental Topics About Tensors With the definition of metric tensor gij ¼ ∂xk ∂xi ∂xk ∂xj ¼ ∂x 1 ∂xi ∂x1 ∂xj þ ∂x 2 ∂xi ∂x2 ∂xj þ ∂x 3 ∂xi ∂x3 ∂xj – i ¼ j ¼ 1 g11 ¼ ∂x1 ∂x1 ∂x1 ∂x1 þ ∂x 2 ∂x1 ∂x2 ∂x1 þ ∂x 3 ∂x1 ∂x3 ∂x1 ¼ cos x2� �2 þ sin x2� �2 þ 0 ¼ 1 – i ¼ j ¼ 2 g22 ¼ ∂x1 ∂x2 ∂x1 ∂x2 þ ∂x 2 ∂x2 ∂x2 ∂x2 þ ∂x 3 ∂x2 ∂x3 ∂x2 ¼ �x1 sin x2� �2 þ x1 cos x2� �2 þ 0 ¼ x1� �2 – i ¼ j ¼ 3 g33 ¼ ∂x1 ∂x3 ∂x1 ∂x3 þ ∂x 2 ∂x3 ∂x2 ∂x3 þ ∂x 3 ∂x3 ∂x3 ∂x3 ¼ 0þ 0þ 1 ¼ 1 – i ¼ 1 , j ¼ 2 g12 ¼ ∂x1 ∂x1 ∂x1 ∂x2 þ ∂x 2 ∂x1 ∂x2 ∂x2 þ ∂x 3 ∂x1 ∂x3 ∂x2 ¼ cos x2 �x1 sin x2� �2 þ sin x2 x1 cos x2� �2 þ 0 ¼ 0 For the other terms g21 ¼ g13 ¼ g31 ¼ g23 ¼ g32 ¼ 0, then ds2 ¼ g11dx1dx1 þ g22dx2dx2 þ g11dx2dx2 ¼ dx1 � �2 þ x1� �2 dx2� �2 þ dx3� �2 ¼ drð Þ2 þ rdθð Þ2 þ dzð Þ2 The metric tensor and its conjugated tensor are given, respectively, by gij ¼ 1 0 0 0 r2 0 0 0 1 264 375 gij ¼ 1 gij ¼ 1 0 0 0 1 r2 0 0 0 1 264 375 and with the base vectors gi ¼ ∂xj ∂xi ej 1.5 Metric Tensor 27 i ¼ 1 ) g1 ¼ ∂xj ∂x1 ej j ¼ 1, 2, 3 ) g1 ¼ ∂x1 ∂x1 e1 þ ∂x 2 ∂x1 e2 ∂x3 ∂x1 e3 g1 ¼ cos x2e1 þ sin x2e2 8>>>>><>>>>>: i ¼ 2 ) g2 ¼ ∂xj ∂x2 ej j ¼ 1, 2, 3 ) g2 ¼ ∂x1 ∂x2 e1 þ ∂x 2 ∂x2 e2 ∂x3 ∂x2 e3 g2 ¼ �x1 sin x2e1 þ x1 cos x2e2 8>>>>><>>>>>: i ¼ 3 ) g3 ¼ ∂xj ∂x3 ej j ¼ 1, 2, 3 ) g3 ¼ ∂x1 ∂x3 e1 þ ∂x 2 ∂x3 e2 ∂x3 ∂x3 e3 g3 ¼ 0þ 0þ 1 � e3 ¼ e3 8>>>>><>>>>>: By means of the dot products gi � gj ¼ δij ei � ej ¼ δij it follows for the components of the metric tensor g11 ¼ g1 � g1 ¼ cos x2e1 þ sin x2e2 � � � cos x2e1 þ sin x2e2� � ¼ 1 g22 ¼ g2 � g2 ¼ �x1 sin x2e1 þ x1 cos x2e2 � � � �x1 sin x2e1 þ x1 cos x2e2� � ¼ x2� �2 g33 ¼ g3 � g3 ¼ e3ð Þ � e3ð Þ ¼ 1 The other components of this tensor are null. Exercise 1.8 Calculate the metric tensor, its conjugated tensor, and the metric for the spherical coordinate system r � x1,φ � x2, θ � x3, �1 r 1, and 0 φ π, where 0 θ 2π, which relations with the Cartesian coordinates are x1 ¼ x1 sin x2 cos x3, x2 ¼ x1 sin x2 sin x3,and x3 � x1 cos x2. With the definition of metric tensor gij ¼ ∂xk ∂xi ∂xk ∂xj ) gij ¼ ∂x1 ∂xi ∂x1 ∂xj þ ∂x 2 ∂xi ∂x2 ∂xj þ ∂x 3 ∂xi ∂x3 ∂xj – i ¼ j ¼ 1 28 1 Review of Fundamental Topics About Tensors g11 ¼ ∂x1 ∂x1 ∂x1 ∂x1 þ ∂x 2 ∂x1 ∂x2 ∂x1 þ ∂x 3 ∂x1 ∂x3 ∂x1 ¼ sin x2 cos x3� � 2 þ sin x2 sin x3� � 2 þ cos x2� � 2 ¼ 1 – i ¼ j ¼ 2 g22 ¼ ∂x1 ∂x2 ∂x1 ∂x2 þ ∂x 2 ∂x2 ∂x2 ∂x2 þ ∂x 3 ∂x2 ∂x3 ∂x2 ¼ x1 cos x2 cos x3� � 2 þ x1 cos x2 sin x3� � 2 þ �x1 sin x2� � 2 ¼ x1� � 2 – i ¼ j ¼ 3 g33 ¼ ∂x1 ∂x3 ∂x1 ∂x3 þ ∂x 2 ∂x3 ∂x2 ∂x3 þ ∂x 3 ∂x3 ∂x3 ∂x3 ¼ �x1 sin x2 sin x3� �2 þ x1 sin x2 cos x3� �2 þ 0 ¼ x1 sin x2� �2 For the other terms g12 ¼ g21 ¼ g13 ¼ g31 ¼ g23 ¼ g32 ¼ 0, then ds2 ¼ g11dx1dx1 þ g22dx2dx2 þ g33dx3dx3 ¼ dx1 � �2 þ x3� �2 dx2� �2 þ x1 sin x2� �2 dx3� �2 ¼ drð Þ2 þ rdφð Þ2 þ r sin θdθð Þ2 The metric tensor and its conjugated tensor are given, respectively, by gij ¼ 1 0 0 0 r2 0 0 0 r2 sin 2φ 264 375 gij ¼ 1 gij ¼ 1 0 0 0 1 r2 0 0 0 1 r2 sin 2φ 266664 377775 Exercise 1.9 Calculate the metric tensor, its conjugated tensor, and the metric for the cylindrical elliptical coordinate system ξ � x1, η � x2, and z � x3, where ξ � 0, 0 η 2π, �1 z 1, which relations with the Cartesian coordinates are x1 ¼ coshx2 cos x2, x2 ¼ sinhx2 sin x2, x3 � x3. With x3 ¼ const:, the elliptical cylinder is x10 ¼ const:: x1 chx10 � �2 þ x 2 shx10 � �2 ¼ cos x2� �2 þ sin x2� �2 ¼ 1 dx1 ¼ sinhx1 cos x2dx1 dx2 ¼ coshx1 sin x2dx1 ds¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi dx1ð Þ2 þ dx2ð Þ2 q ¼ cosh2x1 � cos 2x2� �dx1 g11 ¼ cosh2x1 � cos 2x2 � � 1.5 Metric Tensor 29 Free ebooks ==> www.Ebook777.com With x1 ¼ const: the hyperbolic cylinder is x20 ¼ const:: x1 cos x20 � �2 � x 2 sin x20 � �2 ¼ coshx1� � 2 � sinhx1� � 2 ¼ 1 dx1 ¼ �coshx1 sin x2dx2 dx2 ¼ sinhx1 cos x2dx2 ds¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi dx1ð Þ2 þ dx2ð Þ2 q ¼ cosh2x1 � cos 2x2� �dx2 g22 ¼ cosh2x1 � cos 2x2 � � For x3 � x3 provides dx2 ¼ dx3, whereby g33 ¼ 1, following ds2 ¼ cosh2x1 � cos 2x2� �2 dx1� �2 þ cosh2x1 � cos 2x2� �2 dx2� �2 þ dx3� �2 The metric tensor and its conjugated tensor are given, respectively, by gij ¼ cosh2x1 � cos 2x2 0 0 0 cosh2x1 � cos 2x2 0 0 0 1 264 375 gij ¼ 1 cosh2x1 � cos 2x2 0 0 0 1 cosh2x1 � cos 2x2 0 0 0 1 26664 37775 1.5.2 Dot Product in Metric Spaces Let the vectors u and v contained in the metric space EN defined by the fundamental tensor gk‘. The dot product u � v with u ¼ uiei and v ¼ vjej depends only on the vectors and is independent of the coordinate system in relation to which the same is specified. It is observed that only when the coordinates of the vectors are covariant and contravariant, this product is like to the dot product in Cartesian coordinates. The dot product is invariant in view of the transformation of coordinates u � v¼ uiei � vjej ¼ gijuivj ¼ uiei:vjej ¼ gijuivj ¼ uiei:vjej ¼ gji uivj ¼ uivj ¼ uivj ð1:5:32Þ 30 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com 1.5.2.1 Vector Norm The generalization of the dot product of vectors for a metric space EN allows obtaining the norm of a vector. Let vector v with norm (modulus) vk k ¼ ffiffiffiffiffiffiffiffiv � vp ¼ ffiffiffiffiffiv2p that is equal to the distance between the extreme points, thus, with the expression of the metrics v2 ¼ higk‘vkv‘ results for the norm of the vector in terms of its contravariant components vk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higkkv kvk q In an analogous way for the covariant components vk vk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi higkkvkvk p and for Cartesian coordinates vk k ¼ ffiffiffiffiffiffiffiffi vkvk p If v is a unit vector, the expressions provide higkkv kvk ¼ 1 higkkvkvk ¼ 1 vkvk ¼ 1 The properties of the vectors norm are: (a) vk k � 0, which is a trivial property, for the norm will only cancel itself if v is null. (b) mvk k ¼ mk k vk k, where m is a scalar. (c) uþ vk kuk k þ vk k. (d) u � vk k uk k � vk k, Cauchy–Schwarz inequality. For the case of non-null vectors, the equality of the relation (d) exists only if u ¼ mv, where m is a scalar. Exercise 1.10 Calculate the modulus of vector u(1; 1; 0; 2) in space E4, defined by the metric tensor gij ¼ �1 0 0 0 0 �1 0 0 0 0 �1 0 0 0 0 c2 2664 3775 1.5 Metric Tensor 31 Free ebooks ==> www.Ebook777.com For the line element ds2 ¼ gijuiuj, and developing this expression ds2 ¼ gijuiuj ¼ g11u1u1 þ g22u2u2 þ g33u3u3 þ g44u4u4 ¼ �1� 1� 1� 1� 1� 1þ 0þ c2 � 2� 2 ¼ �2þ 4c2 ds ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2c2 � 1ð Þ p 1.5.2.2 Lowering of a Tensor’s Indexes By means of analysis referent to the transformation of the covariant components of the vector in their contravariant components, and vice versa, it is verified that inner product of a tensor by the metric tensor allows raising or lowering the indexes of this tensor. For multiplying the contravariant tensor of the first order, i.e., the contravariant vector Ti by the tensor gk‘, results in T i k‘ ¼ gk‘Ti, and for the contraction i ¼ ‘, then T iki ¼ gkiTi ¼ Tk that is a covariant vector. The index of the original vector was lowered and its order reduced in two units. 1.5.2.3 Raising of a Tensor’s Indexes Let the covariant vector Tk, which multiplied by g ik, provides as a result the tensor g ikTk, and changing the covariant coordinates of the vector by its contravariant coordinates gikTk ¼ gik gk‘T‘ � � ¼ δ i‘T‘ ¼ Ti that is a contravariant vector. The index of the original vector was raised. Then a covariant vector is obtained by means of the inner product of a contravariant vector, this indexes transformation process as being reciprocal. The vectors Ti and Ti are called associated vectors, and it refers to the contravariant and covariant compo- nents of the vector. For the case of second-order tensors, an analysis is carried out that is analogous to the one developed for the vectors. Let the covariant tensor of the second-order Tk‘ and its associated tensor Tij ¼ gikgj‘Tk‘. It is verified in the general case that these tensors are not conjugated tensors, for example, when Tij ¼ mgij, where m is a scalar, the tensor Tk‘ will be a multiple of gk‘, Tk‘ ¼ gikgj‘Tij ¼ gikgj‘ mgij � � ¼ mgikδ ‘i ¼ mg‘k ¼ mgk‘ 32 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com The raising and lowering operations of the indexes of tensors are carried out adopting, firstly, a point for indicating where the position to be left empty in the index that will be raised or lowered. For example, for the tensor Tji, the empty position is indicated by means of the notation Tj i, and in an equal manner A rs p exists for the tensor Arsp . Let the inner product of the tensor Ti jk by the metric tensor g‘i, g‘iT i jk ¼ T‘jk in which the upper index was lowered, and gijT kj ¼ T k i that had an index lowered, or further, gijgk‘T j‘ ¼ Tik, which two upper indexes were lowered. For raising the indexes, in an analogous way to the raising of an index gijTjk ¼ T i k or gkjTij ¼ T ki , thus gijgk‘Tj‘ ¼ Tik. In the case in which the index is lowered and then raised, the original tensor is obtained gkjT ij ¼ T i k and next gkjT i k ¼ Tij. 1.5.2.4 Tensorial Equation If a term of a tensorial equation contains a dummy index, it can be raised or lowered, i.e., change the position without changing the value of the equation. The following example illustrates this assertion Ai jBi ¼ gkiAkj � � gi‘B ‘ � � ¼ gkigi‘AkjB‘ ¼ δ k‘ AkjB‘ ¼ A‘jB‘ ¼ AijBi where the index i was lowered in one tensor and raised in the tensor. If a free index is a part of the tensorial expression, a new tensorial expression equivalent to this one can be obtained, lowering or raising this index in the members of the original expression. To illustrate this assertion, the following tensorial equation is admitted Tijk ¼ AijBk which is equivalent to gi‘T‘jk ¼ gi‘A‘jBk so it results in T i jk ¼ Ai jBk where the index i was raised. 1.5 Metric Tensor 33 Exercise 1.11 Raise and lower the indexes of vector u, for the metric tensor and its conjugated tensor: (a) uj ¼ 3 4 5 8<: 9=; gij ¼ 1 0 0 0 x1ð Þ2 0 0 0 x1 sin x2ð Þ2 24 35 (b) uj ¼ 5 4 3 8<: 9=; gij ¼ 1 0 0 0 x1ð Þ�2 0 0 0 x1 sin x2ð Þ�2 24 35 (a) Carrying out the following matrix multiplication provides the covariant com- ponents of the vector ui ¼ gijuj ¼ 1 0 0 0 x1ð Þ2 0 0 0 x1 sin x2ð Þ2 24 35 34 5 8<: 9=; ¼ 3 4 x1ð Þ2 5 x1 sin x2ð Þ2 8<: 9=; (b) Carrying out the following matrix multiplication provides the contravariant components of the vector ui ¼ gijuj ¼ 1 0 0 0 x1ð Þ�2 0 0 0 x1 sin x2ð Þ�2 24 35 54 3 8<: 9=; ¼ 5 4 x1ð Þ�2 3 x1 sin x2ð Þ�2 8<: 9=; Exercise 1.12 Given the covariant basis g1 ¼ e1; g2 ¼ e1 þ e2; g3 ¼ e3 and the tensor of the space gij ¼ 1 1 1 1 2 2 1 2 3 24 35, calculate the vectors of the contravariant basis and the conjugated metric tensor. The determinant of the metric tensor is given by g ¼ detgi ¼ 1 1 1 1 2 2 1 2 3 ¼ 1 which indicates that the system is dextrorotary. For the vectors of the contravariant basis, it follows that g1 ¼ g2 � g3 g ¼ e1 e2 e3 1 1 0 1 1 1 ¼ e1 � e2 g2 ¼ g3 � g1 g ¼ e1 e2 e3 1 1 1 1 0 0 ¼ e2 � e3 34 1 Review of Fundamental Topics About Tensors g3 ¼ g1 � g2 g ¼ e1 e2 e3 1 0 0 1 1 0 ¼ e3 then the conjugated metric tensor is given by gij ¼ gi � gj ¼ 2 �1 0 �1 2 �1 0 �1 1 24 35 The verification of the operation is carried out by means of the expression gijg ij ¼ δ ij , thus 1 1 1 1 2 2 1 2 3 24 35 2 �1 0�1 2 �1 0 �1 1 24 35 ¼ 1 0 00 1 0 0 0 1 24 35 1.5.2.5 Associated Tensors The metric tensor gij and its conjugated tensor g ij relate intrinsically to each other, which allows using them for analyzing the relations between the covariant and contravariant components of the vector u ui ¼ gijuj ui ¼ gijuj ð1:5:33Þ These expressions generate two linear equation systems, which unknown quan- tities are u1, u2, u3, and u 1, u2, u3. The solution of the system given by means of Cramer’s rule and with the determinant of the metric tensor g ¼ detgij ¼ g11 g12 g13 g21 g22 g23 g31 g32 g33 and its cofactor Gij ¼ gkm gkn g‘m g‘n thus ui ¼ gijuj ¼ Giju j g ) gij ¼ Gij g 1.5 Metric Tensor 35 In an analogous way for the system given by expression 1.5.33 ui ¼ gijuj ¼ G ijuj g ) gij ¼ G ijuj g The linear operators gij, g ij allow relating the covariant and contravariant com- ponents of vector u. Defining this vector by means of its covariant components and performing the dot product of this vector by the basis unit vectors gi: u � gi ¼ ujgj � � � gi ¼ uj gj � gi� � ¼ gijuj ¼ ui that are the contravariant components of vector u. In an analogous way for the transformation of the contravariant components in the covariant components u � gi ¼ uigj � � � gi ¼ ui gj � gi� � ¼ gijuj ¼ ui These expressions relate to each other in a kind of coordinate as a function of the other, where the tensors gij, g ij are the operators responsible for these transformations. The covariant components and the contravariant components of the vector u are given, respectively, by ui ¼ u1 ¼ g11u1 u2 ¼ g22u2 u3 ¼ g33u3 8<: ui ¼ u 1 ¼ g11u1 u2 ¼ g22u2 u3 ¼ g33u3 8<: ) g11g 11 ¼ 1 g22g 22 ¼ 1 g33g 33 ¼ 1 8<: The linear operators gij, g ij, gji are useful in the explanation of the more general properties of tensors. Exercise 1.13 For the vector v ¼ 4g1 þ 3g2 referenced to a coordinate system, calculate their contravariant and covariant components in the referential system that have the basis vectors g1 ¼ 3g1, g2 ¼ 6g1 þ 8g2, and g3 ¼ g3. The metric tensor of the space is given by gij ¼ gi � gj ¼ g1g1 g1g2 g1g3 g2g1 g2g2 g2g3 g3g1 g3g2 g3g3 24 35 ¼ 9 18 018 100 0 0 0 1 24 35) detgij ¼ 576 For the conjugated metric tensor gijg ij ¼ δij ) gij ¼ gij h i�1 ) gij ¼ 1 576 100 �180 �18 9 0 0 0 576 24 35 36 1 Review of Fundamental Topics About Tensors The vectors of the contravariant basis are given by gi ¼ gijgj gi ¼ g1 g2 g3 8<: 9=; ¼ 1576 100 �18 0 �18 9 0 0 0 576 24 35 3g16g1 þ 8g2 g3 8<: 9=; ¼ 1 3 g1 � 1 4 g2 1 8 g2 g3 8>>><>>>: 9>>>=>>>; With the contravariant components of v vi ¼ v � gi v1 ¼ 4g1 þ 3g2ð Þ � g1 ¼ 7 12 v2 ¼ 4g1 þ 3g2ð Þ � g2 ¼ 3 8 v3 ¼ 4g1 þ 3g2ð Þ � g3 ¼ 0 so v ¼ 7 12 g1 þ 3 8 g2 With the covariant components of v vi ¼ v � gi ¼ 4g1 þ 3g2ð Þ � gi v1 ¼ 4g1 þ 3g2ð Þ:g1 ¼ 12 v2 ¼ 4g1 þ 3g2ð Þ:g2 ¼ 48 v3 ¼ 4g1 þ 3g2ð Þ � g3 ¼ 0 so v ¼ 12g1 þ 48g2 Exercise 1.14 Show that in the space EN exists g‘jgik � g‘igjk � � g‘j ¼ N � 1ð Þgik, where gij is the metric tensor. Developing the given expression g‘jgik � g‘igjk � � g‘j ¼ g‘jgikg‘j � g‘igjkg‘j ¼ g‘jg‘jgik � g‘ig‘jgjk ¼ g‘jg‘jgik � g‘iδ ‘k ¼ g‘jg‘jgik � gki ¼ δ jj gik � gki as δ jj ¼ δ11 þ δ22 þ � � � þ δ nn ¼ N for the space EN, the result is g‘jgik � g‘igjk � � g‘j ¼ Ngik � gki 1.5 Metric Tensor 37 but gik ¼ gki; thus, g‘jgik � g‘igjk � � g‘j ¼ N � 1ð Þgik Q:E:D: Exercise 1.15 Show that in space EN exists g ij ∂gij ∂xk þ gij ∂g ij ∂xk ¼ 0, where gij is the metric tensor. The relation between the metric tensor and its conjugated tensor is given by gijg ij ¼ δ jj ¼ δ11 þ δ22 þ � � � þ δnn ¼ N Differentiating this expression with respect to xk ∂ gijg ij � � ∂xk ¼ ∂gij ∂xk gij þ gij ∂ gijð Þ ∂xk ¼ ∂N ∂xk ¼ 0 so gij ∂gij ∂xk þ gij ∂gij ∂xk ¼ 0 Q:E:D: Exercise 1.16 For the symmetric tensor Tij, that fulfills the condition gijT‘k� gi‘Tjk þ gjkT‘i � gk‘Tij ¼ 0, show that Tij ¼ mgij, where m 6¼ 0 is a scalar. Multiplying the expression given by gij follows gijgijT‘k � gijgi‘Tjk þ gijgjkT‘i � gijgk‘Tij ¼ δ ii T‘k � δ ‘j Tjk þ δ ikT‘i � gijgk‘Tij ¼ 0 As δ jj ¼ δ11 þ δ22 þ � � � þ δnn ¼ N, and for j ¼ ‘ and i ¼ k NTji ¼ gijgijTij As ds2 ¼ gijTij ¼ m1, where m1 is a scalar, and with Tij ¼ Tji follows NTji ¼ m1gij ) Tji ¼ m1 N gij Putting m ¼ m1N Tij ¼ mgij Q:E:D: 38 1 Review of Fundamental Topics About Tensors 1.6 Angle Between Curves The angle between two curves is defined by the angle formed by their tangent unit vectors g1, g2 (Fig. 1.9), by means of the dot product cos α ¼ g1 � g2 g1k k g2k k ¼ g1 � g2 In differential terms this angle is calculated supposing that in the space E3 two curves intersect in a point R, and admitting a third curve that intersects the other two at points A1 and A2, which distances from the point R are, ds(1) and ds(2) (Fig. 1.9). The points M,A1,A2 have coordinates x i, xi þ dx i 1ð Þ and x i þ dx i 2ð Þ, respectively. With the cosine law cos α ¼ lim RA1ð Þ 2 þ RA2ð Þ2 � A1A2ð Þ2 2 RA1ð Þ RA2ð Þ which in differential terms stays cos α ¼ ds 1ð Þ � �2 þ ds 2ð Þ� �2 � ds 3ð Þ� �2 2ds 1ð Þds 2ð Þ and using the basic expressions for the length of the arcs of the curves ds 1ð Þ � �2 ¼ gijdx i1ð Þdxj1ð Þ ds 2ð Þ� �2 ¼ gijdx i2ð Þdxj2ð Þ ds 3ð Þ � �2 ¼ gij xi þ dxi1ð Þ� �� xi þ dx i2ð Þ� �h i2 ¼ gij dx i1ð Þ � dxi2ð Þ� �2 ¼ gij dx i2ð Þ � dx i1ð Þ � � dx j 2ð Þ � dxj1ð Þ � � 1T 2 T 1C 2C R ( )1ds ( )2ds 1A 2A 3C ( )3ds Fig. 1.9 Angle between two curves 1.6 Angle Between Curves 39 Free ebooks ==> www.Ebook777.com then cos α ¼ gij dx i 1ð Þdx j 2ð Þ þ dxj2ð Þdxi1ð Þ � � 2ds 1ð Þds 2ð Þ ¼ gijdx i 1ð Þdx j 2ð Þ ds 1ð Þds 2ð Þ Considering ui ¼ dx i 1ð Þ ds 1ð Þ and vj ¼ dx j 2ð Þ ds 2ð Þ , which are, respectively, the contravariant unit vectors of the tangents T1 and T2 to the curves C1 and C2, respectively, provides cos α ¼ gij dx i 1ð Þ ds 1ð Þ ! dxj 2ð Þ ds 2ð Þ ! If two vectors are orthogonal, then α ¼ π 2 , so the condition of orthogonality for two directions is giju ivj ¼ 0. The necessary and sufficient condition so that a coordinate system is orthonormal is that gij ¼ 0 8i 6¼ j at the points of this space. The null vector has the peculiar characteristic of being normal to itself. Figure 1.10 illustrates the components of the differential element of arc ds with respect to the coordinate system Xi with origin at point P. The lengths of the arc elements measured with respect to the coordinate axes of the referential system are ds 1ð Þ ¼ ffiffiffiffiffiffig11p dx1, ds 2ð Þ ¼ ffiffiffiffiffiffig22p dx2, and ds 3ð Þ ¼ ffiffiffiffiffiffig33p dx3. To prove that α is real and that cos α 1, consider the expression cos α ¼ gijuivj ¼ gijuivj ¼ uivj ¼ uivj where u, v are unit vectors. Admit that these vectors are multiplied by two non-null real numbers ‘,m, originating ‘ui þ mvið Þ, as the metric of the space is positive definite, then for all the values of this pair of numbers gij ‘u i þ mvi� � ‘uj þ mvj� � � 0 1X 2X 3X 2 2 22 xdgds = ( ) 1 111 xdgds = ( ) 3 333 xdgds = P 1 2a 13 a 2 3a ds Fig. 1.10 Components of the differential arc element with respect to the Xi coordinate system 40 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com Developing this inequality ‘2 þ 2‘m cos αþ m2 � 0 for uivj ¼ uivj ¼ cos α, which can be written under the form ‘þ m cos αð Þ2 þ m2 1� cos 2α� � � 0 that will be positive definite if cos 2α 1 or cos αk k 1, so α is real. Let the modulus of a vector in terms of their contravariant components v ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi εgk‘v kv‘ q ð1:6:1Þ and in terms of their covariant components v ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi εgkkvkvk p ð1:6:2Þ thus the angle between two curves is determined when calculating the angle between their tangent unit vector ui, vj, then cos α ¼ giju ivjffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi uiuið Þ vjvj � �q ð1:6:3Þ Exercise 1.17 Let the orthogonal unit vectors ui and vj, calculate the norm of vector wi ¼ ui þ vi. The condition of orthogonality between two vectors is given by giju ivj ¼ 0 and as ui and vj are unit vectors giju iuj ¼ 1 gijvivj ¼ 1 For the vector wi wk k2 ¼ gijwiwj ¼ gij ui þ vjð Þ uj þ við Þ ¼ gijuiuj þ gijuivi þ gijvjuj þ gijvjvi ¼ 1þ 0þ 0þ 1 ¼ 2 then wk k ¼ ffiffiffi 2 p 1.6 Angle Between Curves 41 Exercise 1.18 The vectors ui and vj are orthogonal, and each one of them has modulus ‘, show that gpjgki � gpkgji � � upvjukvi ¼ ‘4. The square of the modulus of the vectors is given by giju iuj ¼ ‘2 gijvivj ¼ ‘2 and the condition of orthogonality between these vectors is given by giju ivj ¼ 0 Developing the given expression gpjgki � gpkgji � � upvjukvi ¼ gpjgkiupvjukvi � gpkgjiupvjukvi ¼ gpjupvj � � gkiu kvi � gpkupuk � � gjiv jvi ¼ 0� ‘2 � ‘2 so gpjgki � gpkgji � � upvjukvi ¼ ‘4 Q:E:D: Exercise 1.19 Given the symmetric tensor Tij and the unit vectors u i and vj orthogonal to the vector wk, show that Tiju i � m1gijui þ n1gijwi ¼ 0 and Tijvi� m2gijv i þ n2gijwi ¼ 0, where m1 6¼ m2 and n1 6¼ n2 are scalars, then these unit vectors are orthogonal. As ui and vj are unit vectors, then giju iuj ¼ 1 gijvivj ¼ 1 and the conditions of orthogonality of these unit vectors with respect to the vector wi are giju iwj ¼ 0 gijviwj ¼ 0 Multiplying by vj both the members of the first expression Tiju ivj � m1gijuivj þ n1gijwivj ¼ 0 ) Tijuivj ¼ m1gijuivj and multiplying by uj both the members of the second expression Tijv iuj � m2gijviuj þ n2gijwiuj ¼ 0 ) Tijviuj ¼ m2gijviuj 42 1 Review of Fundamental Topics About Tensors Free ebooks ==> www.Ebook777.com The indexes i and j are dummies, so their position can be changed Tiju ivj ¼ m2gijuivj thus m1giju ivj ¼ m2gijuivj ) m1 � m2ð Þgijuivj ¼ 0 As by hypothesis m1 6¼ m2, then gijuivj ¼ 0; this shows that the unit vectors ui and vj are orthogonal. 1.6.1 Symmetrical and Antisymmetrical Tensors If the change of position of two indexes, covariant or contravariant, does not modify the tensor’s components, then this is a symmetrical tensor T pqrsijk ¼ T pqrsikj ¼ T pqrsjik ¼ T pqrsjki ¼ T pqrskij ¼ T pqrskji . The symmetry, a priori, does not ensure that the new variety is a tensor.Admit that T pqrsijk‘ ¼ T qprsijk‘ , whereby by the hypothesis of this tensor’s symmetry, it follows that T pqrsijk‘ � T qprsijk‘ ¼ 0. As Tpqrsijk‘ is a tensor, the result of the difference between the two varieties being null, and as the referential system is arbitrary, it is concluded that this result will always be null for any coordinate system, i.e., it always has the tensor null. WritingT pqrsijk‘ þ 0 ¼ T qprsijk‘ , and as the summation of tensors is a tensor, it is concluded that Tpqrsijk‘ is a tensor. A tensor is called antisymmetrical with respect to two of its indexes, if it changes signs on the change of position between these two indexes: Tijk‘ ¼ �T‘jki. The number of independent components of an antisymmetric tensor of order p in the space EN is given by n ¼ N! p! N � pð Þ ! ð1:7:1Þ Let the space EN in which the antisymmetric pseudotensor of the third order ε ijk is defined (a general definition of pseudotensors will be presented in item 1.8), and by the definition of antisymmetry, it provides six components of εijk which are numerically equal: εijk ¼ εjki ¼ εkij ¼ �εikj ¼ �εjik ¼ �εkji This variety has 27 components, having 21 null, for it is verified that only the six components ε123 ¼ ε231 ¼ ε312 ¼ �ε132 ¼ �ε213 ¼ �ε321 are non-null. Let, for example, a linear and homogeneous transformation be applied to the component ε123: 1.6 Angle Between Curves 43 www.Ebook777.com http://www.ebook777.com ε123 ¼ ∂x 1 ∂xi ∂x2 ∂xj ∂x3 ∂xk ε123 ð1:7:2Þ Developing expression (1.7.2) ε123 ¼ ∂x 1 ∂x1 ∂x2 ∂x2 ∂x3 ∂x3 þ ∂x 1 ∂x2 ∂x2 ∂x3 ∂x3 ∂x1 þ ∂x 1 ∂x3 ∂x2 ∂x1 ∂x3 ∂x2 � �∂x 1 ∂x1 ∂x2 ∂x3 ∂x3 ∂x2 � ∂x 1 ∂x3 ∂x2 ∂x2 ∂x3 ∂x1 � ∂x 1 ∂x2 ∂x2 ∂x1 ∂x3 ∂x3 � ε123 ð1:7:3Þ In compact form for the component ε123 in the coordinate system X i ε123 ¼ ∂x k ∂x‘ ε123 ð1:7:4Þ and with gk‘ ¼ ∂xi ∂xk ∂xj ∂x‘ gij It follows by means of product of determinants gk‘j j ¼ ∂xi ∂xk � ∂xj∂x‘ � gij ) gk‘j j ¼ ∂xi∂xk 2 gij ) g ¼ ∂xi∂xk 2g 1ffiffiffi g p ¼ ∂x i ∂xk 1ffiffiffigp ð1:7:5Þ Comparing expression (1.7.5) with the expression (1.7.4) ε123 ¼ 1ffiffiffi g p ð1:7:6Þ So as to generalize expression (1.7.6), this analysis is made for the other contravariant components of the pseudotensor εijk. As this variety assumes the values 0, 1 as a function of the position of their indexes, it is linked to the permutation symbol eijk by means of the following relations: εijk ¼ þ1ffiffiffi g p eijk is an even permutation of the indexes �1ffiffiffi g p eijk is an odd permutation of the indexes 0 when there are repeated indexes 8>>>><>>>>: ð1:7:7Þ 44 1 Review of Fundamental Topics About Tensors Expression (1.7.7) represents the components of the Ricci pseudotensor, also called Levi-Civita pseudotensor. The covariant components of this pseudotensor are obtained by means of the metric tensor, whereby using the approaches presented in item 1.5, it is provided for the lowering of the indexes of the pseudotensor εpqr: εijk ¼ gipgjqgkrεpqr ð1:7:8Þ and with the definition of the determinant of the metric tensor, and with the definition of εijk given by the relations (1.7.7), it follows that εijk ¼ gij 1ffiffiffi g p ¼ ffiffiffigp ð1:7:9Þ In terms of the permutation symbol eijk, it is provided as the covariant coordi- nates of the Ricci pseudotensor εijk ¼ ffiffiffi g p eijk is an even permutation of the indexes � ffiffiffigp eijk is an odd permutation of the indexes 0 when there are repeated indexes 8<: ð1:7:10Þ The definition of the Ricci pseudotensor presented for the space E3 is general- ized for the space EN, in which the contravariant components and covariant of this variety are given, respectively, in terms of the permutation symbol by εi1i2i3���in ¼ þ1ffiffiffi g p ei1i2i3���in is an even permutation of the indexes �1ffiffiffi g p ei1i2i3���in is an odd permutation of the indexes 0 when there are repeated indexes 8>>>><>>>>: ð1:7:11Þ εi1i2i3���in ¼ ffiffiffi g p ei1i2i3���in is an even permutation of the indexes � ffiffiffigp ei1i2i3���in is an odd permutation of the indexes 0 when there are repeated indexes 8<: ð1:7:12Þ The conception of permutation symbol is associated to the value of a determi- nant, with no link to the space EN, whereby it refers only to a symbol that seeks to simplify the calculations. With the definition of the Ricci pseudotensor in terms of this symbol, it is verified that in the relation between these two varieties exists the term ffiffiffi g p linked to the metric of the space. This shows the fundamental difference between the same, for the change of sign of the Ricci pseudotensor as a function of the permutations of their indexes (sign defined by the permutation symbol) indi- cates the orientation of the space. With relation (1.7.10) it follows that εijkεjki ¼ 3! ¼ 6 ð1:7:13Þ 1.6 Angle Between Curves 45 The definitions and deductions presented next seek to complement the relations between the generalized Kronecker delta and the Ricci pseudotensor in the space EN. These expressions are called δ� ε relations. 1.6.1.1 Generalization of the Kronecker Delta The Ricci pseudotensor represents the mixed product of three vectors ∂x i ∂x‘ , ∂xj ∂x‘ , ∂xk ∂x‘, where ‘ ¼ 1, 2, 3 indicates the components of these vectors, which comprise the lines and columns of the determinant that expresses this product, called Gram determinant, that in terms of their covariant components stays εijk ¼ ∂x i ∂x‘ � ∂x j ∂x‘ � � � ∂x k ∂x‘ ¼ ∂xi ∂x1 ∂xj ∂x1 ∂xk ∂x1 ∂xi ∂x2 ∂xj ∂x2 ∂xk ∂x2 ∂xi ∂x3 ∂xj ∂x3 ∂xk ∂x3 ¼ δ1i δ1j δ1k δ2i δ2j δ2k δ3i δ3j δ3k and in terms of their contravariant components εpqr ¼ ∂xp ∂x1 ∂xq ∂x1 ∂xr ∂x1 ∂xp ∂x2 ∂xq ∂x2 ∂xr ∂x2 ∂xp ∂x3 ∂xq ∂x3 ∂xr ∂x3 ¼ δ1p δ1q δ1r δ2p δ2q δ2r δ3p δ3q δ3r The product of these two determinants being given by εijkε pqr ¼ δ1i δ1j δ1k δ2i δ2j δ2k δ3i δ3j δ3k � δ1p δ1q δ1r δ2p δ2q δ2r δ3p δ3q δ3r εijkε pqr ¼ δ1iδ 1p þ δ2iδ2p þ δ3iδ3p � � δ1iδ 1q þ δ2iδ2q þ δ3iδ3q � � δ1iδ 1r þ δ2iδ2r þ δ3iδ3r � � δ1jδ 1p þ δ2jδ2p þ δ3jδ3p � � δ1jδ 1q þ δ2jδ2q þ δ3jδ3q � � δ1jδ 1r þ δ2jδ2r þ δ3jδ3r � � δ1kδ 1p þ δ2kδ2p þ δ3kδ3p � � δ1kδ 1q þ δ2kδ2q þ δ3kδ3q � � δ1kδ 1r þ δ2kδ2r þ δ3kδ3r � � εijkε pqr ¼ δmiδ mp δmiδ mq δmiδ mr δmjδ mp δmjδ mq δmjδ mr δmkδ mp δmkδ mq δmkδ mr εijkε pqr ¼ δpi δ q i δ r i δpj δ q j δ r j δpk δ q k δ r k ð1:7:14Þ 46 1 Review of Fundamental Topics About Tensors With the expressions (1.7.10) and (1.7.7), it follows that εr‘mε rst ¼ δrstr‘m ¼ δst‘m ¼ 1 0 �1 8<: ð1:7:15Þ The contraction of the indexes k and r of the product of two pseudotensors, given by expression (1.7.15), provides εijkε pqr ¼ δpqkijk ¼ δ pi δ q i δ k i δ pj δ q j δ k j δ pk δ q k δ k k and as δ kk ¼ 3 it follows that εijkε pqr ¼ δpqkijk ¼ δ pi δ q i δ k i δ pj δ q j δ k j δ pk δ q k 3 ¼ δ p i δ q i δpj δ q j ¼ δpi δqj � δ qi δpj ð1:7:16Þ Analogously, and with the contraction of the indexes j and p: εijkε pqr ¼ δjqrijk ¼ δ ji δ q i δ k i 3 δqj δ k j δ jk δ q k δ k k ¼ � δ q i δ r i δqk δ r k ¼ δ ri δqk � δqi δ rk The product εijrεpqr ¼ δpqij leads to the generalization of the Kronecker delta that has its value defined as a function of the number of permutations of their indexes. For the covariant components of this operator, δijpq ¼ δpijq is provided, where the number of permutations of the indexes is even, so it is verified that this operator is symmetrical, and δijpq ¼ �δjipq ¼ �δijqp is antisymmetric for an odd number of permutations of the indexes. The deltas with repeated indexes are null, for example, δ11pq ¼ δ22pq ¼ δij33 ¼ 0. This analysis allows defining the generalized Kronecker delta inspace EN: δ j 1 j2j3���jn i1i2i3���in ¼ þ1 is an even permutation of i1i2i3� � �, j1j2j3� � � �1 is an odd permutation of i1i2i3� � �, j1j2j3� � � 0 when there are repeated indexes 8<: ð1:7:17Þ 1.6.1.2 Fundamental Expressions with the Generalized Kronecker Delta The generalized Kronecker delta in terms of the Ricci pseudotensor is given by 1.6 Angle Between Curves 47 εi1i2i3���imε j 1 j2j3���jm ¼ δj1 ���jmjmþ1���jni1���imimþ1���in ¼ δj1i1 δj1i2 � � � δj1in δj2i1 δj2i2 � � � δj2in � � � � � � � � � � � � δjni1 δjni2 δj1i1 δjnin ð1:7:18Þ Various fundamental expressions are obtained with the Kronecker delta δpqij that are useful in Tensor Calculus. Let, for example, the contraction of the indexes j and q of this tensor δpqij ¼ δpjij ¼ δpi δ jj � δ qi δ jj ¼ δpi δ jj � δ ji δpj ¼ 3δpi � δ ji δ pj ¼ 2δ pi whereby δpi ¼ 1 2 δpjij ¼ 1 2 δp1i1 þ δp2i2 þ δp3i3 � � It is also verified for the contractions j ¼ q and k ¼ r δpqrijk ¼ 1 2 δpjkijk ¼ δpi ¼ 1 2 δp12i12 þ δp23i23 þ δp31i31 � � The generalization of these expressions that involve Kronecker deltas for the space EN is given by the following expression: δ p 1 p2p3���pm i1i2i3���im ¼ N � nð Þ ! N � mð Þ ! δ p 1 ���pmpmþ1���pn i1���imimþ1���in ð1:7:19Þ The Kronecker delta tensor δ p 1 p2p3���pm i1i2i3���im provided by expression (1.7.19) is of order 2 n� mð Þ inferior to the order Kronecker delta tensor δp1 ���pmpmþ1���pni1���imimþ1���in , from which it was obtained by means of contractions of the indexes. In this expression for m ¼ 1, n ¼ 3 δpi ¼ 1 N � 2ð Þ N � 1ð Þ δ pjk ijk Putting m ¼ 1, n ¼ 2 in expression (1.7.19) results in δ pi ¼ 1 N � 1ð Þ δ pj ij These two examples show that δpi can be obtained by two contractions of the indexes of the sixth-order tensor δpqrijk or by means of only a contraction of the indexes of the fourth-order tensor δpqij . 48 1 Review of Fundamental Topics About Tensors In expression (1.7.19) for m ¼ 1, i ¼ p δi1i1 ¼ N � nð Þ ! N � 1ð Þ ! δ i1i2���in i1i2���in and as δi1i1 ¼ n it results in n ¼ N � nð Þ ! N � 1ð Þ ! δ i1i2���in i1i2���in ) δi1i2���ini1i2���in ¼ n N � 1ð Þ ! N � nð Þ ! δi1i2���ini1i2���in ¼ n ! N � nð Þ ! ð1:7:20Þ For the inner product of the Ricci pseudotensors εi1i2���in and ε i1i2���in with N ¼ n expression (1.7.20) provides εi1i2���in ε i1i2���in ¼ δi1i2���ini1i2���in ¼ n! ð1:7:21Þ Expression (1.7.19) with N ¼ n provides εi1i2���imimþ1���in ε p1p2���pmpmþ1���pn ¼ N � mð Þ !δp1p2���pni1i2���in ð1:7:22Þ Expression (1.7.22) relates in space EN the inner product of two Ricci pseudotensors with the generalized Kronecker delta tensor. 1.6.1.3 Product of the Ricci Pseudotensor by the Generalized Kronecker Delta The definition of the generalized Kronecker delta shows that δ pijkq123 ¼ 0, for a dummy index will always occur when these vary. With expression (1.7.18) εq123ε pijk ¼ δ pijkq123 ¼ δpq δp1 δp2 δp3 δiq δi1 δi2 δi3 δjq δj1 δj2 δj3 δkq δk1 δk2 δk3 ¼ 0 Developing this determinant in terms of the first column δpqεijkε123 � δiqεpjkε123 þ δjqεpikε123 � δkqεpijε123 ¼ 0 and with εpik ¼ �εipk εijkδpq ¼ εpjkδiq þ εipkδjq þ εijpδkq ð1:7:23Þ 1.6 Angle Between Curves 49 Free ebooks ==> www.Ebook777.com The symmetry of δpq allows changing the position of these indexes εijkδqp ¼ εqjkδip þ εiqkδjp þ εijqδkp ð1:7:24Þ 1.6.1.4 Norm of the Antisymmetric Pseudotensor of the Second Order A vector is represented by an oriented segment of a straight line, and its norm is given by the length of this segment. For an antisymmetric pseudotensor of the second order A associated to an axial vector u provides that its norm is linked to the area of the parallelogram which sides are the vectors that define the vectorial product u ¼ v� w. Let α the angle between the vectors v and w, the square of the modulus of the cross product of these vectors is given by uk k2 ¼ v� wk k2 ¼ vk k2 wk k2 sin 2α ¼ vk k2 wk k2 1� cos 2α� � ¼ vk k2 wk k2 � v � wð Þ2 thus uk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi vk k2 wk k2 � v � wð Þ2 q ð1:7:25Þ This norm can be expressed in terms of the components of the pseudotensor A. Let the components of the vectors v and w in the coordinate system Xk, so with the expression (1.7.25) uk k2 ¼ gi‘viv‘ � � gjmw jwm � � � gimviwmð Þ gj‘vjw‘ � � ¼ gi‘gjm � gimgj‘ � � viv‘wjwm ¼ gi‘ gim gj‘ gjm viv‘wjwm This determinant allows writing gi‘ gim gj‘ gjm viv‘wjwm ¼ 12 gi‘ gimgj‘ gjm viwj � vjwi� �v‘wm and as A‘m ¼ 1 2 v‘wm � vmw‘� � it follows that 50 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com gi‘ gim gj‘ gjm Aijv‘wm ¼ 12 gi‘ gimgj‘ gjm Aij v‘wm � vmw‘� � then gi‘ gim gj‘ gjm viw‘v‘wm ¼ 12 gi‘ gimgj‘ gjm AijA‘m ð1:7:26Þ 1.6.1.5 Generation of Tensors from the Ricci Pseudotensor The Ricci pseudotensor generates an antisymmetric tensor from a pseudotensor (axial vector), and this pseudotensor generates an antisymmetric tensor from a pseudotensor (axial vector). This characteristic of the Ricci pseudotensor in space E3 is generalized for the space EN, where the known antisymmetric tensor A i1i2���in½ � provides Tj1j2���jn�m ¼ 1 m! εj1j2���jn�mi1i2���imA i1i2���im½ � ð1:7:27Þ Tensor Tj1j2���jn�m is generated by the Ricci pseudotensor, which works as an operator applied to the antisymmetric tensor to produce this associated tensor. Multiplying both the members of expression (1.7.27) by εj1j2���jn�mi1i2���im results in εj1j2���jn�mi1i2���imT j1j2���jn�m ¼ 1 m! εj1j2���jn�mi1i2���imε j1j2���jn�mi1i2���imA i1i2���im½ � With expressions (1.7.19), (1.7.21), and (1.7.22), it follows that the expression for the antisymmetric tensor A i1i2���im½ � in terms of the Ricci pseudotensor is given by A i1i2���im½ � ¼ 1 n� mð Þ ! εj1j2���jn�mi1i2���imT j1j2���jn�m ð1:7:28Þ To illustrate the application of expression (1.7.28), let the antisymmetric tensor of the fourth order A[ijk‘] with i, j, k, ‘ ¼ 1, 2, 3, � � �n, to which the following five varieties are associated T ¼ 1 4! εijk‘A ijk‘½ � Ti ¼ 1 4! εijk‘pA jk‘p½ � Tij ¼ 1 4! εijk‘pqA k‘pq½ � Tijk ¼ 1 4! εijk‘pqrA ‘pqr½ � Tijk‘ ¼ 1 4! εijk‘pqrsA pqrs½ � 1.6 Angle Between Curves 51 Free ebooks ==> www.Ebook777.com 1.7 Relative Tensors The tensors defined in the previous items are called absolute tensors. However, other varieties with properties that are analogous to those of these tensors can be defined. The relative tensors are more general varieties, the absolute tensors being a particular case of the same. In solving various problems that involve integration processes, the need of generalizing the concept of tensor is verified. This generalization leads to the concept of relative tensor. To exemplify the concept of the relative tensor, let a covariant tensor of the second order be defined in the space E3, which transforms by means of the expression Tk‘ ¼ ∂x i ∂xk ∂xj ∂x‘ Tij ð1:8:1Þ The determinants of the terms of this function are given by detTk‘, det ∂xi ∂xk � � , det ∂x j ∂x‘ � � , and det Tij. Applying the determinant product rule to the determinant terms of this expression detTk‘ ¼ det ∂x i ∂xk � � det ∂xj ∂x‘ � � detTij ð1:8:2Þ the Jacobian of the inverse transformation of tensor Tk‘ is given by J ¼ det ∂x m ∂xn � � > 0 ð1:8:3Þ so detTk‘ ¼ J2detTij ð1:8:4Þ Expression (1.8.4) shows that detTk‘ of a second-order tensor is not a scalar and also is not a second-order tensor of the type Tpq. This expression is the new transformation. Assuming that detTij > 0 and detTk‘ > 0 provides detTk‘ � �1 2 ¼ J detTij � �1 2 ð1:8:5Þ This shows that the definition of tensors can be expanded introducing the concept of relative tensor. Consider the mixed tensor T ij...p rs...v ¼ Jð ÞW ∂xi ∂xa ∂xj ∂xb � � �∂x p ∂xd ∂xr ∂xe ∂xs ∂xf � � �∂x v ∂xh Tab...def ...h ð1:8:6Þ 52 1 Review of FundamentalTopics About Tensors www.Ebook777.com http://www.ebook777.com that is called relative tensor of weight W or with weighing factor W. This weight is an integer number, and J is the Jacobian of the transformation. For the particular case in which W ¼ 0, an absolute tensor exists. The concept of relative tensor allows distinguishing a relative invariant of a scalar, which is an absolute invariant. To differentiate these concepts, let the relative invariant A of weight W, which transforms according to the expression A ¼ JWA ð1:8:7Þ For the particular case in whichW ¼ 0, an absolute tensor exists A ¼ A that is a scalar. For W ¼ 1 provides the scalar density A ¼ JA. The definition of scalar density will be presented in detail in later paragraphs. To illustrate the concept of relative tensor, let the metric tensor gij with detgij ¼ g. Applying a linear and homogeneous transformation to this tensor eg‘m ¼ ∂xi∂ex‘ ∂xj∂exm gij ð1:8:8Þ with detegij ¼ eg, and by means of the property of the product of determinants, provides the relative scalar of weight W ¼ 2 eg ¼ J2g ) ffiffiffiegp ¼ J ffiffiffigp ð1:8:9Þ For J ¼ 1 provides ffiffiffigp that is a relative tensor of unit weight, being, therefore, an invariant. With expression (1.8.9) and the condition gg ¼ 1, having detgij ¼ g, it is verified that ffiffiffi g p is a relative tensor of weight �1. Let the Jacobian J of weightW ¼ 1, which is an invariant and when changing to a new coordinate system provides for this determinant J ¼ αJ being α a scalar (invariant). Raising both members of this expression to the power W J W ¼ αWJW ð1:8:10Þ where JW is an invariant of weight W, thus αW ¼ JWJ�W ð1:8:11Þ Consider the relative tensor Tijk of weight W that transforms by means of the expression T i jk ¼ αW ∂xm ∂xj ∂xn ∂xk ∂xi ∂x‘ T ‘mn ð1:8:12Þ and substituting in expression (1.8.12), the value of αW given by expression (1.8.11) provides 1.7 Relative Tensors 53 T i jk ¼ J W J�W � �∂xm ∂xj ∂xn ∂xk ∂xi ∂x‘ T ‘mn It follows that J �W T i jk � � ¼ ∂x m ∂xj ∂xn ∂xk ∂xi ∂x‘ J�WT ‘mn � � ð1:8:13Þ As J�WT ‘mn � � is an absolute tensor, and by means of the transformation law of tensors, it is concluded that J �W T i jk is also an absolute tensor. This shows that the transformation of a relative tensor of weight W in an absolute tensor is carried out multiplying it by the invariant of unit weight raised to the power�W. The invariantffiffiffi g p of unit weight is used to carry out this kind of transformation. This systematic allows, for instance, transforming the relative tensor Tkij of weight W into the absolute tensor Akij by means of T kij ffiffiffi g p� ��W ¼ Akij ð1:8:14Þ The operations multiplying by a scalar, addition, subtraction, contraction, outer product, and inner product are applicable to the relative tensors. These operations provide new relative tensors; as a result, the proof is analogous to the demonstra- tions performed for the absolute tensors. Exercise 1.20 Show that δij is an absolute tensor. It is admitted firstly that δij is a relative tensor of unit weight, being detδij ¼ 1, which transforms into the absolute tensor δ∗ij by means of the expression δ∗ij ¼ ffiffiffi g p� ��1 δij Asδ∗ij is an isotropic tensor so δij is also isotropic, then ffiffiffi g p� ��1 ¼ 1, which shows that δij is an absolute tensor. 1.7.1 Multiplication by a Scalar This operation provides as a result a relative tensor of weightW, which components are the components of the original relative tensor multiplied by the scalar. Let, for example, the relative tensor (J )WTij and the scalar m which product Pij is given by Jð ÞWPij ¼ m Jð ÞWTij. To demonstrate that this expression represents a tensor, it is enough to apply the transformation law of tensors to this expression. 54 1 Review of Fundamental Topics About Tensors 1.7.1.1 Addition and Subtraction This operation is defined for relative tensors of the same order and of the same kind, such as in the case of the following mixed tensors Jð ÞWT kij ¼ Jð ÞWAkij þ Jð ÞWBkij Subtraction is defined in the same way as addition, however, admitting that one of the tensors be multiplied by the scalar�1: Jð ÞWT kij ¼ Jð ÞWAkij þ �1ð Þ Jð ÞWBkij . To demonstrate that these expressions represent relative tensors, the transformation law of tensors is applied to this expression. 1.7.1.2 Outer Product This operation is defined in the same way as the outer product of absolute tensors. Let, for example, the relative tensor Jð ÞW1Ak...ij... of variance ( p, q) and weightW1, and the relative tensor Jð ÞW2B...‘m...rs of variance (u, v) and weight W2, which multiplied provide Jð ÞWTk...‘mij...rs ¼ Jð ÞW1Ak...ij... h i Jð ÞW2B...‘m...rs h i that is a relative tensor of variance pþ u, qþ vð Þ and weight W ¼ W1 þW2. To demonstrate that this product is a relative tensor, the transformation law of tensors is applied to this expression. 1.7.1.3 Contraction This operation is defined in the same way as the contraction of the absolute tensors. Let, for example, the relative tensor (J)WTijk‘m, in which contracting the upper index j provides Jð ÞWTijj‘m ¼ Jð ÞWT i‘m that shows that the resulting relative tensor has its order reduced in two, but maintains its weight W. To demonstrate that this contraction is a relative tensor, the transformation law of tensors is applied to this expression. 1.7 Relative Tensors 55 Free ebooks ==> www.Ebook777.com 1.7.1.4 Inner Product This operation is defined in the same manner as the inner product of the absolute tensors. Let, for example, two relative tensors Jð ÞW1Aij and Jð ÞW2B ‘k , whereby it follows for the outer product of these tensors Jð ÞW1þW2P ‘ijk ¼ Jð ÞW1Aij h i Jð ÞW2B ‘k h i that represents a relative tensor of the fourth order and weight W ¼ W1 þW2, and with the contraction of the index ‘ the inner product is given by Jð ÞW1þW2P ‘ij‘ ¼ Jð ÞW1Aij h i Jð ÞW2B ‘‘ h i ¼ Jð ÞW1þW2Pij This shows that the resulting relative tensor is of the second order and weight W ¼ W1 þW2. To demonstrate that this product is a relative tensor, the transfor- mation law of tensors is applied to this expression. 1.7.1.5 Pseudotensor The varieties that present a few tensorial characteristics, for example, when chang- ing the coordinate system they follow a transformation law that differs from the transformation law of tensors by the presence of the Jacobian, are called pseudotensor (relative tensors). However, these varieties are not maintained invari- ant when the coordinate system is transformed. The definitions of the antisymmetric pseudotensors εijk and ε ijk are associated, respectively, to the permutation symbols in the covariant form eijk or in the contravariant form eijk, to which correspond the values þ1 or �1 relative to the even or odd number of permutations, respectively. The Ricci pseudotensors εijk and εijk are associated to the concept of space orientation. These varieties, when changing the coordinate system, transform in the same way as the tensors, but are not invariant after these transformations. This shows that a few characteristics are similar to the tensors but vary with the change of referen- tial, for they assume the values 1, so they are not tensors in the sensu stricto of the term. In expression (1.7.27) it is verified that εj1j2���jn�mi1i2���im has weight þ1, and the tensor Tj1j2���jn�m has weight superior to the weight of the antisymmetric tensor A i1i2���im½ �. This expression illustrates the applying of the pseudotensors. Exercise 1.21 Show that (a) εijk is a covariant pseudotensor of the third order and weight �1. (b) εijk is a contravariant pseudotensor of the third order and weight þ1. (c) The absolute pseudotensors can be obtained from these pseudotensors. 56 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com (a) The definition of determinant allows writing Jεpqr, and as the pseudotensor εijk assume the values 0, 1, on being applied to this variety,it provides a linear and homogeneous transformation Jεpqr ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr εijk ) εpqr ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr J�1εijk ¼ εpqr then εijk is a covariant pseudotensor of the third order and weight �1. (b) In a way that is analogous to the previous case, for defining the determinant Jε‘mn, and for the transformation law of tensors Jεpqr ¼ ∂x p ∂xi ∂xq ∂xj ∂xr ∂xk εijk ) εpqr ¼ ∂x p ∂xi ∂xq ∂xj ∂xr ∂xk J �1 εijk As JJ ¼ 1 it results in εpqr ¼ ∂x p ∂xi ∂xq ∂xj ∂xr ∂xk Jεijk then εijk is a contravariant pseudotensor of the third order and weight +1. (c) As the pseudotensor εijk has weight �1, it follows by the transformation law of relative tensors into absolute tensors, where the upper asterisk indicates the absolute tensor ε*ijk ¼ ffiffiffi g p� ��1h i�1 εijk ¼ ffiffiffigp εijk For the relative pseudotensor εijk the absolute pseudotensor indicated by the lower asterisk exists εijk* ¼ 1ffiffiffi g p εijk Exercise 1.22 Show that gij is an absolute tensor. Rewriting expression (1.8.9) ffiffiffi g p ¼ J ffiffiffigp and with the cofactor of the matrix of tensor gij given by Gij ¼ 1 2 eik‘ejpqgkpg‘q and in terms of Ricci’s pseudotensor 1.7 Relative Tensors 57 Gij ¼ 1 2 εik‘εjpqgkpg‘q it follows that gij ¼ G ij g ¼ 1 2 εik‘εjpqgkpg‘q The term to the right of this expression. is the product of two pseudotensors and the tensors, being gij the inner product of these two varieties. This expression has weight W ¼ 0, then gij is an absolute tensor. 1.7.1.6 Scalar Capacity Let an antisymmetric pseudotensor Cijk in an affine space, for which according to expression (1.7.1) for N ¼ 3 and p ¼ 3, there is only one independent component. Writing Cijk as a function εijk follows Cijk ¼ εijkc where c is a component of the variety and with the change of the coordinate system Cijk ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr C pqr and as the antisymmetry is maintained when the reference system is changed C pqr ¼ εpqrc ð1:8:15Þ Considering the component C123: C123 ¼ ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr C pqr ð1:8:16Þ and substituting expression (1.8.15) in expression (1.8.16) c ¼ εpqr ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr c ð1:8:17Þ Let J ¼ εpqr ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr ð1:8:18Þ 58 1 Review of Fundamental Topics About Tensors results in the following expressions c ¼ Jc ) c ¼ 1 J c ¼ Jc ð1:8:19Þ Function c is the only independent component of the antisymmetric pseudotensor Cijk, which is called scalar capacity. Then a scalar capacity is a pseudotensor of weight �1. To illustrate the concept of scalar capacity, let, for example, the antisymmetric variety dVijk that defines an elementary volume in space E3. This analysis follows the same routine presented when defining the scalar capacity. The elementary volume is obtained by means of the mixed product of three vectors that define the three reference axes in this space dVijk ¼ dx1 0 0 0 dx2 0 0 0 dx3 ¼ dx1dx2dx3 ¼ dV ) dVijk ¼ dxidxjdxk ð1:8:20Þ and with the transformation law of tensors dVijk ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr dxpdxqdxr dV ¼ ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr dxpdxqdxr ð1:8:21Þ The antisymmetry of the pseudotensor is maintained when changing the coor- dinate system dxpdxqdxr ¼ εijkdV ð1:8:22Þ and substituting expression (1.8.21) in expression (1.8.22) dV ¼ εijk ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr dV results in the following expressions dV ¼ JdV ) dV ¼ 1 J dV ∴dx1dx2dx3 ¼ 1 J dx1dx2dx3 ð1:8:23Þ This shows that the elementary volume in an affine space is a pseudoscalar of weight�1. In a more restricted manner, it says that the volume is a scalar capacity. The term capacity comes from the association of the volume (capacity, content) to the variety being analyzed. It is concluded that the integration of expression (1.8.23), which represents a scalar field, is a pseudoscalar. 1.7 Relative Tensors 59 Free ebooks ==> www.Ebook777.com 1.7.1.7 Scalar Density Let the antisymmetric pseudotensor Dijk, for which an analysis analogous to the one developed when defining the scalar capacity is carried out D123 ¼ D ¼ εijk ∂x i ∂x1 ∂xj ∂x2 ∂xk ∂x3 D J ¼ εpqr ∂x 1 ∂xp ∂x2 ∂xq ∂x3 ∂xr ¼ εijk ∂x i ∂x1 ∂xj ∂x2 ∂xk ∂x3 then D ¼ JD ) D ¼ 1 J D ) D ¼ JD Function D is the unique component of the antisymmetric pseudotensor Dijk, which is called scalar density. Then a scalar density is a pseudotensor of weightþ1. To illustrate the concept of scalar density, let, for example, a body of elementary mass dm in the affine space E3. This mass is determined by means of density (specific mass) ρ(x1, x2, x3) and the elementary volume dV, thus dm ¼ ρ x1; x2; x3ð ÞdV. Considering that the mass is invariable (is a scalar) dm ¼ ρ x1; x2; x3� �dV ¼ ρ x1; x2; x3� �dV ð1:8:24Þ and as dV is a scalar capacity of weight �1, substituting expression (1.8.23) in expression (1.8.24) provides ρ x1; x2; x3 � � dV ¼ ρ x1; x2; x3� � 1 J dV ) ρ x1; x2; x3� � ¼ J ρ x1; x2; x3� � ð1:8:25Þ This shows that the density in an affine space is a pseudoscalar of weightþ1, this variety being called scalar density. The term density is not physically correct, for in truth ρ(x1, x2, x3) such as it is presented defines the body’s specific mass. The concepts shown for the elementary volume and for density in the affine space E3 can be generalized for the space EN. The varieties that transform by means of the expressions with structure analogous to the structures of expressions (1.8.23) and (1.8.25) are called, respectively, scalar capacity and scalar density in space EN. 1.7.1.8 Tensorial Capacity Let the space E3 where product of a scalar capacity c by the tensor T k ij exists, which defines a tensorial density Ckij given by 60 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com Ckij ¼ cT kij ð1:8:26Þ and for a new coordinate system C r pq ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk Ckij ð1:8:27Þ it follows that C r pq ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk cT kij ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk JcT kij ) C r pq ¼ JcT rpq ð1:8:28Þ Expression (1.8.28) shows that Ckij transforms in accordance with a law that is similar to the transformation law of scalar capacity; however, it does not represent a relative scalar but a relative tensor of weightþ1. The generalization of the concepts of tensorial capacity for the space EN is immediate. 1.7.1.9 Tensorial Density Let, for example, the space E3 where the product of a scalar densityD by the tensor Tkij exists, which defines a tensorial density D k ij given by Dkij ¼ DT kij ð1:8:29Þ and for a new coordinate system D r pq ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk Dkij ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk DT kij ¼ ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk 1 J DT kij D r pq ¼ J�1DT r pq ð1:8:30Þ Expression (1.8.30) shows that Dkij transforms in accordance with a law that is to the scalar density transformation law. However, it does not represent a relative scalar but a relative tensor of weight �1. The generalization of the concepts of tensorial density for the space EN is immediate. The outer products between these varieties (pseudotensors and tensors) result in Scalar capacity� scalar density ¼ scalar Scalar capacity� tensor ¼ tensorial capacity Scalar density� tensor ¼ tensorial density Pseudotensor� pseudotensor ¼ tensor Tensor� pseudotensor ¼ pseudotensor 1.7 Relative Tensors 61 Free ebooks ==> www.Ebook777.com 1.8 Physical Components of a Tensor In mathematics the approach to the problems, in general, is carried out by means of nondimensional parameters. In physics and engineering the parameters have mag- nitude and dimensions, for example, N/mm2,m/s, etc. The analysis of a physical problem by tensorial means requires that the parameters being studied be invariant when changing the coordinate system. It happens that the axes of the coordinate systems generally do not have the same dimensions. A Cartesian coordinate system has axes that define lengths, but, for example, a spherical coordinate system has two axes that express nondimensional coordinates, the same occurring in thecylindrical coordinate system with one of their axes. Therefore, the components of a tensor have dimensions, and when the coordinate system is changed, these components vary in magnitude and dimension. To express the transformation of tensors in a consistent way (in magnitude and dimension), and that these varieties can be added after a change of the coordinate systems, the same must be expressed in terms of their physical components. 1.8.1 Physical Components of a Vector The concept of geometric vector is associated to the idea of displacement, its transformation law being dxk ¼ ∂xk ∂xi dxi where the coefficients ∂x k ∂xi are constants. With respect to a Cartesian coordinates, the term dxjgj represents a displacement in terms of the unit vectors of the coordinate axes dxjgj ¼ dx1iþ dx2jþ dx3k ð1:9:1Þ However, this term in a curvilinear coordinates does not represent a displace- ment, so gk will not be a unit vector in this coordinate system. This shows that the vector must be written in terms of components that express a displacement, called the physical components of the vector. Consider the vector u with physical components u�j , which can be written in terms of these components and of their contravariant components u ¼ u*kek ¼ ukgk ð1:9:2Þ Comparing expressions (1.9.1) and (1.9.2) dxjgj ¼ dx*jej ð1:9:3Þ where dx* j are the physical components which by analogy correspond to the displacement dxk, and with the unit vectors of base gk, ek, the components u k, u�k 62 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com are obtained in terms of the unit vector gk. Letgi � gj ¼ gij then gj �� �� ¼ ffiffiffiffiffiffiffiffig jjð Þp , where the indexes shown in parenthesis do not indicate a summation in j. As the unit vector ej is collinear with gi, thus gj ¼ ffiffiffiffiffiffiffiffi g jjð Þ p ej ð1:9:4Þ and with expression (1.9.4) in expression (1.9.3) dx*j ¼ ffiffiffiffiffiffiffiffig jjð Þp dxj ð1:9:5Þ In an analogous way, by means of expression (1.9.2) u*k ¼ ffiffiffiffiffiffiffiffiffi g kkð Þ p uk ð1:9:6Þ The physical components u�k have the characteristics of displacement, so they can be added vectorially (parallelogram rule), denoting the contravariant physical components of the vector. These components are not unique. Let another variety of components ũk that represents the projection of vector u on the direction of the unit vector ek. Consider ẽk the reciprocal unit vector of ek, whereby, for this reciprocal basis, euk ¼ u � ek ð1:9:7Þ u ¼ ukgk ¼ eukeek ð1:9:8Þ but ẽk is collinear with g k thus ek ¼ 1eek gk ¼ ffiffiffiffiffiffiffiffiffi g kkð Þ p ek ¼ ffiffiffiffiffiffiffiffiffi g kkð Þ p eek ) eek ¼ ffiffiffiffiffiffiffiffiffi g kkð Þ p gk eek ¼ ffiffiffiffiffiffiffiffiffig kkð Þp gk ð1:9:9Þ where the indexes shown in parenthesis do not indicate summation in k, and with expression (1.9.9) in expression (1.9.7) uk ¼ ffiffiffiffiffiffiffiffiffig kkð Þp euk ð1:9:10Þ The physical components ũk are the covariant components of vector u. Putting uk ¼ g kkð Þuk and with the expressions (1.9.6) and (1.9.10) then in an orthogonal coordinate system u*k ¼ euk. This shows that the distinction between the covariant and contravariant basis disappears when the coordinate system is orthogonal. 1.8 Physical Components of a Tensor 63 Figure 1.11 shows the physical components of vector u in the curvilinear coordinate system Xi. The components ukffiffiffiffiffiffiffi g kkð Þ p (expression (1.9.10)) represent the lengths of the projections which are orthogonal to the coordinate axes of the referential system. The components ffiffiffiffiffiffiffiffiffi g kkð Þ p uk (expression (1.9.6)) represent the lengths the of the sides of the parallelepiped, which diagonal is the vector u. Exercise 1.23 Calculate the contravariant, covariant, and physical components of the velocity vector of a point vi ¼ dxidt , in terms of the cylindrical coordinates of the point xi(r, θ, z). Cartesian to cylindrical Cylindrical to Cartesian x1 ¼ x1 cos x2 ¼ r cos θ x1 ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x1 � � 2 þ x2� � 2q ¼ r x2 ¼ x2 sin x2 ¼ r sin θ x2 ¼ arctg x2 x1 ¼ θ x3 ¼ x3 ¼ z x3 ¼ x3 ¼ z The Cartesian coordinates of the vector are vi ¼ dxidt , and for the cylindrical coordinates vi ¼ ∂x i ∂xj vj ) vi ¼ ∂x i ∂xj dxi dt ¼ dx i dt This shows that the contravariant components of vector v are derivatives with respect to the time of the position vector defined by the coordinates xi, then vi � � ¼ dx1 dt , dx2 dt , dx3 dt � ¼ dr dt , dθ dt , dz dt � For the covariant components in terms of the cylindrical coordinates 1X 2X 3X P 1g 2g 3g u 11 1 g u 22 2 g u 33 3 g u Fig. 1.11 Physical components of the vector u in the curvilinear coordinate system Xi 64 1 Review of Fundamental Topics About Tensors Free ebooks ==> www.Ebook777.com vi ¼ ∂x i ∂xi vj ¼ gijvj gij ¼ 1 0 0 0 r2 0 0 0 1 24 35 Developing the expression of vi v1 ¼ g11v1 þ g12v2 þ g13v3 ¼ v1 ¼ dx1 dt ¼ dr dt v2 ¼ g12v1 þ g22v2 þ g23v3 ¼ x1 � �2 dx2 dt ¼ r2 dθ dt v3 ¼ g31v1 þ g23v2 þ g33v3 ¼ dx3 dt ¼ dz dt whereby vi � � ¼ dx1 dt , x1 � �2 dx2 dt , dx3 dt � ¼ dr dt , r2 dθ dt , dz dt � The vector norm is given by vk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi dx1 dt � �2 þ x1ð Þ2 dx 2 dt �2 þ dx 3 dt � �2s whereby for its physical components vf g ¼ dx 1 dt , x1 dx2 dt , dx3 dt � ¼ dr dt , r dθ dt , dz dt � Exercise 1.24 Let the vector u ¼ 3g1 þ g2 þ 2g3, having g1 ¼ 2e1, g2 ¼ 2e1 þ e2, and g3 ¼ 2e1 þ e2 þ 3e3, where e1, e2, e3 are orthonormal vectors, calculate their contravariant physical components. From the covariant basis g1 � g1 ¼ 2e1:2e1 ¼ 4 ) ffiffiffiffiffiffi g11 p ¼ 2 g2 � g2 ¼ 2e1:2e1 þ e2 � e2 ¼ 4þ 1 ¼ 5 ) ffiffiffiffiffiffi g22 p ¼ ffiffiffi 5 p g3 � g3 ¼ 2e1:2e1 þ e2 � e2 þ 3e3 � 3e3 ¼ 4þ 1þ 9 ¼ 14 ) ffiffiffiffiffiffi g33 p ¼ ffiffiffiffiffi 14 p follows 1.8 Physical Components of a Tensor 65 www.Ebook777.com http://www.ebook777.com u*1 ¼ u1 ffiffiffiffiffiffig11p ¼ 3� 2 ¼ 6 u*2 ¼ u2 ffiffiffiffiffiffig22p ¼ 2� ffiffiffi5p ¼ 2 ffiffiffi5p u*3 ¼ u3 ffiffiffiffiffiffig33p ¼ 1� ffiffiffiffiffi14p ¼ ffiffiffiffiffi14p 1.8.1.1 Physical Components of the Second-Order Tensor The contravariant physical components of the vectors u and v are given by expression (1.9.10) ui ¼ ffiffiffiffiffiffiffiffig iið Þp eui ) eui ¼ uiffiffiffiffiffiffiffiffig iið Þp vj ¼ ffiffiffiffiffiffiffiffig jjð Þp evj ) evj ¼ vjffiffiffiffiffiffiffiffig jjð Þp For the second-order tensor eTij ¼ euievj eTij ¼ uiffiffiffiffiffiffiffiffi g iið Þ p vjffiffiffiffiffiffiffiffi g jjð Þ p ¼ 1ffiffiffiffiffiffiffiffi g iið Þ p ffiffiffiffiffiffiffiffi g jjð Þ p uivj � � ¼ Tijffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi g iið Þg jjð Þ p Tij ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffig iið Þg jjð Þp eTij ð1:9:11Þ In a related manner, for the contravariant physical components *Tij ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffig iið Þg jjð Þp eTij ð1:9:12Þ The obtaining of the physical components of tensors of a higher order follows the analogous way to that of the second-order tensors. 1.9 Tests of the Tensorial Characteristics of a Variety The transformation law of the tensors and the quotient law allow establishing a group of functions Np of the coordinates of the referential system Xi which are the components of a tensor. The tensorial nature of the functions that fulfill these requirements is highlighted by the invariance that this variety has when there is a change of the coordinate system. However, the evaluation if a variety has tensorial characteristics by means of the quotient law is not wholly complete, as it will be shown next applying to the group ofN2 components of a variety Tpq, for which it is desired to search if it has the characteristics of a tensor. Multiplying Tpq by an 66 1 Review of Fundamental Topics About Tensors arbitrary vector vp and admitting by hypothesis that Tpqv pvq ¼ m, where m is a scalar, it provides for a new coordinate system Tijv ivj ¼ m, and as m is an invariant, then m ¼ m, by means of the transformation law of vectors Tpqv pvq ¼ ∂x i ∂xp ∂xj ∂xq Tijv pvq Then Tpq � ∂x i ∂xp ∂xj ∂xq Tij � � vpvq ¼ 0 ð1:10:1Þ The summation rule is applied varying the indexes p and q, so the product vpvq is not, in general, null. Consider the vectors vi with unit components 1, 0, 0� � �0ð Þ, 0, 1, 0� � �0ð Þ, and 0, 0, 0� � �1ð Þ, the term in parenthesis of expression (1.10.1) stays T11 � ∂x i ∂x1 ∂xj ∂x1 Tij � � v1v1 ¼ 0 and as v1v1 6¼ 0 T11 � ∂x i ∂x1 ∂xj ∂x1 Tij ¼ 0 ð1:10:2Þ In an analogous way it results in T22 � ∂x i ∂x2 ∂xj ∂x2 Tij ¼ 0 ð1:10:3Þ and so successively for the other values assumed for the indexes. This shows that for p ¼ q the terms in parenthesis from expression (1.10.1) cancel each other. However, for p 6¼ q the complementary analysis of this expression behavior becomes necessary. Let vector viwith components v1, v2, 0, � � �0ð Þ, whereby from expression (1.10.1) for p, q ¼ 1, 2, it follows that T11 � ∂x i ∂x1 ∂xj ∂x1 Tij � � v1v1 þ T12 � ∂x i ∂x1 ∂xj ∂x2 Tij � � v1v2 þ T21 � ∂x i ∂x2 ∂xj ∂x1 Tij � � v2v1 þ T22 � ∂x i ∂x2 ∂xj ∂x2 Tij � � v2v2 ¼ 0 ð1:10:4Þ Expressions (1.10.2) and (1.10.3) simplify expression (1.10.4), for the coeffi- cients of the terms vpvq are null for p ¼ q. For p 6¼ q with Tij ¼ Tji 1.9 Tests of the Tensorial Characteristics of a Variety 67 ∂xi ∂x1 ∂xj ∂x2 Tij ¼ ∂x i ∂x2 ∂xj ∂x1 Tij and with the hypothesis of symmetry results in ∂xi ∂x1 ∂xj ∂x2 Tij ¼ ∂x i ∂x2 ∂xj ∂x1 Tji Expression (1.10.4) is rewritten as T12 þ T21ð Þ � Tij þ Tji � � ∂xi ∂x1 ∂xj ∂x2 � v1v2 ¼ 0 and as the components v1 and v2 are arbitrary, for v1 ¼ v2 ¼ 1 T12 þ T21 ¼ Tij þ Tji � � ∂xi ∂x1 ∂xj ∂x2 ð1:10:5Þ Generalizing expression (1.10.5) for the variation of the indexes p, q ¼ 1, 2, 3, . . ., it results in Tpq þ Tqp ¼ Tij þ Tji � � ∂xi ∂xp ∂xj ∂xq ð1:10:6Þ Expression (1.10.6) is the transformation law of second-order tensors, for the term Tpq þ Tqp � � represents the symmetric part of tensor 2Tpq. However, the antisymmetric part of this tensor is not contained in this analysis, whereby it cannot be concluded that this portion has tensorial characteristics. It is concluded that only the symmetric part of the N2 components of variety Tpq is a tensor, for when applying the quotient law to this portion it transforms according to the transforma- tion law of second-order tensors. This is the reason why the quotient law must be applied with caution, so as to avoid evaluation errors when checking the tensorial characteristics of a variety. The transformation law of tensors and the consideration of invariance of the variety when having a linear transformation form the criterion that is most appro- priate to evaluate if the Np components of this variety have tensorial characteristics. Problems 1.1 Use the index notation to write: (a) dx1 dt dx2 dt 8><>: 9>=>; ¼ a11 a12a21 a22 � x1 x2 � ; (b) Φ ¼ x21 þ x22 þ 2x1x2 Answer: (a) x, t ¼ aijxj; (b) Φ ¼ xixj. 68 1 Review of Fundamental Topics About Tensors 1.2 Let aij constant 8i, j, calculate ∂ aijxixjð Þ∂xk ) where aij ¼ aji. Answer: ∂ 2aikxið Þ∂x‘ ¼ 2aik 1.3 If aijkx ixjxk ¼ 0 8x1, x2, � � �, xn and aijk are constant values, show that akji þ ajki þ aikj þ aijk þ akij þ ajik ¼ 0. 1.4 Calculate for i, j ¼ 1, 2, 3: (a) δijAi, (b) δijAij, (c) δii, (d) δijδji, (e) δijδjkδk‘, (f) C ¼ aijkaijk Answer: (a) δijAi ¼ Aj, (b) δijAi ¼ Aii ¼ Ajj, (c) δi i ¼ 3, (d) δijδijijji jiji ¼ 3, (e) δijδ ijij jkδ jk jk k‘¼k‘k‘δi‘, and (f) 64. 1.5 Calculate the Jacobian of the linear transformations between the coordinate systems (a) x1 ¼ x1; x2 ¼ x1x2; x3 ¼ x1x2x3 ; (b) x1 ¼ x1 cos x2 sin x3; x2 ¼ x1 sin x2 sin x3; x3 ¼ x1 cos x3. Answer: (a) J ¼ x1� �2x2; (b) J ¼ � x1� � 2 sin x3. 1.6 Given the tensor Tk‘ ¼ 1 0 0 0 2 1 0 1 3 24 35 in the coordinate system Xi, calculate the components of this tensor in the coordinate system X i , with the relations between the coordinates of these systems given by x1 ¼ x1 þ x3, x2 ¼ x1 þ x2, x ¼ x3. Answer: Tij ¼ 3 2 2 2 2 1 2 1 4 24 35 1.7 Given the tensor Tij ¼ 1 1 5 1 2 �1 5 �1 3 24 35 in the coordinate system Xi, calculate the components of this tensor in the coordinate system X i , with the relations between the coordinates of these systems given by x1 ¼ x1 þ 2x2, x2 ¼ 3x3, x ¼ x3. Answer: T ij ¼ 25 8 2 8 4 10 2 10 3 24 35 1.8 Show that (a) tr Tð Þ ¼ tr Sð Þ, where T and S are, respectively, a symmetric and an antisymmetric tensor, both of the second order; (b)Tijk‘ ¼ 0, being Tijk‘ one symmetric tensor in the indexes i, j and antisymmetric in the indexes j, ‘. 1.9 Decompose the second-order tensor in two tensors, one symmetric and another antisymmetric 1.9 Tests of the Tensorial Characteristics of a Variety 69 Free ebooks ==> www.Ebook777.com Tij ¼ �1 2 0 3 0 �2 1 0 1 24 35 Answer: �1 2:5 0:5 2:5 0 �1 0:5 �1 1 24 35 0 �0:5 �0:50:5 0 �1 0:5 1 0 24 35. 1.10 Consider the tensor Tij that satisfies the tensorial equation mTij þ nTji ¼ 0, wherem > 0 and n > 0 are scalars. Prove that if Tij is a symmetric tensor, then m ¼ �n, and m ¼ n if this is an antisymmetric tensor. 1.11 Let the Cartesian coordinate system with basis vectors e1, e2, e3. Calculate the metric tensor of the space with basis vectors g1 ¼ e1, g2 ¼ e1 þ e2, and g3 ¼ e1 þ e2 þ e3. Answer: gij ¼ 1 1 1 1 2 2 1 2 3 24 35 1.12 Let the basis vectors e1, e2 of the coordinate system X i with metric tensor gij and the basis vectors ee1 ¼ 3e1 þ e2 and ee2 ¼ �e1 þ 2e2 of the coordinate system eXi. Calculate the covariant components of the metric tensoregij in terms of the components of gij. Answer: eg11 ¼ 9g11 þ 6g12 þ g22 ; eg12 ¼ eg21 ¼ �3g11 þ 5g12 þ 2g22 ;eg22 ¼ g11 � 4g12 þ 4g22. 1.13 Calculate the contravariant components of the vector u ¼ g1 þ 2g2 þ g3, where the covariant base vectors are g1 ¼ e1, g2 ¼ e1 þ e2, g3 ¼ e3, being e1, e2, e3 base vectors. Answer: u ¼ 4g1 þ 7g2 þ 8g3. 1.14 Let the contravariant base vectors g1 ¼ e1; g2 ¼ e1 þ e2; and g3 ¼ e1 þ e2 þ e3, where e1, e2, e3 are the vectors of the one orthonormal base. Calculate the: (a) Vectors g1, g2, g3 of the contravariant base (b) Metric tensor and the conjugated tensor. Answer: (a) g1 ¼ e1 � e2 g2 ¼ e2 � e3 g3 ¼ e3 8<: ; (b) gij ¼ 1 1 1 1 2 2 1 2 3 24 35 gij ¼ 2 �1 0�1 2 2 0 �1 1 24 35 1.15 Consider the coordinate system x1 ¼ x1 cos x2, x2 ¼ x1 sin x2, x3 ¼ x3. Calcu- late the arc length along the parametric curve x1 ¼ a cos t, x2 ¼ a sin t, x3 ¼ b t in the interval 0 t c, being a, b, c positive constants. Answer: L ¼ c ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi a2 þ b2 p 1.16 Calculate the angle between the vectors (a) u 2;�3; 1ð Þ, v 3;�1;�2ð Þ; (b) u 2; 1;�5ð Þ, v 5; 0; 2ð Þ. Answer: (a) 60o; (b) 90o 1.17 With i, j, k ¼ 1, 2, 3 calculate the following expressions: (a) uivjδji � vkuiδki; (b) δijδji; (c) eijkuiujuk 70 1 Review of Fundamental Topics About Tensors www.Ebook777.com http://www.ebook777.com Answer: (a) zero, (b) 3, (c) zero 1.18 Show that the followings expressions are invariants (a) Tijuivj; (b) Tii; (c) det Tij. 1.19 Let the vector (a) ui show that if A ijuiuj ¼ Bijuiuj, then Aij þ Aji ¼ Bij þ Bij; (b) ui and if A ijuiuj is invariant, show that A ij þ Aji� � is a tensor. 1.20 Let Tpqrs an absolute tensor, show that if Tijk‘ þ Tij‘k ¼ 0 in the coordinate system Xi, then Tijk‘ þ Tij‘k ¼ 0 in another coordinate system Xi. 1.21 Let the vector u ¼ g1 þ 2g2 þ g3, having g1 ¼ e1, g2 ¼ e1 þ e2, and g3 ¼ e1 þ e2 þ e3, where e1, e2, e3 are orthonormal vectors, calculate their contravariant physical components. Answer: u*1 ¼ 1, u*2 ¼ 2 ffiffiffi2p , u*3 ¼ ffiffiffi3p1.22 Calculate the value of the permutation symbol e321546. Answer: e321546 ¼ 1. 1.23 Show that (a) eijkejki ¼ 6; (b) eijkujuk ¼ 0; (c) er‘merst ¼ δrstr‘m ¼ 1 0 �1 8<: r, ‘, m, s, t ¼ 1, 2, 3; (d) er‘me rst ¼ δ s‘δ tm � δ t‘δ sm; (e) eijeij ¼ 2!; (f) eijk‘eijk‘ ¼ 4!; (g) eijk‘���eijk‘��� ¼ N! where N is the index number. 1.24 For i, j, k, ‘ ¼ 1, 2, 3 show that (a) δijk‘ ¼ δ ik δ i ‘ δ jk δ j ‘ ¼ δ ikδ j‘ � δ i‘δ jk, (b) δijkrst ¼ δ i r δ i s δ i t δ jr δ j s δ j t δ kr δ k s δ k t ¼ δijk‘, (c) δijkijk ¼ 6. 1.25 Calculate the determinant by means of the expansion of the permutation symbol 1 1 0 1 �1 0 1 1 0 1 �1 1 1 �1 1 0 ¼ eijk‘a1ia2ja3ka4‘ Answer: �3. 1.26 Show that (a) εijk ¼ ej � ek � � � ei where ei, ej, ek are unit vectors of one coordinate system; (b) if ui and u i are associate tensors and ui ¼ εijkujvk, then ui ¼ εijkujvk; (c) εijkuivjwk ¼ εijkuivjwk. 1.27 Simplify the expression F ¼ εijkεpqrAipAjqAkr. 1.28 Verify if the expression εmnpεmij þ εmnjεmpi ¼ εmniεmpj is correct or false. Justify the answer. 1.29 Write the tensor components Tij ¼ 1 2 εijkAk‘ with i, j, k, ‘ ¼ 1, 2, 3, 4, and show that if ijk‘ is an even permutation for the pair of 1234, then Tij ¼ Ak‘. 1.9 Tests of the Tensorial Characteristics of a Variety 71 Chapter 2 Covariant, Absolute, and Contravariant Derivatives 2.1 Initial Notes The curve represented by a function ϕ(xi) in a closed interval is continuous if this function is continuous in this interval. If the curve is parameterized, i.e., ϕ[xi(t)] being t2 a; b½ �, then it will be continuous if xi(t) are continuous functions in this interval, and it will be smooth if it has continuous and non-null derivatives for a value of t2 a; b½ �. The smooth curves do not intersect, i.e., the conditions xi að Þ ¼ xi bð Þ will only be satisfied if a ¼ b. This condition defines a curve that can be divided into differential elements, forming curve arcs. For the case in which the initial and final points coincide, expressed by condition a ¼ b, the curve is closed. The various differential elements obtained on the curve allow calculating its line integral. The curves can be smooth by part, i.e., they are composed of a finite number of smooth parts (arc elements), connected in their initial and final point. This kind of curve can intersect in one or more points, and if their extreme points coincide, it is called a closed curve. The differentiation condition of a function is associated to the concept of neighborhood and limit. The neighborhood of a point P(xi) is defined admitting that the very small radius ε, with which a sphere is traced, is centered on it. The interior of this sphere is this point’s neighborhood of radius ε. This definition is valid in the plane, changing the sphere for a circle, and is complemented admitting a set of points, which is called an open set. The points interior to the cube shown in Fig. 2.1 form an open set, for in each point P(xi) a sphere of radius ε can be drawn in its interior, which will be contained in the cube’s interior. If the cube’s edges are included, the result is a closed set. © Springer International Publishing Switzerland 2016 E. de Souza Sánchez Filho, Tensor Calculus for Engineers and Physicists , DOI 10.1007/978-3-319-31520-1_2 73 2.2 Cartesian Tensor Derivative The tensors derivative study begins with the research of what happens when a scalar function is differentiated. The behavior of this kind of function leads to the study of more general cases, such as those of the vectorial and tensorial functions. Consider a scalar function ϕ(xi) defined in a coordinate system Xi, which derivative with respect to the variable xi is given by ∂ϕ xið Þ ∂xi ¼ ϕ, i ¼ Gi and in another coordinate system X i has as derivative Gi ¼ ∂ϕ x ið Þ ∂xi ¼ ∂ϕ x ið Þ ∂xi ¼ ∂ϕ x ið Þ ∂xi ∂xi ∂xi ) Gi ¼ ∂x i ∂xi Gi whereby ϕ, i ¼ ∂ϕ xið Þ ∂xi ϕ, i that is the transformation law of vectors, so ∂ϕ xið Þ ∂xi is a vector and defines the gradient of the scalar function. In vectorial notation, it is graphed as u ¼ gradϕ xið Þ. ε P Fig. 2.1 Neighborhood of a point 74 2 Covariant, Absolute, and Contravariant Derivatives Differentiating the scalar function again with respect to the variable xj results in ϕ, ij ¼ ∂xk ∂xj ∂xm ∂xi � � ∂2ϕ ∂xm∂xk and for i ¼ j ϕ, ii ¼ ∂xk ∂xi ∂xm ∂xi � � ∂2ϕ ∂xm∂xk δkm ¼ ∂x k ∂xi ∂xm ∂xi ϕ, ii ¼ δkm ∂2ϕ ∂xm∂xk ¼ ∂ 2ϕ ∂xk∂xk then ϕ(xi),ii is a scalar. 2.2.1 Vectors In the study of vectors, there are two distinct manners to carry out a derivative: the derivative of a vector and the derivative of a point. In this text, a few conceptual considerations are made before calculating the derivative of a vector. Consider the vectorial space EN in which the scalar variable t2 a; bð Þ is defined, where for each value of this variable limited in the open interval there is a vector u (t) embedded in this space which has a metric tensor. Let the vector u(t) defined by a continuous function of the variable t, and that admits continuous derivatives, then by means of an elementary increment Δt this vector will have an elementary increment Δu tð Þ ¼ u tþ Δtð Þ � u tð Þ and when Δt ! 0 a vector v will exist such that lim Δu tð Þ Δt � v � � ! 0 then the vector v is derived from the vector u(t) by means of the variation of the parameter t, whereby v ¼ du dt 2.2 Cartesian Tensor Derivative 75 The rules of Differential Calculus are applicable to this kind of differentiation. The other method of calculating a vector by means of a derivative is associated to the concept of punctual space EN, with respect to which a variable t2 a; bð Þ is associated in a univocal way to an arbitrary point P(t). Taking a fixed point O contained in EN as reference origin, the vector r tð Þ ¼ OP tð Þ is determined. Let the vector r(t) defined by a continuous and derivable function, then the result is the vector dr tð Þ dt that does not depend on the origin O, but only on the point P(t). This statement can be demonstrated admitting a new arbitrary point O* for which the result is the vector OP ¼ OO* þO*P Fixing the vector OO* and calculating the derivative of these vectors with respect to the variable t, the result is d OPð Þ dt ¼ d OO * � � dt þ d O *P � � dt ) d OPð Þ dt ¼ d O *P � � dt This equality proves the independence of the vector dr tð Þ dt with respect to the arbitrated origin. This vector is derived from the point P(t), and in the case of the points defining a smooth curve C contained in the space EN, continuous and differentiable, and will be tangent to the curve in each point for which this derivative was calculated. In the general case of the vector being a function of various scalar variables r xið Þ, i ¼ 1, 2, . . . ,N, the result is by means of the differentiation rules of Differ- ential Calculus ∂2r tð Þ ∂xi∂xk ¼ ∂ 2 r tð Þ ∂xk∂xi For a vectorial function of scalar variables, the vector derivative is given by dr tð Þ ¼ ∂r tð Þ ∂xi dxi If the curve C is a function of the variables xi, and these depend on the variable t2 a; bð Þ, i.e., xi ¼ xi tð Þ, the curve is parameterized, then the hypothesis of Differ- ential Calculus is applicable, and the total differential of the vector r[xi(t)] is dr tð Þ dt ¼ ∂r tð Þ ∂xi dxi dt The rules applicable to the differentiation of vectors, and to the vectors obtained by means of differentiation of a point are the same. The concept of vector calculated by differentiation of one point can be extended to the study of tensors, where the points to be analyzed are contained in the tensorial space EN. 76 2 Covariant, Absolute, and Contravariant Derivatives Exercise 2.1 Show that derivative of the vector r with constant direction maintains its direction invariable. Let r tð Þ ¼ ψ tð Þu where ψ(t) is a parameterized scalar function and u is a constant vector, thus dr tð Þ dt ¼ ψ 0 tð Þuþ ψ tð Þ du dt ¼ ψ 0 tð Þu then r and dr tð Þ dt have thesame direction. 2.2.2 Cartesian Tensor of the Second Order The tensorial functions defined by Cartesian tensors are often found in applications of physics and the areas of engineering. Let, for instance, the derivative of this kind of tensorial function that is a particular case of the derivative of tensors expressed in curvilinear coordinate systems. For the analysis of this derivative a Cartesian tensor of the second order Tij is admitted, which components are functions of the coordi- nates xi. The tensor with this characteristic is a function of the point considered in the space E3. The transformation law of the tensors of the second order is given by Tij ¼ ∂x p ∂xi ∂xq ∂xj Tpq For the Cartesian coordinate systems the coefficients ∂x p ∂xi and ∂x q ∂xj are constants, for they represent the variation rates for the linear transformations, so they do not depend on the point’s coordinates. The derivative of this expression is given by ∂Tij ∂x‘ ¼ ∂x p ∂xi ∂xq ∂xj ∂Tpq ∂xm ∂xm ∂x‘ � � ) ∂Tij ∂x‘ ¼ ∂x p ∂xi ∂xq ∂xj ∂xm ∂x‘ ∂Tpq ∂xm that is the transformation law of tensors of the third order, concluding that the derivative of tensor Tij increased the order of this tensor in one unit. This conclusion is general and applicable to any Cartesian tensor. This kind of derivative is not valid for the more general tensors. The concept of tensors derivative must, therefore, be generalized for tensors which components are given in curvilinear coordinate systems. Exercise 2.2 Show that if Tij is a Cartesian tensor of the second order, then ∂2Tij ∂xk∂xm will be a tensor of the fourth order. 2.2 Cartesian Tensor Derivative 77 The tensor is a function of the coordinates Tij(x1, x2, x3), whereby the result is ∂xi ∂xj ¼ δij ) ∂ 2 xi ∂xj∂xk ¼ 0 and by transformation law Tpq ¼ ∂xi∂xp ∂xj ∂xq Tij then ∂2Tpq ∂xr∂xs ¼ ∂ 2 ∂xr∂xs ∂xi ∂xp ∂xj ∂xq Tij � � ¼ ∂xk ∂xr ∂ ∂xk ∂xm ∂xs ∂ ∂xm ∂xi ∂xp ∂xj ∂xq Tij � � � ¼ ∂xk ∂xr ∂xm ∂xs ∂xi ∂xp ∂xj ∂xq ∂ ∂xk ∂Tij ∂xm � � In a mnemonic manner Tpq, rs ¼ ∂xk∂xr ∂xm ∂xs ∂xi ∂xp ∂xj ∂xq Tij,mk that is the transformation law of tensors of the fourth order as Tij,mk ¼ Tij,km the result is that Tij,km ¼ ∂ 2 Tij ∂xk∂xm is a tensor of the fourth order. 2.3 Derivatives of the Basis Vectors Consider the contravariant vector uk ¼ uk xið Þ defined in terms of the parametric curve xi ¼ xi tð Þ, expressed with respect to the Cartesian coordinate system Xi. By means of the transformation law of vectors, the result for the curvilinear coordinate system X i is u‘ ¼ ∂x i ∂xk uk and with the techniques of differentiation with respect to the parameter t results in du‘ dt ¼ ∂x i ∂xk duk dt þ ∂ 2 xi ∂xk∂x‘ dx‘ dt uk ð2:3:1Þ 78 2 Covariant, Absolute, and Contravariant Derivatives that only represents a contravariant vector if, and only if, xi is a linear function of xk. The first term on the right of this expression represents an ordinary differentiation of a vectorial function expressed in Cartesian coordinates, but the second term contains the derivatives curvilinear coordinates xi, relative to a coordinate system that varies as a function of the points of the space. The study of this term is carried out considering the Cartesian coordinate system Xi and the curvilinear coordinate system X i , with unit vectors ei and gi, respectively (Fig. 2.2). Whereby defining the position vector r of point P with respect to the coordinate system Xi by means of their contravariant components r ¼ xiei the differential total of this vector is given by dr ¼ ∂r ∂xi dxi ð2:3:2Þ As the basis vectors ei do not depend on point P: dr ¼ dxiei so ei ¼ ∂r∂xi ð2:3:3Þ O P 1 X 2 X 3 X 1 X 2 X 3 X 3 g 1 g 2 g 2 e 3 e 1 e Fig. 2.2 Cartesian Xi and curvilinear X i coordinates with basis vectors ei and gi, respectively 2.3 Derivatives of the Basis Vectors 79 With respect to the local coordinate systemX i the result is the differential total of the position vector r: dr ¼ r, i dxi ð2:3:4Þ Whereby the base vectors of the curvilinear coordinate system results gi ¼ r, i ð2:3:5Þ that shows that the unit vectors gi are tangent to the curves that define the curvilinear coordinate system X i , that varies for each point of the vectorial space E3, and as the unit vectors ei do not vary ∂r ∂xk ¼ ∂x i ∂xk ei Comparing with expression (2.3.5) gk ¼ ∂xi ∂xk ei ð2:3:6Þ then ei ¼ ∂x j ∂xi gj ð2:3:7Þ The covariant derivative of the base vector defined by the expression (2.3.6) is given by gk, ‘ ¼ ∂2xi ∂xk∂x‘ ei ð2:3:8Þ and substituting expression (2.3.7) in this expression gk, ‘ ¼ ∂xj ∂xi ∂2xi ∂xk∂x‘ gj Defining the variety Γ jk‘ ¼ ∂xj ∂xi ∂2xi ∂xk∂x‘ ð2:3:9Þ with which the covariant derivatives of the basis vectors of the curvilinear coordi- nate system can be written as linear combination of the base vector gj: gk, ‘ ¼ Γ jk‘gj ð2:3:10Þ 80 2 Covariant, Absolute, and Contravariant Derivatives 2.3.1 Christoffel Symbols The coefficients determined by expression (2.3.9) can be expressed in terms of the derivatives of the metric tensor and its conjugated tensor. For the derivatives of the contravariant basis vectors considering another variety Γ i jm: gi, j ¼ Γ ijmgm ð2:4:1Þ Writing gi � gj � � , k ¼ δ ij � � , k ¼ g i, k � gj þ gj , k � gi ¼ 0 ð2:4:2Þ and substituting the expressions (2.3.10) and (2.4.1) in expression (2.4.2) Γ i jkg k � gj þ Γ ijkgj � gi ¼ 0 ð2:4:3Þ then Γ i jk ¼ �Γ ijk and with the expressions (2.3.9), (2.3.10), (2.4.1) and with the prior expression it follows that Γmij ¼ ∂xm ∂xk ∂2xk ∂2xi∂xj gi, j ¼ Γmij gm gi, j ¼ �Γ ijmgm ð2:4:4Þ The relation between the covariant and contravariant unit vectors is defined by gi ¼ gijgj and the derivative of this expression with respect to the coordinate xk is given by gi, k ¼ gij,kgj þ gijgj, k Replacing expressions (2.3.10) and (2.4.4) in this last expression Γmikgm ¼ gij,kgj � gijΓ jkmgm and with the multiplying by gn 2.3 Derivatives of the Basis Vectors 81 Γmikgm � gn ¼ gij,kgj � gn � gijΓ jkmgm � gn Γmik δ n m ¼ gij,kgjn � gijΓ jkmgmn Γ nik ¼ gij,kgjn � gijgmnΓ jkm The multiplying of this last expression by gnp provides gnpΓ n ik þ gijgmngnpΓ jkm ¼ gij,kgjngnp ) gnpΓ nik þ gijδmp Γ jkm ¼ gij,kδ jp whereby gnpΓ n ik þ gijΓ jkp ¼ gip,k ð2:4:5Þ and with the cyclic permutation of the free indexes i, p, k of expression (2.4.5) gnkΓ n pi þ gpjΓ jik ¼ gpk, i ð2:4:6Þ gniΓ n kp þ gkjΓ jpi ¼ gki,p ð2:4:7Þ Multiplying expression (2.4.5) by �1/2 and expressions (2.4.6) and (2.4.7) by 1/2 and adding � 1 2 gnpΓ n ik þ gijΓ jkp � � þ 1 2 gnkΓ n pi þ gpjΓ jik � � þ 1 2 gniΓ n kp þ gkjΓ jpi � � ¼ 1 2 gpk, i þ gki,p � gip,k � � and with the change of the index n for the index j, and considering the symmetry of the metric tensor gkjΓ j ip ¼ 1 2 gpk, i þ gki,p � gip,k � � The term to the right of the expression shows the existence of coefficients that are functions only of the partial derivatives of the metric tensor that define the Christoffel symbol of first kind p; k½ � ¼ Γip,k ¼ 1 2 gpk, i þ gki,p � gip,k � � ¼ 1 2 ∂gpk ∂xi þ ∂gki ∂xp � ∂gip ∂xk � � ð2:4:8Þ Multiplying expression (2.4.8) by gkm: gkmgkjΓ j ip ¼ 1 2 gkm gpk, i þ gki,p � gip,k � � ) δmj Γ jip ¼ 1 2 gkm gpk, i þ gki,p � gip,k � � 82 2 Covariant, Absolute, and Contravariant Derivatives whereby m ip � ¼ Γmip ¼ 1 2 gkm gpk, i þ gki,p � gip,k � � ¼ 1 2 gkm ∂gpk ∂xi þ ∂gki ∂xp � ∂gip ∂xk � � ð2:4:9Þ The term to the right of this expression shows the existence of coefficients that depend on the partial derivatives of the metric tensor and the conjugate metric tensor. The coefficients represented by Γik‘ given by expressions (2.3.9) and (2.4.9) define Christoffel symbol of second kind. Expression (2.4.9) is more convenient for calculating these coefficients than expression (2.3.9). Multiplying the terms of expression (2.3.10) by gi: gi � gk, ‘ ¼ Γ jk‘gi � gj ¼ δ ijΓ jk‘ Γ ik‘ ¼ gi � gk, ‘ ð2:4:10Þ 2.3.2 RelationBetween the Christoffel Symbols Expressions (2.4.8) and (2.4.9) relate the two Christoffel symbols, i.e.: Γmij ¼ gkmΓij,k ð2:4:11Þ then the Christoffel symbol of second kind is the raising of the third index of the Christoffel symbol of first kind. Expression (2.4.10) written in terms of the Christoffel symbol of first kind is given by Γ kij ¼ gkpΓij, p ¼ gk � gi, j and multiplying the members by gkp gkpg kpΓij, p ¼ gkpgk � gi, j ) Γij, p ¼ gkpgk � gi, j ) Γij, p ¼ gk � gp � gk � gi, j then Γij, p ¼ gp � gi, j ð2:4:12Þ 2.3 Derivatives of the Basis Vectors 83 2.3.3 Symmetry For the Christoffel symbol of first kind Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gik ∂xj � ∂gij ∂xk � � Γji,k ¼ 1 2 ∂gik ∂xj þ ∂gjk ∂xi � ∂gji ∂xk � � and considering the symmetry of the metric tensor gik ¼ gki, gjk ¼ gkj, gij ¼ gji it results in Γij,k ¼ Γji,k then the Christoffel symbol of first kind is symmetrical with respect to the first two indexes, and with considering this symmetry results to the Christoffel symbol of second kind Γmij ¼ gkmΓij,k ¼ gkmΓji,k ¼ Γmji that is symmetrical in regard to a permutation of the lower indexes. 2.3.4 Cartesian Coordinate System For the Cartesian coordinate systems the elements of the metric tensor are gij ¼ δij, whereby for p ¼ 1, 2, . . . ,N it results in ∂gip ∂xp ¼ 0 By means of the definition of the Christoffel symbol of first kind Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gik ∂xj � ∂gij ∂xk � � ¼ 0 It is verified that the Christoffel symbol of second kind is cancelled, for Γ pij ¼ gpkΓij,k ¼ 0 then for the Cartesian, orthogonal, or oblique coordinate systems, all the terms of Γij,k and Γ p ij are null. 84 2 Covariant, Absolute, and Contravariant Derivatives 2.3.5 Notation The oldest notations for the Christoffel symbols are ij k � for the symbol of first kind, and ij k � for the symbol of second kind. Improving this notations Levi-Civita adopted the spelling [ij, k] and {ij, k}, which second symbol was later improved by various authors to k ij � , where the indexes were placed in more adequate and logical positions. This notation is well adopted, using the representation [ij, k] for the Christoffel symbol of first kind. Hermann Weyl used the Greek letter Γ to denote these symbols, which positions of the indexes indicates the kind it represents: Γij,k and Γkij. This symbology is known as the notation of the Princeton School. A few authors invert the position of the indexes and write Γk,ij. The argument for adopting the notations [ij, k] and k ij � is that the Princeton School notation leads to confusing these coefficients with a tensor. However, this argument does not make its adoption valid, for the use of brackets or keys could also lead to confusion with a matrix or column matrix. That is not the case. The use of the notations Γij,k and Γkij, even not being universally accepted, has in its favor the economy of characters in a text with many expressions containing these symbols. Several authors do not use the comma for indicating the differentiation with respect to one of the indexes in the Christoffel symbol of first kind, and write Γijk. 2.3.6 Number of Different Terms For the tensorial space EN where i, j ¼ 1, 2, . . . ,N, it is verified that the metric tensor gij has N 2 terms gij ¼ g11 g12 � � � g1N g21 g22 � � � g2N ⋮ ⋮ ⋮ ⋮ gN1 gN2 � � � gNN 2664 3775 N�N This matrix has N diagonal terms gii, so N 2 � N� � terms remain in their sides. As gij is symmetrical, the result is 1 2 N2 � N� � terms for i 6¼ j. The total of different terms in the metric tensor is N2 þ 1 2 N2 � N� � ¼ N Nþ1ð Þ 2 . For each kind of Christoffel symbol N derivatives of gij are calculated, so the number of different terms for these coefficients is given by N N2þ1ð Þ 2 . 2.3 Derivatives of the Basis Vectors 85 2.3.7 Transformation of the Christoffel Symbol of First Kind Let Γpq,r defined in the coordinate system X i and Γij,k expressed in the coordinate systemX i , then using the expression that defines the Christoffel symbol of first kind, and adopting the notation gij,k ¼ ∂gij∂xk it follows that gij,k ¼ ∂ ∂xk gpq ∂xp ∂xi ∂xq ∂xj � � ¼ ∂gpq ∂xk ∂xp ∂xi ∂xq ∂xj þ gpq ∂2xp ∂xk∂xi ∂xq ∂xj þ gpq ∂xp ∂xi ∂2xq ∂xk∂xj ð2:4:13Þ By the chain rule ∂gpq ∂xk ¼ ∂gpq ∂xr ∂xr ∂xk ¼ gpq, r ∂xr ∂xk ð2:4:14Þ gij,k ¼ gpq, r ∂xp ∂xi ∂xq ∂xj ∂xr ∂xk þ gpq ∂2xr ∂xk∂xi ∂xs ∂xj ! þ gpq ∂2xs ∂xk∂xj ∂xr ∂xi ! ð2:4:15Þ and with cyclic permutation of the indexes in each term of the previous expression and with gqp ¼ gpq gjk, i ¼ gqr,p ∂xq ∂xj ∂xr ∂xk ∂xp ∂xi þ gpq ∂2xq ∂xi∂xj ∂xp ∂xk ! þ gpq ∂2xp ∂xi∂xk ∂xq ∂xj ! ð2:4:16Þ gki, j ¼ grp,q ∂xr ∂xk ∂xp ∂xi ∂xq ∂xj þ gpq ∂2xp ∂xj∂xk ∂xq ∂xi ! þ gpq ∂2xp ∂xj∂xi ∂xq ∂xk ! ð2:4:17Þ Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gki ∂xj � ∂gij ∂xk � � ð2:4:18Þ By substitution Γij,k ¼ ∂x p ∂xi ∂xq ∂xj ∂xr ∂xk Γpq, r þ gpq ∂2xp ∂xi∂xj ∂xq ∂xk ð2:4:19Þ that is the transformation law of the Christoffel symbol of first kind. The second term to the right of this expression shows that these coefficients are not the components of a tensor. 86 2 Covariant, Absolute, and Contravariant Derivatives 2.3.8 Transformation of the Christoffel Symbol of Second Kind Writing the Christoffel symbol of second kind in terms of the components of a new coordinate system X i : Γ i jk ¼ gipΓjk,p ¼ gqr ∂xj ∂xq ∂xp ∂xr � � Γjk,p ð2:4:20Þ As the Christoffel symbol of first kind transforms by mean of expression (2.4.19) Γjk,p ¼ Γ‘m,n ∂x ‘ ∂xj ∂xm ∂xk ∂xn ∂xp þ g‘m ∂2x‘ ∂xj∂xk ∂xm ∂xp ð2:4:21Þ and substituting expression (2.4.21) in expression (2.4.20) it follows that Γ i jk ¼ gqr Γrn ∂x‘ ∂xj ∂xm ∂xk ∂xn ∂xp ∂xi ∂xq ∂xp ∂xr � � þ gqrg‘m ∂2x‘ ∂xj∂xk ∂xm ∂xp ∂xi ∂xq ∂xp ∂xr ! Γ i jk ¼ gqr Γ‘m,nδnr ∂x‘ ∂xj ∂xm ∂xk ∂xi ∂xq þ gqrg‘mδmr ∂2x‘ ∂xj∂xk ∂xi ∂xq ! Γ i jk ¼ gqrΓ‘m,nδnr ∂x‘ ∂xj ∂xm ∂xk ∂xi ∂xq þ gqrg‘mδmr ∂2x‘ ∂xj∂xk ∂xi ∂xq Γ i jk ¼ gqrΓ‘m,nδnr ∂x‘ ∂xj ∂xm ∂xk ∂xi ∂xq þ gqrg‘r ∂2x‘ ∂xj∂xk ∂xi ∂xq gqrΓ‘m, r ¼ Γ q‘m gqrg‘r ¼ δq‘ Γ i jk ¼ Γ q‘m ∂x‘ ∂xj ∂xm ∂xk ∂xi ∂xq þ ∂ 2 x‘ ∂xj∂xk ∂xi ∂xq Replacing the dummy indexes ‘ ! q, q ! p, m ! r results in Γ i jk ¼ Γ pqr ∂xq ∂xj ∂xr ∂xk ∂xi ∂xp þ ∂ 2 xq ∂xj∂xk ∂xi ∂xp ð2:4:22Þ Expression (2.4.22) is the transformation law of the Christoffel symbol of second kind. The second term to the right of this expression shows that these coefficients are not the components of tensor. The Christoffel symbols do not depend only on the coordinate system, but depend also on the rate with which this coordinate system varies in each point of the space. This variation rate is not present in the transformation law of tensors. 2.3 Derivatives of the Basis Vectors 87 2.3.9 Linear Transformations Consider the transformation of coordinates between two coordinate systems given by linear relation xj ¼ aji x i þ b j where a ji and b j are constants, and with the techniques of successive differentiation ∂xj ∂xi ¼ aji ) ∂2xj ∂xi∂xk ¼ 0 then for this kind of transformation of coordinates the Christoffel symbols trans- form as tensors. 2.3.10 Orthogonal Coordinate Systems In the orthogonal coordinate systems, the tensorial space EN is defined by the metric tensor gij 6¼ 0 for i ¼ j and gij ¼ 0 for i 6¼ j. Putting h2i ¼ gii, where gii does not indicate the summation of the terms, with the Christoffel symbol of first kind Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gik ∂xj þ ∂gij ∂xk � � and with the components of the tensor gij given by g‘k ¼ 1 ! i ¼ j 0 ! i 6¼ j ( gij ¼ 1 gij it results for the relation between the Christoffel symbols Γ ‘ij ¼ g‘kΓij,k ¼ gkkΓij,k ! ‘ ¼ k 0 !! ‘ 6¼ k ( Varying the indexes: – i ¼ j ¼ k 88 2 Covariant, Absolute, and Contravariant Derivatives Γii, i ¼ 1 2 ∂gii ∂xj þ ∂gii ∂xj � ∂gii ∂xj � � ¼ 1 2 ∂gii ∂xj ¼ hi ∂hi∂xj Γ kij ¼ Γ iii ¼ giiΓii, i ¼ 1 2gii ∂gii ∂xj ¼ 1 2 ∂ ‘ngiið Þ ∂xj ¼ ∂ ‘n ffiffiffiffiffi gii p� � ∂xj ¼ 1 hi ∂hi ∂xj – i ¼ j 6¼k Γii,k ¼ 1 2 ∂gik ∂xi þ ∂gik ∂xi � ∂gii ∂xk � � as i 6¼ k it implies by definition of the metric tensor that gik ¼ 0, so Γii,k ¼ �1 2 ∂gii ∂xk ¼ �hi ∂hi∂xk Γ kij ¼ Γ kii ¼ gkkΓii,k ¼ � 1 2gkk ∂gii ∂xk ¼ � hi hkð Þ2 ∂hi ∂xk – i ¼ k 6¼ j Γij, i ¼ 1 2 ∂gji ∂xi þ ∂gii ∂xj þ ∂gij ∂xi � � and in an analogous way to the previous case where gji ¼ gij ¼ 0, so Γij, i ¼ 1 2 ∂gii ∂xj ¼ hi ∂hi∂xj Γ kij ¼ Γ iij ¼ giiΓij, i ¼ 1 2gii ∂gii ∂xj ¼ ∂ ‘n ffiffiffiffiffi gii p� � ∂xj ¼ 1 hi ∂hi ∂xj – for i 6¼ j, j 6¼ k, i 6¼ k it results in Γij,k ¼ 0, for by the definition of the metric tensor it implies gij ¼ gij ¼ 0 if i 6¼ j, whereby Γ kij ¼ 0. 2.3.11 Contraction The tensorial expressions at times contain Christoffel symbols. However, the calculation of their components can be avoided, for an expression can be obtained that relates the derivative of the natural logarithm of the metric tensor with these symbols. 2.3 Derivatives of the Basis Vectors 89 Let the Christoffel symbol of second kind Γ jik ¼ 1 2 gmj ∂gkm ∂xi þ ∂gmi ∂xk � ∂gik ∂xm � � and with contraction of the indexes j and k Γ jij ¼ 1 2 gmj ∂gjm ∂xi þ ∂gmi ∂xj � ∂gij ∂xm � � The symmetry of the metric tensor provides gmj ∂gmi ∂xj ¼ gjm ∂gji ∂xm ¼ gmj ∂gij ∂xm where the second equality was obtained by means of indexes interchanging the m $ j. Substituting gmj ∂gmi ∂xj in the expression of the contracted Christoffel symbol Γ jij ¼ 1 2 gjm ∂gjm ∂xi The conjugate metric tensor can be written as gjm ¼ G jm g ) g ¼ Gjmgjm being G jm the cofactor of this matrix and g ¼ detgjm, it follows that Γ jij ¼ 1 2g Gjm ∂gjm ∂xi ¼ 1 2g ∂g ∂xi ¼ 1 2 ∂ ‘ngð Þ ∂xi ¼ ∂ ‘n ffiffiffi g p� � ∂xi and with the contracted form of the Christoffel symbol of second kind Γ jij ¼ 1ffiffiffi g p ∂ ffiffiffi g p� � ∂xi ð2:4:23Þ that is of great use in manipulations of tensorial expressions, for it reduces the algebrism in calculating the Christoffel symbol. For g ¼ jgijj < 0 the analysis is analogous, having only to change the sign of the determinant in the expression shown in the previous demonstration Γ iip ¼ ∂ ‘n ffiffiffiffiffiffiffi�gp� � ∂xp ¼ 1ffiffiffiffiffiffiffi�gp ∂ ffiffiffiffiffiffiffi�gp� � ∂xp ð2:4:24Þ 90 2 Covariant, Absolute, and Contravariant Derivatives 2.3.12 Christoffel Relations Consider the transformation of the Christoffel symbol of second kind from one coordinate system Xi to another coordinate systemX j , whereby rewriting expression (2.4.22) Γ r pq ¼ Γmij ∂xr ∂xm ∂xi ∂xp ∂xj ∂xq þ ∂x r ∂xj ∂2xj ∂xq∂xp and multiplying by ∂x s ∂xr it follows that ∂xs ∂xr Γ r pq ¼ Γmij ∂xs ∂xr ∂xr ∂xm ∂xi ∂xp ∂xj ∂xq þ ∂x s ∂xr ∂xr ∂xj ∂2xj ∂xq∂xp ∂xs ∂xr ∂xr ∂xm ¼ δ sm ∂xs ∂xr ∂xr ∂xj ¼ δ sj ∂xs ∂xr Γ r pq ¼ Γmij δ sm ∂xi ∂xp ∂xj ∂xq þ δ sj ∂2xj ∂xq∂xp ) ∂x s ∂xr Γ r pq ¼ Γ sij ∂xi ∂xp ∂xj ∂xq þ ∂ 2 xs ∂xq∂xp ∂2xs ∂xq∂xp ¼ ∂x s ∂xr Γ r pq � Γ sij ∂xi ∂xp ∂xj ∂xq ð2:4:25Þ Expression (2.4.25) shows that the second derivative of the coordinate xs can be decomposed into terms with the first derivatives of this coordinate and the coordi- nates xi, xj, and with the Christoffel symbols of second kind. This important expression was deducted in 1869 by Elwin Bruno Christoffel. Let an inverse transformation for the Christoffel symbol of second kind of the coordinate system X j to another referential system Xi given by Γ j‘k ¼ Γ r pq ∂xp ∂x‘ ∂xq ∂xk ∂xj ∂xr þ ∂x j ∂xr ∂2xr ∂x‘∂xk and multiplying both members by ∂x m ∂xj and proceeding in a manner that is analogous to the previous one ∂2xm ∂x‘∂xk ¼ ∂x m ∂xj Γ j‘k � Γ m pq ∂xp ∂x‘ ∂xq ∂xk ð2:4:26Þ The transformation of the Christoffel symbols from one coordinate system Xi to another coordinate system X j , and from this one to a third coordinates system eXk is identical to the transformation from Xi directly to eXk, so the transitive property is valid for the transformations of the Christoffel symbols. This shows that these symbols form a group. 2.3 Derivatives of the Basis Vectors 91 The Christoffel relation given by expression (2.4.25) can be written as ∂2xk ∂xj∂xm ∂xs ∂xk ¼ Γ sjm � ∂xs ∂xp ∂xk ∂xj ∂xr ∂xm Γ pkr and contracting the terms in the indexes s and m ∂2xk ∂xj∂xm ∂xm ∂xk ¼ Γmjm � ∂xm ∂xp ∂xk ∂xj ∂xr ∂xm Γ pkr ∂2xk ∂xj∂xm ∂xm ∂xk ¼ Γmjm � δ rp ∂xk ∂xj Γ pkr Γ m jm ¼ ∂xk ∂xj Γ rkr þ ∂2xk ∂xj∂xm ∂xm ∂xk that is the transformation law of the contracted Christoffel symbol of second kind. 2.3.13 Ricci Identity Another usual expression in Tensor Calculus is obtained by means of defining the Christoffel symbol of first kind Γji,k ¼ 1 2 ∂gik ∂xj þ ∂gjk ∂xi � ∂gij ∂xk � � Γki, j ¼ 1 2 ∂gij ∂xk þ ∂gkj ∂xi � ∂gik ∂xj � � The sum of these two expressions provides the Ricci identity ∂gjk ∂xi ¼ Γji,k þ Γki, j ð2:4:27Þ In an analogous way, subtracting the second expression from the first expression of the Christoffel symbol of first kind provides ∂gij ∂xk � ∂gjk ∂xi ¼ Γkj, i � Γij,k ð2:4:28Þ Expressions (2.4.27) and (2.4.28) are very useful in manipulations of tensorial equations. Exercise 2.3 If Tij and gik are the components of a symmetric tensor and the metric tensor, respectively, show that TjkΓij,k ¼ 12 Tjk ∂gjk ∂xi . 92 2 Covariant, Absolute, and Contravariant Derivatives With the Ricci identity ∂gik ∂xi ¼ Γji,k þ Γki, j 1 2 Tjk ∂gjk ∂xi ¼ 1 2 Tjk Γji,k þ Γki, j � � ¼ 1 2 TjkΓji,k þ TjkΓki, j � � Interchanging the indexes j $ k in the last term to the right of the expression and considering the tensor’s symmetry then 1 2 Tjk ∂gjk ∂xi ¼ 1 2 Tjk Γji,k þ Γki, j � � ¼ 1 2 TjkΓji,k þ TkjΓji,k � � ¼ 1 2 � 2TjkΓji,k 1 2 Tjk ∂gjk ∂xi ¼ TjkΓji,k Q:E:D: 2.3.14 Fundamental Relations The derivative of the metric tensor with respect to an arbitrary variable can be placed in terms of Christoffel symbols of second kind and the metric tensor, thus from the definition of this symbol Γ pik ¼ gpjΓik, j Γ pjk ¼ gpiΓjk, i Multiplying these two expressions by gpj and gpi, respectively: gpjΓ p ik ¼ gpjgpjΓik, j ) gpkΓ pik ¼ δpjΓik, j gpiΓ p jk ¼ gpigpiΓjk, i ) gpiΓ pjk ¼ δpiΓjk, i whereby Γik, j ¼ gpjΓ pik Γjk, i ¼ gipΓ pjk Adding these two expressions and considering the Ricci identity ∂gik ∂xi ¼ Γji,k þ Γki, j ð2:4:29Þ and as gip ¼ gpi it results in ∂gij ∂xk ¼ gpjΓ pik þ gipΓ pjk ð2:4:30Þ 2.3 Derivatives of the Basis Vectors 93 With analogous analysis this derivative can be placed in terms of Christoffel symbols of the second kind and the conjugate metric tensor, and with gijgkj ¼ δ ik the derivative is ∂gij ∂xp gkj þ gij ∂gkj ∂xp ¼ 0 ) ∂g ij ∂xp gkj ¼ �gij ∂gkj ∂xp Multiplying both members of this last expression by gkq: gkqgkj ∂gij ∂xp ¼ �gkqgij ∂gkj ∂xp ) δqj ∂gij ∂xp ¼ �gkqgij ∂gkj ∂xp ) ∂g iq ∂xp ¼ �gijgkq ∂gkj ∂xp and with the Ricci identity ∂gkj ∂xp ¼ Γkp, j þ Γjp,k that substituted in the previous expression provides ∂giq ∂xp ¼ �gijgkq Γkp, j þ Γjp,k � � ¼ �gijgkqΓkp, j � gijgkqΓjp,k For the Christoffel symbol of second kind the result is the following relations gkqgijΓkp, j ¼ gkqΓ ikp gijgkqΓjp,k ¼ gijΓ qjp whereby ∂giq ∂xp ¼ �gkqΓ ikp � gijΓ qjp As Γ qjp ¼ Γ qpj, the result is ∂giq ∂xp ¼ �gijΓ qpj � gkqΓ ikp ð2:4:31Þ Expressions (2.4.30) and (2.4.31) are well used in the development of tensorial expressions. Exercise 2.4 Calculate the Christoffel symbols Γijk and Γkij for the polar coordi- nates systems, which metric tensor is given by 94 2 Covariant, Absolute, and Contravariant Derivatives gij ¼ 1 00 x1ð Þ2 � The Christoffel symbol of first kind is given by Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gik ∂xj � ∂gij ∂xk � � so g11 ¼ 1 ) g11,1 ¼ 0, g11,2 ¼ 0 g22 ¼ x1ð Þ2 ) g22,1 ¼ 2x1, g22,2 ¼ 0 g12,1 ¼ g12,2 ¼ g21,1 ¼ g21,2 ¼ g22,2 ¼ 0 It follows that Γ11,1 ¼ Γ11,2 ¼ Γ12,1 ¼ Γ21,1 ¼ Γ22,2 ¼ 0 Γ12,2 ¼ Γ21,2 ¼ x1 Γ22,1 ¼ �x1 In matrix form the result is Γij, 1 ¼ 0 00 �x1 � Γij, 2 ¼ 0 x 1 x1 0 � For the Christoffel symbol of second kind it follows thatgij h i�1 ¼ gij ¼ 1 0 0 1 x1ð Þ2 24 35 Γ k12 ¼ gk2Γ12,2 Γ k22 ¼ gk1Γ22,1 k ¼ 1 ) Γ112 ¼ g12Γ12,2 ¼ 0 Γ122 ¼ g11 �x1ð Þ ¼ �x1 ( In matrix form the result is Γ1ij ¼ 0 0 0 �x1 � Γ2ij ¼ 0 1 x1 1 x1 0 264 375 2.3 Derivatives of the Basis Vectors 95 Exercise 2.5 Calculate the Christoffel symbols for the cylindrical coordinates system, defined by r � x1 , θ � x2 , z � x3, where �1 r 1, 0 θ 2π , �1 z 1, which metric tensor and its conjugated tensor are given, respec- tively, by gij ¼ 1 0 0 0 r2 0 0 0 1 24 35 gij ¼ 1 0 0 0 1 r2 0 0 0 1 2664 3775 Using the expressions deducted for the orthogonal coordinate systems: – i ¼ j ¼ k Γii, i ¼ 1 2 ∂gii ∂xi Γ11,1 ¼ 1 2 g11,1 ¼ 0 Γ22,2 ¼ 1 2 g22,2 ¼ 0 Γ33,3 ¼ 1 2 g33,3 ¼ 0 – i ¼ j 6¼ k Γii,k ¼ �1 2 ∂gii ∂xk Γ11,2 ¼ �1 2 g11,2 ¼ 0 Γ11,3 ¼ � 1 2 g11,3 ¼ 0 Γ22,1 ¼ �1 2 g22,1 ¼ �r Γ22,3 ¼ � 1 2 g22,3 ¼ 0 Γ33,1 ¼ �1 2 g33,1 ¼ 0 Γ33,2 ¼ � 1 2 g33,2 ¼ 0 – i ¼ k 6¼ j Γij, i ¼ 1 2 ∂gii ∂xj Γ12,1 ¼ 1 2 g11,2 ¼ 0 Γ13,1 ¼ 1 2 g11,3 ¼ 0 Γ21,2 ¼ 1 2 g22,1 ¼ r Γ23,2 ¼ 1 2 g22,3 ¼ 0 Γ31,3 ¼ 1 2 g33,1 ¼ 0 Γij, i ¼ 1 2 g33,2 ¼ 0 – i 6¼ j, j 6¼ k, i 6¼ k all the Christoffel symbols are null. Putting these symbols in matrix form, the result is 96 2 Covariant, Absolute, and Contravariant Derivatives Γij, 1 ¼ 0 0 0 0 �r 0 0 0 0 24 35 Γij, 2 ¼ 0 r 0r 0 0 0 0 0 24 35 Γij, 3 ¼ 0½ � For the Christoffel symbols of second kind it follows that Γ pij ¼ gpkΓij,k Γ p22 ¼ gp1Γ22,1 Γ p12 ¼ gp2Γ12,2 – p ¼ 1 Γ122 ¼ g11Γ22,1 ¼ r Γ112 ¼ g12Γ12,2 ¼ 0 – p ¼ 2 Γ222 ¼ g21Γ22,1 ¼ 0 Γ212 ¼ g22Γ12,2 ¼ 1 r – p ¼ 3 Γ322 ¼ g31Γ22,1 ¼ 0 Γ313 ¼ g32Γ12,2 ¼ 0 Putting these symbols in matrix form, the result is Γ1ij ¼ 0 0 0 0 �r 0 0 0 0 264 375 Γ2ij ¼ 0 1 r 0 1 r 0 0 0 0 0 266664 377775 Γ3ij ¼ 0½ � Exercise 2.6 Calculate the Christoffel symbols for the spherical coordinates sys- tem r � x1,φ � x2, θ � x3, where �1 r 1, 0 φ π, 0 θ 2π, which metric tensor and its conjugated tensor are given, respectively, by gij ¼ 1 0 0 0 r2 0 0 0 r2 sin 2φ 24 35 gij ¼ 1 0 0 0 1 r2 0 0 0 1 r2 sin 2φ 266664 377775 Using the expressions deduced for the orthogonal coordinate systems the result is: – i ¼ j ¼ k 2.3 Derivatives of the Basis Vectors 97 Γii, i ¼ 1 2 ∂gii ∂xi Γ11,1 ¼ 1 2 g11,1 ¼ 0 Γ22,2 ¼ 1 2 g22,2 ¼ 0 Γ33,3 ¼ 1 2 g33,3 ¼ 0 – i ¼ j 6¼ k Γii,k ¼ �1 2 ∂gii ∂xk Γ11,2 ¼ �1 2 g11,2 ¼ 0 Γ11,3 ¼ � 1 2 g11,3 ¼ 0 Γ22,1 ¼ �1 2 g22,1 ¼ �r Γ22,3 ¼ � 1 2 g22,3 ¼ 0 Γ33,1 ¼ �1 2 g33,1 ¼ �r sin 2φ Γ33,2 ¼ � 1 2 g33,2 ¼ �r2 sinφ cosφ – i ¼ k 6¼ j Γij, i ¼ 1 2 ∂gii ∂xj Γ12,1 ¼ 1 2 g11,2 ¼ 0 Γ13,1 ¼ 1 2 g11,3 ¼ 0 Γ21,2 ¼ 1 2 g22,1 ¼ r Γ23,2 ¼ 1 2 g22,3 ¼ 0 Γ31,3 ¼ 1 2 g33,1 ¼ r sin 2φ Γij, i ¼ 1 2 g33,2 ¼ r2 sinφ cosφ – i 6¼ j, j 6¼ k, i 6¼ k all the Christoffel symbols are null. Putting these symbols in matrix form, the result is Γij, 1 ¼ 0 0 0 0 �r 0 0 0 r sin 2φ 24 35 Γij, 2 ¼ 0 r 0r 0 0 0 0 �r2 sinφ cosφ 24 35 Γij, 3 ¼ 0 0 r sin 2φ 0 0 r2 sinφ cosφ r sin 2φ r2 sinφ cosφ 0 24 35 For the Christoffel symbols of second kind it follows that Γ pij ¼ gpkΓj,k – p ¼ k ¼ 1 98 2 Covariant, Absolute, and Contravariant Derivatives Γ1ij ¼ g11Γij, 1 Γ122 ¼ g11Γ22,1 ¼ �r Γ133 ¼ g11Γ33,1 ¼ �r sin 2φ – p ¼ k ¼ 2 Γ2ij ¼ g22Γij, 2 Γ212 ¼ g22Γ12,2 ¼ 1 r Γ233 ¼ g22Γ33,2 ¼ � sinφ cosφ – p ¼ k ¼ 3 Γ3ij ¼ g33Γij, 3 Γ313 ¼ g33Γ13,3 ¼ 1 r Γ323 ¼ g33Γ23,3 ¼ cot φ Putting these symbols in matrix form, the result is Γ1ij ¼ 0 0 0 0 �r 0 0 0 �r sin 2φ 264 375 Γ2ij ¼ 0 1 r 0 1 r 0 0 0 0 sinφ cosφ 266664 377775 Γ3ij ¼ 0 0 1 r 0 0 cotφ 1 r cotφ 0 26664 37775 Exercise 2.7 For the antisymmetric tensor Aijk, show that AijkΓ pij ¼ AijkΓ pjk ¼ Aijk Γ pik ¼ 0. The symmetry of the Christoffel symbol of second kind allows writing AijkΓ pij ¼ AijkΓ pji ¼ 0 and replacing the indexes i ! j it follows that AijkΓ pij ¼ �AijkΓ pij the result is AijkΓ pij ¼ 0 2.3 Derivatives of the Basis Vectors 99 Proceeding in an analogous way for AijkΓpjk and A ijkΓpik the equalities of what was enunciated are verified. Exercise 2.8 Given the expression Γ i jk ¼ Γ ijk þ δ ij uk þ δ ikuj, where ui is a covariant vector and Aij is an antisymmetric tensor, show that AjkΓ i jk ¼ 0. The symmetry of the Christoffel symbol of second kind allows writing Γ i jk ¼ Γ ikj Γ i jk ¼ Γ ikj þ δ ikuj þ δ ij uk AjkΓ i jk ¼ AkjΓ ikj and with the consideration of the anti-symmetry Ajk ¼ �Akj verifies that AjkΓ i jk ¼ 0 Q:E:D: 2.4 Covariant Derivative The basic problem treated by the Tensorial Analysis is to research if the derivatives of tensors generate new tensors, which, in general, does not occur. For the case of Cartesian coordinate systems the variation rates of the tensors are expressed by partial derivatives. For instance, the variation rates of a vector’s components indicate the variation of this vector. However, for the curvilinear coordinate systems, the expressions for these variation rates are not expressed only by partial derivatives. Figure 2.3a shows this coordinate systems for the case in which the vector u has constant modulus and directions (Fig. 2.3a), but their components u1 vary. Figure 2.3b shows the behavior of the vectors u with constant modulus and different directions, whereby the three vectors are different, but their radial u1 and tangential u2 ¼ 0 components remain constant. This example indicates the need for AA BB CC u u u uu u 1X 1X 2X 2X OO 1u 1u 1u θ θ a b Fig. 2.3 Polar coordinates: (a) vector uwith constant modulus and direction and (b) vector uwith constant modulus and variable direction 100 2 Covariant, Absolute, and Contravariant Derivatives researching the variation rates of vectors for the curvilinear coordinate systems, because the variation rates of their components do not represent the variation of these vectors. To exemplify this fact let the scalar function ϕ ¼ �mx, where m is a scalar, which generates the potential u ¼ �gradϕ, with Cartesian components u1 ¼ m, u2 ¼ 0. This scalar function in polar coordinates is defined by ϕ ¼ �mr cos θ, which covariant components of its gradient are given by ∂ϕ ∂r ¼ �m cos θ ∂ϕ ∂θ ¼ mr sin θ and its physical components areu*1 ¼ ∂ϕ∂r ¼ �m cos θ andu*2 ¼ 1r ∂ϕ∂θ ¼ m sin θ. These components are not constant. The interpretation of this variation is carried out admitting a polar coordinates point P(r; θ) being displaced to another point nearby P 0 r þ dr; θ þ dθð Þ, so the covariant components of the vector u initially given by u1 ¼ ∂ϕ∂r ¼ �m cos θ u2 ¼ ∂ϕ ∂θ ¼ mr sin θ stay for this new point δu1 ¼ �m sin θdθ δu2 ¼ m sin θdr þ mr cos θdθ The elemental variations of these new components are due to the change of coordinates, and not to the change of vector. This particular indicates the need of defining a kind of derivative that translates the vector’s variation in an invariant manner, and leads to the definition of the covariant derivative. The covariant derivative defines variation rate of parameters that are not depen- dent on the coordinate systems, and because of that it is of extreme importance in the expression of physical models, for it generates a new tensor. The denomination covariant derivative was adopted by Bruno Ricci-Curbastro when conceiving the Tensor Calculus. The term covariant denotes a kind of partial differentiation of tensors that generates new tensors with variance one order above the original tensors. The adjective covariant is used to indicate the tensorial characteristics of the differentiation of tensors, in which the set of Christoffel symbols Γkij are the coefficients of connections of the tensorial space EN. 2.4.1 Contravariant Tensor 2.4.1.1 Contravariant Vector Let the vector u defined by its contravariant components uj: u ¼ ujgj ð2:5:1Þ where the unit vectors gj ¼ gj xjð Þ of the curvilinear coordinate system are functions of the coordinates that define this referential system. 2.4 Covariant Derivative 101 Differentiating the expression (2.5.1)with respect to an arbitrary coordinate xk results in ∂u ∂xk ¼ ∂u j ∂xk gj þ uj ∂gj ∂xk and using expression (2.4.1) ∂gj ∂xk ¼ Γmjkgm then ∂u ∂xk ¼ ∂u j ∂xk gj þ ujΓmjkgm ð2:5:2Þ As j is a dummy index in the first term to the right, it can be changed for the index m: ∂u ∂xk ¼ ∂u m ∂xk gm þ ujΓmjkgm ¼ ∂um ∂xk þ ujΓmjk � � gm This expression shows that the covariant derivative of a contravariant vector is given by the N2 functions ∂ku m ¼ ∂u m ∂xk þ ujΓmjk ð2:5:3Þ whereby ∂u ∂xk ¼ ∂kumð Þgm ð2:5:4Þ For the Cartesian systems the Christoffel symbols are null, so the covariant derivative coincides with the partial derivative ∂u m ∂xk . In expression (2.5.3) the result is the variation rate of the vector u along the axes of the curvilinear coordinate system is given by ∂u j ∂xk, and the variation of the unit vectors gj along the axes of this coordinate system is expressed by ∂gj ∂xk. This physical interpretation of the covariant derivative is associated to the Christoffel symbols Γmjk that are the connection coefficients of the tensorial space. Various notations are found in the literature for the term to the left of expression (2.5.3), the most usual being: ∂kum ¼ Dkum ¼ ∇kum ¼ umjk ¼ umjk ¼ um;k. 102 2 Covariant, Absolute, and Contravariant Derivatives Expressions (2.5.1) and (2.5.4) are analogous, for∂kum has the aspect of a vector. The transformation law of contravariant vectors is admitted to demonstrate that expression (2.5.4) is a tensor, thus ui ¼ ∂x i ∂xp up which differentiated with respect to the coordinate xj provides ∂ui ∂xj ¼ ∂x i ∂xp ∂up ∂xq ∂xq ∂xj � � þ up ∂ 2 xi ∂xq∂xp ∂xq ∂xj ! ð2:5:5Þ and with expression (2.4.26) ∂2xi ∂xq∂xp ¼ Γmpq ∂xi ∂xn � Γ i‘m ∂x‘ ∂xp ∂xm ∂xq ð2:5:6Þ Substituting expression (2.5.6) in expression (2.5.5) ∂ui ∂xj ¼ ∂x i ∂xp ∂up ∂xq ∂xq ∂xj � � þ upΓ npq ∂xi ∂xn ∂xq ∂xj � upΓ i‘m ∂x‘ ∂xp ∂xm ∂xq ∂xq ∂xj ∂ui ∂xj þ upΓ i‘m ∂x‘ ∂xp ∂xm ∂xq ∂xq ∂xj ¼ ∂u p ∂xq ∂xi ∂xp ∂xq ∂xj þ upΓ npq ∂xi ∂xn ∂xq ∂xj The dummy index p in the first term to the right can be changed by the index n: ∂ui ∂xj þ upΓ i‘m ∂x‘ ∂xp ∂xm ∂xq ∂xq ∂xj ¼ ∂u n ∂xq ∂xi ∂xn ∂xq ∂xj þ upΓ npq ∂xi ∂xn ∂xq ∂xj and with expression ∂xm ∂xj ¼ δmj results in ∂ui ∂xj þ upΓ i‘m ∂x‘ ∂xp δmj ¼ ∂un ∂xq þ upΓ npq � � ∂xi ∂xn ∂xq ∂xj With the transformation law of contravariant vectors u‘ ¼ up ∂x ‘ ∂xp 2.4 Covariant Derivative 103 the above expression becomes ∂ui ∂xj þ u‘Γ i‘j ¼ ∂un ∂xq þ upΓ npq � � ∂xi ∂xn ∂xq ∂xj ð2:5:7Þ It is verified that in expression (2.5.7) the variety in parenthesis transforms as a mixed second-order tensor, then the covariant derivative of a contravariant vector is a mixed second-order tensor, i.e., of variance (1, 1). 2.4.2 Contravariant Tensor of the Second-Order The transformation law of contravariant tensors of the second-order is given by T pq ¼ Tij ∂x p ∂xi ∂xq ∂xj The derivative of this expression with respect to coordinate x‘ is given by ∂T pq ∂x‘ ¼ ∂T ij ∂xk ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj þ Tij ∂ 2 xp ∂xk∂xi ∂xk ∂x‘ ∂xq ∂xj þ Tij ∂x p ∂xi ∂2xq ∂xk∂xj ∂xk ∂x‘ and with expression (2.4.26) ∂2xp ∂xk∂xi ¼ Γ rki ∂xp ∂xr � Γ p‘m ∂x‘ ∂xi ∂xm ∂xk ∂2xq ∂xk∂xj ¼ Γ rkj ∂xq ∂xr � Γ q‘m ∂x‘ ∂xj ∂xm ∂xk then ∂T pq ∂x‘ ¼ ∂T ij ∂xk ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj þ Tij Γ rki ∂xp ∂xr � Γ p‘m ∂x‘ ∂xi ∂xm ∂xk � � ∂xk ∂x‘ ∂xq ∂xj þ Tij Γ rkj ∂xq ∂xr � Γ q‘m ∂x‘ ∂xj ∂xm ∂xk � � ∂xk ∂x‘ ∂xp ∂xi ∂T pq ∂x‘ ¼ ∂T ij ∂xk ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj þ TijΓ rki ∂xp ∂xr ∂xk ∂x‘ ∂xq ∂xj � TijΓ p‘m ∂x‘ ∂xi ∂xm ∂xk ∂xk ∂x‘ ∂xq ∂xj þ TijΓ rkj ∂xq ∂xr ∂xk ∂x‘ ∂xp ∂xi � TijΓ q‘m ∂x‘ ∂xj ∂xm ∂xk ∂xk ∂x‘ ∂xp ∂xi With δm‘ ¼ ∂xm ∂xk ∂xk ∂x‘ 104 2 Covariant, Absolute, and Contravariant Derivatives it follows that ∂T pq ∂x‘ ¼ ∂T ij ∂xk ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj þ TijΓ rki ∂xp ∂xr ∂xk ∂x‘ ∂xq ∂xj � TijΓ p‘m ∂x‘ ∂xi ∂xq ∂xj þ TijΓ rkj ∂xq ∂xr ∂xk ∂x‘ ∂xp ∂xi � TijΓ q‘m ∂x‘ ∂xj ∂xp ∂xi In the second term to the right interchanging the indexes i$ r and, likewise, in the same fourth term with the permutation of the indexes j$ r, it results in ∂T pq ∂x‘ ¼ ∂T ij ∂xk ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj þ Trj ∂x p ∂xi ∂xk ∂x‘ ∂xq ∂xj Γ ikr � Tij ∂x‘ ∂xi ∂xq ∂xj Γ p ‘m þ Tir ∂x q ∂xj ∂xk ∂x‘ ∂xp ∂xi Γ jkr � Tij ∂x‘ ∂xj ∂xp ∂xi Γ q ‘m and with the expressions T ‘q ¼ Tij ∂x ‘ ∂xi ∂xq ∂xj T ‘p ¼ Tij ∂x ‘ ∂xj ∂xp ∂xi it follows that ∂T pq ∂x‘ ¼ ∂T ij ∂xk þ TrjΓ ikr þ TirΓ jkr � � ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj � T‘qΓ p‘m � T ‘p Γ q ‘m ∂T pq ∂x‘ þ T‘qΓ p‘m þ T ‘p Γ q ‘m ¼ ∂Tij ∂xk þ TrjΓ ikr þ TirΓ jkr � � ∂xk ∂x‘ ∂xp ∂xi ∂xq ∂xj ∂‘T pq ¼ ∂kTij ∂x k ∂x‘ ∂xp ∂xi ∂xq ∂xj ð2:5:8Þ where ∂kT ij ¼ ∂T ij ∂xk þ TrjΓ ikr þ TirΓ jkr ð2:5:9Þ is the covariant derivative of the contravariant tensor of the second order. Expression (2.5.8) indicates that the covariant derivative of a contravariant tensor of the second order is a mixed tensor of the third order, twice contravariant and once covariant, i.e., variance (2, 1). For the Cartesian coordinates the Christoffel symbols are null, so the covariant derivative coincides with the partial derivative ∂T ij ∂xk . 2.4 Covariant Derivative 105 2.4.2.1 Contravariant Tensor of Order Above Two To generalize expression (2.5.9) for tensors of order above two, i.e., for instance, the covariant derivative of the contravariant tensor of the third order, which expression may be developed by means of the following steps: (a) The basic structure of its expression is written considering the expression obtained for the covariant derivative of a contravariant tensor of the second order ∂pT ijk ¼ ∂T ijk ∂xp þ T Γ þ T Γ þ T Γ (b) The indexes of the Christoffel symbols corresponding to the coordinate with respect to which the differentiation is being carried out are placed ∂pT ijk ¼ ∂T ijk ∂xp þ T Γ p þ T Γ p þ T Γ p (c) The tensor indexes sequence must be obeyed on placing the contravariant indexes of the Christoffel symbol ∂pT ijk ¼ ∂T ijk ∂xp þ T Γ i p þ T Γ j p þ T Γ k p (d) The dummy index q is placed on the Christoffel symbol and in sequential form in the tensors ∂pT ijk ¼ ∂T ijk ∂xp þ Tq Γ iqp þ T q Γ jqp þ T qΓ kqp (e) The remaining indexes are placed in the same sequence in which they appear on the tensor that is being differentiated ∂pT ijk ¼ ∂T ijk ∂xp þ TqjkΓ iqp þ TiqkΓ jqp þ TijqΓ kqp This tensor generated by the differentiation of a variance tensor (3, 0) has a variance (3, 1). Expression (2.5.9) can be generalized by adopting this indexes placement systematic for a contravariant tensor of order p > 3, then the variance of this new tensor will always be ( p, 1). Exercise 2.9 Calculate the covariant derivative of the contravariant components of vector u expressed in polar coordinates. In Exercise 2.4 the Christoffel symbols were calculated for the polar coordinates, given by 106 2 Covariant, Absolute, and Contravariant Derivatives Γij, 1 ¼ 0 00 �x1 � Γij, 2 ¼ 0 x 1 x1 0 � Γ1ij ¼ 0 0 0 �x1 � Γ2ij ¼ 0 1 x1 1 x1 0 264 375 The expression for the derivative of the contravariant components of vector u is: ∂ku m ¼ ∂u m ∂xk þ ujΓmjk – m ¼ 1 ∂ku 1 ¼ ∂u 1 ∂xk þ ujΓ1jk k ¼ 1 ) ∂1u1 ¼ ∂u 1 ∂x1 þ ujΓ1j1 ) ∂1u1 ¼ ∂u1 ∂x1 þ u1Γ111 þ u2Γ121 ∂1u 1 ¼ ∂u 1 ∂x1 þ 0þ 0 ¼ ∂u 1 ∂x1 k ¼ 2 ) ∂2u1 ¼ ∂u 1 ∂x2 þ ujΓ1j1 ) ∂2u1 ¼ ∂u1 ∂x2 þ u1Γ112 þ u2Γ122 ∂2u 1 ¼ ∂u 1 ∂x2 þ u 1 x1 þ 0 ¼ ∂u 1 ∂x2 þ u 1 x1 – m ¼ 2 ∂ku 2 ¼ ∂u 2 ∂xk þ ujΓ2jk k ¼ 1 ) ∂1u2 ¼ ∂u 2 ∂x1 þ ujΓ2j1 ) ∂1u2 ¼ ∂u2 ∂x1 þ u1Γ211 þ u2Γ221 ∂1u 2 ¼ ∂u 2 ∂x1 þ 0þ u 2 x1 ¼ ∂u 2 ∂x1 þ u 2 x1 k ¼ 2 ) ∂2u2 ¼ ∂u 2 ∂x2 þ ujΓ2j2 ) ∂2u2 ¼ ∂u2 ∂x2 þ u1Γ212 þ u2Γ222 ∂2u 2 ¼ ∂u 2 ∂x2 þ u 1 x1 þ 0 ¼ ∂u 2 ∂x2 þ u 1 x1 Exercise 2.10 Show that ∂jTij ¼ 1ffiffigp ∂ Tij ffiffigpð Þ∂xj þ TjpΓ ijp. The expression of the covariant derivative of a contravariant tensor of thesecond order is given by 2.4 Covariant Derivative 107 ∂jT ij ¼ ∂T ij ∂xk þ TmjΓ ikm þ TimΓ jmk and assuming k ¼ j ∂jT ij ¼ ∂T ij ∂xk þ TmjΓ ijm þ TimΓ jmj In the study of the contraction of the Christoffel symbol, it was verified that Γ imj ¼ ∂ ‘n ffiffiffi g p� � ∂xr Substituting this expression in the previous expression ∂jT ij ¼ ∂T ij ∂xk þ TmjΓ ijm þ Tim ∂ ‘n ffiffiffi g p� � ∂xm As m is a dummy index, it can be changed by the index j in the third term to the right ∂jT ij ¼ ∂T ij ∂xk þ TmjΓ ijm þ Tij ∂ ‘n ffiffiffi g p� � ∂xj ) ∂jTij ¼ ∂T ij ∂xk þ Tij ∂ 1 2 ‘ng � � ∂xj � þ TmjΓ ijm and multiplying and dividing the two terms between brackets by ffiffiffi g p ∂jT ij ¼ 1ffiffiffi g p ffiffiffigp ∂Tij ∂xk þ Tij 1 2 ffiffiffi g p ∂g ∂xj � � þ TmjΓ ijm Changing the indexes j ! p and m ! j in the last term ∂jT ij ¼ 1ffiffiffi g p ffiffiffigp ∂Tij ∂xk þ Tij 1 2 ffiffiffi g p ∂g ∂xj � � þ TjpΓ ipj ) ∂jTij ¼ 1ffiffiffi g p ∂ T ij ffiffiffigp� � ∂xj þ TjpΓ ipj By means of the symmetry of the Christoffel symbol it results ∂jT ij ¼ 1ffiffiffi g p ∂ T ij ffiffiffigp� � ∂xj þ TjpΓ ipj Q:E:D: 108 2 Covariant, Absolute, and Contravariant Derivatives 2.4.3 Covariant Tensor 2.4.3.1 Covariant Vector Let the vector u defined by their covariant components uj: u ¼ uigj ð2:5:10Þ where gj ¼ gj xj � � are the basis vectors of the curvilinear coordinate system, which are functions of the coordinates that define this referential system. Differentiating the expression (2.5.10) with respect to an arbitrary coordinate xk: ∂u ∂xk ¼ ∂ui ∂xk gi þ ui ∂g i ∂xk ð2:5:11Þ and substituting expression (2.4.4) ∂gi ∂xk ¼ �Γ ikjgj in expression (2.5.11) the result is ∂u ∂xk ¼ ∂ui ∂xk gi � uiΓ ikjgj ð2:5:12Þ As i is a dummy index in the first term to the right of expression (2.5.12), it can be changed by j: ∂u ∂xk ¼ ∂uj ∂xk � uiΓ ikj � � gj ð2:5:13Þ thus the covariant derivative of a covariant vector is given by the N2 functions ∂kuj ¼ ∂uj∂xk � u iΓ ikj ð2:5:14Þ whereby ∂u ∂xk ¼ ∂kuj � � gj ð2:5:15Þ For the Cartesian coordinate systems the Christoffel symbols are null, so in these referential systems the covariant derivative of a covariant vector coincides with the partial derivative ∂uj ∂xk. 2.4 Covariant Derivative 109 Expression (2.5.15) has the aspect of a vector, and to demonstrate that this expression is a tensor let the transformation law of covariant vectors up ¼ ∂x i ∂xp ui that differentiated with respect to the coordinate xq provides ∂up ∂xp ¼ ∂ui ∂xk ∂xk ∂xq ∂xi ∂xq þ ui ∂ 2 xi ∂xq∂xp ð2:5:16Þ Expression (2.4.25) can be written as ∂2xi ∂xq∂xp ¼ ∂x i ∂xs Γ s pq � Γ ijk ∂xj ∂xp ∂xk ∂xq and substituting this expression in expression (2.5.15) ∂up ∂xq ¼ ∂ui ∂xk ∂xk ∂xq ∂xi ∂xp þ ui ∂x i ∂xs Γ s pq � Γ ijk ∂xj ∂xp ∂xk ∂xq � � ∂up ∂xq � ui ∂x i ∂xs Γ s pq ¼ ∂ui ∂xk ∂xk ∂xq ∂xi ∂xp � ui ∂x j ∂xp ∂xk ∂xq Γ ijk Replacing the indexes i ! ‘, j ! i in the second term to the right of the expression, and with us ¼ ui ∂x i ∂xs this expression becomes ∂up ∂xq � usΓ spq ¼ ∂ui ∂xk � u‘Γ ‘ik � � ∂xi ∂xp ∂xk ∂xq Putting ∂qup ¼ ∂up∂xq � usΓ s pq the result is ∂qup ¼ ∂kuið Þ ∂x i ∂xp ∂xk ∂xq ð2:5:17Þ 110 2 Covariant, Absolute, and Contravariant Derivatives Then the covariant derivative of a covariant vector is a covariant tensor of the second order, i.e., of variance (0, 2). Various notations are found in the literature for the covariant derivative. For the covariant vector, the most usual ones are: ∂kum ¼ Dkum ¼ ∇kum ¼ um kj ¼ um;k. 2.4.3.2 Covariant Tensor of the Second Order The transformation law of covariant tensors of the second order is given by Tpq ¼ Tij ∂x i ∂xp ∂xj ∂xq and differentiating with respect to the coordinate xr ∂Tpq ∂xr ¼ ∂ 2 xi ∂xr∂xp ∂xj ∂xq Tij þ ∂x i ∂xp ∂2xj ∂xr∂xq Tij þ ∂x i ∂xp ∂xj ∂xq ∂Tij ∂xk ∂xk ∂xr ∂2xi ∂xr∂xp ¼ ∂x i ∂xs Γ s rp � Γ i‘m ∂x‘ ∂xp ∂xm ∂xr ∂2xj ∂xr∂xp ¼ ∂x j ∂xs Γ s rq � Γ j‘m ∂x‘ ∂xr ∂xm ∂xq Substituting these two expressions in the expression of the covariant derivative ∂Tpq ∂xr ¼ ∂x i ∂xs Γ s rp � Γ i‘m ∂x‘ ∂xp ∂xm ∂xr � � ∂xj ∂xq Tij þ ∂x i ∂xp ∂xj ∂xs Γ s rq � Γ j‘m ∂x‘ ∂xr ∂xm ∂xq � � Tij þ ∂x i ∂xp ∂xj ∂xq ∂Tij ∂xk ∂xk ∂xr ∂Tpq ∂xr ¼ Tij ∂x i ∂xs ∂xj ∂xq Γ s rp � Tij ∂xj ∂xq ∂x‘ ∂xp ∂xm ∂xr Γ i‘m þ Tij ∂xj ∂xs ∂xi ∂xp Γ s rq � Tij ∂x i ∂xp ∂x‘ ∂xr ∂xm ∂xq Γ j‘m þ ∂Tij ∂xk ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr In the second term to the right replacing the dummy index m ! k, and interchanging the indexes i $ ‘, and in the fourth term replacing the indexes ‘ ! k and interchanging the indexes j $ m results in ∂Tpq ∂xr ¼ Tij ∂x i ∂xs ∂xj ∂xq Γ s rp � T‘j ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr Γ ‘ik þ Tij ∂xi ∂xp ∂xj ∂xs Γ s rq � Tim ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr Γmkj þ ∂Tij ∂xk ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr 2.4 Covariant Derivative 111 and with the transformation law of covariant tensors of the second order Tsq ¼ Tij ∂x i ∂xs ∂xj ∂xq Tps ¼ Tij ∂x i ∂xp ∂xj ∂xs ∂Tpq ∂xr � TsqΓ srp � TpsΓ srq ¼ ∂Tij ∂xk � T‘jΓ ‘ik � TimΓmkj � � ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr Replacing the dummy indexes m ! ‘: ∂Tpq ∂xr � TsqΓ srp � TpsΓ srq ¼ ∂Tij ∂xk � T‘jΓ ‘ik � Ti‘Γ ‘kj � � ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr whereby ∂kTpq ¼ ∂kTij � � ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr therefore the covariant derivative of a covariant tensor of the second order is a covariant tensor of the third order, i.e., of variance (0, 3). Whereby the covariant derivative of a covariant tensor of the second order is given by ∂kTij ¼ ∂Tij∂xk � T‘jΓ ‘ ik � Ti‘Γ ‘kj ð2:5:18Þ For the Cartesian coordinates the Christoffel symbols are null, so in these referential systems the covariant derivative of the tensor Tij coincides with the partial derivative ∂Tij ∂xk . 2.4.3.3 Covariant Tensor of Order Above Two To generalize expression (2.5.18) for tensors of order above two, i.e., for instance, the covariant derivative of the covariant tensor of the third order, which expression may be developed by means of the following steps: (a) The basic structure of its expression is written considering the expression obtained for the covariant derivative of a covariant tensor of the second order ∂pTijk ¼ ∂Tijk∂xp þ T Γ þ T Γ þ T Γ (b) The indexes of the Christoffel symbols corresponding to the coordinate with respect to which the differentiation is being carried out are placed ∂pTijk ¼ ∂Tijk∂xp þ T Γ p þ T Γ p þ T Γ p 112 2 Covariant, Absolute, and Contravariant Derivatives (c) The covariant indexes of the Christoffel symbols must be completed obeying the sequence of the indexes of the tensor that is being differentiated ∂pTijk ¼ ∂Tijk∂xp þ T Γ ip þ T Γ jp þ T Γ kp (d) The dummy index q is placed on the Christoffel symbols and in sequential form in the tensors ∂pTijk ¼ ∂Tijk∂xp þ Tq Γ q ip þ T q Γ qjp þ T qΓ qkp (e) The remaining indexes are placed in the same sequence in which they appear on the tensor that is being differentiated ∂pTijk ¼ ∂Tijk∂xp þ TqjkΓ q ip þ TiqkΓ qjp þ TijqΓ qkp This tensor generated by the differentiation of a variance tensor (0, 4). Expres- sion (2.5.18) can be generalized by adopting this indexes placement systematic for a covariant tensor of order q > 3, and the variance of this new tensor will always be 0, qþ 1ð Þ. 2.4.4 Mixed Tensor Consider the transformation law of the mixed tensors of the second T m n ¼ T ij ∂xm ∂xi ∂xj ∂xn that can be written as T m n ∂xi ∂xm ¼ T ij ∂xj ∂xn which derivative with respect to coordinate xr is given by ∂T m n ∂xr ∂xi ∂xm þ Tmn ∂2xi ∂xr∂xm ¼ ∂T i j ∂xk ∂xk ∂xr ∂xj ∂xn þ T ij ∂2xj ∂xr∂xn and with the following expressions ∂2xi ∂xr∂xm ¼ ∂x i ∂xs Γ s rm � Γ i‘j ∂x‘ ∂xm ∂xj ∂xr ∂2xj ∂xr∂xm ¼ ∂x j ∂xs Γ s mr � Γ j‘p ∂x‘ ∂xn ∂xp ∂xr 2.4 Covariant Derivative 113 this expression becomes ∂T m n ∂xr ∂xi ∂xm þ Tmn ∂xi ∂xs Γ s rm � Γ i‘j ∂x‘ ∂xm ∂xj ∂xr � � ¼ ∂T i j ∂xk ∂xk∂xr ∂xj ∂xn þ T ij ∂xj ∂xs Γ s mr � Γ j‘p ∂x‘ ∂xn ∂xp ∂xr � � As T m n ¼ T pq ∂xm ∂xp ∂xq ∂xn T ij ¼ T m e ∂xi ∂xm ∂xe ∂xj it follows that ∂T m n ∂xr ∂xi ∂xm þ Tmn ∂xi ∂xs Γ s rm � T pq ∂xm ∂xp ∂xq ∂xn ∂x‘ ∂xm ∂xj ∂xr Γ i‘j ¼ ∂T i j ∂xk ∂xk ∂xr ∂xj ∂xn þ Tme ∂xi ∂xm ∂xe ∂xj ∂xj ∂xs Γ s mr � T ij ∂x‘ ∂xn ∂xp ∂xr Γ j‘p ∂T m n ∂xr ∂xi ∂xm þ Tmn ∂xi ∂xs Γ s rm � T pq δ ‘p ∂xq ∂xn ∂xj ∂xr Γ i‘j ¼ ∂T i j ∂xk ∂xk ∂xr ∂xj ∂xn þ Tms ∂xi ∂xm δ es Γ s mr � T ij ∂x‘ ∂xn ∂xp ∂xr Γ j‘p ∂T m n ∂xr ∂xi ∂xm þ Tmn ∂xi ∂xs Γ s rm � T ‘q ∂xq ∂xn ∂xj ∂xr Γ ipj ¼ ∂T i j ∂xk ∂xk ∂xr ∂xj ∂xn þ Tms ∂xi ∂xm Γ s mr � T ij ∂x‘ ∂xn ∂xp ∂xr Γ j‘p Interchanging the indexes in the second term on the left m $ s, in the last term on the right, interchanging the indexes j $ ‘ and replacing the indexes p ! k results in ∂T m n ∂xr ∂xi ∂xm þ T sn ∂xi ∂xm Γ m rs � T ‘q ∂xq ∂xn ∂xj ∂xr Γ i‘j ¼ ∂T ij ∂xk ∂xk ∂xr ∂xj ∂xn þ Tms ∂xj ∂xm Γ s mr � T i‘ ∂xj ∂xn ∂xk ∂xr Γ ‘jk and replacing the indexes j ! k and q ! j in the last term on the left 114 2 Covariant, Absolute, and Contravariant Derivatives ∂T m n ∂xr ∂xi ∂xm þ T sn ∂xi ∂xm Γ m rs � T ‘j ∂xj ∂xn ∂xk ∂xr Γ i‘k ¼ ∂T ij ∂xk ∂xk ∂xr ∂xj ∂xn þ Tms ∂xj ∂xm Γ s mr � T i‘ ∂xj ∂xn ∂xk ∂xr Γ ‘jk that can be written as ∂T m n ∂xr þ T snΓmrs � T m s Γ s mr � � ∂xi ∂xm ¼ ∂T i j ∂xk þ T ‘j Γ i‘k � T i‘Γ ‘jk ! ∂xj ∂xn ∂xk ∂xr then ∂T m n ∂xr þ T snΓmrs � T m s Γ s mr ¼ ∂T ij ∂xk þ T ‘j Γ i‘k � T i‘Γ ‘jk ! ∂xm ∂xi ∂xj ∂xn ∂xk ∂xr ð2:5:19Þ Putting ∂rT m n ¼ ∂T m n ∂xr þ T snΓmrs � T m s Γ s mr ð2:5:20Þ ∂rT i j ¼ ∂T ij ∂xr þ T ‘j Γ i‘k � T i‘Γ ‘jk ð2:5:21Þ the result is the expressions that represent the covariant derivative of the mixed tensors of the second-order T m n and T i j, whereby ∂rT m n ¼ ∂rT ij ∂xm ∂xi ∂xj ∂xn ∂xk ∂xr ð2:5:22Þ Expression (2.5.22) shows that the derivative of a mixed tensor of the second order is a mixed tensor of the third order, once contravariant and twice covariant, i.e., of variance (1, 2). The covariant derivative of a mixed tensor of variance ( p, q) generates a variance tensor p, qþ 1ð Þ. To generalize expression (2.5.22) for mixed tensors of order above two, assume as an example the covariant derivatives of a mixed tensor of the third order of variance (1, 2) and of a mixed tensor of fifth order of variance (3, 2), which are given, respectively, by the expressions ∂kT j p‘ ¼ ∂T jp‘ ∂xk � T jq‘Γ qpk � T jpqΓ q‘k þ T qp‘Γ jkq ∂kT j‘m rs ¼ ∂Tj‘mrs ∂xk � Tj‘mqs Γ qrk � Tj‘mrq Γ qsk þ Tq‘mrs Γ jkq þ Tjqmrs Γ ‘kq þ Tj‘qrs Γmkq 2.4 Covariant Derivative 115 2.4.5 Covariant Derivative of the Addition, Subtraction, and Product of Tensors Expression (2.5.21) shows that the covariant derivative of a mixed tensor comprises a partial derivative of this tensor and the terms containing Christoffel symbols, which are always linear in the components of the original tensor. This characteristic indicates that the covariant differentiation follows the same rules of the ordinary differentiation of Differential Calculus. To stress the properties of the covariant derivative let the scalar ϕ(xi) which ordinary derivative is equal to its covariant derivative, that can be written as the dot product of the vectors ui and vi expressed in Cartesian coordinates ϕ xi � � ¼ uivi and differentiating ∂kϕ x i � � ¼ ∂k uivi� � ¼ d uivið Þ dxk ¼ du i dxk vi þ ui dvi dxk As the covariant and ordinary derivatives are equal, it results in ∂kϕ x i � � ¼ ∂k ui� �vi þ ui∂k við Þ Substituting the expressions of the covariant derivatives of contravariant and covariant vectors ∂kϕ x i � � ¼ ∂ui ∂xk þ uiΓ ikj � � vi þ ui ∂vi∂xk � viΓ i kj � � ¼ ∂u i ∂xk vi þ ui ∂vi∂xk This expression suggests that the covariant derivative of an inner product of tensors behaves in a manner that is similar to the ordinary derivative. To prove this assumption, let, for instance, the tensors Aij and Bij for which the following properties of the covariant derivative are admitted a priori as valid: (a) ∂k Aij þ Bij � � ¼ ∂kAij þ ∂kBij; (b) ∂k Aij � Bij � � ¼ ∂kAij � ∂kBij; (c) ∂k AijBij � � ¼ ∂kAij� �Bij þ Aij ∂kBij� �. To demonstrate property (a) let the tensor Cij ¼ Aij þ Bij, so 116 2 Covariant, Absolute, and Contravariant Derivatives ∂k Aij þ Bij � �¼ ∂kCij ¼ ∂Cij∂xk � C‘jΓ ‘ik � Ci‘Γ ‘kj ¼ ∂ Aij þ Bij � � ∂xk � A‘j þ B‘j � � Γ ‘ik � Ai‘ þ Bi‘ð ÞΓ ‘kj ¼ ∂Aij ∂xk � A‘jΓ ‘ik � Ai‘Γ ‘kj � � þ ∂Bij ∂xk � B‘jΓ ‘ik � Bi‘Γ ‘kj ¼ ∂kAij þ ∂kBij In an analogous way, it is possible to prove property (b), replacing only the addition sign for the subtraction sign in the previous demonstration. To demonstrate property (c) let the inner product AijB‘m ¼ Cij‘m that generates a covariant tensor of the fourth order ∂k AijB‘m � � ¼ ∂kCij‘m ¼ ∂Cij‘m∂xk � Cpj‘mΓ pki � Cip‘mΓ pkj � CijpmΓ pk‘ � Cij‘pΓ pkm Substituting the expressions of the tensor of the fourth order in terms of the inner product ∂k AijB‘m � �¼ ∂ AijB‘m� � ∂xk � ApjB‘mΓ pki � AipB‘mΓ pk‘ � AijBpmΓ pk‘ � AijB‘pΓ pkm ¼ ∂Aij ∂xk � ApjΓ pki � AipΓ pkj � � B‘m þ Aij ∂B‘m∂xk � BpmΓ p k‘ � B‘pΓ pkm � � The terms in parenthesis are the covariant derivatives of the covariant tensors of the second-order, whereby ∂k AijB‘m � � ¼ ∂kAij� �B‘m þ Aij ∂kB‘mð Þ thus the covariant derivative of an inner product of tensors follows the same rule as the derivative of the product of functions in Differential Calculus. 2.4.6 Covariant Derivative of Tensors gij, g ij, δij Ricci’s Lemma The metric tensor behaves as a constant when calculating the covariant derivative. 2.4 Covariant Derivative 117 The covariant derivative of the metric tensor gij is calculated to demonstrate this lemma, thus ∂kgij ¼ ∂gij ∂xk � gpjΓ pik � gipΓ pkj ∂kgij ¼ ∂gij ∂xk � Γik, j þ Γkj, i � � and with the Ricci identity ∂gij ∂xk ¼ Γik, j þ Γjk, i and by the symmetry Γjk, i ¼ Γkj, i ∂kgij ¼ ∂gij ∂xk � ∂gij ∂xk ¼ 0 In an analogous way the conjugate metric tensor gij is given by ∂kg ij ¼ ∂g ij ∂xk þ gpjΓ ikp þ gipΓ jkp ð2:5:23Þ Since gijg jp ¼ δpi ) ∂ gijg jp � � ∂xk ¼ 0 ) ∂gij ∂xk gjp þ gij ∂gjp ∂xk ¼ 0 and multiplying by giq giqgjp ∂gij ∂xk þ giqgij ∂gjp ∂xk ¼ 0 ) giqgjp ∂gij ∂xk þ δ qj ∂gjp ∂xk ¼ 0 ) ∂g qp ∂xk ¼ �giqgjp ∂gij ∂xk it follows that ∂gqp ∂xk ¼ �giqgjp Γik, j þ Γjk, i � � ¼ �giqgjpΓik, j þ�giqgjpΓjk, i ¼ �giqΓ pik � giqΓ qjk Replacing the indexes i ! p, q ! i, and p ! j: ∂gqp ∂xk ¼ �gpiΓ jpk � gpjΓ ipk 118 2 Covariant, Absolute, and Contravariant Derivatives and substituting this expression in expression (2.5.23) ∂kg ij ¼ �gpiΓ jpk � gpjΓ ipk � � þ gpjΓ ikp þ gipΓ jkp and with the symmetry of gij and the Christoffel symbol of second kind ∂kg ij ¼ �gipΓ jkp � gpjΓ ikp � � þ gpjΓ ikp þ gipΓ jkp ¼ 0 Following the same systematic it implies for the covariant derivative of the Kronecker delta ∂kδ i j ¼ ∂δ ij ∂xk þ δ pj Γ ipk � δ ipΓ pjk ¼ 0þ Γ ijk � Γ ijk ¼ 0 These deductions show that the conjugate metric tensor gij and the Kronecker delta δij also behave as constants in calculating the covariant derivative. Exercise 2.11 Show that ∂kT ij ¼ gim ∂kT ij � � . Expressing the mixed tensor by T ij ¼ gimTmj the result for its covariant derivative is ∂kT i j ¼ ∂kgim � � T ij þ gim ∂kT ij � � As ∂kgim ¼ 0, it results in ∂kT i j ¼ gim ∂kT ij � � Q:E:D: Exercise 2.12 Show that ∂ui∂xj � ∂uj ∂xi � � is a covariant tensor of the second order, being ui a covariant vector. The covariant derivative of a covariant vector is given by ∂jui ¼ ∂ui∂xj � upΓ p ij ) ∂ui ∂xj ¼ ∂jui þ upΓ pij and replacing the indexes i ! j results in ∂uj ∂xi ¼ ∂iuj þ upΓ pji 2.4 Covariant Derivative 119 Carrying out the subtraction presented in the enunciation ∂ui ∂xj � ∂uj ∂xi � � ¼ ∂jui þ upΓ pji � � � ∂jui þ upΓ pij � � and with the symmetry Γ pij ¼ Γ pji ∂ui ∂xj � ∂uj ∂xi � � ¼ ∂jui � ∂iuj As the covariant derivative of a covariant vectoris a tensor of the second order, then this expression represents a tensor of variance (0, 2). Exercise 2.13 Show that Γ pij ¼ 12 Tpq ∂Tik∂xj þ ∂Tjk ∂xi � ∂Tij ∂xk � � , being Tij a symmetric tensor and detTij 6¼ 0, and with covariant derivative ∂kTij ¼ 0. The tensor Tpk can be written under the form Tpk ¼ gipgjkTij For the tensor Tij the covariant derivative is given by ∂kTij ¼ ∂Tij∂xk � TpjΓ p ik � TipΓ pjk ¼ 0 ) ∂Tij ∂xk ¼ TpjΓ pik þ TipΓ pjk Interchanging the indexes i, j, k cyclically ∂Tjk ∂xi ¼ TpkΓ pji þ TjpΓ pki ∂Tki ∂xj ¼ TpiΓ pkj þ TkpΓ pij and adding these two expressions and subtracting the one that comes before them, and considering the tensor’s symmetry ∂Tjk ∂xi þ ∂Tki ∂xj � ∂Tij ∂xk ¼ TpkΓ pji þ TjpΓ pki � � þ TpiΓ pkj þ TkpΓ pij � � � TpjΓ pik þ TipΓ pjk � � ¼ 2TkpΓ pij The dummy index p can be changed by the index q, so ∂Tjk ∂xi þ ∂Tki ∂xj � ∂Tij ∂xk ¼ 2TkqΓ qij ) 1 2 ∂Tjk ∂xi þ ∂Tki ∂xj � ∂Tij ∂xk � � ¼ TkqΓ qij and multiplying by Tpk 1 2 Tpq ∂Tjk ∂xi þ ∂Tki ∂xj � ∂Tij ∂xk � � ¼ TpkTkqΓ qij 120 2 Covariant, Absolute, and Contravariant Derivatives Free ebooks ==> www.Ebook777.com and with the contraction TpkTkq ¼ δpq it follows that 1 2 Tpq ∂Tjk ∂xi þ ∂Tki ∂xj � ∂Tij ∂xk � � ¼ δpqΓ qij ¼ Γ pij Q:E:D: 2.4.7 Particularities of the Covariant Derivative To exemplify a particularity of the covariant derivative let the vector u defined by its covariant components uj ¼ gijui, then ∂kuj ¼ ∂k gijui � � ¼ ∂kgij � � ui þ gij∂kui and with Ricci’s lemma ∂k giju i � � ¼ gij∂kui The covariant derivative of the contravariant vector is given by ∂ku i ¼ ∂u i ∂xk þ u‘Γ i‘k so by substitution ∂k giju i � � ¼ gij ∂ui ∂xk þ u‘Γ i‘k � � The contravariant components of the vector can be expressed in terms of their covariant components ∂k giju i � � ¼ gij ∂ gi‘u‘ � � ∂xk þ giju‘Γ i‘k ¼ gij ∂ gi‘u‘ � � ∂xk þ u‘Γ‘k, j ¼ gij ∂gi‘ ∂xk u‘ þ gijgi‘ ∂u‘ ∂xk þ u‘Γ‘k, j Rewriting expression (2.4.31) ∂gi‘ ∂xk ¼ �g‘mΓ imk � gimΓ ‘mk 2.4 Covariant Derivative 121 www.Ebook777.com http://www.ebook777.com which substituted in the previous expression provides ∂k giju i � � ¼ gij �g‘mΓ imk � gimΓ ‘mk � � u‘ þ gijgi‘ ∂u‘ ∂xk þ u‘Γ‘k, j ¼ gij �g‘mu‘Γ imk � gimu‘Γ ‘mk � �þ δ ‘j ∂u‘∂xk þ u‘Γ‘k, j ¼ �gijumΓ imk � δmj u‘Γ ‘mk þ ∂uj ∂xk þ u‘Γ‘k, j ¼ umΓmk, j � u‘Γ ‘mk þ ∂uj ∂xk þ u‘Γ‘k, j Replacing the dummy indexes ‘ ! m: ∂k giju i � � ¼ �umΓmk, j � u‘Γ ‘mk þ ∂uj ∂xk þ umΓmk, j ) ∂kuj ¼ ∂k gijui � � ¼ ∂uj ∂xk � u‘Γ ‘mk then the covariant derivative of a covariant vector is equal to the covariant deriv- ative of the product of the metric tensor by the contravariant components of this vector. This characteristic of the covariant derivative can be generalized for tensors of order above one, for instance, for a contravariant tensor of the second order the result is ∂k gipgjqT pq � � ¼ ∂kTij Another particularity of the covariant derivative is its successive differentiation of a scalar function. Let a scalar function ϕ that represents an invariant, so its derivative with respect to its coordinate xi is a covariant vector given by ϕ, i ¼ ∂ϕ ∂xi ¼ ∂iϕ Taking the derivative of this function again, nowwith respect to the coordinate xj: ϕ, ij ¼ ∂2ϕ ∂xj∂xi ¼ ∂j ∂iϕð Þ ¼ ∂ 2ϕ ∂xj∂xi � ∂ϕ ∂xm Γmij The dummy index m can be changed, and as the Christoffel symbol is symmet- ric, it results in ∂j ∂iϕð Þ ¼ ∂i ∂jϕ � � Then the covariant derivative of an invariant is commutative. 122 2 Covariant, Absolute, and Contravariant Derivatives 2.5 Covariant Derivative of Relative Tensors The covariant derivative of relative tensors has characteristics that differ from the covariant derivative of absolute tensors. For studying the derivatives of these varieties in a progressive manner, a scalar density of weight W with respect to the coordinate system X i is admitted, given by JWϕ xið Þ. Taking the derivative of this function ∂ JWϕ � � ∂xj ¼ JW ∂ϕ ∂xk ∂xk ∂xj þWJW�1 ∂J ∂xj ϕ ð2:6:1Þ The second parcel on the right shows that the gradient of a scalar density is not a vector. It is verified that for W ¼ 0 the result is a scalar function and ∂ϕ ∂xj ¼ ∂ϕ ∂xk ∂xk ∂xj is the transformation law of the vectors. Let the Jacobian cofactor Cmk ¼ ∂xk ∂xm or ∂xr ∂xj Cmr ¼ Jδ rr ) Cmr ¼ J ∂xm ∂xk it follows that ∂J ∂xj ¼ ∂ ∂xj ∂xk ∂xm � � � Cmk ) ∂J ∂xj ¼ J ∂ 2 xk ∂xj∂xm ∂xm ∂xk The substitution of this expression in expression (2.6.1) provides ∂ JWϕ � � ∂xj ¼ JW ∂ϕ ∂xk ∂xk ∂xj þW ∂ 2 xk ∂xj∂xm ∂xm ∂xk ϕ ! ð2:6:2Þ that is the transformation law of the pseudoscalar JWϕ(xi). Using expression (2.4.25) the second term in parenthesis can be written as ∂2xk ∂xj∂xm ∂xm ∂xk ¼ Γmj‘ � ∂xm ∂xp ∂xk ∂xj ∂xq ∂x‘ Γ pkq 2.5 Covariant Derivative of Relative Tensors 123 The contraction in the indexes m and ‘ provides ∂2xk ∂xj∂xm ∂xm ∂xk ¼ Γmj‘ � ∂x‘ ∂xp ∂xk ∂xj ∂xq ∂x‘ Γ pkq and with δqp ¼ ∂x‘ ∂xp ∂xq ∂x‘ the result is ∂2xk ∂xj∂xm ∂xm ∂xk ¼ Γmj‘ � ∂xk ∂xj Γ qkq The substitution of this expression in expression (2.6.2) provides ∂ JWϕ � � ∂xj ¼ JW ∂ϕ ∂xk ∂xk ∂xj þWJWΓmj‘ϕ�WJW ∂xk ∂xj Γ qkqϕ Let a scalar density which transformation law is given by ϕ ¼ JWϕ it results in ∂ JWϕ � � ∂xj �WΓmj‘ϕ ¼ JW ∂xk ∂xj ∂ϕ ∂xk �WΓ qkqϕ � � The term in parenthesis to the right represents a covariant pseudovector of weight W. This expression shows that the covariant derivative of a scalar density presents an additional term in its expression, in which the factor multiplies the contracted Christoffel symbol. ForW ¼ 0 this expression is reduced to the gradient expression of the scalar function ϕ(xi) ∂ϕ ∂xj ¼ ∂ϕ ∂xk ∂xk ∂xj For a contravariant pseudovector of weight W it follows by means of this expression that is analogous to the one shown for a scalar density, the next expression ∂ku j ¼ ∂u j ∂xk þ uqΓ jkq �Wu jΓ qkq ð2:6:3Þ 124 2 Covariant, Absolute, and Contravariant Derivatives and the contraction of the indexes j and k provides ∂ju j ¼ ∂u j ∂xj þ uqΓ jjq �Wu jΓ qjq The dummy index j in the third term to the right can be changed by the index q: ∂ju j ¼ ∂u j ∂x j þ 1�Wð ÞuqΓ qjq If the pseudovector has weightW ¼ 1 this expression is simplified for∂ju j ¼ ∂u j∂x j, which represents the divergence of vector u j. The generalization of expression (2.6.3) for a relative tensor of weight W and variance (1, 1) is given by ∂rT i j ¼ ∂T ij ∂xr þ T ‘j Γ j‘k � T i‘Γ ‘jk �WT ijΓ qrq ð2:6:4Þ For a relative tensor Ti���j��� of weight W and variance ( p, q) it results in ∂rT i��� j��� ¼ ∂Ti���j��� ∂xr þ T ‘j Γ j‘k þ � � �� � �� � �� � �|fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} terms relative to the contravariance �T i‘Γ ‘jk � � � �� � �� � �� � �|fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} terms relative to the covariance �WT ijΓ qrq ð2:6:5Þ By means of the considerations presented in the first paragraph of item (2.5.4), and adding that the parcel WTijΓ q rq in expression (2.6.5) is linear in terms of the original tensor, it implies that the rules of ordinary differentiation of Differential Calculus are applicable to the covariant differentiation of relative tensors. 2.5.1 Covariant Derivative of the Ricci Pseudotensor The covariant derivative of the Ricci pseudotensor in its contravariant form is given by ∂iε jk‘ ¼ ∂i e jk‘ffiffiffi g p � � ¼ ∂ e ijkffiffi g p � � ∂xi þ Γ jip epk‘ffiffiffi g p þ Γ kip ejp‘ffiffiffi g p þ Γ ‘ip ejkpffiffiffi g p ∂ e jk‘ffiffi g p � � ∂xi ¼ 1ffiffiffi g p ∂e jk‘ ∂xi þ ejk‘ ∂ 1ffiffigp� � ∂xi ¼ ejk‘ ∂ 1ffiffigp� � ∂xi ¼ � e jk‘ 2g 2 3 ∂g ∂xi 2.5 Covariant Derivative of Relative Tensors 125 The contraction of the Christoffel symbol provides ∂g ∂xi ¼ 2gΓ ppi whereby ∂ e jk‘ffiffi g p � � ∂xi ¼ � e jk‘ffiffiffi g p Γ ppi Substituting this expression in the expression of covariant derivative ∂iε jk‘ ¼ ∂i e jk‘ffiffiffi g p � � ¼ � e jk‘ffiffiffi g p Γ ppi þ Γ jip epk‘ffiffiffi g p þ Γ kipejp‘ffiffiffi g p þ Γ ‘ip ejkpffiffiffi g p The conditions for which this pseudotensor is non-null are that the first three indexes be different, i.e., j 6¼ k 6¼ ‘, then for j ¼ 1, k ¼ 2, ‘ ¼ 3: ∂iε jk‘ ¼ ∂i e jk‘ffiffiffi g p � � ¼ � e 123ffiffiffi g p Γ ppi þ Γ jip ep23ffiffiffi g p þ Γ kip e1p3ffiffiffi g p þ Γ ‘ip e12pffiffiffi g p With p ¼ 1, 2, 3: ∂iε jk‘ ¼ ∂i e jk‘ffiffiffi g p � � ¼ � e 123ffiffiffi g p Γ11i þ Γ22i þ Γ33i � �þ Γ ji1 e123ffiffiffigp þ Γ ki2 e123ffiffiffigp þ Γ ‘i3 e123ffiffiffigp and with the symmetry of the Christoffel symbol it results in ∂iε jk‘ ¼ ∂i e jk‘ffiffiffi g p � � ¼ 0 With an analogous expression for the covariant form of the Ricci pseudotensor εijk ¼ ffiffiffigp eijk it results for its covariant derivative ∂iεijk ¼ ∂i ffiffiffigp eijk� � ¼ ∂ ffiffiffigp eijk� �∂xi � ffiffiffigp epjkΓ pi‘ � ffiffiffigp eipkΓ pj‘ � ffiffiffigp eijpΓ pk‘ The partial derivative referent to the first term to the right is given by ∂ ffiffiffi g p eijk � � ∂x‘ ¼ ∂ ffiffiffi g p� � ∂x‘ eijk þ ffiffiffigp ∂ eijk� �∂x‘ but ∂ eijk � � ∂x‘ ¼ 0 126 2 Covariant, Absolute, and Contravariant Derivatives it results in ∂ ffiffiffi g p eijk � � ∂x‘ ¼ ∂ ffiffiffi g p� � ∂x‘ eijk Expression (2.4.23) can be written as ∂ ffiffiffi g p� � ∂x‘ ¼ ffiffiffigp Γ pp‘ ) ∂ ffiffiffigp eijk � � ∂x‘ ¼ ffiffiffigp eijkΓ pp‘ Substituting this expression in the expression of the covariant derivative ∂iεijk ¼ ∂i ffiffiffigp eijk� � ¼ ffiffiffigp eijkΓ pp‘ � ffiffiffigp epjkΓ pi‘ � ffiffiffigp eipkΓ pj‘ � ffiffiffigp eijpΓ pk‘ The conditions for which this pseudotensor is non-null are that the first three indexes be different, i.e., j 6¼ k 6¼ ‘, then for j ¼ 1, k ¼ 2, ‘ ¼ 3: ∂iεijk ¼ ∂i ffiffiffigp eijk� � ¼ ffiffiffigp e123Γ pp‘ � ffiffiffigp ep23Γ p1‘ � ffiffiffigp e1p3Γ p2‘ � ffiffiffigp e12pΓ p3‘ With p ¼ 1, 2, 3: ∂iεijk ¼ ∂i ffiffiffigp eijk� � ¼ ffiffiffigp e123 Γ11‘ þ Γ22‘ þ Γ33‘� �� ffiffiffigp e123Γ11‘ � ffiffiffigp e123Γ22‘ � ffiffiffigp e123Γ33‘ whereby ∂iεijk ¼ ∂i ffiffiffigp eijk� � ¼ 0 These derivatives show that ∂i δ ijk pqr � � ¼ ∂i εijkεpqr � � ¼ ∂i ε ijk� �εpqr þ εijk∂i εpqr� � ¼ 0 The covariant derivatives of the Ricci pseudotensors εijk, εpqr and the generalized Kronecker delta δijkpqr being null, it implies that these varieties behave as constants in the calculation of the covariant derivative. As an example of an application of this characteristic, let the tensorial expression εijk∂juk, which covariant derivative is given by ∂i ε ijk∂juk � � ¼ εijk∂i ∂juk� �þ ∂juk∂i ∂jε ijk� � but with ∂iε ijk ¼ 0 this expression becomes 2.5 Covariant Derivative of Relative Tensors 127 ∂i ε ijk∂juk � � ¼ εijk∂i ∂juk� � 2.6 Intrinsic or Absolute Derivative The absolute derivative of a variety is calculated when the coordinates xi vary as a function of time, i.e., xi ¼ xi tð Þ. A covariant derivative of an invariant ϕ(xk) is given by ∂kϕ x k � � ¼ ∂ϕ xk� � ∂xk which is equal to its partial derivative. For the absolute derivative δϕ xk � � δt ¼ ∂ϕ x k � � ∂t þ ∂kϕ xk � � dxk dt ¼ ∂ϕ x k � � ∂t þ ∂ϕ x k � � ∂xk dxk dt ¼ dϕ x k � � dt then this derivative is equal to its total derivative. For the vector u(xi) where xi varies as a function of time, which is expressed by means of its contravariant coordinates, or u ¼ uk xi tð Þ, t� �gk xi tð Þ� � The derivative with respect to time is given by du dt ¼ d u kgk � � dt ¼ du k dt gk þ uk ∂gk ∂t dxi dt ð2:7:1Þ and with duk dt ¼ ∂u k ∂t þ ∂u k ∂xi dxi dt ð2:7:2Þ The following expression (item 2.3) ∂gk ∂xi ¼ gmΓmki substituted in expression (2.7.1) provides du dt ¼ du k dt gk þ ukΓmki dxi dt gm Replacing the indexes k ! m in the first term to the right 128 2 Covariant, Absolute, and Contravariant Derivatives du dt ¼ du m dt þ ukΓmki dxi dt � � gm thus the absolute derivative of a vector generates a vector. The covariant derivative of the contravariant vector is written as δum δ t ¼ du m dt þ ukΓmki dxi dt ð2:7:3Þ and substituting expression (2.7.2) in expression (2.7.3) δum δ t ¼ ∂u k ∂t þ ∂u k ∂xi dxi dt þ ukΓmki dxi dt ) δu m δt ¼ ∂u k ∂t þ ∂u k ∂xi þ ukΓmki � � dxi dt The covariant derivative of the contravariant vector is given by ∂iu k ¼ ∂u k ∂xi þ ukΓmki or in vectorial form ∂u ∂xi ¼ ∂iuk � � gk whereby for the absolute derivative of vector u it results that δum δ t ¼ ∂u k ∂t þ ∂iuk dx i dt or in vectorial form du dt ¼ δu m δ t � � gm The vector u in terms of their covariant components is given by u ¼ uk xi tð Þ, t � � gk xi tð Þ� � and with an analogous analysis to the one shown for the contravariant vectors, and with gk, i ¼ �Γ kimgm it results for the absolute derivative of vector u 2.6 Intrinsic or Absolute Derivative 129 δuk δ t ¼ ∂um ∂t þ ∂iuk dx i dt where ∂iuk is the covariant derivative of the covariant vector. These expressions can be generalized for the tensors δTij δ t ¼ ∂T ij ∂t þ ∂kTij dx k dt ð2:7:4Þ δTij δ t ¼ ∂Tij ∂t þ ∂kTij dx k dt ð2:7:5Þ δTijm δ t ¼ ∂T ij m ∂t þ ∂kTijm dxk dt ð2:7:6Þ The differentiation rules of Differential Calculus are applicable to absolute differentiation, which can be proven, for instance, for two tensors Aij and Bij, which algebraic addition generates the tensors Cij ¼ Aij Bij, and which product results in AijBij. Calculating the absolute derivative of this sum δCij δ t ¼ ∂kCij dx k dt ¼ ∂k Aij þ Bij � � dxk dt ¼ ∂kAij dx k dt þ ∂kBij dx k dt ¼ δAij δ t þ δBij δ t Calculating the absolute derivative of the product of the tensors δ AijBij � � δ t ¼ ∂k AijBij � � dxk dt ¼ ∂kAij � � Bij dxk dt þ Aij ∂kBij � � dxk dt ¼ ∂kAij dx k dt Bij þ Aij∂kBij dx k dt ¼ δAij δ t Bij þ Aij δBij δ t The absolute derivative of vector u calculated along the curve xi ¼ xi tð Þ can be defined by means of the inner product of its covariant derivative by the tangent vector to this curve dx i dt . For a tensor of order above the unit, and with an analogous way, the absolute derivative is the inner product of this tensor by the vector tangent to a curve, then δTijpqr δt ¼ ∂kTijpqr dxk dt This definition in conjunction with the considerations made in the first paragraph of item 2.5.4 indicates that the absolute derivative follows the rules of Differential Calculus, such as shown for the addition and product of two tensors. The derivative of the metric tensor gij is given by 130 2 Covariant, Absolute, and Contravariant Derivatives Free ebooks ==> www.Ebook777.com δgij δ t ¼ ∂gij ∂t þ ∂kgij dxk dt Ricci’s lemma shows that ∂kgij ¼ 0, then δgij δ t ¼ ∂gij ∂t As the metric tensor is independent of time it implies that ∂gij ∂t ¼ 0, whereby it results that δgij δ t ¼ 0, i.e., its absolute derivative is null. For the tensors gij and δij, which have the same characteristics of the metric tensor, developing an analysis analogous to the one shown for this tensor it results in δgij δt ¼ ∂g ij ∂t þ ∂kgij dx k dt ¼ ∂g ij ∂t ¼ 0 δδ ij δt ¼ ∂δ i j ∂t þ ∂kδ ij dxk dt ¼ ∂δ i j ∂t ¼ 0 2.6.1 Uniqueness of the Absolute Derivative The covariant derivative of a Cartesian tensor coincides with its partial derivative, then the absolute derivative of this variety, calculated along a curve xi ¼ xi tð Þ, can be defined by means of the scalar product of this derivative by the vector tangent to this curve dx i dt . For instance, for a Cartesian tensor of variance (2, 3) it results in δTijpqr δ t ¼ ∂T ij pqr ∂t dxk dt As the partial derivative of a Cartesian tensor is unique, and the scalar product that defines the absolute derivative generates an invariant, it is possible to conclude that this derivative is also unique. This analysis can be generalized for arbitrary tensors. Exercise 2.14 Calculate the absolute derivative of: (a) giju ivj; (b) giju iuj; (c) vector ui knowing that δui δ t ¼ 0. (a) The expression giju ivj represents a scalar, and taking the derivative 2.6 Intrinsic or AbsoluteDerivative 131 www.Ebook777.com http://www.ebook777.com δ giju ivj � � δt ¼ d giju ivj � � dt δ giju ivj � � δ t ¼ δ gij � � δ t uivj þ gij δ uið Þ δt vj þ gijui δ vjð Þ δt ¼ gij δ uið Þ δt vj þ gijui δ vjð Þ δt (b) The change of vector v j by vector u j in the expression calculated in the previous item provides δ giju iuj � � δt ¼ gij δ uið Þ δt uj þ gijui δ ujð Þ δt Interchanging the indexes i $ j in the first term to the right, and with the symmetry of the metric tensor results in δ giju iuj � � δt ¼ gji δ ujð Þ δt ui þ gijui δ ujð Þ δt ¼ 2gijui δ ujð Þ δt As giju iuj ¼ uk k2, it implies that δ giju iujð Þ δt ¼ 0, which indicates that δu j δt ¼ 0. (c) The covariant components of the vector are given by uj ¼ gijui whereby differentiating δuj δ t ¼ δ giju i � � δ t ¼ δgij δ t ui þ gij δui δ t ¼ gij δui δ t ¼ 0 Exercise 2.15 Show that δ δ t dxi dt � � ¼ d 2xi dt2 þ Γ ijk dxj dt dxk dt . Putting ui ¼ dx i dt results for the absolute derivative of this vector δui δt ¼ ∂kui dx k dt ¼ ∂u i ∂xk þ ujΓ ijk � � dxk dt ¼ ∂u i ∂xk dxk dt þ ujΓ ijk dxk dt and with 132 2 Covariant, Absolute, and Contravariant Derivatives uj ¼ dx j dt it implies δui δt ¼ ∂u i ∂xk dxk dt þ Γ ijk dxj dt dxk dt It follows that δ δt dxi dt � � ¼ d dt dxi dt � � þ Γ ijk dxj dt dxk dt δ δt dxi dt � � ¼ d 2xi dt2 þ Γ ijk dxj dt dxk dt Q:E:D: 2.7 Contravariant Derivative The contravariant derivative is defined considering the tensorial nature of the covariant derivative, for the raising of the index of tensor ∂k . . . the result is ∂‘ . . . ¼ gk‘∂k . . . ð2:8:1Þ It is promptly verified with Ricci’s lemma that∂kgij ¼ 0, as well as∂kgij ¼ 0and ∂kδ ij ¼ 0. These relations show that the tensors gij, gij, δij behave as constants in the calculation of the contravariant derivative. For the variance tensors ( p, q) the result by means of the expression (2.8.1) is ∂kT������ ¼ gkj∂jT������ ¼ ∂j gkjT������ � � ð2:8:2Þ Then the contravariant derivative is equivalent to the raising of the indexes of tensor ∂k . . ., or the covariant derivative of tensor gkjT������. For instance, for the covariant vector uk: ∂kuk ¼ gkj∂juk ¼ ∂j gkjuk � � ¼ ∂juj Problems 2.1 Calculate the Christoffel symbols for the coordinates Xi which metric tensor is given by 2.7 Contravariant Derivative 133 gij ¼ 1 0 0 1 x2ð Þ 2 24 35 Answer: Γij, 1 ¼ 0 for i, j ¼ 1, 2 Γij, 2 ¼ 0 0 0 � 1 x2ð Þ2 24 35 Γ1ij ¼ 0 for i, j ¼ 1, 2 Γ2ij ¼ 0 0 0 �1 � 2.2 Calculate the Christoffel symbols for the coordinate system Xi which metric tensor and its conjugated metric tensor are given by gij ¼ 1 0 0 0 x1ð Þ 2 0 0 0 x1 sin x2ð Þ 2 264 375 gij ¼ 1 0 0 0 1 x1ð Þ 2 0 0 0 1 x1 sin x2ð Þ 2 2666664 3777775 Answer: Γij, 1 ¼ 0 0 0 0 �x1 0 0 0 �x1 sin x2ð Þ 2 264 375 Γij, 2 ¼ 0 x 1 0 x1 0 0 0 0 � x1ð Þ 2 sin x2 cos x2 264 375 Γij, 3 ¼ 0 0 x1 sin x2ð Þ 2 0 0 x1ð Þ2 sin x2 cos x2 x1 sin x2ð Þ 2 x1ð Þ 2 sin x2 cos x2 0 2664 3775 Γ1ij ¼ 0 0 0 0 �x1 0 0 0 �x1 sin x2ð Þ 2 264 375 Γ2ij ¼ 0 1 x1 0 1 x1 0 0 0 0 � x1ð Þ 2 sin x2 cos x2 2666664 3777775 Γ3ij ¼ 0 0 1 x1 0 0 cot x2 1 x1 cot x2 0 266664 377775 134 2 Covariant, Absolute, and Contravariant Derivatives 2.3 Calculate the Christoffel symbols of the second kind, where F(x1; x2) is a function of the coordinates, for the referential system which metric tensor is gij ¼ 1 00 F x1; x2ð Þ � Answer: Γ1ij ¼ 1 0 0 �1 2 ∂F ∂x1 24 35 Γ2ij ¼ 1 1 2F ∂F ∂x1 1 2F ∂F ∂x1 1 2F ∂F ∂x2 2664 3775 2.4 Calculate the Christoffel symbols for the space defined by the metric tensor gij ¼ �1 0 0 0 0 �1 0 0 0 0 �1 0 0 0 0 e�x 4 2664 3775 Answer: Γ44,4 ¼ �12 e�x 4 ; Γ444 ¼ �12. 2.5 Calculate the covariant derivative of the inner product of the tensors Ajk and B ‘m n with respect to coordinate xp. Answer: ∂pA j k � � B‘mn þ Ajk ∂pB‘mn � � 2.6 Show that ∂ ffiffi g p gijð Þ ∂xi þ Γ jpq ffiffiffi g p gpq ¼ 0. 2.7 Contravariant Derivative 135 Chapter 3 Integral Theorems 3.1 Basic Concepts The integral theorems and the concepts presented in this chapter are treated in Differential and Integral Calculus of multiple variables. The approach of this subject is carried in a concise and direct manner, and seeks solely to provide theoretical subsides so that the gradient, divergence, and curl differential operators can be physically interpreted. 3.1.1 Smooth Surface The surface S, open or closed, with upward normal n unique in each point, which direction is a continuous function of its points, is classified as a smooth surface. For instance, the surface of a sphere is closed smooth, and the surface of a cube is closed smooth by parts, for it can be decomposed into six smooth surfaces. 3.1.2 Simply Connected Domain For every closed curve C defined in the domain D, the region formed by C and its interior is fully contained in D. This curve defines a region R � D, and D is called simply connected domain (Fig. 3.1a). The interior of a circle and the interior of a sphere are simply connected regions. Two concentric spheres define a simply connected region. © Springer International Publishing Switzerland 2016 E. de Souza Sánchez Filho, Tensor Calculus for Engineers and Physicists, DOI 10.1007/978-3-319-31520-1_3 137 3.1.3 Multiply Connected Domain Multiply Connected Domain is the domain D that contains a region R with N “holes” (Fig. 3.1b). A circle excluded its center defines a simply connected domain, and the “hole” is reduced to a point, but the region between two coaxial cylinders is multiply connected. 3.1.4 Oriented Curve The closed smooth curve C that limits a region R is counterclockwise oriented if this region stays to its left, i.e., this curve is positively oriented. 3.1.5 Surface Integral Consider S a smooth surface by parts with upward unit normal vector n, and ϕ(xi) a function that represents a smooth curve C over this surface (Fig. 3.2). Dividing this finite area surface, defined by the function ϕ(xi) in N elementary areas dSi, i ¼ 1, 2, . . . ,N, where the elementary area contains the point P(xi), and carrying out the sum XN i¼1 ϕ xi � � dSi and for N ! 1, thus dSi ! 0, implies the limitðð S ϕ xi � � dS that represents the integral of surface S. This limit exists and is independent of the number of divisions made. For a vectorial function, it results in a similar way ðð S udS. C C D D R 1C 2C R a bFig. 3.1 Domain: (a) simply connected and (b) multiply connected 138 3 Integral Theorems 3.1.6 Flow Let the vectorial function u dependent on point P(xi) located on the surface S. The component of u in the direction of the unit normal vector to the surface in this point is given by the scalar product u � n. With this dot product for all the points located in the surface elements dS, and carrying out the sum XN i¼1 u � ndS, and for N ! 1, and dS ! 0 implies the integral F ¼ ðð S u � ndS ð3:1:1Þ that defines the flow of the vectorial function u on the surface S (Fig. 3.3). The surface area element dS is associated to the area vector dS, with modulus dS and same direction of n, then C n u S dS ( )x iP ( )φ α x i Fig. 3.2 Smooth surface S Flow of u S dSnu ⋅ dS a b Fig. 3.3 Flow: (a) through the surface S and (b) component of the vectorial function u in the direction normal to the surface S 3.1 Basic Concepts 139 dS ¼ ndS ð3:1:2Þ Expression (3.1.1) is written as F ¼ ðð S u � ndS ¼ ðð S u � dS ð3:1:3Þ and the integration shown in this expression is independent of the coordinate system, because the dot product u � n is invariant. In terms of the components of u, it follows that F ¼ ðð S uinidS ð3:1:4Þ where ni are the direction cosines of the unit normal vector n. 3.2 Oriented Surface Let S a surface oriented by means of its upward unit normal vector n, then its outline C is oriented positively if S stays to its left, thus this curve is anticlockwise oriented. Figure 3.4 shows a smooth surface S with upwardunit normal vector n, defined in a Cartesian coordinate system. This surface is expressed by the function z ¼ ϕ x; yð Þ, which orthogonal projection in plane OX3 determines the region R ¼ S12. The unit 1X O dS ixP S 2X 2dx 1dx k C 12C 12S 1X O dS ixP S X 2dx 1dx k C 12C 12S 3X Fig. 3.4 Smooth surface S with upward unit normal vector n which outline is a curve closed smooth C 140 3 Integral Theorems normal vector n forms an angle α with the axis OX3, being cosα its direction cosine. The orthogonal projection of the area element dS is given by dS ¼ dx 1dx2 cos α The dot product of the unit vectors n and k is given by n � k ¼ nk k kk k cos α so n � kk k ¼ cos α and therefore dS ¼ dx 1dx2 n � kk k Substituting this expression in expression (3.1.3) results in F ¼ ðð S u � ndS ¼ ðð S u � n dx 1dx2 n � kk k ð3:1:5Þ then the surface integral can be calculated as a double integral defined in the region R. The algebraic value of the flow depends on the field’s orientation. If α < π 2 thenF > 0, i.e., the flow “is outward,” and ifα > π 2 thenF < 0, i.e., “the flow is inward.” 3.2.1 Volume Integral Consider the closed smooth surface S that contains a volume V, and ϕ(xi) a function of position defined on this volume. Dividing V into elementary volumes dVi, then for the point P(xi) situated over S implies ϕ P xið Þ½ � ¼ ϕ xið Þ. Carrying out the sum of elementary volumes XN i¼1 ϕ xi � � dVi and for N ! 1, thus dVi ! 0, results the limitððð V ϕ xi � � dV that represents the volume integral. This limit exists and is indepen- dent on the number of divisions. If the function is vectorial, it results in a similar way ððð V udV. 3.2 Oriented Surface 141 3.3 Green’s Theorem Consider R a region in the plane OX1X2 involved by the closed smooth curve C with R to its left. Let the real continuous functions F1(x 1; x2) and F2(x 1; x2), with continuous partial derivatives in R [ C. Thenðð R ∂F2 ∂x1 � ∂F1 ∂x2 � � dx1dx2 ¼ þ C F1dx 1 þ F2dx2 � � ð3:2:1Þ This theorem is due to George Green (1793–1841) and deals with a generalization of the fundamental theorem of Integral Calculus for two dimensions. Figure 3.5 shows the region R involved by the closed smooth curve C, in which there are lines parallel to the coordinate axes that are tangent to this curve. It is assumed as a premise that C is intersect by straight lines parallel to the coordinate axes in a maximum of two points. The region R is defined by a x1 b, f x1ð Þ x2 g x1ð Þ c x2 d, p x2ð Þ x1 q x2ð Þ Let C ¼ AEB [ BFA, with AEB given by x2 ¼ f x1ð Þ, and BFA by x2 ¼ g x1ð Þ. In an analogous way resultsC ¼ FAE [ EBF, with FAE given by x1 ¼ p x2ð Þ, and EBF by x1 ¼ q x2ð Þ. With ðð R ∂F1 ∂x2 dx1dx2 ¼ ðb a ðg x1ð Þ f x1ð Þ ∂F1 ∂x2 dx2 2664 3775dx1 ¼ ð b a F1 x 1; x2 � � g x1ð Þ f x1ð Þ dx1 ¼ � ðb a F1 x 1; f x1 � �� � dx1 � ðb a F1 x 1; g x1 � �� � dx1 b d C A B F E R 2 X 1 XO Fig. 3.5 Simply connected region 142 3 Integral Theorems The two right members are the line integrals, thenðð R ∂F1 ∂x2 dx1dx2 ¼ � ð BFC F1 x 1; x2 � � dx1 � ð AEB F1 x 1; x2 � � dx1 ¼ þ C F1 x 1; x2 � � dx1 If the segment of curve C is parallel to axis OX2, the results of the integrals are not modified (Fig. 3.6). The integral ð F1dx 1 is cancelled in segment GH, for x1 ¼ constant then dx1 ¼ 0. The same occurs for segment PQ. With the segmentQGgiven byx2 ¼ f x1ð Þ, and the segmentHPgivenbyx2 ¼ g x1ð Þ: � ðð R ∂F1 ∂x2 dx1dx2 ¼ þ C F1 x 1; x2 � � dx1 ð3:2:2Þ and in the same way � ðð R ∂F2 ∂x1 dx1dx2 ¼ þ C F2 x 1; x2 � � dx2 ð3:2:3Þ Adding expressions (3.2.2) and (3.2.3) results inðð R ∂F2 ∂x1 � ∂F1 ∂x2 � � dx1dx2 ¼ þ C F1dx 1 þ F2dx2 � � Q:E:D: To prove the validity of this theorem for the more general cases being the region R ¼ R1 [ R2, in which the integrals are calculated for each subregion (Fig. 3.7). C 2 X 1XO P Q H G Fig. 3.6 Region simply connects with segments parallel to one of the coordinate axes 3.3 Green’s Theorem 143 In the segment ST the line integrals are calculated twice, but as they are of different direction they cancel each other when they are added, henceþ TS F1dx 1 þ F2dx2 � �þ þ ST F1dx 1 þ F2dx2 � � ¼ 0 Therefore, the expression of Green’s theorem is valid for the subdivision of region R (Fig. 3.7). This ascertaining is generalized for a finite region R ¼ R1 [ R2 � � �RN comprising N simple regions, with the outline curvesCi, i ¼ 1, 2, . . . ,N, thenðð R ∂F2 ∂x1 � ∂F1 ∂x2 � � dx1dx2 ¼ XN i¼1 þ Ci F1dx 1 þ F2dx2 � � The consequence of this division of region R into parts is that this theorem can be applicable to multiply connected regions (Fig. 3.8). The region involved by the curve TSBSTAT is simply connected, so Green’s theorem is valid for this region, hence ðð R ∂F2 ∂x1 � ∂F1 ∂x2 � � dx1dx2 ¼ þ TSBSTAT F1dx 1 þ F2dx2 � � To demonstrate the validity of Green’s theorem for this kind of region, let the line integrals written in a symbolic wayð TS þ ð C2 þ ð ST þ ð C1 ¼ ð C2 þ ð C1 ¼ þ C C X 2 X 1O T S R2 R1 Fig. 3.7 Division of the simply connected region into two simply connected regions 144 3 Integral Theorems for ð TS ¼ � ð ST therefore ðð R ∂F2 ∂x1 � ∂F1 ∂x2 � � dx1dx2 ¼ þ C F1dx 1 þ F2dx2 � � proves the previous statement. With the condition ∂F2 ∂x1 ¼ ∂F1 ∂x2 in the region R it follows by Green’s theorem þ C F1dx 1 þ F2dx2 � � ¼ 0 thus the line integral is independent of the path on the closed curveC. To demonstrate that the admitted condition is necessary and sufficient being the segments C1 and C2 of the curve C shown in Fig. 3.9, for the line integral it follows thatþ ADBEA F1dx 1 þ F2dx2 � � ¼ 0 Writing the line integrals of the various segments of curve C under symbolic form A 2X 1XO T R 1C 2C 1 R S B Fig. 3.8 Multiply connected regions 3.3 Green’s Theorem 145 ð ADB þ ð BEA ¼ 0 ð ADB ¼ � ð BEA ¼ ð AEB then þ C1 F1dx 1 þ F2dx2 � � ¼ þ C2 F1dx 1 þ F2dx2 � � where by ∂F2 ∂x1 ¼ ∂F1 ∂x2 is the necessary and sufficient condition for this independence. To admit that a parallel straight line of a coordinated axis intersects the region R in only two points is not essential, because R can be divided into a number of subregions which separately fulfill this property. In vectorial notation with the function F ¼ F1iþ F2 j and the position vector r ¼ x1iþ x2j, and in differential form dr ¼ dx1iþ dx2j, the line integral along the curve C is given by þ C F1dx 1 þ F2dx2 � � ¼ þ C F � dr A R 1 C 2 C B D E Fig. 3.9 Segments C1 and C2 of the closed curve C 146 3 Integral Theorems 3.4 Stokes’ Theorem Consider the surface S with upward unit normal vector n involved by a closed smooth curve C with S to its left, which direction cosines are ni > 0. Let the continuous real functions F1(x 1; x2; x3), F2(x 1; x2; x3), F3(x 1; x2; x3) with continuous partial derivatives in S [ C. Thenðð S ∂F3 ∂x2 � ∂F2 ∂x3 � � n1 þ ∂F1∂x3 � ∂F3 ∂x1 � � n2 þ ∂F2∂x1 � ∂F1 ∂x2 � � n3 � dS ¼ þ C F1dx 1 þ F2dx2 þ F3dx3 � � ð3:3:1Þ To demonstrate this theorem admit that a line parallel to axis OX3 intersects S only in a point, then the projection of S on the plane OX1X2 will be the region S12 involved by the closed smooth curve C12 oriented positively (Fig. 3.10), then dS12 ¼ n3dS ð3:3:2Þ and n3 > 0. The equation of surface S is given explicitly by x3 ¼ ϕ x1; x2ð Þ, which allows substituting the line integral along the curve C by the line integral along curve C12:þ C F1 x 1; x2; x3 � � dx1 ¼ þ C12 F1 x 1; x2;ϕ x1; x2 � �� � dx1 In the term to the right the coordinate x2 appears twice, in a direct way and in the function that represents the surface S. Applying Green’s theorem it follows thatþ C F1 x 1; x2;ϕ x1; x2 � �� � dx1 ¼ � ðð S12 ∂F1 ∂x2 þ ∂F1 ∂ϕ ∂ϕ ∂x2 � � dx1dx2 where dS12 ¼ dx1dx2 and using expression (3.3.2) 3.4 Stokes’ Theorem 147 þ C F1 x 1; x2; x3 � � dx1 ¼ � ðð S ∂F1 x1; x2;ϕx1; x2ð Þ½ � ∂x2 þ ∂F1 x 1; x2;ϕ x1; x2ð Þ½ � ∂ϕ ∂ϕ x1; x2ð Þ ∂x2 � n3dS ð3:3:3Þ In Integral Calculus of Multiple Variables when studying the surface integrals of x3 ¼ ϕ x1; x2ð Þ the following expressions are deducted for the direction cosines of its upward unit normal vector n: n1 ¼ ∂ϕ ∂x1 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1þ ∂ϕ∂x1 � �2 þ ∂ϕ∂x2 � �2r ð3:3:4Þ 1X O S 2X C 12 C 12 S S 3X a b c 1C 2C S B A Fig. 3.10 Stokes theorem: (a) projection of the smooth surface S with upward unit normal vector n on the plane OX1X2; (b) surface delimited by the closed smooth curve C; and (c) surface with outline delimited by more than one curve 148 3 Integral Theorems n2 ¼ ∂ϕ ∂x2 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1þ ∂ϕ∂x1 � �2 þ ∂ϕ∂x2 � �2r ð3:3:5Þ n3 ¼ 1 � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1þ ∂ϕ∂x1 � �2 þ ∂ϕ∂x2 � �2r ð3:3:6Þ As the direction cosines are positive, expression (3.3.5) provides n3 ∂ϕ x1; x2ð Þ ∂x2 ¼ �n3 and substituting this expression in expression (3.3.3)þ C F1 dx 1 ¼ � ðð S n2 ∂F1 ∂x3 � n3 ∂F1∂x2 � � dS In an analogous way for the projections of S on the planes OX2X3 and OX3X1 it follows that þ C F2 dx 2 ¼ � ðð S n3 ∂F2 ∂x1 � n1 ∂F2∂x3 � � dS þ C F3 dx 3 ¼ � ðð S n1 ∂F3 ∂x2 � n2 ∂F3∂x1 � � dS Adding these three expressions results inðð S ∂F3 ∂x2 � ∂F2 ∂x3 � � n1 þ ∂F1∂x3 � ∂F3 ∂x1 � � n2 þ ∂F2∂x1 � ∂F1 ∂x2 � � n3 � dS ¼ þ C F1dx 1 þ F2dx2 þ F3dx3 � � Q:E:D: Admit that a line parallel to one of the coordinate axis cuts the surface S only in a point is not an essential premise. Figure 3.10b, c shows two kinds of surface that do not fulfill this condition. In this case the surfaces must be divided into a finite number of subsurfaces, which separately fulfills this hypothesis, allowing Stokes’ theorem to be applied to these subsurfaces, and add the partial results obtained. Then the line integrals referent to the outlines common to the projections of these surfaces on a plane of the coordinate system cancel each other, for they are integrated twice, but with the signs changed. 3.4 Stokes’ Theorem 149 For a surface formed by several closed curves it is also possible to apply Stokes’ theorem. Figure 3.10c shows a surface S limited by the closed and smooth curves C1 and C2. The section S along the curve AB generates a new surface, which outlines are the curves C1, C2 and AB, considered in opposite directions. Then the line integral referent to curve AB is calculated twice, but with opposite signs, whereby it cancels itself, leaving only the results referent to the line integrals of the curves C1 and C2. The Stokes theorem is a generalization of Green’s theorem for the tridimensional space. In vectorial notation with the function F ¼ F1iþ F2 jþ F3k and the vector r ¼ x1iþ x2jþ x3k, which differential is dr ¼ dx1iþ dx2jþ dx3k, the line integral along the curve C is given byþ C F1dx 1 þ F2dx2 þ F3dx3 � � ¼ þ C F � dr ð3:3:7Þ The surface integrals that are present in Stokes’ theorem also have a vectorial interpretation (item 4.4). 3.5 Gauß–Ostrogradsky Theorem Consider the volume V with upward unit normal vector n involved by a closed and smooth surface S, which direction cosines are ni > 0. Let the continuous real functions F1(x 1; x2; x3), F2(x 1; x2; x3), F3(x 1; x2; x3) with continuous par- tial derivatives in V [ S. Thenððð V ∂F1 ∂x1 þ ∂F2 ∂x2 þ ∂F3 ∂x3 � � dx1dx2dx3 ¼ ðð S F1n1 þ F2n2 þ F3n3ð ÞdS ð3:4:1Þ Consider a line parallel to axis OX2 that intersects the surface S in a maximum of two points P and P0, with upward unit normal vector n(P) and n(P0), respectively (Fig. 3.11). Then the projection of S on OX3X1 will be S31, it follows thatððð V ∂F2 ∂x2 dx1dx2dx3 ¼ ðð S31 ð ∂F2 ∂x2 dx2 � � dS31 ¼ ðð S31 F2 P 0 � � � F2 Pð Þ h i dS31 150 3 Integral Theorems For the area element in this plane and with the direction cosines of the upward normal n(P) and n(P0): dS31 ¼ dS Pð Þn2 Pð Þ ¼ �dS P0 � � n2 P 0 � � Substituting it results for the point P on S:ððð V ∂F2 ∂x2 dx1dx2dx3 ¼ ðð S F2 Pð Þn2 Pð ÞdS In an analogous way, for the projections of S on the planes OX1X2 and OX2X3:ððð V ∂F3 ∂x3 dx1dx2dx3 ¼ ðð S F3 Pð Þn3 Pð ÞdS ððð V ∂F1 ∂x1 dx1dx2dx3 ¼ ðð S F1 Pð Þn1 Pð ÞdS 3 X O V S Pn 31 dS C ' Pn 1 X 2 X 31 S j Fig. 3.11 Volume V with upward unit normal vector n, which outline is a closed and smooth surface S 3.5 Gauß–Ostrogradsky Theorem 151 Free ebooks ==> www.Ebook777.com The addition of these three expressions resultsððð V ∂F1 ∂x1 þ ∂F2 ∂x2 þ ∂F3 ∂x3 � � dx1dx2dx3 ¼ ðð S F1n1 þ F2n2 þ F3n3ð ÞdS Q:E:D: One of the premises adopted in the proof of the theorem of Carl Friedrich Gauß and Mikhail Vasilievich Ostrogradsky (1801–1861) is that the surface S has two sides, with a single upward and inward normal in each point. To admit that a straight line parallel to a coordinate axis intersects the volume V in only two points is not an essential hypothesis, for V can be divided into a number of subvolumes that separately fulfill the property admitted initially, allowing the Gauß–Ostrogradsky theorem to be applied to these subvolumes and adding the partial results obtained. Figure 3.12a shows the volume V cut in more than two points by a straight line parallel to axis OX2. The division of V into two volumes V1 and V2, separated by surface S*, with opposite unit normal vector n1 and n2, being V1 involved by S1 [ S*, and V2 by S2 [ S*. Then the surface integrals referent to this part common to the two volumes cancel each other, remaining the integrals on the surfaces S1 and S2. This makes the applying of this theorem valid to volume V. If the closed surface S that involves volume V is not smooth, it can be divided into a finite number of smooth surfaces represented by the functions ϕ(xi), which have continuous partial derivatives, each one involving a subvolume. This proce- dure allows applying the Gauß–Ostrogradsky theorem to these subvolumes and adding the results obtained. 3X 1X O 2X 2V 1V 2n 1n V V 1V 2V S S 2S 1S * 1S * 2S *S *S a b Fig. 3.12 Gauß–Ostrogradsky theorem: (a) volume cut in more than two points by a straight line parallel to a coordinated axis and (b) volume V with voids V1 and V2 152 3 Integral Theorems www.Ebook777.com http://www.ebook777.com Figure 3.12b shows the volume V involved by the closed surface S with empty volumes V1 and V2, with which are involved, respectively, by the smooth closed surfaces S1 and S2. In this case it is necessary to cut the total volume and the volumes of the voids by a plane π and the surfaces of their outlines to project in this plane, originating the surfaces S*, S�1 and S � 2, and then apply the Gauß–Ostrogradsky theorem considering these surfaces. In vectorial notation with F ¼ F1iþ F2 jþ F3k resultsððð V ∂F1 ∂x1 þ ∂F2 ∂x2 þ ∂F3 ∂x3 � � dx1dx2dx3 ¼ ðð S F � ndS ð3:4:2Þ The volume integral that is present in Gauß–Ostrogradsky theorem also has a vectorial interpretation (item 4.3). 3.5 Gauß–Ostrogradsky Theorem 153 Chapter 4 Differential Operators 4.1 Scalar, Vectorial, and Tensorial Fields 4.1.1 Initial Notes The study of the scalar, vectorial, and tensorial fields is strictly related with the differential operators which are applied to the analytic functions that represent these fields. In this chapter the differential operators gradient, divergence, and curl will be defined, and their physical interpretations, as well as various fundamental relations with these operators, will be presented. These expressions form the mathematical backbone for the practical applications of the Field Theory. The conception of fields is of fundamental importance to the formulation of Tensor Calculus, and allows definingvarious concepts and deducing several expressions which form the frame- work for the study of the tensors contained in the tensorial space that defines the field. The scalar, vectorial, and tensorial fields are formulations carried out on a point xi2D, the domain D � EN being an open subset and embedded in the ordinary geometric space. In these three kinds of fields the formulations are the functions smooth, continuous, and derivable. By defining an arbitrary origin in the space EN a biunivocal correspondence is determined for each domain with a variety, scalar, vector, or tensor that defines the kind of field. The scalar and vectorial fields are particular cases of the tensorial fields. The behavior of a tensorial field is measured by the variation rate of the tensor in the points contained in the field. In the literature it is usual to call this variation rate as tensor derivative, which is incorrect, for what exists is the variation rate of the field defined by this variety, so the proper denomination is variation of the tensorial field. However, on account of being customary by use, the denomination tensor deriv- ative will be used in this text to express this variation. © Springer International Publishing Switzerland 2016 E. de Souza Sánchez Filho, Tensor Calculus for Engineers and Physicists, DOI 10.1007/978-3-319-31520-1_4 155 4.1.2 Scalar Field Let a scalar be associated to a point in the Euclidian space E3 given by a function of the coordinates xi, which is defined asϕ ¼ ϕ xi; tð Þ, i ¼ 1, 2, 3, where t is the time in the instant in which the scalar is measured. A scalar field is defined by the function of field ϕ(xi; t), and if the time variable t is constant, the level surface of the field ϕ xið Þ ¼ C is defined, where C is a constant. For several values of C there is a family of level surfaces, which characterize the field geometrically. These surfaces do not intersect, for if they did the function ϕ(xi) would have to assume various values, which is impossible, for this function has only one value for each xi. As an example of scalar field there is a point in the interior of a reservoir containing liquid, where each particle of this fluid is submitted to a pressure proportional to the distance of this particle up to the top of the free surface. Another example is the field of temperatures due to a heat source, where the isotherms are spherical surfaces, with the temperature decreasing to the points farthest from this source. 4.1.3 Pseudoscalar Field If the field function defines a pseudoscalar then the field is pseudoscalar. The specific mass ρ(xi) of the points of a solid body is an example of this sort of field. 4.1.4 Vectorial Field If the vector u(xi, t) is associated with the point P(xi) of the space EN, then a vectorial field is defined, and if t¼ constant the field is homogeneous. For the space E3 which points are referenced to a Cartesian coordinate system there are three scalar functions of these points, f 1 x ið Þ, f 2 xið Þ, f 3 xið Þ, i ¼ 1, 2, 3, which express the field vector u xi � � ¼ f 1 xi� �iþ f 2 xi� �jþ f 3 xi� �k Field lines are defined for a vectorial field determined by the vectorial function u(xi), in which for each point P(xi) the field vectors are collinear with the vectors tangents to these lines (Fig. 4.1). The condition of collinearity between the vector u(xi) and the tangent vector t(xi) is given by the nullity of cross product εijkujdxk ¼ 0 156 4 Differential Operators Developing provides: – i ¼ 1 ε1jkujdxk ¼ 0 ) ε12ku2dxk ¼ 0 ) ε123u2dx3 ¼ 0 ) u2dx3 ¼ 0 ε13ku3dx2 ¼ 0 ) ε132u3dx2 ¼ 0 ) �u3dx2 ¼ 0 – i ¼ 2 ε2jkujdxk ¼ 0 ) ε213u1dx3 ¼ 0 ) �u1dx3 ¼ 0 ε2jkujdxk ¼ 0 ) ε231u3dx1 ¼ 0 ) u3dx1 ¼ 0 – i ¼ 3 ε3jkujdxk ¼ 0 ) ε312u1dx2 ¼ 0 ) u1dx2 ¼ 0 ε3jkujdxk ¼ 0 ) ε321u2dx1 ¼ 0 ) �u2dx1 ¼ 0 Thus the following system results u3dx2 ¼ u2dx3 u1dx2 ¼ u2dx1 u3dx1 ¼ u1dx3 8><>: ) u3 u2 ¼ dx3 dx2 u1 u2 ¼ dx1 dx2 u3 u1 ¼ dx3 dx1 8>>>>>><>>>>>>: Field line s t u P a b Fig. 4.1 Vectorial field: (a) field lines and field vector and (b) field vectors 4.1 Scalar, Vectorial, and Tensorial Fields 157 For a flat vectorial field the condition of collinearity between the field vector and the vector tangent to the field lines is given by u2dx1 � u1dx2 ¼ 0 The gravitational, the electric, and the magnetic are examples of vectorial fields. 4.1.5 Tensorial Field The fundamental problem of Tensor Calculus is associated to the concept of tensorial field. If the tensorial field is fixed the tensor T(xi) is a function of the coordinates of a point P(xi) situated in the tensorial space EN. When this tensor is function of xi and other parameters then the tensorial field is variable. For tensor T xið Þ which components are defined with respect to a curvilinear coordinates X i which origin is the point P(xi), a few difficulties arise in the calculation of its derivatives, because the local coordinate system varies as a function of the point. The study of the tensorial fields in a tensorial space EN, considering curvilinear local coordinate systems, is associated to the basis of this space. Exercise 4.1 Calculate the parametric equation of the lines of the vectorial field u ¼ �x2iþ x1 jþ mk that contains the point of coordinates (1; 0; 0) where m is a scalar. The differential equations of the field lines are dx1 �x2 ¼ dx2 x1 ¼ dx3 m Following with the first two differential relations x1dx1 þ x2dx2 ¼ 0 ) ð x1dx1 þ ð x2dx2 ¼ C0 ) x1ð Þ2 þ x2ð Þ2 ¼ C1; C1 > 0 and introducing a parameter t x1 ¼ ffiffiffiffiffiffi C1 p cos t x2 ¼ ffiffiffiffiffiffi C1 p sin t so dx2 ¼ ffiffiffiffiffiffi C1 p cos t � � dt and with the differential relations 158 4 Differential Operators dx2 x1 ¼ dx3 m it follows that ffiffiffiffiffiffi C1 p cos t � � dtffiffiffiffiffiffi C1 p cos t ¼ dx3 m ) dx3 ¼ mdt ) x3 ¼ mtþ C2 As the field line contains the point of coordinates (1; 0; 0), then 1 ¼ ffiffiffiffiffiffi C1 p cos t ) t ¼ 2kπ; k ¼ 0, 1, . . . – k ¼ 0 ! C1 ¼ 1; – k ¼ 0 ! t ¼ 0 so C2 ¼ 0; – verifying that 0 ¼ mtþ C2 for t ¼ 0. The parametric equations of the field lines represent a circular helix given by x1 ¼ cos t; x2 ¼ sin t; x3 ¼ mt 4.1.6 Circulation Consider the field defined by the vectorial function u and the point P(xi) located on an open curve C, continuous by parts, smooth, and derivable, which is the hodograph of the position vector r(s), where s is the curvilinear abscissa, and admitting that this point varies in the interval a xi b, then the line integral of this curve is given by I ¼ ðb a u � dr ð4:1:1Þ where line integral defines the circulation of the vectorial function u on the curve C. Let u � dr be the differential total of the function ϕ(xi), thus I ¼ ðb a dϕ xi � � ¼ϕ xi� � b a ¼ ϕ bð Þ � ϕ að Þ ð4:1:2Þ The value of this integral depends only on the extreme points of the interval for which the function ϕ(xi) is defined, regardless of the integration path. Expression (4.1.2) is a generalization of the fundamental theorem of the Integral Calculus. 4.1 Scalar, Vectorial, and Tensorial Fields 159 For a closed curve the extreme points of this interval are coincident, which allows concluding that þ C u � dr ¼ 0 ð4:1:3Þ This expression defines the circulation of vector u along the closed curve C. The line integral of an open curve C defined in a certain interval will be independent of the path adopted in this calculation, and will be null if the curve C is closed. Figure 4.2 shows two types of closed paths of curves defined in domain D—the single closed path in which there are no self-intersection points and the closed path with self-intersection points. For the closed spatial curves defined in the domain D self-intersecting in a finite number of points, the line integral is calculated dividing the path in a finite number of single closed paths. For an infinite number of intersections, a reasonable approx- imation is obtained with the integrals on paths which arepolygonal segments, using a limit process to achieve a finite number of intersections. 4.2 Gradient In item 2.2 the gradient of a scalar field was defined, by a function ϕ(xi) which differential is given by dϕ ¼ ∂ϕ ∂xi dxi ð4:2:1Þ that is called differential parameter of the first order of Beltrami. Expression (4.2.1) shows that there is no difference between the total differential dϕ and the absolute differential, which allows adopting the notation ϕ,i for the partial derivatives of this function. It was also shown that the gradient is a vector Path Path Path Path 1C 1C 2C 2C b b C C D D a b Fig. 4.2 Closed curve paths: (a) with no self-intersection and (b) with a finite number of self- intersections 160 4 Differential Operators obtained by means of applying a vectorial operator to the scalar function ϕ(xi), that with respect to a coordinate system Xi is given by ∇� � � ¼ ei ∂� � � ∂xi For a curvilinear coordinate system X j by the chain rule ∂� � � ∂xi ¼ ∂x j ∂xi ∂� � � ∂xj and with the transformation law of unit vectors ei ¼ gk ∂x i ∂xk it follows that ∇� � � ¼ ei ∂� � � ∂xi ¼ gk ∂x i ∂xk ∂� � � ∂xi ¼ gk ∂x i ∂xk ∂xj ∂xi ∂� � � ∂xj ¼ gkδ jk ∂� � � ∂xj ¼ gk ∂� � � ∂xk The several notations for the gradient vector are gradϕ xi � � ¼ G ϕð Þ ¼ ∇ ϕð Þ ¼ gk ∂ϕ xið Þ ∂xk ¼ gkϕ, k ð4:2:2Þ This comma notation will hereafter be used in some special case for derivatives with respect to coordinates. The classic notation for the operator that defines the gradient of a scalar function is gradϕ(xi), and was introduced by Maxwell, Rie- mann, and Weber. The other notation is an inverted delta, called nabla operator (in Greek να 0 βλα ¼ harp), del, atled (inverted delta), expressed as ∇� � � ¼ gk ∂���∂xk. This notation was designed by Hamilton in 1837, initially was not used to represent the gradient of a function, but was written with the rotated delta ⊲, and represented symbolically the Laplace operator ddx � �2 þ ddy� �2 þ ddz� �2 that was already well used at the time, thereby the denomination Hamilton operator, or Hamiltonian operator. Another interpretation for the name nabla is due to Maxwell, who remarks that the rotated delta calls to cuneiform writing, which name in Hebrew would be this one. The use of the nabla operator has many advantages with respect to the spelling grad, especially in the development of expressions, for it reinforces the tensorial characteristics of the gradient. This symbolic vector enables making the spelling for the differential operators uniform, and complies with the Vectorial Algebra rules. Figure 4.3a shows schematically a scalar field defined by a function ϕ(xi), where in a field line contained in the level surface the ϕ xið Þ ¼ C, being C¼ constant, a point P is defined, and with an arbitrary origin O for the coordinate system Xi, 4.2 Gradient 161 results in the position vector r, which derivative is the vector dr ¼ dxkgk tangent to the field line, and denotes the line element. The differential of the scalar function that represents this field is given by the dot product dϕ ¼ dr �∇ϕ ¼ dxkgk � g‘ ∂ϕ ∂x‘ ¼ δ ‘k ∂ϕ ∂x‘ dxk ¼ ∂ϕ ∂xk dxk The field represented by the gradient of a function is conservative, thus this function is called potential, or field gradient. As the operator nabla is a vector, it is invariant for a change in the coordinate system, which can be proven admitting ∇ for a coordinate system Xj, and ∇ for a coordinate system Xi, so by means of the vectors transformation law ∂� � � ∂xk ¼ ∂� � � ∂x‘ � � ∂x‘ ∂xk ∇� � � ¼ gk ∂� � � ∂xk ¼ ∂x k ∂xm ∂x‘ ∂xk gm ∂� � � ∂x‘ ¼ δ ‘mgm ∂� � � ∂x‘ ¼ g‘ ∂� � � ∂x‘ ¼ ∇� � � therefore the operator ∇ is a vector. The gradient for the function ϕ ¼ xk, where xk represents a coordinate of the referential system, is given by ∇ϕ ¼ ∂xk∂xi gi, then the gradient is the unit vector for the coordinate axis. For the product of two scalar functions ϕ(xi) and ψ(xi) the result is P Cxi =φ rd P td rd φ∇ O b (P)u element of curve line C a b Fig. 4.3 Scalar field: (a) gradient and (b) line element 162 4 Differential Operators grad ϕψð Þ ¼ ∇ ϕψð Þ ¼ gk ∂ ϕψð Þ ∂xk ¼ gk ∂ϕ ∂xk ψ þ ϕgk ∂ψ ∂xk ¼ ∇ϕð Þψ þ ϕ ∇ψð Þ In this demonstration it is observed that the nabla operator acts on each parcel of the expression in a distinct way, maintaining a parcel variable and the other constant. If it comes before the parcel it acts with a variable, if it comes after, it acts as a constant. The gradient can be defined by means of the Gauß-Ostrogradsky theorem. Let the field be determined by the vectorial function u ¼ vϕ xið Þ, v being a constant vector, then ∂u1 ∂xi ¼ v � ∂ϕ x ið Þ ∂xi ∂u1 ∂x1 þ ∂u 2 ∂x2 þ ∂u 3 ∂x3 ¼ v � ∂ϕ ∂x1 þ ∂ϕ ∂x2 þ ∂ϕ ∂x3 � � ¼ v �∇ϕ it follows that ððð V ∂u1 ∂x1 þ ∂u 2 ∂x2 þ ∂u 3 ∂x3 � � dV ¼ ðð S u � ndS v � ððð V ∇ϕdV ¼ v � ðð S u � ndS v � ððð V ∇ϕdV � ðð S u � n dS 0@ 1A ¼ 0 For the point P in the scalar field ϕ(xi) contained in an elementary volume, and with the component ∂ϕ∂x1 of ∇ϕ by the mean value theorem of the Integral Calculusððð V ∂ϕ ∂x1 dV ¼ ∂ϕ ∂x1 � � P* V where P* is the midpoint of volume V. Applying the Gauß-Ostrogradsky theorem ∂ϕ ∂x1 � � P* ¼ 1 V ðð S ϕn1dS where n1 is the direction cosine of the angle between the upward normal unit vector n and the coordinate axis OX1. 4.2 Gradient 163 When the point P approaches the point P*, the volume V and the surface S also come close to P, and with continuous function ϕ(xi) and its partial derivatives it results in ∂ϕ ∂x1 � � P ¼ lim V!0 1 V ðð S ϕn1dS For the coordinates x2 and x3 the result with analogous formulations is ∂ϕ ∂x2 � � P ¼ lim V!0 1 V ðð S ϕn2dS ∂ϕ ∂x3 � � P ¼ lim V!0 1 V ðð S ϕn3dS If these limits exist, the gradient of the scalar function ϕ(xi) in point P is determined by ∇ϕ xi � � ¼ lim V!0 1 V ðð S ϕ xi � � ndS ð4:2:3Þ that is valid for any coordinate system, which shows that the gradient is independent of the coordinate system. 4.2.1 Norm of the Gradient The norm of the gradient is given by ∇ϕk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ∇ϕ �∇ϕk k p ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi gij ∂ϕ ∂xi ∂ϕ ∂xj r ð4:2:4Þ A few authors use the spelling Δ1ϕ ¼ gij ∂ϕ∂xi ∂ϕ∂xj to designate the first differential parameter of Beltrami. For the orthogonal coordinate systems gij ¼ δij: ∇ϕk k ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ∂ϕ ∂xi � �2s ð4:2:5Þ 164 4 Differential Operators 4.2.2 Orthogonal Coordinate Systems Consider the point P(xi) be coincident with the origin of the curvilinear orthogonal coordinate X j , and r its position vector with respect to the Cartesian coordinate Xj. Rewriting expression (2.3.4) the result for this vector’s differential is dr ¼ ∂r ∂xi dxi with ∂r ∂xi ¼ higi then hi ¼ ∂r ∂xi ���� ���� where gi are the unit vectors of the coordinate system X j , and hi ¼ ffiffiffiffiffiffiffiffig iið Þp are the metric tensor coefficients that represent scale factors of the magnitudes of the vectors tangents to the curves of this coordinate system, where the indexes in parenthesis do not indicate summation. They are called Lamé coefficients for orthogonal coordinate systems. The differential of the scalar function ϕ(xi) is given by dϕ ¼ ∂ϕ ∂xi dxi but dϕ ¼ ∇ϕ � dr dr ¼ hidxigi then dϕ ¼ ∂ϕ ∂xi dxi � � gi � dr whereby ∇ϕ ¼ 1 hi ∂ϕ ∂xi � � gi ð4:2:6Þ 4.2 Gradient 165 http://dx.doi.org/10.1007/978-3-319-31520-1_2 that provides the components of the gradient vector in a curvilinear orthogonal coordinate system. The physical components of the vector ∇ϕ are given by ∇ϕð Þ* ¼ 1 hi ∂ϕ ∂xi ð4:2:7Þ 4.2.3 Directional Derivative of the Gradient Figure 4.3b shows the differential element of the line contained in the level surface ϕ xið Þ ¼ C and its tangent unit vector t, collinear with vector dr, which allows writing for the line element dr ¼ tds. The geometric interpretationof the gradient of a scalar field is given by the dot product dϕ ds ¼ t �∇ϕ ð4:2:8Þ that defines the field directional derivative. The symbol∇ϕ characterizes the field, and unit vector t being independent of the function ϕ(xi), this indicates the direction in which the derivative is calculated. If∇ϕ xið Þ exists in the point P, defined by the field ϕ(xi), it will be possible to calculate the directional derivative of this function in all the directions of the field. Then the field ϕ(xi) is non-homogeneous. Let α be the angle between the two vectors from expression (4.2.8), the dot product provides dϕ ds ¼ tk k ∇ϕk k cos α ¼ ∇ϕk k cos α As ϕ¼ constant, it results in dϕ ¼ 0, so dr �∇ϕ ¼ 0, then the vector ∇ϕ is perpendicular to the vector dr. The variation rate of the field defined by the function ϕ is maximum in the direction of ∇ϕ, for α ¼ 0 results in cos α ¼ 1, then dϕ dn � � max ¼ ∇ϕk k > 0 The directional derivative is calculated in the direction of the unit normal vector n to the level surface ϕ xið Þ ¼ C (Fig. 4.4), thus ∇ϕ ¼ dϕ dn n ð4:2:9Þ 166 4 Differential Operators 4.2.4 Dyadic Product The nabla operator applied to a vectorial function u ¼ ukgk results in the dyadic product ∇� u ¼ ∇u ¼ T T ¼ gi ∂u ∂xi ¼ gi ∂ u kgk � � ∂xi ¼ gi ∂u k ∂xi gk þ uk ∂gk ∂xi � � and with expression (2.3.10) ∂gk ∂xi ¼ Γmkigm it follows that T ¼ gi ∂u k ∂xi gk þ ukΓmkigm � � Interchanging the indexes m $ k in the second member to the right T ¼ ∂u k ∂xi þ umΓ kmi � � gi � gk The covariant derivative of a contravariant vector results in a variance tensor (1, 1) tP Cxi =φ Fig. 4.4 Interpretation of the gradient as a vector normal to surface 4.2 Gradient 167 http://dx.doi.org/10.1007/978-3-319-31520-1_2 T ki ¼ ∂iuk ¼ ∂uk ∂xi þ umΓ kmi then T ¼ T ki gi � gk ð4:2:10Þ This analysis shows that the gradient and the covariant derivative represent a same concept, i.e., they represent the derivative of a scalar, vectorial, or tensorial function, increasing their variance from one unit. Formulating an analogous analysis for a covariant vector T ¼ ∇� u ¼ ∇u ¼ gi ∂u ∂xi T¼ gi ∂ ukg k � � ∂xi ¼ gi ∂uk ∂xi gk þ uk ∂g k ∂xi � � ¼ gi ∂uk ∂xi gk � umΓmkigm � � ¼ gi ∂uk ∂xi � ukΓ kmi � � gi � gk then ∇� u ¼ ∇u ¼ ∂iukð Þgi � gk ¼ Tikgi � gk ð4:2:11Þ The differential of a vectorial field is a vector, for the differential of vector u: du ¼ dr �∇� u ¼ dxigi � � ∂iukg i � gk� � ¼ dxi∂iukgi � gi � gk whereby du ¼ dxi∂iukgk ð4:2:12Þ For the fields vectorial there is the directional derivative du ds ¼ t �∇� u ð4:2:13Þ The same considerations formulated for the directional derivative of a scalar field ϕ(xi) are applicable to the vectorial fields. The physical components for the gradient of vector u* are obtained considering the physical components of the second-order tensor. 168 4 Differential Operators 4.2.5 Gradient of a Second-Order Tensor The generalization of the concept of gradient for an arbitrary tensorial field is immediate. For a coordinate system Xi with unit vector g‘, and with the tensor T defining the tensorial field ∇� T ¼ gradT ¼ g‘ ∂T ∂x‘ ð4:2:14Þ Then the gradient of a tensor T is calculated by nabla operator ∇ ¼ g‘ ∂���∂x‘ applying to this tensor. This operator is defined for a contravariant base. The tensorial product of the nabla operator by the second-order tensor T is given by ∇� T ¼ ∇T ¼ gm ∂ T kigk � gi � � ∂xm ¼ gm ∂T ki ∂xm gk � gi þ Tki ∂gk ∂xm � gi þ Tkigk ∂gi ∂xm � � and with the expressions ∂gk ∂xm ¼ Γ pkmgp ∂gi ∂xm ¼ Γ pimgp it follows that ∇� T ¼ gm ∂T ki ∂xm gk � gi þ TkiΓ pkmgp � gi þ TkiΓ pimgk � gp � � Interchanging the indexes p $ k in the second term to the right, and the indexes p $ i in the third term ∇� T ¼ ∂T ki ∂xm þ TkiΓ pkm þ TkiΓ pim � � gm � gk � gi this expression becomes ∇� T ¼ ∂mTkigm � gk � gp ð4:2:15Þ and shows that the gradient of a second-order tensor is a variance tensor (1, 2). The other components of the gradient of tensor T are given by expressions (2.5.18) and (2.5.21). The generalization of the definition of the gradient of a third-order tensor T is immediate. The components of the fourth-order tensor that result from applying this operator to tensor T being given by their covariant derivatives, for instance, for tensor Tijk: 4.2 Gradient 169 http://dx.doi.org/10.1007/978-3-319-31520-1_2 http://dx.doi.org/10.1007/978-3-319-31520-1_2 ∇� T ¼ ∂Tijk ∂x‘ � TmjkΓmi‘ � TimkΓmj‘ � TijmΓmk‘ � � g‘ � gi � g j � gk For a tensor T of order p the variety ∇� T is a tensor of order pþ 1ð Þ. The differential of a tensorial field is a tensor, is gives by then the differential of the second-order tensor T thus: dT¼ dr �∇� T ¼ dxjgj � ∂mTkigm � gk � gp ¼ dxj∂mTkigj � gm � gk � gp ¼ dxj∂mTkiδmj gk � gp whereby dT ¼ dxj∂jTkjgk � gp ð4:2:16Þ The physical components for the gradient of tensor T* are obtained considering the physical components of the tensor of the third order. The same considerations formulated for the directional derivative of a scalar field ϕ(xi) and of a vectorial field are applicable to the tensorial fields, where dT ds ¼ t �∇� T ð4:2:17Þ 4.2.6 Gradient Properties The ascertaining achieved in the previous paragraphs allow establishing the condi- tions so that a vector is gradient of a scalar function, for if the vector u(xi) defined in a single or multiply connected region, and if the line integral ð C u � dr is independent of the path, then a scalar function ϕ(xi) exists and fulfills the condition u ¼ ∇ϕ xið Þ in all of this region of the space. The gradient operator applied to the addition of two tensors provides ∇� Tþ Að Þ ¼ g‘ ∂ Tþ Að Þ ∂x‘ ¼ g‘ ∂T ∂x‘ þ g‘ ∂A ∂x‘ ¼ ∇� Tþ∇� A The applying of this operator to the multiplication of the scalar m by the tensor T provides ∇� mTð Þ ¼ g‘ ∂ mTð Þ ∂x‘ ¼ mg‘ ∂T ∂x‘ ¼ m∇� T These two demonstrations prove that the gradient is a linear operator, which is already implicit, because it is a vector. 170 4 Differential Operators Exercise 4.2 Calculate: (a) v �∇u; (b) ∇ u � vð Þ. (a) The gradient for the field defined by a vectorial function is given by ∇u ¼ ∂iukgi � gk With v ¼ v‘g‘ it follows that v �∇u ¼ v‘g‘ � ∂iukgi � gk ¼ v‘∂iukg‘ � gi � gk ¼ v‘∂iukδ i‘gk ¼ vi ∂iukð Þgk Thus for the Cartesian coordinates v �∇u ¼ vi ∂ukð Þ∂xi gk (b) The gradient of the scalar field represented by the dot product of the vectorial functions u and v is given by ∇ u � vð Þ ¼ gk ∂ u ivið Þ ∂xk ¼ ∂kui � � vi þ ui ∂kvið Þ � � gk and with the expressions ∂ku i ¼ ∂u i ∂xk þ umΓ imk ∂kvi ¼ ∂vi ∂xk � vmΓmik it follows that ∇ u � vð Þ ¼ ∂u i ∂xk vi þ umΓ imkvi þ ui ∂vi ∂xk � uivmΓmik � � gk In the last term in parenthesis interchanging the indexes i $ m: ∇ u � vð Þ ¼ ∂u i ∂xk vi þ umΓ imkvi þ ui ∂vi ∂xk � umviΓ imk � � gk then ∇ u � vð Þ ¼ ∂u i ∂xk vi þ ui ∂vi∂xk � � gk 4.2 Gradient 171 For the Cartesian coordinates ∇ u � vð Þ ¼ ∂ui ∂xk vi þ ui ∂vi∂xk � � gk Another way of expressing ∇ u � vð Þ is to use the expression between the covariant derivative of a covariant vector and the covariant derivative of a contravariant vector, which is given by ∂kum ¼ gim∂kui The multiplying of this expression by gin provides gin∂kum ¼ gingim∂kui ¼ δnm∂kui ) ∂kui ¼ gim∂kumm whereby vi∂ku i ¼ vigim∂kum ¼ vm∂kum Replacing the dummy index m ! i: vi∂ku i ¼ vi∂kui and by substitution ∇ u � vð Þ ¼ vi∂kui þ ui∂kvi � � gk Adding and subtracting the terms vi∂iuk and ui∂ivk ∇ u � vð Þ ¼ vi ∂kui � ∂iukð Þ þ vi∂iuk þ ui ∂kvi � ∂ivkð Þ þ ui∂ivk � � gk and with the expressions v� ∇� uð Þ ¼ vi ∂kui � ∂iukð Þgk u� ∇� vð Þ ¼ ui ∂kvi � ∂ivkð Þgk v �∇u ¼ vi ∂iukð Þgk u �∇v ¼ ui ∂ivkð Þgk it results in ∇ u � vð Þ ¼ v �∇uþ v� ∇� uð Þ þ u �∇vþ u� ∇� vð Þ For the particular case in which u ¼ v: v �∇v ¼ 1 2 ∇v2 � v� ∇� vð Þ 172 4 Differential Operators Free ebooks ==> www.Ebook777.com Exercise 4.3 Calculate the gradient of the scalar function ϕ(xi) expressed in cylindrical coordinates. Forthe cylindrical coordinates ffiffiffiffiffiffi g11 p ¼ ffiffiffiffiffiffig33p ¼ 1, ffiffiffiffiffiffig22p ¼ r, then ∇ϕ ¼ 1ffiffiffiffiffi gii p ∂ϕ ∂xi gk it follows that ∇ϕ ¼ ∂ϕ ∂r gr þ 1 r ∂ϕ ∂θ gθ þ ∂ϕ ∂z gz Exercise 4.4 Calculate the gradient of the scalar function ϕ(x1) expressed in spherical coordinates. For the spherical coordinates ffiffiffiffiffiffi g11 p ¼ 1, ffiffiffiffiffiffig22p ¼ r, ffiffiffiffiffiffig33p ¼ r sinϕ, then ∇ϕ ¼ ∂ϕ ∂r gr þ 1 r ∂ϕ ∂ϕ gϕ þ 1 r sin ϕ ∂ϕ ∂xθ gθ Exercise 4.5 Show that ∂2ϕ xið Þ ∂xi∂xj is a second-order tensor, where ϕ(xi) is a scalar function. Putting Tij ¼ ∂ 2ϕ ∂xi∂xj ¼ ϕ, ij for the coordinate system X i ∂2ϕ ∂xi∂xj ¼ ∂ϕ ∂xi ∂ϕ ∂xk ∂xk ∂xk � � ¼ ∂ϕ ∂xm ∂ϕ ∂xk ∂xk ∂xj � � � ∂xm ∂xi ¼ ∂x k ∂xj ∂xm ∂xi � � ∂2ϕ ∂xm∂xk This transformation law proves that ∂2ϕ xið Þ ∂xi∂xj is a second-order tensor. Exercise 4.6 Calculate the directional derivative of the function ϕ x, yð Þ ¼ x2 þ y2 � 3xy3, at the point P(1; 2) in the direction of vector u ¼ 1 2 e1 þ ffiffi 3 p 2 e2, being e1(1; 0), e2(0; 1). The gradient of the scalar field is given by ∇ϕ ¼ 2x� 3y3� �e1 þ 2y� 9xy2� �e2 4.2 Gradient 173 www.Ebook777.com http://www.ebook777.com and in point P(1; 2) ∇ϕ ¼ �22e1 � 32e2 The vector u is a unit vector, for uuk k ¼ 12 e1 þ ffiffi 3 p 2 e2, whereby it follows that for the directional derivative ∇ϕ � u ¼ �22� 1 2 � 32� ffiffiffi 3 p 2 ¼ �11� 16 ffiffiffi 3 p 4.3 Divergence The analysis of the flow magnitude for the vectorial field u that passes through the volume V involved by the closed surface S with respect to this volume leads to the conception of a differential operator (Fig. 4.5a). In volume V let the elementary parallelepiped with sides dx1, dx2, dx3, and the vectorial function u continuous with continuous partial derivatives (Fig. 4.5b). The study of the flow of u that passes through the volume V with respect to this volume is carried out considering the point P(x1; x2; x3) at the center of an elemen- tary parallelepiped (Fig. 4.5). For the face dS1, with upward normal unit vector n (1; 0; 0), the component of u in the direction of axis OX1 is given by u � n ¼ u1 1X O n n n 2X 12S * 12S P 1dx 2dx 3dx u u 3X V S P a b Fig. 4.5 Flow of the vectorial function u: (a) that passes through the volume V and (b) in the elementary parallelepiped 174 4 Differential Operators The center of the elementary area dS1 ¼ dx1dx2 has coordinates x1 þ dx12 ; � dx2; dx3Þ, whereby it follows that for the surface integral in this face of the elementary parallelepiped dVðð dS1 u � ndS ffi u1 x1 þ dx 1 2 ; dx2; dx3 � � dx2dx3 for the area considered is elementary, which allows calculating approximately the surface integral as the dot product u:ndS ¼ u1dx2dx3 For the elementary face dS*1 ¼ dx2dx3 with upward normal unit vector n �1; 0; 0ð Þ centered in the midpoint x1 � dx1 2 ; dx2; dx3 � � it follows that in an analogous way ðð dS*1 u:ndS ffi �u1 x1 � dx 1 2 ; dx2; dx3 � � dx2dx3 Adding these contributionsðð dS1þdS*1 u:ndS ffi u1 x1 þ dx 1 2 ; dx2; dx3 � � � u1 x1 � dx 1 2 ; dx2; dx3 � � � dx2dx3 ffi u1 x1� �dx2dx3 The component u1 in the point of coordinate dx1 varies according to the rate du1 ¼ ∂u 1 ∂x1 dx1 then ðð dS1þdS*1 u � ndS ffi ∂u 1 ∂x1 dx1dx2dx3 ffi ∂u 1 ∂x1 dV and the same way for the components u2 and u3 in the other faces of the parallel- epiped the result is, respectively 4.3 Divergence 175 ðð dS2þdS*2 u � n dS ffi ∂u 2 ∂x2 dV ðð dS3þdS*3 u � ndS ffi ∂u 3 ∂x3 dV Adding these three expressions results for the six faces of the elementary parallelepiped ðð S u � n dS ¼ ððð V ∂u1 ∂x1 þ ∂u 2 ∂x2 þ ∂u 3 ∂x3 � � dV Putting divu ¼ ∂u 1 ∂x1 þ ∂u 2 ∂x2 þ ∂u 3 ∂x3 ð4:3:1Þ this analysis leads to the following definition for the divergence of the vectorial function u at point P(xi) divu ¼ lim V!0 1 V ðð S u � ndS ð4:3:2Þ that can be interpreted as the dot product between the nabla operator and the vectorial function u, thus divu ¼ ∇ � u ¼ ∂� � � ∂xj � � gj � uigi To demonstrate that expression (4.3.2) represents the divergence of the vectorial function u, consider the sphere of radius R > 0, of surface S(R) and volume V(R), centered at point P located in the vectorial space E3. For the vectorial field u acting in the space divu Pð Þ ¼ lim R!0 1 V Rð Þ ðð S Rð Þ u � ndS ð4:3:3Þ Let g Pð Þ ¼ divu, and admitting that g(xi) is a continuous function that can be written as g xi � � ¼ g Pð Þ þ h xi� � where h xi � � xi!P ¼ 0 176 4 Differential Operators Applying the divergence theorem to the vectorial field 1 V Rð Þ ðð S Rð Þ u:n dS ¼ 1 V Rð Þ ððð V Rð Þ h xi � � dV ¼ 1 V Rð Þ ððð V Rð Þ h Pð ÞdV þ 1 V Rð Þ ððð V Rð Þ h xi � � dV As g Pð Þ ¼ divu: 1 V Rð Þ ððð V Rð Þ g Pð ÞdV ¼ 1 V Rð Þ g Pð Þ ððð V Rð Þ dV ¼ 1 V Rð Þ g Pð ÞV Rð Þ ¼ g Pð Þ For the function h(xi) the result when R ! 0 is 1 V Rð Þ ððð V Rð Þ h xi � � dV ������� ������� ¼ Maxxi�Pk k R h xi � ��� �� 1 V Rð Þ ððð V Rð Þ dV Max xi�Pk k R h xi � ��� �� The maximum value of this function fulfills the condition h xið Þk k ! 0 when xi � Pk k ! 0, then the expression (4.3.2) represents the divergence of the vectorial function u at the point P. This expression is valid for any kind of coordinate system, which shows that the divergence is independent of the referential system. This analysis was formulated for a Cartesian coordinate system for a question of simplicity, being that for the case of curvilinear coordinate systems it was enough to adopt the local trihedron with unit vectors (g1; g2; g3), and one elementary parallelepiped of volume dV. The scalar field generated by the applying of the divergence to the vectorial field defined by the vectorial function u is called solenoidal or vorticular field, when divu ¼ 0, where u is a solenoidal vector, and this field is called field without source. 4.3.1 Divergence Theorem The divergence allows writing the Gauß-Ostrogradsky theorem asðð S u � ndS ¼ ððð V ∇ � udV ð4:3:4Þ which is called the divergence theorem. The symbology adopted in expression (4.3.4) does not change the characteristics and properties shown in item 3.4. Let a solenoidal field acting in a region R be located between the two closed surfaces S1 and S2 (Fig. 4.6), then 4.3 Divergence 177 ðð S1 u � ndS ¼ ðð S2 u � ndS To demonstrate this equality consider the closed surface S1 with upward normal unit vector n, with R to the left of the outline of this surface that involves the volume V1, and the closed surface S2 with unit downward unit normal vector n, involving the volume V2. Applying the divergence theorem divu ¼ ðð S1 u � ndS� ðð S2 u � ndS ¼ 0 ) ðð S1 u � n dS ¼ ðð S2 u � ndS then it is enough to calculate only the integral of a surface. For a field represented by the vectorial function u ¼ ϕ∇ψ , where ϕ and ψ are scalar functions ∇ � u ¼ ∇ ϕ∇ψð Þ ¼ ϕ∇ �∇ψ þ∇ϕ �∇ψ ¼ ϕ∇2ψ þ∇ϕ �∇ψ The component of u in the direction of the normal unit vector n is given by u � n ¼ ϕn �∇ψ ¼ ϕ∂ψ ∂n and applying the divergence theoremðð S u � ndS ¼ ððð V ∇ � udV 1S 2S 21 VVR −= 1V2V Fig. 4.6 Solenoidal field in a region R between two volumes 178 4 Differential Operators results in ðð S ϕ ∂ψ ∂n dS ¼ ððð V ϕ∇2ψ þ∇ϕ �∇ψ� �dV that is called Green’s first formula. If the vectorial function is given by u ¼ ϕ∇ψ þ ψ∇ϕ, in an analogous way ∇ � u ¼ ϕ∇2ψ þ ψ∇2ϕ u � n ¼ ϕ∂ψ ∂n � ψ ∂ϕ ∂n whereby ðð S ϕ ∂ψ ∂n � ψ ∂ϕ ∂n dS ¼ ððð V ϕ∇2ψ þ ψ∇2ϕ� �dV that is called Green’s second formula. 4.3.2 Contravariant and Covariant Components The vectorial function u can be expressed by means of their contravariant or covariant components, so it is necessary to calculate this function’s divergence for these components. For the vector’s contravariant coordinates divu ¼ ∇ � u ¼ g j � ∂u igi ∂xj � � ð4:3:5Þ and for its covariant coordinates divu ¼ ∇ � u ¼ gj � ∂uigi ∂xj � � ð4:3:6Þ The terms in parenthesis in theseexpressions indicate that this definition can be amplified considering the vector’s covariant derivatives, expressed in their contravariant and covariant coordinates. Let the covariant derivative of the contravariant vector ui be: ∂ku i ¼ ∂u i ∂xk þ u jΓ ijk 4.3 Divergence 179 that generates a tensor which contraction for i ¼ k provides ∂iu i ¼ ∂u i ∂xi þ ujΓ iji and rewriting the expression (2.4.23) Γ iji ¼ ∂ ‘n ffiffiffi g p� � ∂xj ¼ 1ffiffiffi g p ∂ ffiffiffi g p� � ∂xj The use of this expression is more adequate, for it abbreviates the calculation of the Christoffel symbol. Substituting this expression in the previous expression ∂iu i ¼ ∂u i ∂xi þ u j 1ffiffiffi g p ∂ ffiffiffi g p� � ∂xj ¼ 1ffiffiffi g p ffiffiffigp ∂ui ∂xi þ u jffiffiffi g p ∂ 1 2 ‘ng � � ∂xj � and replacing the indexes i ! j of the first term to the right ∂iu i ¼ 1ffiffiffi g p ffiffiffigp ∂u j ∂xj þ u j 2 ffiffiffi g p ∂ 1 2 ‘ng � � ∂xj � or in a compact form ∂iu i ¼ 1ffiffiffi g p ∂ ffiffiffi g p uj � � ∂xj ð4:3:7Þ It is verified that expression (4.3.7), deducted by means of the contravariant vector ui, represents a scalar, for it was obtained by means of contraction of the second-order tensor. The other way of formulating this analysis is by means of the covariant derivative of their covariant components. Let the covariant derivative of the covariant vector ui be: ∂iui ¼ ∂i gijuj � � that developed leads to the following expression ∂iui ¼ ∂i gij � � uj þ gij∂i uj � � Ricci’s lemma shows that ∂i gijð Þ ¼ 0 whereby ∂iui ¼ gij∂i uj � � ¼ ∂iu j 180 4 Differential Operators http://dx.doi.org/10.1007/978-3-319-31520-1_2 and the contraction of this tensor for i ¼ j provides ∂iui ¼ ∂iui ¼ divui ¼ 1ffiffiffi g p ∂ ffiffiffi g p ui � � ∂xi ð4:3:8Þ Expressions (4.3.7) and (4.3.8) provide the same result, i.e., ∂iui ¼ ∂iui. Then the covariant derivative of a vector is independent of the type of the component. The divergence defined by expressions (4.3.5) and (4.3.6) is the dot product of the nabla operator by the vector to which it is applied. The development of the derivatives indicated in these expressions leads to the same results of expressions (4.3.7) and (4.3.8), whereby these last expressions represent the divergence of a vectorial function. For the Cartesian coordinates ∇ � u ¼ ∂ui ∂xi ð4:3:9Þ 4.3.3 Orthogonal Coordinate Systems Consider the elementary parallelepiped with sides ds1, ds2, ds3, defined in the curvilinear orthogonal coordinates OXj, by means of which the flow of the field is represented by the vectorial function u (Fig. 4.7). The divergence of this field is given by ∇ � u ¼ lim V!0 1 V ðð S u � ndS Let dsi ¼ hidxi O 1X 2X 3X 1ds 2ds 3ds ( )321 ;; xxxu 2gn = ( )3221 ;; xxxx ∂+u2gn −= Fig. 4.7 Divergence of the vectorial function u in the curvilinear orthogonal coordinates 4.3 Divergence 181 dV ¼ ds1ds2ds3 ¼ h1h2h3dx1dx2dx3 there is, respectively, for the face with upward normal unit vector n ¼ �g2 and n ¼ g2 �u � g2ds1ds3 ¼ �u2h1h3dx1dx3 ¼ �u2 h1h3 þ ∂ u2h1h3ð Þ ∂x2 dx2 � dx1dx3 In an analogous way, for the other faces �u � g1ds2ds3 ¼ �u1h2h3dx2dx3 ¼ �u1 h2h3 þ ∂ u1h2h3ð Þ ∂x1 dx2 � dx2dx3 �u � g2ds1ds3 ¼ �u2h1h3dx1dx3 ¼ �u2 h1h3 þ ∂ u2h1h3ð Þ ∂x2 dx1 � dx1dx3 �u � g3ds1ds2 ¼ �u3h1h2dx1dx2 ¼ �u3 h1h2 þ ∂ u3h1h2ð Þ ∂x3 dx3 � dx1dx2 Adding the expressions relative to the six faces of the parallelepipedðð S u � ndS ¼ ∂ u 1h2h3ð Þ ∂x1 þ ∂ u 2h1h3ð Þ ∂x2 þ ∂ u 3h1h2ð Þ ∂x3 � dx1dx2dx3 but dx1dx2dx ¼ dV h1h2h3 then ∇ � u ¼ lim V!0 ðð S u � ndS ¼ 1 h1h2h3 ∂ u1h2h3ð Þ ∂x1 þ ∂ u 2h1h3ð Þ ∂x2 þ ∂ u 3h1h2ð Þ ∂x3 � The result for the orthogonal coordinate system is ∇ � u ¼ divu ¼ 1 h1h2h3 ∂ ∂xi h1h2h3u i hi � � ð4:3:10Þ where hi ¼ ffiffiffiffiffiffiffiffig iið Þp are the components of the metric tensor, and the indexes in parenthesis do not indicate summation. 182 4 Differential Operators Free ebooks ==> www.Ebook777.com 4.3.4 Physical Components With expression (4.3.6) the physical components of the divergence of a vector takes the form ∇ � u* ¼ divu* ¼ 1 h1h2h3 ∂ ∂xi h1h2h3u *i hi � � ð4:3:11Þ where u* i are the vector’s physical components. 4.3.5 Properties As the divergence is the dot product of the nabla operator for a vectorial function, the distributive property of the dot product is valid. For the sum of two vectorial functions u and v: div uþ vð Þ ¼ ∇ � uþ vð Þ ¼ ∇ � uþ∇ � v ð4:3:12Þ and in terms of the covariant derivative ∇ � uþ vð Þ ¼ ∂iui þ ∂ivi ð4:3:13Þ and for the Cartesian coordinates the result is ∇ � uþ vð Þ ¼ ∂u i ∂xi þ ∂v i ∂xi ð4:3:14Þ Considering the vectorial function mu, where m is a scalar, the result of the dot product of vectors is div muð Þ ¼ ∇ � muð Þ ¼ m∇ � u ð4:3:15Þ These two demonstrations prove that the divergence, for these cases, is a linear operator. In general, the divergence is not a linear operator, as it will be shown in Exercise 4.7. 4.3.6 Divergence of a Second-Order Tensor The generalization of the divergence theorem for tensorial fields is immediate. Consider, for example, the field represented by the tensorial function of the second 4.3 Divergence 183 www.Ebook777.com http://www.ebook777.com order T(r) in space E3, which components depend on the position vector, i.e., Tij ¼ Tij rð Þ. For the surface S smooth and continuous by parts in its two faces, with normal unit vector n(n1; n2; n3) varying on each point of the surface, the flow of this tensorial function through S is given by the vector v of components vi ¼ ðð S TijnjdS; i, j ¼ 1, 2, 3 or vi ¼ ðð S TjinjdS; i, j ¼ 1, 2, 3 In absolute notation for the flow v the result is v ¼ ðð S T� ndS ð4:3:16Þ The flow of the unit tensor δij through the closed surface S is given by the components of vector n: vi ¼ ðð S δijnjdS ¼ ðð S nidS or in absolute notation vi ¼ ðð S ndS The comparison with expression (4.2.3) ∇ϕ xi � � ¼ lim V!0 1 V ðð S ϕ xi � � ndS shows that ϕ xið Þ ¼ 1, i.e., ϕ xið Þ ¼ constant, so ∇ϕ xið Þ ¼ 0, which indicates that v ¼ 0. Concluding that for a unitary tensorial field the flow through the closed surface S is null. The concept of a field’s divergence can be extended to the tensorial fields, for it is enough that the tensors be contravariant. In the case of covariant tensors their indexes must be raised by means of the metric tensor, next they must be derived and contracted. 184 4 Differential Operators There are distinct divergences, depending on the index to be contracted. For Tij there are two divergences:∂iTij and∂jTij. If the tensor is symmetrical Tij ¼ Tji then ∂iTij ¼ ∂jTij, i.e., the divergence is unique. The divergence components of a contravariant second-order tensor are given by divTij ¼ ∂jTij ð4:3:17Þ and the covariant derivative of the components for this tensor is ∂kT ij ¼ ∂T ij ∂xk þ TmkΓ imk þ TimΓ jmk � � With k ¼ j: ∂jT ij ¼ ∂T ij ∂xj þ TmjΓ imj þ TimΓ jmj � � being Γ jmj ¼ ∂ ‘n ffiffiffi g p� � ∂xm it follows that ∂jT ij ¼ ∂T ij ∂xj þ TmjΓ imj þ Tim ∂ ‘n ffiffiffi g p� � ∂xm � The change of the indexes m ! j in the last term in brackets provides ∂kT ij ¼ ∂T ij ∂xj þ TmjΓ imj þ Tij ∂ ‘n ffiffiffi g p� � ∂xj � or ∂kT ij ¼ TmjΓ imj þ 1ffiffiffi g p ffiffiffigp ∂Tij ∂xj þ Tij ∂ ffiffiffi g p� � ∂xj � then divTij ¼ TmjΓ imj þ 1ffiffiffi g p ∂ ffiffiffi g p Tij � � ∂xk ð4:3:18Þ that shows that the divergence of a second-order tensor is a vector. 4.3 Divergence 185 For a mixed second-order tensor the result is divT ij ¼ ∂iT ij and rewriting expression (2.5.21) ∂kT i j ¼ ∂T ij ∂xk þ Tmj Γ imk � T imΓmjk Assuming i ¼ k: ∂iT i j ¼ ∂T ij ∂xk þ Tmj Γ imi � T imΓmji and with Γ imi ¼ ∂ ‘n ffiffiffi g p� � ∂xm then ∂iT i j ¼ ∂T ij ∂xk þ Tmj ∂ ‘n ffiffiffi g p� � ∂xm � T imΓmji " # The change of indexes m ! i in the second term in brackets provides ∂iT i j ¼ ∂T ij ∂xk þ T ij ∂ ‘n ffiffiffi g p� � ∂xi � T imΓmji " # or∂iT i j ¼ ffiffiffi g p ∂T ij ∂xk þ 1ffiffiffi g p T ij ∂ ffiffiffi g p� � ∂xi � T imΓmji then ∂iT i j ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � T imΓmji ð4:3:19Þ The generalization of the definition of the divergence for a third-order tensor is immediate 186 4 Differential Operators http://dx.doi.org/10.1007/978-3-319-31520-1_2 ∇ � T ¼ g j � ∂jT ¼ g j � ∂jTkimgk � gi � gm ¼ ∂jTkim � � g j � gk � gi � gm ¼ ∂jTkim � � δ jkgi � gm thus ∇ � T ¼ ∂jTjimgi � gm ð4:3:20Þ This expression shows that ∇ � T is a second-order tensor. For a tensor T of order p then ∇ � T is a tensor of order p� 1ð Þ. In absolute or invariant notation the result for the divergence of tensor T is ∇ � T ¼ ∇� Tð Þ�G ð4:3:21Þ where G is the metric tensor. In the particular case in which divT ¼ 0 the tensor T defines a tensorial solenoidal field. The divergence theorem also applies to a tensorial field. Let the field be defined by u ¼ Tv, which in terms of the components of vectors and tensor is given by ui ¼ Tikvk, being v an arbitrary and constant vector. Applying the divergence theorem to this field ðð S u � ndS ¼ ððð V ∇ � udV where ∇ � u ¼ ∇ Tvð Þ ¼ ∇T � v This vector has components ∂Tik ∂xk vi, and the component of vector u in the direction of the normal unit vector n is given by the dot product u � n ¼ Tikvið Þnk then ðð S Tikvið Þnk dS ¼ ððð V ∂Tik ∂xk vidV whereby in terms of the tensor components the result isðð S Tik nk dS ¼ ððð V ∂Tik ∂xk dV ð4:3:22Þ 4.3 Divergence 187 and in absolute notation this expression becomesðð S T� ndS ¼ ððð V ∇ � TdV ð4:3:23Þ Exercise 4.7 Calculate: (a) ∇ � ϕuð Þ; (b) ∇ � u� vð Þ. (a) The field divergence defined by the product of a scalar function ϕ(xi) by a vector u is given by div ϕuð Þ ¼ ∇ � ϕuð Þ ¼ gm ∂ ϕu kgk � � ∂xm ¼ gm ∂ϕ ∂xm ukgk þ ϕ ∂uk ∂xm gk þ ϕuk ∂gk ∂xm � � and substituting (2.3.10) ∂gk ∂xm ¼ Γ pkmgp in the previous expression div ϕuð Þ ¼ gm ∂ϕ ∂xm ukgk þ ϕ ∂uk ∂xm gk þ ϕukΓ pkmgp � � The permutation of the indexes p $ k in the third member in parenthesis provides div ϕuð Þ ¼ ∂ϕ ∂xm uk þ ϕ ∂u k ∂xm þ ϕupΓ kpm � � gm � gk ¼ ∂ϕ ∂xm uk þ ϕ ∂u k ∂xm þ ϕupΓ kpm and with ∂mu k ¼ ∂ϕ ∂xm uk þ ϕ ∂u k ∂xm ) div ϕuð Þ ¼ ∂ϕ ∂xm uk þ ϕ∂muk Putting ∂mu k ¼ ∇uk ∂ϕ ∂xm ¼ ∇ϕ thus ∇ � ϕuð Þ ¼ ∇ϕð Þ � uþ ϕ ∇uð Þ 188 4 Differential Operators http://dx.doi.org/10.1007/978-3-319-31520-1_2 or div ϕuð Þ ¼ gradϕ � uþ ϕgradu In this case the divergence is not a linear operator, but for ϕ ¼ m, where m is a constant, the result is expression (4.3.15), verifying the linearity of this operator. (b) The field represented by the vectorial function generated by the cross product w ¼ u� v is given by w ¼ wpgp ¼ εpqruqvrgp The divergence of this function is given by ∇ � u� vð Þ ¼ gi � ∂ w pgp � � ∂xi ¼ ∂w p ∂xi gp þ wp ∂gp ∂xi � � � gi and the expression ∂gp ∂xi ¼ Γ jpigj substituted in the previous expression provides ∇ � u� vð Þ ¼ ∂w p ∂xi gp þ wpΓ jpigj � � � gi Interchanging indexes p $ j in the second term in parenthesis it follows that ∇ � u� vð Þ ¼ ∂w p ∂xi þ wjΓ pji � � gi � gp ¼ δ ip∂iwp or ∂pw p ¼ ∂p εpqruqvr � � ¼ ∂pεpqr� �uqvr þ εpqr∂p uqvr� � and with ∂pw p ¼ εpqr∂p uqvr � � thus ∇ � u� vð Þ ¼ εpqr∂p uqvr � � 4.3 Divergence 189 whereby ∇ � u� vð Þ ¼ εpqr ∂puq � � vr þ uq∂pvr � � With the εpqr ¼ εqpr and εpqr ¼ �εrpq the results for the terms to the right are εpqr ∂puq � � vr ¼ εrpq ∂puq � � vr ¼ v �∇u εpqruq ∂pvr � � ¼ �εqpruq ∂pvr� � ¼ �u �∇v whereby ∇ � u� vð Þ ¼ v �∇� u� u �∇� v ) div u� vð Þ ¼ v � rotu� u � rotv For the Cartesian coordinates ∇ � u� vð Þ ¼ εijk ∂ ujvk � � ∂xi Exercise 4.8 Let T ij and Tij be associated tensors, write div T i j in terms of the symmetrical tensor Tij. The divergence of a second-order tensor is given by divT ij ¼ ∂iT ij ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � T imΓmji and with Γmij ¼ gmkΓij,k thus divT ij ¼ ∂iT ij ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � T imgmkΓij,k Let Γij,k ¼ 1 2 ∂gjk ∂xi þ ∂gik ∂xj þ ∂gij ∂xk � � T img mk ¼ Tik then ∂iT i j ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � Tik1 2 ∂gjk ∂xi þ ∂gik ∂xj þ ∂gij ∂xk � � 190 4 Differential Operators or ∂iT i j ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � 1 2 Tik ∂gjk ∂xi � 1 2 Tik ∂gik ∂xj � 1 2 Tik ∂gij ∂xk Interchanging the indexes i $ j in the last term to the right ∂iT i j ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � 1 2 Tik ∂gjk ∂xi � 1 2 Tik ∂gik ∂xj � 1 2 Tki ∂gkj ∂xk As gjk ¼ gkj Tik ¼ Tki thus ∂iT i j ¼ 1ffiffiffi g p ∂ T ij ffiffiffi g p� � ∂xi � 1 2 Tik ∂gik ∂xj Exercise 4.9 Calculate the divergence of vector ui expressed in cylindrical coordinates. For the cylindrical coordinates ffiffiffi g p ¼ r, and with the contravariant components of vector (ur, uθ, uz) it follows that divui ¼ 1ffiffiffi g p ∂ ffiffiffi g p ui � � ∂xi ¼ 1 r ∂ ruið Þ ∂xi � ¼ 1 r ∂ rurð Þ ∂r þ ∂ ru θ � � ∂θ þ ∂ ru zð Þ ∂z � ¼ ∂u r ∂r þ ∂u θ ∂θ þ ∂u z ∂z þ ur In an analogous way in terms of the vector’s covariant components divuj ¼ ∂ur∂r þ 1 r2 ∂uθ ∂θ þ ∂uz ∂z þ ur Exercise 4.10 Calculate the divergence of vector ui expressed in spherical coordinates. For the spherical coordinates ffiffiffi g p ¼ r2 sinϕ, and with the contravariant compo- nents of vector (ur, uϕ, uθ) it follows that 4.3 Divergence 191 divui ¼ 1ffiffiffi g p ∂ ffiffiffi g p ui � � ∂xi ¼ 1 r2 sinϕ ∂ r2 sin ϕuið Þ ∂xi � ¼ 1 r2 sinϕ ∂ r2 sinϕurð Þ ∂r þ ∂ r 2 sinϕuϕ � � ∂ϕ þ ∂ r 2 sin ϕuθ � � ∂θ � ¼ ∂u r ∂r þ ∂u ϕ ∂ϕ þ ∂u θ ∂θ þ 2u r r þ cotϕð Þuϕ For the vector’s covariant components the result is divuj ¼ ∂ur∂r þ 1 r2 ∂uϕ ∂ϕ þ 1 r2 sin 2ϕ ∂uθ ∂θ þ 2ur r þ cotϕð Þ r2 uϕ Exercise 4.11 Let r be the position vector of the points in the space E3, show that: (a) div r ¼ 3; (b) div rn rð Þ ¼ nþ 3ð Þ rn; (c) div rr3 � � ¼ 0; (d) div rr� � ¼ 2r. (a) With the definition of divergence ∇ � r ¼ gi ∂ ∂xi � � � r ¼ gi � ∂r ∂xi but ∂r ∂xi ¼ gi then ∇ � r ¼ gi � gi For i ¼ 1, 2, 3 the result is div r ¼ 3 Q:E:D: (b) Let div ϕuð Þ ¼ ϕdivuþ ugradϕ and putting u ¼ r ϕ ¼ rn thus div rnrð Þ ¼ rn div rþ rgrad rn ¼ 3rn þ r � nrn�11 r r � � ¼ 3rn þ nrn�2r2� � 192 4 Differential Operators Free ebooks ==> www.Ebook777.com then div rn rð Þ ¼ nþ 3ð Þ rn Q:E:D: (c) Putting div r r3 � � ¼ div r�3r� � it follows that div r�3rð Þ ¼ r�3div rþ r � grad r�3 ¼ 3r�3 þ r � �3r�4grad rð Þ ¼ 3r�3 þ r � �3r�4r r � � ¼ 3r�3 þ r2 �3r�41 r � � whereby div r�3r � � ¼ 0 Q:E:D: This conclusion shows that r�3r is a solenoidal vectorial function. (d) Putting div r r � � ¼ div 1 r r � � r ¼ xiþ yjþ zk it follows that div 1 r r � � ¼ div x r iþ y r jþ z r k � � ¼ ∂ ∂x x r � � þ ∂ ∂y y r � � þ ∂ ∂z z r � � ¼ 1 r � x r2 ∂r ∂x � � þ 1 r � y r2 ∂r ∂y � � þ 1 r � z r2 ∂r ∂z � � and with r2 ¼ x2 þ y2 þ z2 ∂r ∂x ¼ x r ∂r ∂y ¼ y r ∂r ∂z ¼ z r thus div 1 r r � � ¼ 3 r � x r2 x r þ y r2 y r þ z r2 z r � � 4.3 Divergence 193 www.Ebook777.com http://www.ebook777.com whereby div r r � � ¼ 2 r Q:E:D: 4.4 Curl The vector product of the nabla operator by a vector generates a differential operator linked to the direction of rotation of the coordinate system defining the curl, also called rotation or whirl. In absolute notation this operator is written as ∇� u ¼ rot u ¼ v ð4:4:1Þ In English literature the notation curl u is used to designate rotational of vector u, which was adopted firstly by Maxwell. The term curl literally means ring, and it designates the pseudovector ∇� u. With v ¼ gi � ∂ ujg j � � ∂xi ¼ gi � ∂uj ∂xi g j þ gi � uj ∂g j ∂xi and rewriting expression (2.4.4) ∂g j ∂xi ¼ �Γ jkigm it follows for the second term of the member to the right of expression (4.4.1) ujg i � ∂g m ∂xi ¼ �ujΓ jkigi � gm The cross product of these vectors is given by gi � gm ¼ e imkffiffiffi g p gk ¼ þ1 for an even number of permutationsof the indexes �1 for an odd number of permutations of the indexes 0 when there are repeated indexes 8>>>>>><>>>>>>: and substituting results in ujg i � ∂g m ∂xi ¼ � ujffiffiffi g p Γ jki � Γ jik � � gk ¼ 0 194 4 Differential Operators http://dx.doi.org/10.1007/978-3-319-31520-1_2 whereby ∇� u ¼ v ¼ ∂uj ∂xi gi � gm As gi � g j ¼ e ijkffiffiffi g p gk in tensorial terms ∇� u ¼ e ijkffiffiffi g p ∂uj ∂xi gk ð4:4:2Þ As a function of Ricci’s pseudotensor ∇� u ¼ εijk ∂uj ∂xi gk ð4:4:3Þ and with εijk ¼ e ijkffiffiffi g p the expression (4.4.3) after a cyclic permutation of the indexes i, j, k ¼ 1, 2, 3 takes the form ∇� u ¼ 1ffiffiffi g p ∂uj ∂xi � ∂ui ∂xj � � gk ð4:4:4Þ In a space provided with metric, the curl of a vectorial function can also be defined by means of its contravariant components, for these relate with its covariant components by means of the metric tensor. In an analogous way, the results for the contravariant coordinates are ∇� u ¼ g‘ � ∂ u kgk � � ∂x‘ ¼ g‘ � ∂u k ∂x‘ gk þ uk ∂gk ∂x‘ � � ∂gk ∂x‘ ¼ Γmk‘gm ∇� u ¼ g‘ � ∂u k ∂x‘ gk þ ukΓmk‘gm � � ¼ ∂u k ∂x‘ þ umΓ km‘ � � g‘ � gk 4.4 Curl 195 ∇� u ¼ ∂‘uk g‘ � gk g‘ ¼ g‘jgj ∇� u ¼ ∂‘uk g‘jgj � gk gj � gk ¼ εijkgi ∇� u ¼ εijk ∂‘uk � � g‘jgi ð4:4:5Þ A vectorial field is called an irrotational field when ∇� u ¼ 0, then ∂uj ∂xi ¼ ∂ui ∂xj In space E3 the curl∇� u is an axial vector (vectorial density), so it is associated to an antisymmetric second-order tensor, which components are Aij ¼ 0 ∂u2 ∂x1 � ∂u 1 ∂x2 ∂u3 ∂x1 � ∂u 1 ∂x3 ∂u1 ∂x2 � ∂u 2 ∂x1 0 ∂u3 ∂x2 � ∂u 2 ∂x3 ∂u1 ∂x3 � ∂u 3 ∂x1 ∂u2 ∂x3 � ∂u 3 ∂x2 0 26666664 37777775 ð4:4:6Þ For the space EN the curl ∇� u has 12N N � 1ð Þ independent components. In space E2 the curl is a pseudoscalar. For the Cartesian coordinates ∇� u ¼ eijk ∂uj∂xi gk ð4:4:7Þ or in a determinant form ∇� u ¼ i j k ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 u1 u2 u3 ð4:4:8Þ 4.4.1 Stokes Theorem In expression (3.3.7) with F ¼ u, and with the expression (4.4.8) Stokes theorem in vectorial notation is given byðð S n �∇� u dS ¼ þ C u � dr ð4:4:9Þ 196 4 Differential Operators http://dx.doi.org/10.1007/978-3-319-31520-1_3 A more consistent definition of the curl can be formulated analyzing the circu- lation of the vectorial field u in a closed surface S with upward unit normal vector n (Fig. 4.8a). Consider the elementary rectangle dS determined in the orthogonal curvilinear coordinate system X j , with sides h1dx 1 and h2dx 2 located in the plane OX 2 X 3 , with the point P x1; x2; x3 � � located in its center. Locally, the coordinate system X j is considered as a Cartesian orthogonal system (Fig. 4.8b), with the scale factors hi, i ¼ 1, 2, 3. The line integral þ C u � dr along the perimeter of this rectangle is carried out dividing this perimeter into segments C1,C2,C3,C4. The center of segment C1 of perimeter of the rectangle is given by the coordinates x1; x2; x3 � dx3 2 � � then u � dr ¼ u2dx2. As this length is elementary its contribution to the line integral is given by þ C1 u � dr ffi u2 x1; x2; x3 � dx 3 2 � � h2dx 2 For segment C3 with center x 1; x2; x3 þ dx3 2 � � : þ C3 u � dr ffi �u2 x1; x2; x3 þ dx 3 2 � � h2dx 2 where the negative sign indicates that the direction of the path is contrary to the coordinate axis. n O P 3C 4C iS P i C S 3X 1X 2X 1g 2g 3g 3 32 xdhC = 2 21 xdhC = a b Fig. 4.8 Concept of curl: (a) circulation in a closed surface and (b) elementary rectangle 4.4 Curl 197 Adding the contributions of these two segmentsþ C1þC3 u � dr ffi u2 x1; x2; x3 � dx 3 2 � � � u2 x1; x2; x3 þ dx 3 2 � � � h2dx 2 The component u2 varies according to the rate du2 ¼ �∂u 2 ∂x3 dx3 where the negative sign indicates that this variation decreases in the positive direction of axis OX 1 , it follows thatþ C1þC3 u � dr ffi �∂u 2 ∂x3 dx3h2dx 2 and dividing by dS ¼ h2h3dx2dx3 1 dS þ C1þC3 u � dr ffi � 1 h2h3 ∂ u2h2ð Þ ∂x3 Adopting analogous formulations for the other segmentsþ C2þC4 u � dr ffi u3 x1; x2; x3 þ dx 3 2 � � � u3 x1; x2; x3 � dx 3 2 � � � h3dx 3 ffi ∂u 3 ∂x2 dx2h3dx 3 1 dS þ C2þC4 u � dr ffi 1 h2h3 ∂ u3h3ð Þ ∂x2 Adding these contributions the result when dS ! 0 is ∂ h3u3ð Þ ∂x2 � ∂ h2u 2ð Þ ∂x3 � ¼ lim dS!0 1 dS þ C u � dr or e1 �∇� u ¼ ∂ h3u 3ð Þ ∂x2 � ∂ h2u 2ð Þ ∂x3 198 4 Differential Operators For the components u1 and u2 the result is, respectively, e2 �∇� u ¼ ∂ h1u 1ð Þ ∂x3 � ∂ h3u 3ð Þ ∂x1 e3 �∇� u ¼ ∂ h2u 2ð Þ ∂x1 � ∂ h1u 1ð Þ ∂x2 Concluding that the curl components of the vectorial field u in the direction of the upward unit normal vector to the closed surface S are given by n �∇� u ¼ lim S!0 1 S þ C u � dr ð4:4:10Þ This expression is valid for any type of referential system, which shows that the curl is independent of the coordinate system. For demonstrating that expression (4.4.9) represents the Stokes theorem, let the surface S which outline is curve C, and the field represented by the vectorial function u(r), continuous and with continuous partial derivatives in S [ C. Dividing S in N cells Si, i ¼ 1, 2, . . .N, which components of the upward normal unit vectors are ni, with closed outline curves Ci (Fig. 4.8a), and with expression (4.4.10) the result for each cell of S is ni �∇� u ¼ lim Si!0 1 Si þ Ci u � dr Applying this expression to point P contained in cell Si with boundary Ci, the result when the area of this cell is reduced approaching the outline P is ni �∇� uð ÞSi ¼ þ Ci u � drþ h xi� �Si where h xið Þk k > 0 is a function with very small value, which decreases with the reduction of size of Si. With the division of the surface S into N parts the result is thatN hi x ið Þ½ � > 0, then Max 1 i N hi x ið Þ < h xið Þ. For N ! 1 the result is h xið Þ ! 0, so ni �∇� uð ÞSi � XN i¼1 þ Ci u � dr ������� ������� < h xi � �XN i¼1 Si ¼ h xi � � S 4.4 Curl 199 and with XN i¼1 þ Ci u � dr ¼ þ C u � dr for the outlines of the cells Si are calculated twice, but in opposite directions, whereby these parcels cancel each other, leaving only the parcel of boundary C of S. Then ni �∇� uð ÞSi � þ C u � dr ������ ������ < h xi� �S As N ! 1 the result is lim N!1 ni �∇� uð ÞSi ¼ þ C u � dr whereby the result of the expression of Stokes theorem isðð S n �∇� u dS ¼ þ C u � dr This theorem is a particular case of the divergence theorem. To demonstrate this assertion let the vectorial function u ¼ v� w, and an arbitrary and constant vector, then ∇ � v� wð Þ ¼ w �∇� v and n � v� wð Þ ¼ w � n� v Applying the divergence theorem to the function u it is written asðð S u:ndS ¼ ððð V ∇ � udV The substitution of the previous expressions in this expression shows thatððð V ∇ � udV ¼ ðð S v� wð Þ�n dS ¼ ðð S w � n� vð ÞdS 200 4 Differential Operators whereby w � ððð V ∇�v ¼ w � ðð S n� vdS As w is arbitrary it results inððð V ∇�v ¼ ðð S n� v dS The concept of curl of a vector u can be generalized for a space EN, in which the vector is associated to an antisymmetric tensor A, and its order depends on the dimension of the space. This tensor is generated by means of the dot product between the Ricci pseudotensor and the vector’s covariant derivative Ai1 i2���ip�2 ¼ εi1 i2���ip�2 j k∂juk ð4:4:11Þ 4.4.2 Orthogonal Curvilinear Coordinate Systems With the expressions used in the previous item to demonstrate expression (4.4.10), there is in index notation for the curl coordinates of vector u in a curvilinear orthogonal coordinate system ∇� u ¼ hk h1h2h3 ∂hjuj ∂xi � ∂hiu i ∂xj � � gk ð4:4:12Þ where hi ¼ ffiffiffiffiffiffiffiffig iið Þp , hj ¼ ffiffiffiffiffiffiffiffig jjð Þp , hk ¼ ffiffiffiffiffiffiffiffiffig kkð Þp are the components of the metric tensor, and the indexes in parenthesis indicate no summation. In a determinant form the result is ∇� u ¼ 1 h1h2h3 h1g1 h2g2 h3g3 ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 h1u 1 h2u 2 h3u 3 ð4:4:13Þ and with thephysical components of vector u*i it follows that ∇� u* ¼ hk h1h2h3 ∂hju*j ∂~xi � ∂hiu *i ∂xj � � gk ð4:4:14Þ 4.4 Curl 201 4.4.3 Properties As the curl is the cross product of the nabla operator by a vectorial function, the properties of this vector product are valid. For the sum of two vectorial functions u and v: ∇� uþ uð Þ ¼ ∇� uþ∇� v ð4:4:15Þ and the successive applying of the curl to this sum provides ∇�∇� uþ uð Þ ¼ ∇�∇� uþ∇�∇� v ð4:4:16Þ Considering the vectorial function mu, where m is a scalar, the result of the cross product ∇� muð Þ ¼ m∇� u ð4:4:17Þ Expressions (4.4.16) and (4.4.17) show that the curl, for these cases, is a linear operator, which is valid for the general case as will be shown in item 4.5. 4.4.4 Curl of a Tensor The concept of curl of a vector in space EN is developed in an analogous way. For instance, for the second-order tensor Tk1k2 in space EN exists the curl of order p� 3ð Þ, given by the cross product between the Ricci pseudotensor of order p and the tensor’s covariant derivative, then Ai1 i2���ip�3 ¼ εi1 i2���ip�3 j k1k2∂jTk1k2 ð4:4:18Þ Expression (4.4.18) shows that the Ricci pseudotensor is the generator of the antisymmetric tensor that represents the rotational of the tensor. In absolute notation the result is ∇� T ¼ ∇� Tð Þ�E ð4:4:19Þ where E is the Ricci pseudotensor. In the particular case of the space E4 the curl for the second-order tensor Tk‘ is given by the components of vector Ai: Ai ¼ εijk‘∂jTk‘ 202 4 Differential Operators Free ebooks ==> www.Ebook777.com Assuming that the second-order tensor is decomposed into two tensors, one symmetric and the other antisymmetric T ¼ Sþ A then rotT ¼ rotSþ rotA The components of the curl of the symmetric tensor are given by εi1 i2���ip�3 j k1k2∂jSk1k2 , i.e., are obtained by means of the dot product of the Ricci pseudotensor (antisymmetric) by the symmetric tensor which is null, whereby rotS ¼ 0. Concluding that the curl of a symmetric tensor is null, and that only the antisymmetric tensor A generates the rotational of tensor T. In the particular case in which rotT ¼ 0 the tensor T defines an irrotational tensorial field. The definition of curl of a second-order tensor can be applied to a tensor of order p > 2, whereby Ai1 i2���iq�p�1 ¼ εi1 i2���iq�p�1 j k1���kp∂jTk1���kp ð4:4:20Þ being q� 1ð Þ the order of the Ricci pseudotensor, and q� p� 1ð Þ the order of the antisymmetric tensor that represents the curl of the tensor. Exercise 4.12 Calculate: (a) ∇� ϕu; (b) u� ∇� vð Þ; (c) ∇� u� vð Þ. (a) The curl of the field defined by the product of a scalar function ϕ(xi) for a vectorial function u is given by ∇� ϕu ¼ g j � ∂ ϕukg k � � ∂xj ¼ g j � ∂ϕ ∂xj ukg k þ ϕ∂uk ∂xj gk þ ϕuk ∂g k ∂xj � � Substituting expression (2.4.4) ∂gk ∂xj ¼ �Γ kmjgm in this expression ∇� ϕu ¼ g j � ∂ϕ ∂xj ukg k þ ϕ∂uk ∂xj gk � ϕukΓ kmjgm � � The permutation of the indexes k $ m in the last term provides ∇� ϕu ¼ ∂ϕ ∂xj uk þ ϕ∂uk∂xj � ϕumΓ m kj � � g j � gk 4.4 Curl 203 www.Ebook777.com http://dx.doi.org/10.1007/978-3-319-31520-1_2 http://www.ebook777.com and with expressions g j � gk ¼ εijkgi ∂juk ¼ ∂uk ∂xj � umΓmkj it follows that ∇� ϕu ¼ ∂ϕ ∂xj uk þ ϕ∂juk � � εijkgi ¼ ∂ϕ ∂xj ukε ijkgi þ ϕ ∂juk � � εijkgi Putting ∂juk ¼ ∇uk ∇� ϕu ¼ ∇ϕ� uþ ϕ∇� u ) rotϕu ¼ gradϕ� uþ ϕ rotu For the Cartesian coordinates ∇� ϕu ¼ εijk uk ∂ϕ∂xj þ εijkϕ ∂uk ∂xj � � gi It is verified for this case that the curl is not a linear operator. Forϕ ¼ m, where m is a constant, the result with expression (4.4.17) is that this operator’s linearity is valid by this particular case. (b) The curl ∇� v is given by ∇� v ¼ w ¼ εkmn∂mvngk then u� w ¼ εijku jwkgi whereby substituting u� ∇� vð Þ ¼ εijku jεkmn∂mvngi ¼ εijkεkmnu j∂mvngi and with εijkε kmn ¼ δmnij δmnij ¼ δmi δ nj � δmj δni is it follows that u� ∇� vð Þ ¼ δmi δnj u j∂mvn � δmj δ ni u j∂mvn h i gi ¼ u j∂ivj � u j∂jvi � � gi 204 4 Differential Operators For the Cartesian coordinates the result is u� ∇� vð Þ ¼ uj ∂vj∂xi � uj ∂vi ∂xj � � gi (c) The cross product u� v ¼ w is given by u� v ¼ w ¼ w‘g‘ ¼ ε‘mnumvng‘ thereby ∇� u� vð Þ ¼ εij‘∂jw‘gi ¼ εij‘∂j ε‘mnumvnð Þgi ¼ δijmn∂j umvnð Þgi ¼ δ imδ jn � δ inδ jm � � ∂j umvnð Þgi ¼ ∂j uiv jð Þ � ∂j u jvið Þ � � gi For the Cartesian coordinates ∇� u� vð Þ ¼ vj ∂u i ∂xj þ ui ∂v j ∂xj � vi ∂u j ∂xj � uj ∂v i ∂xj � � gi ¼ v �∇� u� u �∇� v Exercise 4.13 Calculate ∇� u for the vector u expressed in cylindrical coordinates. For the cylindrical coordinates the result is h1 ¼ 1, h2 ¼ r, h3 ¼ 1 and (ur, uθ, uz). The determinant is given by the expression (4.4.13) ∇� u ¼ 1 r gr rgθ gz ∂ ∂xr ∂ ∂xθ ∂ ∂xz ur ruθ uz which development provides ∇� u ¼ 1 r ∂uz ∂θ � ∂uθ ∂z � � gr þ ∂ur ∂z � ∂uz ∂r � � gθ þ 1 r ∂ruθ ∂r � 1 r ∂ur ∂θ � � gz Exercise 4.14 Calculate∇� u for the vector u expressed in spherical coordinates. For the spherical coordinates the result is h1 ¼ 1, h2 ¼ r, h3 ¼ r sin ϕ and (ur, uϕ, uθ). The determinant given by expression (4.4.13) 4.4 Curl 205 ∇� u ¼ 1 r2 sin ϕ gr rgκ r sin ϕgθ ∂ ∂xr ∂ ∂xϕ ∂ ∂xθ ur ruϕ r sin ϕuθ which development provides ∇� u ¼ 1 r sin ϕ ∂uθ sin ϕ ∂ϕ � ∂uϕ ∂θ � � gr þ 1 r sin ϕ ∂uθr sin ϕ ∂r � 1 r ∂ur ∂θ � � gϕ þ 1 r ∂ruϕ ∂r � ∂ur ∂ϕ � � gθ Exercise 4.15 Let r be the position vector of the point in space E3, show that: (a) ∇� r ¼ 0; (b) ϕ(r)r is irrotational. (a) With the definition of curl ∇� r ¼ gi ∂ ∂xi � � � r ¼ gi � ∂r ∂xi but ∂r ∂xi ¼ gi then ∇ � r ¼ gi � gi ¼ 0 Q:E:D: (b) A condition that a vectorial function must fulfill so that the field that it represents is irrotational is ∇� ϕ rð Þr½ � ¼ 0 and putting ϕ rð Þ ¼ ψ it follows that ∇� ϕ rð Þr½ � ¼ gradϕ� rþ ϕ∇� r ¼ ϕ0 rð Þgrad r h i � rþ ϕ rð Þ �∇� r 206 4 Differential Operators but ∇� r ¼ 0 then ∇� ϕ rð Þ r½ � ¼ ϕ0 rð Þ1 r r � � r as r� r ¼ 0 thus ∇� ϕ rð Þ r½ � ¼ 0 Q:E:D: 4.5 Successive Applications of the Nabla Operator The operator∇ can be applied successively to a field. The number of combinations of two out of the three differential operators, the gradient, the divergence, and the curl, are 32 ¼ 9 types of double operators. The combinations ∇ � ∇ � uð Þ and ∇� ∇ � uð Þ have no mathematical meaning. 4.5.1 Basic Relations (1) ∇ � ∇� uð Þ The curl of a vectorial function is given by ∇� u ¼ εk‘m∂‘umgk ¼ w ¼ wkgk ð4:5:1Þ wk ¼ εk‘m∂‘um ð4:5:2Þ 4.5 Successive Applications of the Nabla Operator 207 then ∇ � ∇� uð Þ ¼ gi � ∂ w kgk � � ∂xi ¼ gi � ∂w k ∂xi gk þ wk ∂gk ∂xi � � and with expression ∂gk ∂xi ¼ Γ nkign it follows that ∇ � ∇� uð Þ ¼ gi � ∂w k ∂xi gk þ wkΓ nkign � � The permutation of indexes n $ k in the second member in parenthesis provides ∇ � ∇� uð Þ ¼ ∂w k ∂xi þ wnΓ kni � � gi � gk and with gi � gk ¼ δ ik the result is ∇ � ∇� uð Þ ¼ ∂w k ∂xk þ wnΓ knk � � ¼ ∂kwk Substituting expression (4.5.2) ∇ � ∇� uð Þ ¼ ∂k εk‘m∂‘um � � ¼ εk‘m∂k ∂‘umð Þ ¼ ek‘mffiffiffi g p ∂k ∂‘umð Þ and interchanging the indexes i, j, k ¼ 1, 2, 3 cyclically ∇ � ∇� uð Þ ¼ 1ffiffiffi g p ∂k∂‘um � ∂‘∂kumð Þ ð4:5:3Þ whereby ∇ � ∇� uð Þ ¼ 0 ð4:5:4Þ 208 4 Differential Operators Vector∇� u represents a vectorial field associated to the vectorial function u. Expression (4.5.4) defines the condition of existence for this function. The property of the field defined by the curl of the vectorial function u shows that the divergence of this field is null, i.e., the field is solenoidal. In a reciprocal way for a solenoidal field ∇ � ∇� uð Þ ¼ 0 a solenoidal vector v can be determined, such that v ¼ ∇� uð Þ. In this case the vector v derives from the potential function u, being linked to this function. (2) ∇� ∇ϕð Þ The gradient of a scalar function ϕ(xi) is given by ∇ϕð Þ ¼ u ¼ ∂ϕ ∂xk gk ¼ ukgk ð4:5:5Þ then ∇� ∇ϕð Þ ¼ ∇u ¼ εijk∂jukgi it follows that ∇� ∇ϕð Þ ¼ εijk∂j ∂ϕ∂xk � � gi ¼ εijk ∂2ϕ ∂xj∂xk ! gi ¼ eijkffiffiffi g p ∂ 2ϕ ∂xj∂xk ! gi Interchangingthe indexes i, j, k ¼ 1, 2, 3 cyclically ∇� ∇ϕð Þ ¼ 1ffiffiffi g p ∂ 2ϕ ∂xj∂xk � ∂ 2ϕ ∂xk∂xj ! gi As ∂2ϕ ∂xj∂xk ¼ ∂ 2ϕ ∂xk∂xj it results in ∇� ∇ϕð Þ ¼ 0 ð4:5:6Þ The field that fulfills the condition given by expression (4.5.6) is called a conservative field, i.e., every vectorial field with potential is an irrotational field. Let ∇ϕ ¼ u the result is ∇� ∇ϕð Þ ¼ ∇� u ¼ 0, then 4.5 Successive Applications of the Nabla Operator 209 ∂ui ∂xj ¼ ∂uj ∂xi As uidx i is an exact differential it follows that for a scalar function ϕ(xi): ϕ xi � � ¼ ∂ϕ ∂xi dxi ) ui � ∂ϕ∂xi � � dxi ¼ 0 whereby ui ¼ ∂ϕ∂xi This analysis shows that the vector u can be considered as the gradient of a scalar function ϕ(xi) as long as it fulfills the condition ∇� u ¼ 0. Expression (4.5.6) can be demonstrated changing only the order of the opera- tions, for ∇�∇ð Þϕ ¼ 0, where the term in parenthesis indicates the cross product of a vector by itself, which results in the null vector. The condition ∇� u ¼ 0 being u ¼ ∇ϕ xið Þ, where the scalar field defined by the function ϕ(xi) is divided into families of level surfaces ϕ xið Þ ¼ C, which do not intersect, so they form level surface “layers,” leads to the denomination of lamellar field. (3) ∇� ∇� uð Þ For ∇� ∇� uð Þ using the Grassmann identity u� v� wð Þ ¼ u � wð Þv� u � vð Þw whereby ∇� ∇� uð Þ ¼ ∇ ∇ � uð Þ �∇ � ∇uð Þ ð4:5:7Þ In terms of the vector coordinates it follows that ∇� u ¼ εtjk ∂‘uk � � g‘jgt ¼ w ¼ wtgt wt ¼ εtjk ∂‘uk � � g‘j ∇� w ¼ gs � ∂ wtg tð Þ ∂xs ¼ gs � ∂wt ∂xs gt þ wt ∂g t ∂xs � � ∂gt ∂xs ¼ �Γ tsngn 210 4 Differential Operators ∇� w ¼ gs � ∂wt ∂xs gt � wtΓ tsngn � � ¼ ∂wt ∂xs � wnΓ nst � � gs � gt gs � gt ¼ εrstgr ∇� w ¼ εrst ∂wt ∂xs � wnΓ nst � � gr ∂sε rst ¼ 0 ∇� w ¼ εrst ∂swtð Þgr ¼ εrstεijk ∂s∂‘ukg‘j � � gr ∂sg ‘j ¼ 0 ∇� w ¼ εrstεtjk ∂s∂‘uk � � g‘j gr and with εrstεtjk ¼ δrsjk δrsjk ¼ δ rj δ sk � δ sj δ rk it follows that ∇� ∇� uð Þ ¼ δ rj δ sk ∂s∂‘uk � � g‘j � δ sj δ rk ∂s∂‘uk � � g‘j h i gr whereby ∇� ∇� uð Þ ¼ ∂k∂‘uk � � g‘r � ∂j∂‘ur � � g‘j � � gr ð4:5:8Þ ∇� ∇� uð Þ ¼ ∂k∂ruk � � g‘r � ∂j∂jur � �h i gr ð4:5:9Þ For the Cartesian coordinates the result is ∇� ∇� uð Þ ¼ ∂ 2 uk ∂xk∂xr � ∂ 2 ur ∂2xj ! gr ð4:5:10Þ (4) ∇ ∇ � uð Þ For the gradient of a vector ∇ � u ¼ ∂u i ∂xi þ ujΓ iji � � 4.5 Successive Applications of the Nabla Operator 211 then ∇ ∇ � uð Þ ¼ gm ∂ ∂xm ∂ui ∂xi þ ujΓ iji � � The development provides ∇ ∇ � uð Þ ¼ gm ∂ 2 ui ∂xm∂xi þ ∂u j ∂xm Γ iji þ uj ∂Γ iji ∂xm ! or ∇ ∇ � uð Þ ¼ ∂m ∂iui � � gm ð4:5:11Þ and with u ¼ ϕ0 rð Þ du dr ¼ ϕ0 rð Þ it follows that ∇ ∇ � uð Þ ¼ gmk∂m ∂iui � � gk whereby ∇ ∇ � uð Þ ¼ ∂k ∂iui � � gk ð4:5:12Þ Exercise 4.16 Let ϕ(xi) and ψ(xi) be scalar functions, show that: (a) ∇� ψ∇ϕþ ϕ∇ψð Þ ¼ 0; (b) ∇ � ∇ϕ�∇ψð Þ ¼ 0; (c) tr ∇� uð Þ ¼ ∇ � u. (a) Putting ∇ϕ ¼ u ∇ψ ¼ v then ∇� ψ∇ϕþ ϕ∇ψð Þ ¼ ∇� ψuþ ϕvð Þ ¼ ∇� ψuþ∇� ϕv and with the expression shown in Exercise 4.12 it follows that ∇� ψu ¼ ∇ψ � uþ ψ �∇u ¼ ∇ψ �∇ϕþ ψ∇�∇ϕ ∇� ϕv ¼ ∇ϕ� vþ ϕ�∇v ¼ ∇ϕ�∇ψ þ ϕ∇�∇ψ 212 4 Differential Operators and with expression (4.5.6) ∇�∇ϕ ¼ ∇�∇ψ ¼ 0 and ∇ϕ�∇ψ ¼ �∇ψ �∇ϕ then ∇� ψ∇ϕþ ϕ∇ψð Þ ¼ 0 Q:E:D: (b) Putting ∇ϕ ¼ u ∇ψ ¼ v ∇� ∇ϕ�∇ψð Þ ¼ 0 then ∇ � ∇ϕ�∇ψð Þ ¼ ∇ � u� vð Þ With expression deducted in Exercise 4.7b it follows that ∇ � u� vð Þ ¼ v �∇� u� u �∇� v ∇� ∇ϕ�∇ψð Þ ¼ 0 and with expression (4.5.6) ∇�∇ϕ ¼ ∇�∇ψ ¼ 0 then ∇ � ∇ϕ�∇ψð Þ ¼ 0 Q:E:D: (c) With expression (4.2.11) ∇� u ¼ ∂iukð Þgi � gk the result for i ¼ k is tr ∇� uð Þ ¼ ∂iui1 and comparing this result with expression (4.3.6) 4.5 Successive Applications of the Nabla Operator 213 ∇ � u ¼ ∂iui it is verified that tr ∇� uð Þ ¼ ∇ � u Q:E:D: 4.5.2 Laplace Operator The combination of the divergence and the gradient, in this order, defines the Laplace operator or Laplacian ∇2 ¼ ∇ �∇ ¼ Δ ¼ Dk � Dk ¼ ∂k∂k ¼ divgrad ¼ lap ð4:5:13Þ A few authors denominate this operator of differential parameter of the second order of Beltrami, and use the spelling Δ 2 to represent it. With the expression the contravariant derivative ∂k ¼ gkj∂j it follows that for the Laplacian of an arbitrary tensor ∂k∂ k T������ ¼ ∂k ∂kT������ � � ¼ ∂k gkj∂jT������ � � ¼ gkj∂k ∂jT������� � ¼ ∂j∂jT ������ that shows that the Laplacian operator is commutative. For Cartesian coordinates the covariant and contravariant derivatives are equal ∂k ¼ ∂� � �∂xk ∂ k ¼ ∂� � � ∂xk ∂k ¼ ∂k resulting for the Laplacian ð dϕ rð Þ ¼ ð m1 r2 dr 4.5.2.1 Laplacian of a Scalar Function The Laplacian of the scalar function ϕ(xi) expresses in a curvilinear coordinate system, with covariant derivative given by 214 4 Differential Operators ϕ rð Þ ¼ m1 r þ m2 thus H � � �ð Þ ¼ gi∇� gj∇ � � �ð Þ The development of the covariant derivative of the term in parenthesis provides ϕ xi � � The contracted Christoffel symbol Γ kmk ¼ 1ffiffiffi g p ∂ ffiffiffi g p� � ∂xm provides ∇2ϕ ¼ ∂ ∂xk gkj ∂ϕ ∂xj � � þ gmj ∂ϕ ∂xj 1ffiffiffi g p ∂ ffiffiffi g p� � ∂xm whereby H ϕð Þ ¼ ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 � gi � gj ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 � ϕ ¼ ∂2ϕ ∂x1∂x1 ∂2ϕ ∂x1∂x2 ∂2ϕ ∂x1∂x3 ∂2ϕ ∂x2∂x1 ∂2ϕ ∂x2∂x2 ∂2ϕ ∂x2∂x3 ∂2ϕ ∂x3∂x1 ∂2ϕ ∂x3∂x2 ∂2ϕ ∂x3∂x3 2666666664 3777777775 gi � gj or ∇2ϕ ¼ gik ∂jkϕ � � ð4:5:14Þ it follows that ∇2ϕ ¼ gik ∂ 2ϕ ∂xj∂xk � ∂ϕ ∂xm Γmjk ! ð4:5:15Þ 4.5 Successive Applications of the Nabla Operator 215 In vectorial notation ∇ � ∇ϕð Þ ¼ ∇2ϕ ð4:5:16Þ or div gradϕð Þ ¼ ∇2ϕ ð4:5:17Þ In space E3 and in orthogonal Cartesian coordinates the result is gij ¼ δij, then the Laplacian of a scalar function is the sum of its derivatives of the second order ∇2ϕ ¼ ∂ 2ϕ ∂xj∂xj ð4:5:18Þ 4.5.3 Properties The Laplacian of the sum of two scalar functions ϕ(xi) and ψ(xi) provides ∇2 ϕþ ψð Þ ¼ ∇ �∇ ϕþ ψð Þ ¼ ∇ � ∇ϕþ∇ψð Þ ¼ ∇ �∇ϕþ∇ �∇ψ whereby ∇2 ϕþ ψð Þ ¼ ∇2ϕþ∇2ψ For the function mϕ(xi), where m is a scalar, this operator provides ∇2 mϕð Þ ¼ ∇ �∇ mϕð Þ ¼ ∇ � m∇ ϕð Þ ¼ m∇ �∇ ϕð Þ whereby ∇2 mϕð Þ ¼ m∇ �∇ ϕð Þ These two demonstrations prove that the Laplacian is a linear operator. The gradient of the product of two scalar functions is given by ∇ ϕψð Þ ¼ ψ∇ ϕð Þ þ ϕ∇ ψð Þ then ∇ �∇ ϕψð Þ ¼ ∇ � ψ∇ ϕð Þ þ ϕ∇ ψð Þ½ � 216 4 Differential Operators Putting ∇ϕ ¼ u ∇ψ ¼ v thus ∇ � ψ∇ ϕð Þ þ ϕ∇ ψð Þ½ � ¼ ∇ � ψuþ ϕvð Þ Applying the distributive property of the divergence to this expression, and using the expression deducted in Exercise 4.7a it follows that ∇ � ψuþ ϕvð Þ ¼ ∇ � ψuþ∇ � ϕv ∇ � ψu ¼ ∇ψ � uþ ψ∇u ∇ � ϕv ¼ ∇ϕ � vþ ϕ∇v then ∇2 ϕψð Þ ¼ ∇ψ � uþ ψ∇uð Þ þ ∇ϕ � vþ ϕ∇vð Þ ¼ v � uþ ψ∇uð Þ þ u � vþ ϕ∇vð Þ ¼ ψ∇uþ ϕ∇vþ 2 v � uð Þ but ∇u ¼ ∇∇ϕ ¼ ∇2ϕ ∇v ¼ ∇∇ψ ¼ ∇2ψ whereby substituting ∇2 ϕψð Þ ¼ ψ∇2 ϕð Þ þ ϕ∇2 ψð Þ þ 2∇ϕ∇ψ An equation involving the Laplacian of a scalar function that appears in various problems of physics and engineering, the Laplace equation, is given by ∇2ϕ xi � � ¼ 0 ð4:5:19Þ The functionϕ xið Þ ¼ x4zþ 3xy2 � zxyþ 1 that fulfills this equation is said to be harmonic. In addition to satisfying the Laplace equation it must be regular in the domain D, with partial derivatives of the first order continuous in the interior and in the boundary of D, and derivatives of the second order also continuous in D, which can be discontinuous in the boundary of this domain. The successive applying of the Laplacian to a scalar function (r; z sin θ; eθ cos z ) results in the bi-harmonic equation ϕ xi � � ¼ xyþ yzþ xz ð4:5:20Þ 4.5 Successive Applications of the Nabla Operator 217 For ∂� � � ∂t2 ð4:5:21Þ where ψ(xi) is a scalar function, this partial differential equation is called Poisson’s equation. As a consequence of the definition of the Laplacian the result is that ∇2m ¼ 0, where m is a scalar. The Laplacian of a scalar function ϕ(xi) is a scalar, then its physical components are equal to its ordinary components. 4.5.4 Orthogonal Coordinate Systems With the gradient of the scalar function ϕ(xi): ∇ϕ xi � � ¼ gi ∂ϕ ∂xi ¼ u and the orthogonalcomponents of the vectorial function u given by ∇ � u ¼ 1 h1h2h3 ∂ ∂xi uih1h2h3 hi � � results for the Laplacian of this function expressed in an orthogonal coordinate system ∇2ϕ ¼ 1 h1h2h3 ∂ ∂xi h1h2h3 hi ∂ϕ ∂xi � � ð4:5:22Þ where h1, h2, h3 are the components of the metric tensor. 4.5.5 Laplacian of a Vector With expression (4.5.7) ∇2u ¼ ∇ ∇ � uð Þ �∇� ∇� uð Þ and substituting expressions (4.5.12) and (4.5.9) this expression becomes 218 4 Differential Operators ∇2u ¼ gk∂k ∂iui � �� ∂k∂ruk� �� ∂j∂jur� �h igr ð4:5:23Þ The change of the indexes k ! r in the first term to the right and indexes k ! i in the second term to the right provides ∇2u ¼ ∂r∂iui � �� ∂i∂rui� �þ ∂j∂jur� �h igr As ∂r∂iu i ¼ ∂i∂rui thus ∇2u ¼ ∂j∂jur � � gr ð4:5:24Þ 4.5.6 Curl of the Laplacian of a Vector The curl of the Laplacian of a vector is∇�∇2u, and it can be developed by means of the Grassmann formula ∇2u ¼ divgradu ¼ ∇∇ � u�∇�∇� u ð4:5:25Þ or ∇�∇� u ¼ ∇∇ � u�∇2u The curl of this expression is given by ∇�∇�∇� u ¼ ∇�∇∇ � u�∇�∇2u ð4:5:26Þ or ∇�∇�∇� u ¼ ∇∇ � ∇� uð Þ �∇ �∇ ∇� uð Þ ð4:5:27Þ Expressions (4.5.4) and (4.5.6) show, respectively, that ∇ � ∇� uð Þ ¼ 0 ∇�∇ϕ ¼ 0 4.5 Successive Applications of the Nabla Operator 219 whereby the result for expression (4.5.26) is ∇�∇�∇� u ¼ ∇�∇∇ � u�∇�∇2u ¼ ∇�∇ϕ�∇�∇2u ¼ �∇�∇2u and for expression (4.5.27) ∇�∇�∇� u ¼ �∇ �∇ ∇� uð Þ The result of these two expressions is ∇�∇2u ¼ ∇2 ∇� uð Þ ð4:5:28Þ or rot lapu ¼ lap rotu ð4:5:29Þ It is concluded that the operators ∇2 and∇� are commutative when applied to vector u. 4.5.7 Laplacian of a Second-Order Tensor The gradient of a second-order tensor is given by ∇� T ¼ ∂mTijgm � gi � gp and the divergence of the tensor defined by the previous expression stays ∇ �∇� T ¼ gk � ∂k∂mTijgm � gi � gp ¼ ∂k∂mTijgm � gi � gp � gk ¼ ∂k∂mTijgm � giδ kp whereby ∇ �∇� T ¼ ∂p∂mTijgm � gi ð4:5:30Þ is a second-order tensor. Exercise 4.17 Calculate∇2ϕ for the scalar function ϕ(xi) expressed in cylindrical coordinates. The tensorial expression that defines the Laplacian is 220 4 Differential Operators ∇2ϕ ¼ 1ffiffiffi g p ∂ ∂xk ffiffiffi g p gkj ∂ϕ ∂xj � � and for the cylindrical coordinates g11 ¼ 1 g22 ¼ 1 r2 g33 ¼ 1 ∇ϕ ¼ ∂ϕ ∂r gr þ 1 r ∂ϕ ∂θ gθ þ ∂ϕ ∂z gz it follows that ∇2ϕ ¼ 1 r ∂ ∂r r ∂ϕ ∂r � � þ ∂ ∂θ 1 r ∂ϕ ∂ϕ � � þ ∂ ∂z r ∂ϕ ∂z � � � then ∇2ϕ ¼ 1 r ∂ϕ ∂r þ ∂ 2ϕ ∂r2 þ 1 r2 ∂2ϕ ∂θ2 þ ∂ 2ϕ ∂z2 Exercise 4.18 Calculate ∇2ϕ for the scalar function ϕ(xi) expressed in spherical coordinates. The tensorial expression that defines the Laplacian is ∇2ϕ ¼ 1ffiffiffi g p ∂ ∂xk ffiffiffi g p gkj ∂ϕ ∂xj � � and for the spherical coordinates g11 ¼ 1 g22 ¼ 1 r2 g33 ¼ 1 r2 sin 2ϕ ∇ϕ ¼ ∂ϕ ∂r gr þ 1 r ∂ϕ ∂ϕ gϕ þ 1 r sin ϕ ∂ϕ ∂xθ gθ it follows that ∇2ϕ¼ 1 r2 sinϕ ∂ ∂r r2 sinϕ ∂ϕ ∂r � � þ ∂ ∂ϕ r2 sinϕ 1 r2 ∂ϕ ∂ϕ � � þ ∂ ∂θ r2 sinϕ 1 r2 sin2ϕ ∂ϕ ∂θ � � � then 4.5 Successive Applications of the Nabla Operator 221 ∇2ϕ ¼ 2 r ∂ϕ ∂r þ ∂ 2ϕ ∂r2 þ 1 r2 sin ϕ ∂ ∂ϕ sin ϕ ∂ϕ ∂ϕ � � þ 1 r2 sin 2ϕ ∂2ϕ ∂θ2 Exercise 4.19 Let r be the position vector of the point in space E3, show that: (a) ∇2 xr3 � � ¼ 0; (b)∇2 rnrð Þ ¼ n nþ 3ð Þ rn�2r; (c)∇2ϕ rð Þ ¼ ϕ00 rð Þ þ 2r ϕ0 rð Þ; (d) for∇2 ϕ rð Þ ¼ 0 the result is ϕ rð Þ ¼ m1r þ m2, where m1,m2 are constant. (a) With the definition of Laplacian ∇2 x r3 � � ¼ ∂ 2 ∂x2 þ ∂ 2 ∂y2 þ ∂ 2 ∂z2 ! x r3 � � and for the derivative with respect to the variable x ∂2 ∂x2 x r3 � � ¼ ∂ ∂x ∂ ∂x x r3 � � � ¼ ∂ ∂x 1 r3 � 3x r4 ∂r ∂x � but 2r ∂r ∂x ¼ 2x then ∂2 ∂x2 x r3 � � ¼ ∂ ∂x 1 r3 � 3x r4 x r � ¼ � 3 r4 ∂r ∂x � 6x r5 þ 15x 2 r6 ∂r ∂x ¼ � 9x r5 þ 15x 3 r2 In an analogous way for the other derivatives it follows that ∂2 ∂y2 y r3 � � ¼ � 3x r5 þ 15xy r7 ∂2 ∂z2 y r3 � � ¼ � 3x r5 þ 15xz r7 Then ∇2 x r3 � � ¼ � 9x r5 þ 15x 3 r2 � 3x r5 þ 15xy r7 � 3x r5 þ 15xz r7 ∇2 x r3 � � ¼ 0 Q:E:D: (b) Putting ∇2 rnrð Þ ¼ ∇ ∇ � rnrð Þ½ � 222 4 Differential Operators Free ebooks ==> www.Ebook777.com it follows that ∇2 rnrð Þ ¼ ∇ ∇ rnð Þ � rþ rn∇ � r½ � ¼ ∇ nrn�3r� � � rþ 3rn� � ¼ ∇ nrn�3r2� �þ 3rn� � ¼ nþ 3ð Þ∇rn then ∇2 rnrð Þ ¼ nþ 3ð Þnr�2r Q:E:D: (c) With the definition of Laplacian ∇2ϕ rð Þ ¼ ∇ � ∇ϕ rð Þ½ � it follows that ∇ � ∇ϕ rð Þ½ � ¼ ∇ � ϕ0 rð Þ∇r h i ¼ ∇ � 1 r ϕ 0 rð Þr � ¼ 1 r ϕ 0 rð Þ∇ � rþ r �∇ 1 r ϕ 0 rð Þ � but ∇ � r ¼ 3 so ∇ � ∇ϕ rð Þ½ � ¼ 3 r ϕ 0 rð Þ∇ � rþ r � d dr 1 r ϕ 0 rð Þ � ∇r � ¼ 3 r ϕ 0 rð Þ þ r � � 1 r2 ϕ 0 rð Þ þ 1 r ϕ 00 rð Þ � 1 r r � ¼ 3 r ϕ 0 rð Þ þ � 1 r2 ϕ 0 rð Þ þ 1 r ϕ 00 rð Þ � 1 r � r � r ¼ 3 r ϕ 0 rð Þ þ � 1 r2 ϕ 0 rð Þ þ 1 r ϕ 00 rð Þ � 1 r � r2 then ∇ � ∇ϕ rð Þ½ � ¼ 2 r ϕ 0 rð Þ þ ϕ00 rð Þ (d) Let ∇2ϕ rð Þ ¼ ϕ00 rð Þ þ 2 r ϕ 0 rð Þ ) ϕ00 rð Þ þ 2 r ϕ 0 rð Þ ¼ 0 ) ϕ 00 rð Þ ϕ 0 rð Þ ¼ � 2 r 4.5 Successive Applications of the Nabla Operator 223 www.Ebook777.com http://www.ebook777.com Putting u ¼ ϕ0 rð Þ ) du dr ¼ ϕ0 rð Þ then du u ¼ �2 r dr and integrating ð du u ¼ � ð 2 r dr it follows that ‘n uð Þ ¼ �‘n r2� �þ ‘n m1ð Þ ¼ ‘n m1 r2 � � or ‘nϕ 0 rð Þ ¼ ‘n m1 r2 � � ) ϕ0 rð Þ ¼ m1 r2 Integrating ð dϕ rð Þ ¼ ð m1 r2 dr then ϕ rð Þ ¼ m1 r þ m2 Q:E:D: 4.6 Other Differential Operators 4.6.1 Hesse Operator The operator defined on a scalar field, given by the tensorial product of two nabla operators applied to the scalar function that field represents 224 4 Differential Operators H � � �ð Þ ¼ gi∇� gj∇ � � �ð Þ ð4:5:31Þ In matrix form the scalar function ϕ(xi) in Cartesian coordinates in the space E3 is H ϕð Þ ¼ ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 � gi � gj ∂ ∂x1 ∂ ∂x2 ∂ ∂x3 � ϕ ¼ ∂2ϕ ∂x1∂x1 ∂2ϕ ∂x1∂x2 ∂2ϕ ∂x1∂x3 ∂2ϕ ∂x2∂x1 ∂2ϕ ∂x2∂x2 ∂2ϕ ∂x2∂x3 ∂2ϕ ∂x3∂x1 ∂2ϕ ∂x3∂x2 ∂2ϕ ∂x3∂x3 266666664 377777775 gi � g j ð4:5:32Þ This operator is a symmetric second-order tensor, is called Hessian or Hesse operator in homage to Ludwig Otto Hesse (1881–1874). 4.6.2 D’Alembert Operator The differential operator defined by the expression □ ¼ ∇2� � � þ 1 c2 ∂� � � ∂t2 ð4:5:33Þ where c is a scalar and ∂� � � ∂t2 denotes the differentiation with respect to the time t, is called D’Alembert or D’Alembertian operator in homage to Jean Le Rond d’Alembert (1717–1783). The applying of this operator to a field represented by the scalar function that depends on the position vector and the time provides as a result the scalar function □ ϕ xi; t � � ¼ ∇2ϕ xi; t� �þ 1 c2 ∂ϕ xi; tð Þ ∂t2 ð4:5:34Þ If the field is represented by a vectorial function the result is the vector □ u xi; t � � ¼ ∇2u xi; t� �þ 1 c2 ∂u xi; tð Þ ∂t2 ð4:5:35Þ The notation □. . . was initially applied by Cauchy to represent the Laplacian. The D’Alembertian is the four-dimensional equivalent to the Laplacian. 4.6 Other Differential Operators 225 Problems 4.1 Calculate the gradient of the scalar functions: (a) ϕ xið Þ ¼ xyþ yzþ xz; (b) ϕ xið Þ ¼ xex2þy2 . Answer: (a) yþ zð Þiþ xþ zð Þjþ xþ yð Þk; (b) 1þ 2x2ð Þex2þy2 iþ 2xyex2þy2 j 4.2 Calculate the directional derivative of the scalar function ϕ ¼ 2 x1ð Þ2 þ 3 x2ð Þ 2 þ x3ð Þ2 at the point (2; 1; 3) in the direction of vector u 1; 0;�2ð Þ. Answer: �1:789. 4.3 Calculate div ui with the vector u expressed in cylindrical coordinates by its covariant components (r; z sin θ; eθ cos z). Answer: divui ¼ 2þ 1rð Þ 2 z cos θ � eθ sin z 4.4 Show that∇ � ϕψuð Þ ¼ ϕ∇ψ � uþ ψ∇ϕ � uþ ϕψ∇ � u, where ϕ,ψ are scalar functions and u is a vectorial function. 4.5 Calculate the curl of the following vectorial fields: (a) y2iþ z2jþ x2k; (b) xyz xiþ yjþ zkð Þ. Answer: (a) �2 ziþ xjþ ykð Þ; (b) xz2 � xy2ð Þ iþ x2y� yz2ð Þ jþ y2z� x2zð Þk. 4.6 Calculate the Laplacian of the function ϕ xið Þ ¼ x4zþ 3xy2 � zxyþ 1. Answer: ∇2ϕ ¼ 12x2z� 6x. 226 4 Differential Operators Chapter 5 Riemann Spaces 5.1 Preview The space provided with metric is called Riemann space, for which the tensorial formalism is based on the study inits first fundamental form, being complemented by the definition of curvature and by the concept of geodesics, which allows expanding the basic conceptions of the Euclidian geometry for this type of space with N dimensions. In the Riemann spaces the covariant derivatives of tensors are equal to the partial derivatives when the coordinates are Cartesian, but the problem arises of researching how these derivatives behave when the coordinate system is curvilin- ear. The analysis of this derivative leads to the definition of curvature of the space, which is the fundamental parameter for the development of a consistent study of the Riemann spaces EN. The concepts and expressions of Tensor Calculus are essential for the formula- tion of the Theory of General Relativity, and it is for this theory just as the Integral and Differential Calculus is for the Classic Mechanics. 5.2 The Curvature Tensor The Euclidian geometry is grounded on the basic concepts of point, straight line, and plane, and in various axioms. In this geometry a curved line is defined in the Euclidian space E2 as the one that is not a straight line, and in the Euclidian space E3 a curved surface is defined as the one that is not a plane. The curvature is an intrinsic characteristic of the space, so it is not a property measurable by comparison between distinct spaces. © Springer International Publishing Switzerland 2016 E. de Souza Sánchez Filho, Tensor Calculus for Engineers and Physicists, DOI 10.1007/978-3-319-31520-1_5 227 The conception of a Riemann geometry for the space EN is grounded on the basic concepts of the Euclidian geometry in space E3, which generalization is carried out by means of defining the metric for the space EN, given by ds2 ¼ εgijdxidxj where ε ¼ 1 is a functional indicator. The space in which metric can be writed as an Euclidian metric, positive and definite, is called a flat space, otherwise it is called space with curvature. The concept of curvature of the space EN was firstly conceived by Riemann as a generalization of the study of a surface’s curvature developed by Gauß. Riemann presented his results in a paper in 1861, published only in 1876. Christoffel in 1869 and R. Lipschitz in four papers published in 1869, 1870 (two articles), and 1877 obtained the same results as Riemann when studying the transformation of the quadratic differential formula gijdx idxj to the Euclidian metric ds2 ¼ X i dxi � �2 . The curvature analysis of the Riemann space EN was carried out by Ricci-Curbastro and Levi-Civita who deducted the expression of the curvature tensor in a very formal and concise approach, which was also obtained by Christoffel, whose deduction has an extensive algebrism. In 1917 Tulio Levi-Civita, and after Jan Arnoldus Schouten (1918) and Karl Hessenberg found independently an interpre- tation for the curvature tensor associating it to the concept of parallel transport of vectors. 5.2.1 Formulation The covariant derivative of a tensor is a tensor, just as when repeating this differentiation will provide a new tensor. However, the differentiation order with respect to the variables must be considered in this analysis. For a function ϕ(xi) of class C2 that represents a scalar field exists the derivative ∂ϕ xið Þ ∂xk that represents a covariant vector. Differentiating again with respect to the variable xj results by means of the partial differentiation rule of Differential Calculus ∂2ϕ xið Þ ∂xk∂xj ¼ ∂ 2ϕ xið Þ ∂xj∂xk In this case, the covariant derivative is commutative. However, for tensors which components are functions of class C2 represented in curvilinear coordinate systems this independence of the differentiation order in general is not verified. It is concluded that only the condition of the functions being class C2 is not enough to ensure this independence. 228 5 Riemann Spaces For the case of a covariant vector ui the result for its covariant derivative is the tensor with variance (0, 2): ∂jui ¼ ∂ui∂xj � u‘Γ ‘ ij ð5:2:1Þ and with ∂jui ¼ Tij it follows that for the covariant derivative of this tensor with respect to the variable xk ∂kTij ¼ ∂Tij∂xk � T‘jΓ ‘ ik � Ti‘Γ ‘jk ð5:2:2Þ The substitution of expression (5.2.1) provides ∂kTij ¼ ∂ ∂jui � � ∂xk � ∂ju‘ � � Γ ‘ik � ∂‘uið ÞΓ ‘jk ¼ ∂ ∂xk ∂ui ∂xj � u‘Γ ‘ij � � � ∂u‘ ∂xj � umΓm‘j � � Γ ‘ik � ∂ui ∂x‘ � umΓmi‘ � � Γ ‘jk it follows that ∂kTij ¼ ∂j∂kui ¼ ∂ 2 ui ∂xk∂xj � ∂u‘ ∂xk Γ ‘ij � u‘ ∂Γ ‘ij ∂xj � ∂u‘ ∂xj Γ ‘ik þ umΓm‘jΓ ‘ik � ∂ui ∂x‘ Γ ‘jk þ umΓmi‘Γ ‘jk ð5:2:3Þ that represents a tensor with variance (0, 3). The inversion of the differentiation order provides ∂k∂jui ¼ ∂ 2 ui ∂xj∂xk � ∂u‘ ∂xj Γ ‘ik � u‘ ∂Γ ‘ik ∂xj � ∂u‘ ∂xk Γ ‘ij þ umΓm‘kΓ ‘ij � ∂ui ∂x‘ Γ ‘kj þ umΓmi‘Γ ‘kj ð5:2:4Þ In Differential Calculus the differentiation order does not change the result obtained then ∂2ui ∂xj∂xk ¼ ∂ 2 ui ∂xk∂xj 5.2 The Curvature Tensor 229 Subtracting expression (5.2.4) from expression (5.2.3) and considering the symmetry of the Christoffel symbols ∂j∂kui � ∂k∂jui ¼ u‘ ∂Γ ‘ ik ∂xj � ∂Γ ‘ ij ∂xk ! þ um Γm‘jΓ ‘ik � Γm‘kΓ ‘ij � � and with the permutation of the dummy indexes m $ ‘ in the second term to the right ∂j∂kui � ∂k∂jui ¼ u‘ ∂Γ ‘ ik ∂xj � ∂Γ ‘ ij ∂xk ! þ Γ ‘mjΓmik � Γ ‘mkΓmij � �" # Putting R ‘ijk ¼ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ Γ ‘mjΓmik � Γ ‘mkΓmij ð5:2:5Þ results in ∂j∂kui � ∂kjui ¼ u‘R ‘ijk The quotient law is used for verifying if the variety R‘ijk is a tensor, carrying out the inner product of vector u‘ by R ‘ ijk: R ‘ijku‘ ¼ R ‘ijk‘ ¼ Rijk The transformation law of tensors to the variety Rijk is given by Rpqr ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr Rijk for the vector u‘ the result of the transformation law is um ¼ ∂x ‘ ∂xm u‘ that substituted in previous expression provides Rpqr ¼ ∂x i ∂xp ∂xj ∂xq ∂xk ∂xr ∂x‘ ∂xm R ‘ijku‘ In the coordinate system X i the variety Rpqr is given by 230 5 Riemann Spaces Rpqr ¼ R ‘pqru‘ whereby R ‘ pqr � ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr ∂x‘ ∂xm R ‘ijk � � u‘ ¼ 0 As �u‘ is an arbitrary vector it results in R ‘ pqr ¼ ∂xi ∂xp ∂xj ∂xq ∂xk ∂xr ∂x‘ ∂xm R ‘ijk that represents the transformation law of tensor with variances (1, 3), as R‘ijk is a tensor. The tensor defined by expression (5.2.5) is called Riemann–Christoffel curvature tensor, Riemann–Christoffel mixed tensor, or Riemann–Christoffel ten- sor of the second kind, or simply curvature tensor. This tensor defines a tensorial field that depends only on the metric tensor and its derivatives up to the second order, and classifies the space, for thus R ‘ijk 6¼ 0 the result is a space with curvature. 5.2.2 Differentiation Commutativity The formulation of an analogous analysis for a contravariant vector ui, which generates a mixed tensor with variance (1, 1), is carried out by calculating firstly the covariant derivative of this vector with respect to the coordinate xj: ∂ju i ¼ ∂u i ∂xj þ u‘Γ ij‘ ¼ T ij ð5:2:6Þ The covariant derivative of the second order of this vector with respect to the coordinate xk is given by ∂k ∂ju i � � ¼ ∂kT ij ¼ ∂T ij∂xk þ T ‘j Γ i‘k � T i‘Γ ‘jk Substituting expression (5.2.6) in this expression ∂k ∂ju i � � ¼ ∂ ∂xk ∂ui ∂xj þ u‘Γ ij‘ � � þ ∂u ‘ ∂xj þ umΓ ‘jm � � Γ i‘k � ∂ui ∂x‘ þ umΓ i‘m � � Γ ‘jk whereby 5.2 The Curvature Tensor 231 ∂k ∂ju i � � ¼ ∂2ui ∂xk∂xj þ ∂u ‘ ∂xk Γ ij‘ þ u‘ ∂Γ ij‘ ∂xk þ ∂u ‘ ∂xj Γ i‘k þ umΓ ‘jmΓ i‘k � ∂ui ∂x‘ Γ ‘jk � umΓ i‘mΓ ‘jk ð5:2:7Þ The inversion of the differentiation is obtained interchanging indexes j $ k, so ∂j ∂ku i � � ¼ ∂2ui ∂xj∂xk þ ∂u ‘ ∂xj Γ ik‘ þ u‘ ∂Γ ik‘ ∂xj þ ∂u ‘ ∂xk Γ i‘j þ umΓ ‘kmΓ i‘j � ∂ui ∂x‘ Γ ‘kj � umΓ i‘mΓ ‘kj ð5:2:8Þ As in the partial derivative the order of differentiation does not change the result ∂2ui ∂xk∂xj ¼ ∂ 2 ui ∂xj∂xk and subtracting expression (5.2.7) from expression (5.2.8) ∂k ∂ju i � �� ∂j ∂kui� � ¼ ∂2ui∂xk∂xj þ ∂u‘∂xk Γ ij‘ þ u‘ ∂Γ i j‘ ∂xk þ ∂u ‘ ∂xj Γ i‘k þ umΓ ‘jmΓ i‘k � ∂ui ∂x‘ Γ ‘jk �umΓ i‘mΓ ‘jk � � ∂ 2 ui ∂xj∂xk þ ∂u ‘ ∂xj Γ ik‘ þ u‘ ∂Γ ik‘ ∂xj þ ∂u ‘ ∂xk Γ i‘j þ umΓ ‘kmΓ i‘j � ∂ui ∂x‘ Γ ‘kj � umΓ i‘mΓ ‘kj � and with the symmetry of the Christoffel symbols ∂k ∂ju i � �� ∂j ∂kui� � ¼ u‘ ∂Γ ij‘∂xk � ∂Γ ik‘∂xj ! þ um Γ ‘jmΓ i‘k � Γ ‘kmΓ i‘j � � The permutation of indexes ‘ $ m in the last two terms provides ∂k ∂ju i � �� ∂j ∂kui� � ¼ ∂Γ ij‘∂xk � ∂Γ ik‘∂xj þmj‘ Γ imk � Γmk‘Γ imj ! u‘ 232 5 Riemann Spaces Free ebooks ==> www.Ebook777.com and putting Ri‘kj ¼ ∂Γ ij‘ ∂xk � ∂Γ i k‘ ∂xj þ Γmj‘Γ imk � Γmk‘Γ imj ð5:2:9Þ it results in ∂k ∂ju i � �� ∂j ∂kui� � ¼ Ri‘kju‘ ð5:2:10Þ The permutation of indexes j $ k provides ∂j ∂ku i � �� ∂k ∂jui� � ¼ Ri‘jku‘ where Ri‘jk ¼ ∂Γ ik‘ ∂xj � ∂Γ i j‘ ∂xk þ Γmk‘Γ imj � Γmj‘Γ imk ð5:2:11Þ This analysis shows that Ri‘jk ¼ 0 ) ∂j∂kuk ¼ ∂k∂juk, i.e., the space is flat. The necessary and sufficient condition so that the differentiation commutativity be valid is that the tensor Ri‘jk be null. 5.2.3 Antisymmetry of Tensor Ri‘jk The comparison of expressions (5.2.9) and (5.2.11) shows that the Riemann– Christoffel curvature tensor is antisymmetric with respect to the last two indexes Ri‘kj ¼ �Ri‘jk 5.2.4 Notations for Tensor Ri‘jk Putting the indexes in the sequence i, j, k, ‘ the result in tensorial notation is R ¼ R ‘ijkg‘ � gk � gj � gi ð5:2:12Þ and rewriting the Riemann–Christoffel curvature tensor as 5.2 The Curvature Tensor 233 www.Ebook777.com http://www.ebook777.com R ‘ijk ¼ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ ΓmikΓ ‘mj � Γmij Γ ‘mk the result in symbolic form by means of determinants is R ‘ijk ¼ ∂ ∂xj ∂ ∂xk Γ ‘ij Γ ‘ ik þ Γmik Γ m ij Γ ‘mk Γ ‘ mj ð5:2:13Þ 5.2.5 Uniqueness of Tensor R‘ijk The metric tensor gij and its conjugated tensor g ij are unique in a Riemann space, then their partial derivatives of the first and second order the Christoffel symbols of this space are unique at pointxi2EN . Thus it is verified that expression (5.2.11) does not ensure that tensor Ri‘jk is the only tensor that can be expressed by the derivatives of the first and second order of the metric tensor. However, the covariant derivatives of a contravariant vector with respect to the coordinates of a referential system are unique at point xi2EN , and having the Riemann–Christoffel curvature tensor with variance (1, 3) obtained by means of these derivatives, it is concluded that it is unique in the point being considered. Expressions (5.2.5) and (5.2.11) obtained in distinct manners indicate this tensor’s uniqueness. For the points xi2EN in which the Christoffel symbols are null, it is verified that Ri‘jk is expressed by means of a linear combination of the derivatives of the second order of the metric tensor. 5.2.6 First Bianchi Identity The Riemann–Christoffel curvature tensor R ‘ijk ¼ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ ΓmikΓΓ ‘mj � Γmij Γ ‘mk and the cyclic permutations of indexes i, j, k generate the expressions R ‘jki ¼ ∂Γ ‘ji ∂xk � ∂Γ ‘ jk ∂xi þ Γmji Γ ‘mk � ΓmjkΓ ‘mi 234 5 Riemann Spaces R ‘kij ¼ ∂Γ ‘kj ∂xi � ∂Γ ‘ ki ∂xj þ ΓmkjΓ ‘mi � ΓmkiΓ ‘mj The sum of these three expressions provides the first Bianchi identity for the Riemann–Christoffel curvature tensor R ‘ijk þ R ‘jki þ R ‘kij ¼ 0 ð5:2:14Þ 5.2.7 Second Bianchi Identity The covariant derivative of a tensor with variance (1, 3) is given by ∂kT j p‘m ¼ ∂T jp‘m ∂xk � T jq‘mΓ qpk � T jpqmΓ q‘k � T jp‘qΓ qmk þ T qp‘mΓ jkq whereby for the Riemann–Christoffel curvature tensor yields it follows that R ‘ijk ¼ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ ΓmikΓ ‘mj � Γmij Γ ‘mk The covariant derivative with respect to the coordinate xp is given by ∂pR ‘ijk ¼ ∂2Γ ‘ik ∂xp∂xj � ∂ 2Γ ‘ij ∂xp∂xk þ ∂Γ m ik ∂xp Γ ‘mj þ Γmik ∂Γ ‘mj ∂xp � ∂Γ m ij ∂xp Γ ‘mk � Γmij ∂Γ ‘mk ∂xp þRmijkΓ ‘mp � R ‘mjkΓmip � R ‘imkΓmjp � R ‘ijmΓmkp and with the cyclic permutation of indexes j, k, p it follows that ∂jR ‘ikp ¼ ∂2Γ ‘ip ∂xj∂xk � ∂ 2Γ ‘ik ∂xj∂xp þ ∂Γ m ip ∂xj Γ ‘mk þ Γmip ∂Γ ‘mk ∂xj � ∂Γ m ik ∂xj Γ ‘mp � Γmik ∂Γ ‘mp ∂xj þ RmikpΓ ‘mj � R ‘mkpΓΓmij � R ‘impΓmkj � R ‘ikmΓmpj ∂kR ‘ipj ¼ ∂2Γ ‘ij ∂xk∂xp � ∂ 2Γ ‘ip ∂xk∂xj þ ∂Γ m ij ∂xk Γ ‘mp þ Γmij ∂Γ ‘mp ∂xk � ∂Γ m ip ∂xk Γ ‘mj � Γmip ∂Γ ‘mj ∂xk þ RmipjΓ ‘mk � R ‘mpjΓmik � R ‘imjΓmpk � R ‘ipmΓmjk The sum of these three expressions provides 5.2 The Curvature Tensor 235 ∂pR ‘ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ ∂2Γ ‘ik ∂xp∂xj � ∂ 2Γ ‘ij ∂xp∂xk þ ∂Γ m ik ∂xp Γ ‘mj þ Γmik ∂Γ ‘mj ∂xp � ∂Γ m ij ∂xp Γ ‘mk �Γmij ∂Γ ‘mk ∂xp þ RmijkΓ ‘mp � R ‘mjkΓmip � R ‘imkΓmjp � R ‘ijmΓmkp þ ∂ 2Γ ‘ip ∂xj∂xk � ∂ 2Γ ‘ik ∂xj∂xp þ ∂Γ m ip ∂xj Γ ‘mk þ Γmip ∂Γ ‘mk ∂xj � ∂Γ m ik ∂xj Γ ‘mp �Γmik ∂Γ ‘mp ∂xj þ RmikpΓ ‘mj � R ‘mkpΓmij � R ‘impΓmkj � R ‘ikmΓmpj þ ∂ 2Γ ‘ij ∂xk∂xp � ∂ 2Γ ‘ip ∂xk∂xj þ ∂Γ m ij ∂xk Γ ‘mp þ Γmij ∂Γ ‘mp ∂xk � ∂Γ m ip ∂xk Γ ‘mj �Γmip ∂Γ ‘mj ∂xk þ RmipjΓ ‘mk � R ‘mpjΓmik � R ‘imjΓmpk � R ‘ipmΓmjk and with the equalities ∂2Γ ‘ik ∂xp∂xj ¼ ∂ 2Γ ‘ik ∂xj∂xp ∂2Γ ‘ij ∂xp∂xk ¼ ∂ 2Γ ‘ij ∂xk∂xp ∂2Γ ‘ip ∂xj∂xk ¼ ∂ 2Γ ‘ip ∂xk∂xj the previous expression stays ∂pR ‘ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ ∂Γmik ∂xp Γ ‘mj þ Γmik ∂Γ ‘mj ∂xp � ∂Γ m ij ∂xp Γ ‘mk � Γmij ∂Γ ‘mk ∂xp þRmijkΓ ‘mp � R ‘mjkΓmip � R ‘imkΓmjp � R ‘ijmΓmkp þ∂Γ m ip ∂xj Γ ‘mk þ Γmip ∂Γ ‘mk ∂xj � ∂Γ m ik ∂xj Γ ‘mp � Γmik ∂Γ ‘mp ∂xj þRmikpΓ ‘mj � R ‘mkpΓmij � R ‘impΓmkj � R ‘ikmΓmpj þ∂Γ m ij ∂xk Γ ‘mp þ Γmij ∂Γ ‘mp ∂xk � ∂Γ m ip ∂xk Γ ‘mj � Γmip ∂Γ ‘mj ∂xk þRmipjΓ ‘mk � R ‘mpjΓmik � R ‘imjΓmpk � R ‘ipmΓmjk Putting the Christoffel symbols in evidence and considering the antisymmetry of the Riemann–Christoffel curvature tensor, i.e., R ‘imk ¼ �R ‘ikm, R ‘ijm ¼ �R ‘imj, R ‘imp ¼ �R ‘ipm, and the symmetry of the Christoffel symbols, i.e., Γmjp ¼ Γmpj , Γmkp ¼ Γmpk, Γmkj ¼ Γmjk it follows that 236 5 Riemann Spaces ∂pR ‘ikjþ∂jR ‘ikpþ∂kR ‘ipj ¼ Γ ‘mp Rmijk� ∂Γmik ∂xj þ∂Γ m ij ∂xk � � þΓ ‘mj Rmikpþ ∂Γmik ∂xp �∂Γ m ip ∂xk � � þΓ ‘mk Rmipj� ∂Γmij ∂xp þ∂Γ m ip ∂xj � � �Γmip R ‘mjk� ∂Γ ‘mk ∂xj þ∂Γ ‘ mj ∂xk ! �Γmij R ‘mkpþ ∂Γ ‘mk ∂xp �∂Γ ‘ mp ∂xk ! �Γmik R ‘mpj� ∂Γ ‘mj ∂xp þ∂Γ ‘ mp ∂xj ! The expressions of the tensors are given by Rmijk ¼ ∂Γmik ∂xj � ∂Γ m ij ∂xk þ Γ qikΓmqj � Γ qijΓmqk Rmikp ¼ ∂Γmip ∂xk � ∂Γ m ik ∂xp þ Γ qipΓmqk � Γ qikΓmqp Rmipj ¼ ∂Γmij ∂xp � ∂Γ m ip ∂xj þ Γ qijΓmqp � Γ qipΓmqj R ‘mjk ¼ ∂Γ ‘mk ∂xj � ∂Γ ‘ mj ∂xk þ Γ qmkΓ ‘qj � Γ qmjΓ ‘qk R ‘mkp ¼ ∂Γ ‘mp ∂xk � ∂Γ ‘ mk ∂xp þ Γ qmpΓ ‘qk � Γ qmkΓ ‘qp R ‘mpj ¼ ∂Γ ‘mj ∂xp � ∂Γ ‘ mp ∂xj þ Γ qmjΓ ‘qp � Γ qmpΓ ‘qj that substituted in previous expression provide ∂pR ‘ikjþ∂jR ‘ikpþ∂kR ‘ipj ¼ Γ ‘mp ∂Γmik ∂xj �∂Γ m ij ∂xk þΓ qikΓmqj �Γ qijΓmqk� ∂Γmik ∂xj þ∂Γ m ij ∂xk � � þΓ ‘mj ∂Γmip ∂xk �∂Γ m ik ∂xp þΓ qipΓmqk�Γ qikΓmqpþ ∂Γmik ∂xp �∂Γ m ip ∂xk � � þΓ ‘mk ∂Γmij ∂xp �∂Γ m ip ∂xj þΓ qijΓmqp�Γ qipΓmqj � ∂Γmij ∂xp þ∂Γ m ip ∂xj � � �Γmip ∂Γ ‘mk ∂xj �∂Γ ‘ mj ∂xk þΓ qmkΓ ‘qj�Γ qmjΓ ‘qk� ∂Γ ‘mk ∂xj þ∂Γ ‘ mj ∂xk ! �Γmij ∂Γ ‘mp ∂xk �∂Γ ‘ mk ∂xp þΓ qmpΓ ‘qk�Γ qmkΓ ‘qpþ ∂Γ ‘mk ∂xp �∂Γ ‘ mp ∂xk ! �Γmik ∂Γ ‘mj ∂xp �∂Γ ‘ mp ∂xj þΓ qmjΓ ‘qp�Γ qmpΓ ‘qj� ∂Γ ‘mj ∂xp þ∂Γ ‘ mp ∂xj ! 5.2 The Curvature Tensor 237 Simplifying ∂pR ‘ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ Γ ‘mp Γ qikΓmqj � Γ qijΓmqk � � þΓ ‘mj Γ qipΓmqk � Γ qikΓmqp � � þΓ ‘mk Γ qijΓmqp � Γ qipΓmqj � � �Γmip Γ qmkΓ ‘qj � Γ qmjΓ ‘qk � � �Γmij Γ qmpΓ ‘qk � Γ qmkΓ ‘qp � � �Γmik Γ qmjΓ ‘qp � Γ qmpΓ ‘qj � � and with the permutation of the dummy indexes m $ q in the first six terms it follows that ∂pR ‘ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ Γ ‘qpΓmikΓ qmj � Γ ‘qpΓmij Γ qmk þΓ ‘qjΓmipΓ qmk � Γ ‘qjΓmikΓ qmp þΓ ‘qkΓmij Γ qmp � Γ ‘qkΓmipΓ qmj �ΓmipΓ qmkΓ ‘qj þ ΓmipΓ qmjΓ ‘qk �Γmij Γ qmpΓ ‘qk þ Γmij Γ qmkΓ ‘qp �ΓmikΓ qmjΓ ‘qp þ ΓmikΓ qmpΓ ‘qj whereby ∂pR ‘ ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ 0 ð5:2:15Þ that is called second Bianchi identity. 5.2.8 Curvature Tensor of Variance (0, 4) The Riemann–Christoffel curvature tensor generates a curvature tensor expressed in covariant components. With the multiplying of tensor R‘ijk by the metric tensorgp‘ it follows that 238 5 Riemann Spaces gp‘R ‘ ijk ¼ gp‘ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ ΓmikΓ ‘mj � Γmij Γ ‘mk ! or gp‘R ‘ ijk ¼ ∂ gp‘Γ ‘ ik � � ∂xj � ∂gp‘ ∂xj Γ ‘ik � ∂ gp‘Γ ‘ ij � � ∂xk þ ∂gp‘ ∂xk Γ ‘ij þ gp‘ΓmikΓΓ ‘mj � gp‘Γmij Γ ‘mk Ricci’s identity allows writing ∂gp‘ ∂xj ¼ Γpj, ‘ þ Γ‘j, p ∂gp‘ ∂xk ¼ Γpk, ‘ þ Γ‘k, p then gp‘R ‘ ijk ¼ ∂ gp‘Γ ‘ ik � � ∂xj � ∂ gp‘Γ ‘ ij � � ∂xk þ Γ ‘ik Γpj, ‘ þ Γ‘j, p � �þ Γ ‘ij Γpk, ‘ þ Γ‘k, p� � �Γmk, pΓmij þ Γmj, pΓmik ¼ ∂Γik, p ∂xj � ∂Γij, p ∂xk � Γ ‘ik Γpj, ‘ þ Γ‘j, p � �þ Γ ‘ij Γpk, ‘ þ Γ‘k, p� � �Γmk, pΓmij þ Γmj, pΓmik and replacing indexes m ! ‘ in the last two terms gp‘R ‘ ijk ¼ ∂Γik,p ∂xj �∂Γij,p ∂xk �Γ ‘ik Γpj,‘þΓ‘j,p � �þΓ ‘ij Γpk,‘þΓ‘k,p� ��Γ‘k,pΓ ‘ijþΓ‘j,pΓ ‘ik ¼∂Γik,p ∂xj �∂Γij,p ∂xk þΓ ‘ijΓpk,‘�Γ ‘ikΓpj,‘ whereby the result for the Riemann–Christoffel curvature tensor with variance (0, 4) or Riemann–Christoffel of first tensor type is Rpijk ¼ ∂Γik, p∂xj � ∂Γij, p ∂xk þ Γ ‘ijΓpk, ‘ � Γ ‘ikΓpj, ‘ ð5:2:16Þ which in tensorial notation is written as R ¼ Rpijkgp � gi � gj � gk ð5:2:17Þ and in symbolic form by means of determinants stays 5.2 The Curvature Tensor 239 Rpijk ¼ ∂ ∂xj ∂ ∂xk Γij, p Γik, p þ Γ ‘ ij Γ ‘ ik Γpj, ‘ Γpk, ‘ ð5:2:18Þ In tensorial notation the Riemann–Christoffel tensors, mixed and covariant, are represented by R. 5.2.9 Properties of Tensor Rpijk For the Riemann–Christoffel covariant tensor the first Bianchi identity provides g‘p R ‘ ikj þ R ‘jki þ R ‘kij � � ¼ 0 whereby the following cyclic property results Rpikj þ Rpjki þ Rpkij ¼ 0 ð5:2:19Þ Considering the antisymmetry of the Riemann–Christoffel tensor with variance (1, 3) the result is gp‘R ‘ ijk ¼ �gp‘R ‘ikj ) Rpijk ¼ �Rpikj then the Riemann–Christoffel tensor with variance (0, 4) is antisymmetric in the last two indexes. Rewriting expression (5.2.16) Rpijk ¼ ∂Γik, p∂xj � ∂Γij, p ∂xk þ Γ ‘ijΓpk, ‘ � Γ ‘ikΓpj, ‘ and with expressions Γik, p ¼ 1 2 ∂gpk ∂xi þ ∂gip ∂xk � ∂gik ∂xp � � Γij, p ¼ 1 2 ∂gjp ∂xi þ ∂gip ∂xj � ∂gij ∂xp � � Γpk, ‘ ¼ gq‘Γ qpk Γpj, ‘ ¼ gq‘Γ qpj it follows that Rpijk ¼ ∂∂xj 1 2 ∂gpk ∂xi þ ∂gip ∂xk � ∂gik ∂xp � � � � ∂ ∂xk 1 2 ∂gjp ∂xi þ ∂gip ∂xj � ∂gij ∂xp � � � þ gq‘Γ qpkΓ ‘ij � gq‘Γ qpjΓ ‘ik 240 5 Riemann Spaces Rpijk ¼ 1 2 ∂2gik ∂xj∂xp þ ∂ 2 gpk ∂xj∂xi � ∂ 2 gji ∂xk∂xp � ∂ 2 gpj ∂xk∂xi ! þ gq‘ Γ qpkΓ ‘ij � Γ qpjΓ ‘ik � � ð5:2:20Þ The expression (5.2.20) allows calculating the components of the tensor Rpijk directly in terms of the metric tensor. With the permutation of indexes i $ p in expression (5.2.20) Ripjk ¼ 1 2 ∂2gpk ∂xj∂xi þ ∂ 2 gik ∂xj∂xp � ∂ 2 gjp ∂xk∂xi � ∂ 2 gij ∂xk∂xp ! þ gq‘ Γ qpkΓ ‘ij � Γ qpjΓ ‘ik � � and with the permutation of the dummy indexes q $ ‘ this expression becomes Ripjk ¼ 1 2 ∂2gpk ∂xj∂xi þ ∂ 2 gik ∂xj∂xp � ∂ 2 gjp ∂xk∂xi � ∂ 2 gij ∂xk∂xp ! þ g‘q Γ ‘ikΓ qpj � Γ ‘ijΓ qpk � � Considering the symmetry of the metric tensor it is verified that the term to the right represents the components �Rpijk, then Ripjk ¼ �Rpijk, i.e., the tensor is antisymmetric in the first two indexes. These analyses show that the tensor Rpijk is antisymmetric in the first two and the last two indexes. The permutation of indexes p $ j, i $ k in expression (5.2.20) leads to Rpijk ¼ 1 2 ∂2gji ∂xp∂xk � ∂ 2 gki ∂xp∂xj � ∂ 2 gjp ∂xi∂xk þ ∂ 2 gpk ∂xi∂xj ! þ gq‘ Γ qjiΓ ‘kp � Γ qjpΓ ‘ki � � The symmetry of the metric tensor gives Rpijk ¼ Rjkpi. It is concluded that the tensor Rpijk is symmetric for the permutation of the pair of initial indexes for the pair of final indexes. 5.2.10 Distinct Algebraic Components of Tensor Rpijk The number of components of tensor Rpijk in the Riemann space EN cannot be obtained counting the equations Rpikj þ Rpjki þ Rpkij ¼ 0 and considering the com- ponents antisymmetricRpijk ¼ �Ripjk,Rpijk ¼ �Rpikj and the symmetric components Rpijk ¼ Rjkpi, because these two equations overlap. The methodology used to carry out this counting is given by means of classifying the tensor components into four groups, as a function of the number of repeated indexes: (a) The four indexes are equal Riiii (b) The initial pair of indexes is equal to the second pair Ripip 5.2 The Curvature Tensor 241 (c) One index is repeated Rppik (d) The four indexes are different Rpijk Case (a) must fulfill the antisymmetry of tensor Rpijk that provides Riiii ¼ �Riiii then Riiii ¼ 0. The components are null when the four indexes are equal. For case (b) only two indexes are different: Ripip having that these components differ from the components Rippi solely in the sign, and by the antisymmetry the result is Rpipi ¼ �Rippi ¼ � �Ripip � � ¼ Ripip. There is a number of components for Ripip as many as the different pair of indexes, i.e., i 6¼ p. For index i there are N distinct combinations, and for index p there are N � 1ð Þ distinct combinations, and considering the antisymmetry of the tensor for these last indexes N 2 N � 1ð Þ different combinations result. This number of combinations corresponds to the number of the N 2 N � 1ð Þ distinct combinations. There is no reduction of com- ponents due to the symmetry Rpijk ¼ Rjkpi. The first Bianchi identity is satisfied, for Rpipi þ Rppii þ Rpiip ¼ Rpipi þ 0� Rpipi ¼ 0 does not reduce the number of components. Therefore, in this case only N 2 N � 1ð Þ independent components are non-null. Case (c) has components of the kind Rppik. In this case there are N combinations for the index p, N � 1ð Þ combinations for index i, and N � 2ð Þ combinations for index k. The number of combinations for the indexes provides the number of tensor components. The antisymmetry does not reduce the number of components, for Rppik ¼ 0 and Rpipk ¼ 0, and the first Bianchi identity is satisfied. Considering the symmetryRpipk ¼ Rpkpi the number of components is reduced by half, whereby there are N 2 N � 1ð Þ N � 2ð Þ independent and non-null components. Admitting the four indexes different there are, for example, the components R1234,R2314,R3124. With methodology analogous to the previous case, it is verified that the indexes p, i, j, k can be selected in N N � 1ð Þ N � 2ð Þ N � 3ð Þ modes. Considering the antisymmetries Rpijk ¼ �Ripjk and Rpijk ¼ �Rpikj, the combination of indexes is reduced to N 4 N � 1ð Þ N � 2ð Þ N � 3ð Þ modes. The symmetry Rpijk ¼ Rjkpi reduces to half these combinations, then having N 8 N � 1ð Þ N � 2ð Þ N � 3ð Þ modes. The first Bianchi identity is given by Rpikj þ Rpjki þ Rpkij ¼ 0 ) Rpikj ¼ � Rpjki þ Rpkij � � that shows that the different combinations of the indexes are related among themselves, for a component can be expressed in terms of the other two. Therefore, the total number of combinations of indexes is reduced in 2 3 , and the total number of non-null independent components for this case is 2 3 N 8 N � 1ð Þ N � 2ð Þ N � 3ð Þ. The consideration of all the cases that were analyzed leads to 242 5 Riemann Spaces 0þ N 2 N � 1ð Þ þ N 2 N � 1ð Þ N � 2ð Þ þ N 12 N � 1ð Þ N � 2ð Þ N � 3ð Þ whereby there are N 2 12 N2 � 1� � independent and non-null components for the tensor Rpijk. The expressions that provide the Christoffel symbols for the orthogonal coordi- nate systems are Γ kij ¼ 0 Γ kii ¼ � 1 2gkk ∂gii ∂xk Γ iij ¼ ∂ ‘n ffiffiffiffiffi gii p� � ∂xj Γ iii ¼ ∂ ‘n ffiffiffiffiffi gii p� � ∂xi and with expression (5.2.20) that defines the Riemann–Christoffel curvature tensor with variance (0, 4) it results for the components of this tensor, where the indexes p, i, j, k indicate no summation: – Four different indexes Rpijk ¼ 0 ð5:2:21Þ – i ¼ j and the other three indexes different Rpiik ¼ ffiffiffiffiffigiip ∂2 ffiffiffiffiffigiip∂xp∂xk � ∂ ffiffiffiffiffi gii p ∂xp ∂ ‘n ffiffiffiffiffiffigppp� � ∂xk � ∂ ffiffiffiffiffi gii p ∂xk ∂ ‘n ffiffiffiffiffiffi gkk p� � ∂xp 0@ 1A ð5:2:22Þ – p ¼ k, i ¼ j, p 6¼ i (two different indexes) Rkiik ¼ ffiffiffiffiffigiip ffiffiffiffiffiffigkkp ∂∂xk 1ffiffiffiffiffiffigkkp∂ ffiffiffiffiffi gii p ∂xk � � þ ∂ ∂xi 1ffiffiffiffiffi gii p ∂ ffiffiffiffiffiffi gkk p ∂xi � � þ 1 gmm ∂ ffiffiffiffiffi gii p ∂xm ∂ ffiffiffiffiffiffi gkk p ∂xm � ð5:2:23Þ withm 6¼ pandm ¼ i to fulfill the condition of having two different pairs of indexes, with the summation carried out only for the index m. Table 5.1 shows four Riemann spaces EN and the independent and non-null components of tensor Rpijk. For the Riemann space E1 the only component of tensor Rpijk is R1111, which by means of its antisymmetry will always be null. Expression N 2 12 N2 � 1� � proves this nullity. It is concluded that this tensor express only the internal properties of the space and not the way how this space is embedded in the Riemann spaces EN, N > 1, for this characteristic verifies that in E1 a curved line has null curvature, seen that R1111 ¼ 0. 5.2 The Curvature Tensor 243 For the Riemann space E2 the tensor Rpijk has null components when three or more indexes are equal. Only one component cannot be null: R1212. By means of the symmetry and the antisymmetry it is verified that R1212 ¼ �R2112 ¼ �R1221 ¼ R2121. This component is given by R1212 ¼ 1 2 2 ∂2g12 ∂x1∂x2 � ∂ 2 g11 ∂x2∂x2 � ∂ 2 g22 ∂x1∂x1 ! þ gq‘ Γ q12Γ ‘12 � Γ q11Γ ‘22 � � ð5:2:24Þ For the Riemann space E3 the six components of tensor Rpijk are: – Three components with two repeated indexes R1212 R1313 R2323 – Three components with only one index repeated (three indexes are different) R1213 R1223(¼R2123) R1323(¼R3132) For the Riemann space E4 there are 21 non-null components of tensor R‘ijkwhich are: – Six components with two repeated indexes R1212 R1313 R1414 R2323 R2424 R3434 – Twelve components with only one index repeated (three indexes are different) R1213 R1214 R1223 R1224 R1314 R1323 R1334 R1424 R1434 R2324 R2334 R2434 Table 5.1 Independent and non-null components of tensor Rpijk Dimension of space EN 2 3 4 5 Number of components 16 81 256 625 Independent and non-null components of Rpijk 1 6 20 50 Kinds of components R1212 Riþ1 iþ2 jþ1 jþ2 Rpipi,Rppik, Rpijk Rpipi,Rppik, Rpijk 244 5 Riemann Spaces – Three components with only one index repeated (three indexes are different) R1234 R1324 R1423 having that R1234 þ R1423 � R1324 ¼ 0, then there are 20 independent non-null components. The non-null components of tensor R‘ijk for the Riemann space E5 are: – Ten components with two repeated indexes R1212 R1313 R1414 R1515 R2323 R2424 R2525 R3434 R3535 R4545 – Thirty components with only one index repeated (three indexes are different) R1213 R1214 R1215 R1314 R1315 R1415 R2123 R2124 R2125 R2324 R2325 R2425 R3132 R3134 R3135 R3234 R3235 R3435 R4142 R4143 R4145 R4243 R4245 R4345 R5152 R5153 R5154 R5253 R5254 R5354 – Ten components in which all the indexes are different R1234 R1235 R1245 R1345 R2345 R1324 R1325 R1425 R1435 R2435 5.2.11 Classification of Spaces As a function of the values assumed by the Riemann–Christoffel tensors the spaces are classified as: (a) flat: R ‘ijk ¼ Rijkm ¼ 0; (b) curved space R ‘ijk 6¼ 0; Rijkm 6¼ 0. The condition Ri‘jk ¼ Rijkm ¼ 0 indicates that the space is flat with the compo- nents of its metric tensor gij being constant. If the metric ds 2 ¼ gijdxidxj is definite positive, i.e., gij > 0, this space is Euclidian, then it is possible to carry out a linear transformation of the coordinates xi to the coordinates xi for which the result is gij ¼ δ ij , so the metric is ds2 ¼ δ ij dxidxj ¼ dx1dx1 þ dx2dx2 þ � � � þ dxmdxm ð5:2:25Þ 5.2 The Curvature Tensor 245 The vectors of base ei of this new coordinate system X i form a set of orthogonal directions, thus δ ij ¼ ei � ei, and define an Euclidian space EM. Consider the Rie- mann space EN with the coordinates x i, i ¼ 1, 2, . . .N, EN � EM, with M > N, which coordinates are xk, k ¼ 1, 2, . . . ,M. Let the functions M be independent in terms of the coordinates xk, so as to have the metric ds2 ¼ gijdxidxj ¼ dxk � � 2 ) gijdxidxj ¼ dxkdxk By means of the transformation law for coordinates it follows that dxk ¼ ∂x k ∂xi dxi dxk ¼ ∂x k ∂xj dxj gijdx idxj ¼ ∂x k ∂xi dxi � � ∂xk ∂xj dxj � � ) gij � ∂xk ∂xi ∂xk ∂xj � � dxidxj ¼ 0 As dxi and dxj are arbitrary, provides gij ¼ ∂xk ∂xi ∂xk ∂xj that defines N 2 N þ 1ð Þ independent differential equations as a function of M unknowns xk. In this case M < N 2 N þ 1ð Þ is the condition in order to have EN � EM. For N ¼ 1 the result is M � N2. 5.3 Riemann Curvature 5.3.1 Definition The study of the Riemann space EN is carried by means of the definition of the Riemann K curvature, which is more effective for the formulations of analyses than the Riemann–Christoffel curvature tensor Rpijk, for it considers the directions of the space. For establishing a general formulation, valid for the Riemann spaces EN with undefined metric, with the unit vectors ui and vi, linearly independents, defined in a point xi2EN , and the expression wi ¼ aui þ bvi ð5:3:1Þ that defines a coplanar vector with these two unit vectors, where a, b, are scalars that assume arbitrary values. The elementary displacements in the directions defined by the vectors wi determine a plane π that contains the point xi2EN . 246 5 Riemann Spaces It is admitted that ui and vi define coplanar vectors wi ¼ a1ui þ b1vi ð5:3:2Þ ri ¼ a2ui þ b2vi ð5:3:3Þ where a1, b1, a2, b2 are scalars, and putting ε uð Þ ¼ 1 and ε vð Þ ¼ 1 as functional indicators of these unit vectors, and having wi and ri vectors mutually orthogonal it follows that ε wð Þ ¼ gk‘wkw‘ ¼ a21uku‘ þ b21vkv‘ ¼ a21ε uð Þ þ b21ε vð Þ ε rð Þ ¼ gk‘rkr‘ ¼ a22ε uð Þ þ b22ε vð Þ and with the condition of orthogonality gk‘w kr‘ ¼ ε uð Þa1a2 þ ε vð Þb1b2 ¼ 0 whereby ε wð Þε rð Þ ¼ a21ε uð Þ þ b21ε vð Þ � � a22ε uð Þ þ b22ε vð Þ � �� ε uð Þa1a2 þ ε vð Þb1b2½ �2 ¼ ε uð Þε vð Þ a1b2 � a2b1ð Þ2 ð5:3:4Þ As the functional indicators assume the values 1: a1b2 � a2b1 ¼ 1 ð5:3:5Þ whereby ε uð Þε vð Þ ¼ ε wð Þε rð Þ ð5:3:6Þ Consider two orthogonal unit vectors u and v that determine the plane π that contains the point xi2EN , thus the Riemann curvature is defined by K ¼ ε uð Þε vð ÞRk‘mnukv‘umvn ð5:3:7Þ 5.3.2 Invariance For the other pair of orthogonal vectors w and r coplanar with u and v, there is in an analogous way for the Riemann curvature eK ¼ ε wð Þε rð ÞRk‘mnwkr‘wmrn 5.3 Riemann Curvature 247 and with expressions (5.3.4)–(5.3.6) it follows that ε uð Þε vð Þ a1b2 � a2b1ð ÞRk‘mnwkr‘wmrn ¼ ε uð Þε vð ÞRk‘mnukv‘umveK ¼ K thus the Riemann curvature does not depend on the pair of unit vectors used to define it, then K is an invariant. 5.3.3 Normalized Form The obtaining of an expression for the Riemann curvature can be carried out admitting that the Riemann space EN is isotropic, in which the isotropic tensor is defined by Tij‘m ¼ Agijg‘m þ Bgi‘gjm þ Cgimgj‘ where A,B,C are scalars that depend on the point xi2EN . Assuming that tensor Tij‘m is the curvature tensor Rij‘m the result is Rij‘m ¼ Agijg‘m þ Bgi‘gjm þ Cgimgj‘ ð5:3:8Þ and the antisymmetry of tensor Rij‘m allows writing Riiii ¼ 0, Riijj ¼ 0, Rii‘m ¼ Rij‘‘ ¼ 0, and with expression (5.3.8) it follows that Riiii ¼ Agiigii þ Bgiigii þ Cgiigii ¼ g2ii Aþ Bþ Cð Þ ¼ 0 Aþ Bþ C ¼ 0 ) Bþ C ¼ �A Riijj ¼ Agiigjj þ Bgijgij þ Cgijgij ¼ Agiigjj þ Bþ Cð Þgijgij ¼ A giigjj � g2ij � � ¼ 0 ð5:3:9Þ Rij‘‘ ¼ Agijg‘‘ þ Bgi‘gj‘ þ Cgi‘gj‘ ¼ A gijg‘‘ � gi‘gj‘ � � ¼ 0 ð5:3:10Þ The minors of det gi‘ cannot all be simultaneously null, then in expressions (5.3.9) and (5.3.10) the result is A ¼ 0 and B ¼ �C, whereby Rij‘k ¼ B gi‘gjm � gimgj‘ � � ð5:3:11Þ 248 5 Riemann Spaces Let K ¼ ε uð Þε vð ÞRk‘mnukv‘umvn or K ¼ Rk‘mnukv‘umvn ð5:3:12Þ and substituting expression (5.3.11) it is concluded that B ¼ K is the Riemann curvature in xi2EN , so Rij‘m ¼ K gi‘gjm � gimgj‘ � � ð5:3:13Þ The expression of the Riemann curvature for the isotropic space EN, withN > 2, in terms of the generalized Kronecker delta and the Ricci pseudotensor εi1i2...imimþ1...inεp1p2...pmpmþ1...pn ¼ N � 2ð Þ!δp1p2...pni1i2...in takes the form Rij‘k ¼ K εi1i2...imimþ1...inε p1p2...pmpmþ1...pn N � 2ð Þ! ¼ Kδ p1p2...pn i1i2...in ð5:3:14Þ The normalized Riemann curvature is established admitting that the vectors u and v form an angle α and define a tangent plane π in point xi2EN . The norm of the vector perpendicular to this plane is given by u� vk k2 ¼ uk k2 vk k2 sin 2α and with the square of the dot product of these two vectors it follows that u � vð Þ2 ¼ uk k2 vk k2 cos 2α ¼ cos 2α u� vk k2 ¼ uk k2 vk k2 1� cos 2α� � ¼ uk k2 vk k2 � u � vð Þ2 In terms of the components of these vectors uk k2 ¼ gkmukum vk k2 ¼ g‘nv‘vn then u� vk k2 ¼ gkmukumg‘nv‘vn � gknukvngm‘umv‘ ¼ ukv‘umvn gkmg‘n � gkngm‘ð Þ Expression (5.3.12) in its normalized form is 5.3 Riemann Curvature 249 K xi; u; v � � ¼ Rk‘mnukv‘umvn gkmg‘n � gkngm‘ð Þukv‘umvn ð5:3:15Þ or K xi; u; v � � ¼ Rk‘mnAk‘Amn gkmg‘n � gkngm‘ð ÞAk‘Amn ð5:3:16Þ where Ak‘ ¼ ukv‘,Amn ¼ umvn represent the plane π defined by the vectors u, v. This expression highlights that the Riemann curvature K(xi;u, v) of the Riemann space EN relative to the plane π defined by the vectors u and v depends on the point xi2π � EN . In the numerator of expression (5.3.15) the product Rk‘mnu kv‘umvn is an invariant. Putting Gk‘mn ¼ gkmg‘n � gkngm‘ it is verified thatGk‘mn is a tensor, for it is obtained by means of algebraic operations with the metric tensor. The permutation of the tensor indexes Gk‘mn shows that this tensor has the same properties of symmetry and antisymmetry as tensor Rk‘mn. For an orthogonal coordinate system exists gij ¼ 0 for i 6¼ j, and the non-null compo- nents of this tensor are given by Gijij ¼ giigjj, where the indexes do not indicate summation. The inner product Gk‘mnu kv‘umvn generates a scalar, then expression (5.3.15) represents an invariant, highlighting the demonstration that eK ¼ K. 5.4 Ricci Tensor and Scalar Curvature The Riemann–Christoffel curvature tensor Rpijk allows obtaining tensors of lower order by means of theirs various contractions. To obtain a non-null tensor first an index of a pair of indexes are contracted with an index of another pair of indexes, being possible the contractions: 1–3; 1–4; 2–3; 2–4. The contraction of this tensor generates the Ricci tensor, thus the multiplying of tensor Rpijk by g mp provides gmpRpijk ¼ gmpgp‘R ‘ijk ¼ δm‘ R ‘ijk ¼ Rmijk and with the contraction m ¼ k the result is Rkijk ¼ Rij 250 5 Riemann Spaces Then the Riemann–Christoffel curvature tensor with variance (1, 3) provides two Ricci tensors, one of variance (0, 2) and another of variance (1, 1). The second contraction gives a scalar with important properties, called scalar curvature. The Ricci tensor is essentially the only contraction of the Riemann–Christoffel tensor. 5.4.1 Ricci Tensor with Variance (0, 2) The contraction of the curvature tensor R ‘ijk ¼ ∂Γ ‘ik ∂xj � ∂Γ ‘ ij ∂xk þ ΓmikΓ ‘mj � Γmij Γ ‘mk in indexes ‘ ¼ k provides Rij ¼ ∂Γ ‘ i‘ ∂xj � ∂Γ ‘ ij ∂x‘ þ Γmi‘Γ ‘mj � Γmij Γ ‘m‘ ð5:4:1Þ In determinants form the result is Rij ¼ ∂ ∂xj ∂ ∂x‘ Γ ‘ij Γ ‘ i‘ þ Γmi‘ Γ m ij Γ ‘m‘ Γ ‘ mj ð5:4:2Þ and with the expressions Γ ‘i‘ ¼ ∂ ‘n ffiffiffi g p� � ∂xi Γ ‘m‘ ¼ ∂ ‘n ffiffiffi g p� � ∂xm it follows that Rij ¼ ∂∂xj ∂ ‘n ffiffiffi g p� � ∂xi � � ∂Γ ‘ ij ∂x‘ þ Γmi‘Γ ‘mj � Γmij ∂ ‘n ffiffiffi g p� � ∂xm whereby for the Ricci tensor with variance (0, 2) the result is Rij ¼ ∂2 ‘n ffiffiffi g p� � ∂xj∂xi � ∂Γ ‘ ij ∂x‘ þ Γmi‘Γ ‘mj � Γmij ∂ ‘n ffiffiffi g p� � ∂xm or 5.4 Ricci Tensor and Scalar Curvature 251 Rij ¼ 1 2 ∂2 ‘ngð Þ ∂xj∂xi � ∂Γ ‘ ij ∂x‘ þ Γmi‘Γ ‘mj � 1 2 Γmij ∂ ‘ngð Þ ∂xm ð5:4:3Þ If g < 0 it is enough to change g for �g in the expression (5.4.3). The permutation of indexes j $ i leads to Rji ¼ 1 2 ∂2 ‘ngð Þ ∂xi∂xj � ∂Γ ‘ ji ∂x‘ þ Γmj‘Γ ‘mi � 1 2 Γmji ∂ ‘ngð Þ ∂xm As the Christoffel symbols are symmetric and the order of differentiation in the first term of the previous expression is independent of the sequence in which it is carried out, it is concluded that the Ricci tensor Rij is symmetric, so it has N 2 N þ 1ð Þ distinct components. The contractions that can be carried out in tensor R‘ijk are: R ‘ ‘jk,R ‘ i‘k,R ‘ ij‘. Con- sidering the antisymmetry of curvature tensor R ‘ijk ¼ �R ‘ikj and with k ¼ ‘ the result isR ‘ij‘ ¼ �R ‘i‘j, whereby Rij ¼ �R ‘i‘j. The contraction R‘i‘k generates the Ricci tensor Rij with sign changed, then it is enough to consider only the contraction R ‘ ij‘ to obtain tensor Rij, which contains components independent of R ‘ ijk in the more adequate form of a symmetric tensor. The contraction of tensor R‘ijk in the indexes i ¼ ‘ is given by R ‘‘jk ¼ ∂Γ ‘‘k ∂xj � ∂Γ ‘ ‘j ∂xk þ Γm‘kΓ ‘mj � Γm‘jΓ ‘mk and with the expressions Γ ‘‘k ¼ ∂ ‘n ffiffiffi g p� � ∂xk Γ ‘‘j ¼ ∂ ‘n ffiffiffi g p� � ∂xj the result is R ‘‘jk ¼ ∂ ∂xj ∂ ‘n ffiffiffi g p� � ∂xk � � ∂ ∂xk ∂ ‘n ffiffiffi g p� � ∂xj � þ Γm‘kΓ ‘mj � Γm‘jΓ ‘mk The permutation of indexes ‘ $ m in the last term and the symmetry of the Christoffel symbols allow writing R ‘‘jk ¼ ∂ ∂xj ∂ ‘n ffiffiffi g p� � ∂xk � � ∂ ∂xk ∂ ‘n ffiffiffi g p� � ∂xj � þ Γm‘kΓ ‘mj � Γ ‘mjΓm‘k and as 252 5 Riemann Spaces Free ebooks ==> www.Ebook777.com ∂ ∂xj ∂ ‘n ffiffiffi g p� � ∂xk � ¼ ∂ ∂xk ∂ ‘n ffiffiffi g p� � ∂xj � then R ‘‘jk ¼ Rjk ¼ 0 It is concluded that the contraction of the Riemann–Christoffel curvature tensor R‘ijk in the indexes ‘ ¼ i generates the null tensor. 5.4.2 Divergence of the Ricci Tensor with Variance Ricci (0, 2) The calculation of the divergence of tensor R‘ijk is carried out considering the second Bianchi identity ∂‘R ‘ ijk þ ∂jR ‘ik‘ þ ∂kR ‘i‘j ¼ 0 in which the contraction of the indexes ‘ ¼ k provides ∂‘R k ijk þ ∂jRkik‘ þ ∂kRki‘j ¼ 0 ) ∂‘Rij þ ∂jRi‘ þ divRki‘j ¼ 0 whereby divRki‘j ¼ � ∂‘Rij þ ∂jRi‘ � � and with the ordination of the indexes divR ‘ijk ¼ � ∂jRik þ ∂kRij � � ð5:4:4Þ 5.4.3 Bianchi Identity for the Ricci Tensor with Variance (0, 2) An identity analogous to the second Bianchi identity can be obtained for the Ricci tensor. Rewriting expression (5.2.15) ∂pR ‘ ikj þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ 0 5.4 Ricci Tensor and Scalar Curvature 253 www.Ebook777.com http://www.ebook777.com and with the relations g ‘i Rkj ¼ R ‘ikj g ‘i Rkp ¼ R ‘ikp g ‘i Rpj ¼ R ‘ipj it follows that g ‘i ∂pRkj ¼ ∂pR ‘ikj g ‘i ∂iRkp ¼ ∂iR ‘ikp g ‘i ∂kRpj ¼ ∂kR ‘ipj The sum of these three expressions provides g ‘i ∂pRkj þ ∂iRkp þ ∂kRpj � � ¼ ∂pR ‘ikj þ ∂iR ‘ikp þ ∂kR ‘ipj As the term to the right is the second Ricci identity it results in ∂pRkj þ ∂iRkp þ ∂kRpj ¼ 0 The changes of the indexes j ! i, k ! j, p ! k allow the ordination of the same, then ∂kRij þ ∂iRjk þ ∂jRki ¼ 0 ð5:4:5Þ that is called Bianchi identity for the Ricci tensor of covariant components. 5.4.4 Scalar Curvature The multiplying of the Ricci tensor Rij by the conjugate metric tensor g ij provides R ¼ gijRij ð5:4:6Þ that defines the scalar curvature, which is the trace of the Ricci tensor, also called Ricci curvature or invariant curvature of the Riemann space EN. 5.4.5 Geometric Interpretation of the Ricci Tensor with Variance (0, 2) Let the Riemann curvature K xi; u; v � � ¼ Rk‘mnukv‘umvn gkmg‘n � gkngm‘ð Þukv‘umvn 254 5 Riemann Spaces where u, v are orthogonal unit vectors, the result thereof is gkmg‘nu kv‘umvn ¼ gkmukum � � g‘nv ‘vn � � gkngm‘u kv‘umvn ¼ gknukvn � � gm‘u mv‘ � � but gkmu kum ¼ g‘nv‘vn ¼ 1 gknukvn ¼ gm‘umv‘ ¼ 0 then Kuv ¼ K xi; u; v � � ¼ Rk‘mnukv‘umvn 1� 1� 0 ¼ Rk‘mnu kv‘umvn where the notation Kuv is adopted by convenience of graphic representation. If the unit vectors u, v are linearly dependent, the result is K ¼ 0. The summation of all the N components of vector u is given by XN vj¼1 Kuv ¼ XN vj¼1 Rk‘mnu kv‘umvn ¼ ukum XN vj¼1 Rk‘mnv ‘vn but XN vj¼1 v‘vn ¼ g‘n whereby the contraction R‘i‘j generates the Ricci tensor Rij with the signchanged, then XN vj¼1 Kuv ¼ �ukumg‘nRk‘mn ¼ �ukumRnkmn ¼ �ukumRkm Putting Ku ¼ XN vj¼1 Kuv ¼ �ukumRkm ð5:4:7Þ where Ku is the sum of the Riemann curvature for the space EN determined by the components of vector u and each N � 1ð Þ directions which are mutually orthogonal to them. This expression is independent of these directions and defines the mean curvature of EN in the direction of this vector. 5.4 Ricci Tensor and Scalar Curvature 255 In expression (5.4.7) when carrying out the summation on the N directions mutually orthogonal, it follows that XN ui¼1 Ku ¼ � XN ui¼1 ukumRkm XN ui¼1 ukum ¼ gkm XN ui¼1 Ku ¼ �gkmRkm ¼ �R ð5:4:8Þ Expression (5.4.8) shows that the sum of the mean curvatures in the Riemann space EN for mutually orthogonal directions are independent of the directions defined by the vectors u, v, being equal to the scalar curvature. 5.4.6 Eigenvectors of the Ricci Tensor with Variance (0, 2) The Ricci tensor Rij is symmetric and has in each point of the Riemann space EN a system of linearly independent equations that define principal directions (eigenvectors). Let the Riemann curvature K xi; u; v � � ¼ Rk‘mnukv‘umvn gkmg‘n � gkngm‘ð Þukv‘umvn where the vectors are orthogonal and only v is a unit vector, so gkmg‘nu kv‘umvn ¼ gkmukum � � g‘nv ‘vn � � gkngm‘u kv‘umvn ¼ gknukvn � � gm‘u mv‘ � � ¼ 0 then K xi; u; v � � ¼ Rk‘mnukv‘umvn gkmu kumð Þ g‘nv‘vnð Þ but as v is a unit vector the result is g‘nv ‘vn ¼ 1 ) v‘vn ¼ g‘n 256 5 Riemann Spaces whereby K xi; u; v � � ¼ �Rk‘mng‘nukum gkmu kum thereof Ku ¼ K xi; u; v � � ¼ �Rkmukum gkmu kum ð5:4:9Þ is the normalized mean curvature, where the index indicates that u is not unit vector. The calculation of the eigenvalues is carried out by means of the equations system Rkm þ Kugkmð Þukum ¼ 0 with extreme values given by the condition ∂ ∂uk Rkm þ Kugkmð Þukum � � ¼ 0 which developed stays 2 Rkm þ Kugkmð Þum þ ∂Rkm ∂uk ukum þ ∂Ku ∂uk gkmu kum ¼ 0 and as the Ricci tensor Rij does not depend on vector u k the result is 2 Rkm þ Kugkmð Þum þ ∂Ku ∂uk gkmu kum ¼ 0 For the extreme values of Ku the result is ∂Ku ∂xk ¼ 0, whereby the equations system Rkm þ Kugkmð Þum ¼ 0 allows determining the principal directions (eigenvectors) of the Ricci tensor Rij. 5.4.7 Ricci Tensor with Variance (1, 1) The Ricci tensor in terms of its mixed components is given by Rij ¼ gimRmj ð5:4:10Þ 5.4 Ricci Tensor and Scalar Curvature 257 An important expression that relates the Ricci tensor with variance (1, 1) with the derivative of the scalar curvature can be obtained by means of the second Bianchi identity ∂pR ‘ ijk þ ∂jR ‘ikp þ ∂kR ‘ipj ¼ 0 where with the antisymmetry R ‘ikp ¼ �R ‘ipk the result is ∂pR ‘ ijk � ∂jR ‘ipk þ ∂kR ‘ipj ¼ 0 The contraction of these tensors in indexes ‘ ¼ k provides ∂pRij � ∂jRip þ ∂kRkipj ¼ 0 Multiplying by gip it follows that gip∂pRij � gip∂jRip þ gip∂kRkipj ¼ 0 ∂pg ipRij � ∂jgipRip þ ∂kgipRkipj ¼ 0 ∂pR p j � ∂R ∂xj þ ∂kRkj ¼ 0 ) ∂R ∂xj ¼ ∂pRpj þ ∂kRkj The change of the dummy indexes p ! k provides ∂R ∂xj ¼ 2∂kRkj whereby ∂kR k j ¼ 1 2 ∂R ∂xj ð5:4:11Þ For the Riemann space EN, with N > 2, multiplying expression (5.4.5) by g ij the result is gij∂kRij þ gij∂iRjk þ gij∂jRki ¼ 0 ) ∂kgijRij þ ∂igijRjk þ ∂jgijRki ¼ 0 and having curvature R a scalar function at its partial derivative is equal to its covariant derivative, then ∂R ∂xk þ ∂iR ik þ ∂jR jk ¼ 0 258 5 Riemann Spaces and with the change of indexes j ! i the result is ∂R ∂xk þ 2∂iR ik ¼ 0 and with ∂iR i k ¼ 1 2 ∂R ∂xk it follows that ∂R ∂xk þ 2 � 1 2 ∂R ∂xk ¼ 0 ) ∂R ∂xk ¼ 0 then the scalar curvature is constant for this kind of space. The purpose of the suppositionN > 2will be clarified by expression (5.6.10), obtained when analyzing the scalar curvature in the Riemann space E2. Exercise 5.1 For the tensorial expression T ij ¼ Rij þ δ ij αRþ βð Þ, where α, β are scalars, calculate the value of α so that the covariant derivative ∂iT ij is null. The null covariant derivative ∂iT ij is given by ∂iT i j ¼ ∂iR ij þ ∂i δ ij αRþ βð Þ h i having ∂iδ ij ¼ 0 it follows that ∂iT i j ¼ ∂iR ij þ α∂iR ¼ 0 With the expression (5.4.11) ∂iR i j ¼ 1 2 ∂R ∂xj ) ∂iT ij ¼ 1 2 þ α � � ∂R ∂xj ¼ 0 for ∂iR ¼ ∂R∂xj, and as this derivative assumes any values the result is α ¼ �12. 5.4.8 Notations In Table 5.2, in which the Tulio Levi-Civita notation was inserted, there is a compilation of the evolution of the notation for the Riemann–Christoffel curvature tensors and for the Ricci tensor. The notations that make use of (,) or (;) seek to 5.4 Ricci Tensor and Scalar Curvature 259 indicate the properties of symmetry and antisymmetry of the Riemann–Christoffel tensors. In the case of using (.) it indicates the index, or the position and the index that will be lowered or raised. The only difference between the two notations of Christoffel and Bianchi is the change of the point and comma (;) for the comma (,). Currently these two forms of spelling were abandoned. It is stressed that several authors have opted for different positioning of the indexes. The Weyl notation, with the change of the letter F for R (Riemann), was the one that became consecrated in the current literature. Exercise 5.2 In a coordinates system let Γ ijk ¼ δ ij ∂ϕ∂xk þ δ ik ∂ψ∂xj, where ϕ,ψ are functions of position. Calculate: (a) Rijk‘; (b) Rjk for ψ ¼ �‘n aixið Þ. (a) Substituting the expression Γ ijk ¼ δ ij ∂ϕ ∂xk þ δ ik ∂ψ ∂xj in the expression of the Riemann–Christoffel curvature tensor Rijk‘ ¼ ∂Γ ij‘ ∂xk � ∂Γ i jk ∂x‘ þ Γ irkΓ rj‘ � Γ ir‘Γ rjk Table 5.2 Notations for the Riemann–Christoffel curvature tensors and Ricci tensor Author Riemann–Christoffel curvature tensor Ricci tensor Mixed variance components (1, 3) Covariant components (0, 4) Brillouin Rij; k‘ Rij, k‘ Rj‘ ¼ X m Rmj,m‘ Appe-Thiry Ri j k‘ Rijk‘ Rjk ¼ X m Rm jkm Weyl Fijk‘ Fijk‘ Rj‘ ¼ X m Fmjm‘ Eddington- Becquerel Bij k‘ Bjk‘i Gjk ¼ X m Bmjmk Galbrun Rij‘k Rij‘k Rjk ¼ X m Rmjmk Juvet R ij ‘k Rji‘k Rjk ¼ X m R mj mk Cartan Rij;‘k Rji,‘k Rj‘ ¼ X m Rmj‘m Christoffel and Bianchi ( ji; k‘) ( ji, k‘) – Levi-Civita {ji, k‘} ( ji, k‘) αj‘ 260 5 Riemann Spaces it follows that Rijk‘ ¼ ∂ ∂xk δ ij ∂ϕ ∂x‘ þ δ i‘ ∂ψ ∂xj � � � ∂ ∂x‘ δ ij ∂ϕ ∂xk þ δ ik ∂ψ ∂xj � � þ δ ij ∂ϕ ∂xk þ δ ik ∂ψ ∂xr � � δ rj ∂ϕ ∂x‘ þ δ r‘ ∂ψ ∂xj � � � δ ir ∂ϕ ∂x‘ þ δ i‘ ∂ψ ∂xr � � δ rj ∂ϕ ∂xk þ δ rk ∂ψ ∂xj � � ¼ δ ij ∂2ϕ ∂xk∂x‘ þ δ i‘ ∂2ψ ∂xk∂xj � δ ij ∂2ϕ ∂x‘∂xk � δ ik ∂2ψ ∂x‘∂xj þ δ ij δ rj ∂ϕ ∂xk ∂ϕ ∂x‘ þ δ ij δ r‘ ∂ϕ ∂x‘ ∂ψ ∂xj þδ ikδ rj ∂ψ ∂xr ∂ϕ ∂x‘ þ δ ikδ r‘ ∂ψ ∂xr ∂ψ ∂xj � δ irδ rj ∂ϕ ∂x‘ ∂ϕ ∂xk � δ irδ rk ∂ϕ ∂x‘ ∂ψ ∂xj � δ i‘δ rj ∂ψ ∂xr ∂ϕ ∂xk �δ i‘δ rk ∂ψ ∂xr ∂ψ ∂xj R ijk‘ ¼ δ ij ∂ϕ ∂xk ∂ϕ ∂x‘ þ δ ik ∂ψ ∂xj ∂ψ ∂x‘ þ δ ik ∂ϕ ∂x‘ ∂ψ ∂xj þ δ i‘ ∂ϕ ∂xk ∂ψ ∂xj � δ ij ∂ϕ ∂x‘ ∂ϕ ∂xk �δ i‘ ∂ψ ∂xj ∂ψ ∂xk � δ i‘ ∂ϕ ∂xk ∂ψ ∂xj � δ ik ∂ϕ ∂x‘ ∂ψ ∂xj þ δ ij ∂2ϕ ∂x‘∂xk þ δ i‘ ∂2ψ ∂xj∂xk �δ ij ∂2ϕ ∂xk∂x‘ � δ ik ∂2ψ ∂xj∂x‘ Rijk‘ ¼ δ ik ∂ψ ∂xj ∂ψ ∂x‘ � ∂ 2ψ ∂xj∂x‘ ! � δ i‘ ∂ψ ∂xj ∂ψ ∂xk � ∂ 2ψ ∂xj∂xk ! then Rijk‘ only depends on the function ψ . (b) For ψ ¼ �‘n aixið Þ the partial derivatives result ∂ψ ∂xj ¼ � aj aixi ) ∂ 2ψ ∂xj∂x‘ ¼ aja‘ aixið Þ2 ∂ψ ∂x‘ ¼ � a‘ aixi ) ∂ψ ∂xj ∂ψ ∂x‘ ¼ aja‘ aixið Þ2 and substituting this derivatives in the expression obtained in item (a) it follows that 5.4 Ricci Tensor and Scalar Curvature 261 Rijk‘ ¼ δ ik ∂ψ ∂xj ∂ψ ∂x‘ � ∂ 2ψ ∂xj∂x‘ ! � δ i‘ ∂ψ ∂xj ∂ψ ∂xk � ∂ 2ψ ∂xj∂xk ! Rijk‘ ¼ δ ik aja‘ aixið Þ2 � aja‘ aixið Þ2 " # � δ i‘ ajak aixið Þ2 � ajak aixið Þ2 " # ¼ 0 whereby Rijki ¼ Rjk ¼ 0 Q:E:D: 5.5 Einstein Tensor The tensor Rijk‘, the second Bianchi identity, the Ricci tensor Rij and the scalar curvature R allow obtaining a second-order tensor with peculiar characteristics. Let the second Bianchi identity ∂mRijk‘ þ ∂kRij‘m þ ∂‘Rijmk ¼ 0 and with the antisymmetry of the Riemann–Christoffel curvaturetensor Rijk‘ ∂mRijk‘ � ∂kRijm‘ � ∂‘Rjimk ¼ 0 and multiplying by gi‘ and gjk it follows that gi‘gjk∂mRijk‘ � gi‘gjk∂kRijm‘ � gi‘gjk∂‘Rjimk ¼ 0 gjk∂mR ‘ jk‘ � gjk∂kR ‘jm‘ � gi‘∂‘Rkimk ¼ 0 whereby in terms of the Ricci tensor gjk∂mRjk � gjk∂kRjm � gi‘∂‘Rim ¼ 0 The change of the dummy index ‘ ! k in the last term provides gjk∂mRjk � gjk∂kRjm � gik∂kRim ¼ 0 ) ∂mRjkjk � ∂kRjkjm � ∂kRikim ¼ 0 The contractions of the curvature tensors provide ∂mR� ∂kRkm � ∂kRkm ¼ 0 ) ∂mR ¼ 2∂kRkm 262 5 Riemann Spaces whereby ∂kR k m ¼ 1 2 ∂mR ð5:5:1Þ is the divergence of a tensor, which can be written under the form ∂k R k m � 1 2 δ kmR � � ¼ 0 ð5:5:2Þ where the terms in parenthesis define the Einstein tensor with variance (1, 1) Gkm ¼ Rkm � 1 2 δ kmR ð5:5:3Þ The Einstein tensor can be written as a function of its covariant components, so Gij ¼ gikGkj ¼ gik Rkj � 1 2 δ kj R � � ð5:5:4Þ thus Gij ¼ Rij � 1 2 gijR ð5:5:5Þ By means of this expression it is verified that the Einstein tensor is generated only by the metric tensor and the Ricci tensor. As Rij and gij are two symmetric tensors then Einstein tensor is symmetric. For the contravariant components of this tensor the result is Gij ¼ Rij � 1 2 gijR ð5:5:6Þ The divergence of the Einstein tensor is given by ∂iG j i ¼ ∂iR ij � δ ji 1 2 ∂iR ¼ ∂iR ij � 1 2 ∂jR but ∂iR i j ¼ 1 2 ∂jR then ∂iG i j ¼ 0 ð5:5:7Þ 5.5 Einstein Tensor 263 Thus for any Riemann space the divergence of the Einstein tensor is null, and with the contraction of this tensor it follows that Gii ¼ Rii � 1 2 δ ii R ¼ R� 1 2 NR G ¼ �1 2 N � 2ð ÞR ð5:5:8Þ For the Riemann space E2 it is verified that G ¼ 0. Exercise 5.3 Show that the tensor of the kind T ij ¼ Rij þ δ ij m, being m a scalar function, has the characteristics of an Einstein tensor. The divergence of this tensor given by ∂jT ij ¼ 0 stays ∂jT i j ¼ ∂jR ij þ δ ij∂jm ¼ ∂j R ij þ m � � ¼ 0 and with expression (5.4.11) ∂jR i j ¼ 1 2 ∂jR substituted in this expression ∂jT i j ¼ ∂j 1 2 Rþ m � � ¼ 0 ) 1 2 Rþ m ¼ k1 ) m ¼ k1 � 1 2 R where k1 is a constant. The substitution of this expression in the expression of tensor Tij provides T ij ¼ Rij � δ ij 1 2 Rþ k2 � � where k2 ¼ �k1. Thus this tensor has the same characteristics of the Einstein tensor defined by expression (5.5.3). 5.6 Particular Cases of Riemann Spaces Some kinds of Riemann spaces will be analyzed in this item with specific charac- teristics that make them important: the Riemann space E2, the Riemann space with constant curvature, the Minkowski space, and the conformal space. 264 5 Riemann Spaces 5.6.1 Riemann Space E2 In the Riemann space E2 the Ricci tensor Rij is defined by its components Rij ¼ R11 R12 R21 R22 " # as R12 ¼ R21 and the metric tensor in matrix form is given by gij ¼ g11 g12 g21 g22 " # where g12 ¼ g21. The Ricci tensor written in terms of the Riemann–Christoffel curvature tensor with variance (0, 4), and considering the symmetry and the metric tensor is given by Rij ¼ gkpRpijk ¼ gpkRipkj and the development provides Rij ¼ g11Ri11j þ g12Ri12j þ g21Ri21j þ g22Ri22j whereby the result for component R11 is R11 ¼ g11R1111 þ g12R1121 þ g21R1211 þ g22R1221 As the tensor Rpijk is antisymmetric in the first two and the last two indexes, i.e., Rpijk ¼ �Ripjk and Rpijk ¼ �Rpikj it follows that R11 ¼ 0þ 0þ 0þ g22R1221 Let g ¼ detgij and G22 the cofactor of g22: g22 ¼ G 22 g ¼ g11 g whereby R11 ¼ g11 g �R1212ð Þ ) R11 g11 ¼ �R1212 g Proceeding in an analogous way for component R22: 5.6 Particular Cases of Riemann Spaces 265 R22 ¼ g11R2112 þ g12R2122 þ g21R2212 þ g22R2222 R22 ¼ g11R2112 þ 0þ 0þ 0 R22 ¼ �g11R1212 g11 ¼ G 11 g ¼ g22 g R22 ¼ �g22 g R1212 whereby R22 g22 ¼ �R1212 g For component R12, it follows that R12 ¼ g11R1112 þ g12R1122 þ g21R1212 þ g22R1222 ¼ 0þ 0þ g21R1212 þ 0 R12 ¼ g21R1212 g21 ¼ G 21 g ¼ g12 g R12 ¼ �g12 g R1212 thus R12 g12 ¼ �R1212 g and with the symmetries Rij ¼ Rji and gij ¼ gji the result for component R21 is R21 g21 ¼ �R1212 g The analysis developed shows that K ¼ R11 g11 ¼ R22 g22 ¼ R12 g12 ¼ R21 g21 ¼ �R1212 g These equalities indicate that in the Riemann space E2 the components of the Ricci tensor Rij are proportional to the components of the metric tensor gij and to its derivatives, and are independent of the directions considered. It is verified that the Riemann curvature does not vary with the orientation considered, then all the points 266 5 Riemann Spaces of the space E2 are isotropic. This, in general, is not valid for spaces with dimension N > 2. The scalar K in Riemann space E2 is called Gauß curvature. This analysis allows writing the components of the Ricci tensor as a function of the component R1212 and of the metric tensor, thus Rij ¼ �R1212 g gij ð5:6:1Þ 5.6.2 Gauß Curvature Expression (5.6.1) is valid only for the Riemann space E2. The knowledge of the properties of the surfaces in the Euclidian space E3 is not useful for understanding the properties of the Riemann spaces EN, with N > 3. For N ¼ 2 several simplifi- cations are admitted in the formulation of the expression of Rij, so the conclusions obtained for the Riemann space E2 cannot be generalized for the spaces of dimen- sions N > 3. The scalar curvature allows expressing the Riemann–Christoffel tensor Rpijk as a function of the components of the metric tensor. With the non-null components R1212, ¼ �R2121, ¼ �R1221 ¼ R2112, and the expression of the scalar curvature it follows that R ¼ gijRij ¼ �gijgij R1212 g ¼ �δ ii R1212 g ¼ �2 g R1212 ) R1212 ¼ �R 2 g and the development provides R1212 ¼ �R 2 g11 g12 g21 g22 ¼ �R2 g11g22 � g12g21ð Þ The other non-null components are obtained by means of the indexes in this expression, and considering the symmetry of tensor Rpijk it follows that R2121 ¼ �R 2 g22g11 � g21g12ð Þ R1221 ¼ �R 2 g12g21 � g11g22ð Þ R2121 ¼ �R 2 g21g12 � g22g11ð Þ 5.6 Particular Cases of Riemann Spaces 267 then Rijk‘ ¼ �R 2 gikgj‘ � gi‘gjk � � ð5:6:2Þ or Rijk‘ ¼ �K gikgj‘ � gi‘gjk � � ð5:6:3Þ The Gauß curvature, that in general depends on the coordinates of the point considered, is determined by K ¼ 1 2 R ð5:6:4Þ that can be obtained as a function of the Riemann–Christoffel curvature tensor with variance (0, 4), and with the Ricci pseudotensor for the Riemann space E2 εij ¼ ffiffiffigp eij εij ¼ eijffiffiffi g p and with the expression K ¼ R1212 g then Rijk‘ ¼ Kεijεk‘ ð5:6:5Þ The multiplication of both members of this expression by εijεk‘ provides εijεk‘Rijk‘ ¼ Kεijεk‘εijεk‘ and as εijε ij ¼ δ ii ¼ 2 thus K ¼ 1 4 Rijk‘ε ijεk‘ ð5:6:6Þ this expression shows that the Gauß curvature is an invariant. 268 5 Riemann Spaces 5.6.3 Component R1212 in Orthogonal Coordinate Systems For the orthogonal coordinate systems in the Riemann space EN expression (5.2.24) provides the component R1212 ¼ 1 2 2 ∂2g12 ∂x1∂x2 � ∂ 2 g11 ∂x2∂x2 � ∂ 2 g22 ∂x1∂x1 ! þ gq‘ Γ q12Γ ‘12 � Γ q11Γ ‘22 � � or more explicitly R1212 ¼�1 2 ∂2g11 ∂x2∂x2 þ ∂ 2 g22 ∂x1∂x1 ! þg11 Γ112Γ112�Γ111Γ122 � �þg22 Γ212Γ212�Γ211Γ222� � The Christoffel symbols for these coordinates systems are given by – i ¼ j ¼ k ) Γ kij ¼ Γ iii ¼ 12gii ∂gii ∂xj ) Γ111 ¼ 1 2g11 ∂g11 ∂x1 Γ222 ¼ 1 2g22 ∂g22 ∂x2 8>>><>>>: – i ¼ j 6¼ k ) Γ kij ¼ Γ kii ¼ � 12gkk ∂gii ∂xk ) Γ211 ¼ � 1 2g22 ∂g11 ∂x2 Γ122 ¼ � 1 2g11 ∂g22 ∂x1 8>>><>>>: – i ¼ k 6¼ j ) Γ kij ¼ Γ iij ¼ 12gii ∂gii ∂xj ) Γ112 ¼ 1 2g11 ∂g11 ∂x2 Γ212 ¼ 1 2g22 ∂g22 ∂x1 8>>><>>>: – For i 6¼ j, j 6¼ k, i 6¼ k it results in Γij,k ¼ 0 so R1212 ¼ �1 2 ∂2g11 ∂x2∂x2 þ ∂ 2 g22 ∂x1∂x1 ! þ 1 4g11 ∂g11 ∂x2 � �2 þ ∂g11 ∂x1 ∂g22 ∂x1 " # þ 1 4g22 ∂g22 ∂x1 � �2 þ ∂g11 ∂x2 ∂g22 ∂x2 " # ¼ � 1 2 ffiffiffiffiffiffiffiffiffiffiffiffi g11g22 p ∂ ∂x1 1ffiffiffiffiffiffiffiffiffiffiffiffi g11g22 p ∂g22 ∂x1 � � þ ∂ ∂x2 1ffiffiffiffiffiffiffiffiffiffiffiffi g11g22 p ∂g11 ∂x2 � � � 5.6 Particular Cases of Riemann Spaces 269 or R1212 ¼ � 12 ffiffiffi g p ∂ ∂x1 1ffiffiffi g p ∂g22 ∂x1 � � þ ∂ ∂x2 1ffiffiffi g p ∂g11 ∂x2 � � � ð5:6:7Þ Exercise 5.4 Calculate the components of tensors Rijk‘, Rij, and the Gauß curvature for the space E2 defined by the fundamental form ds 2 ¼ c2 dx1ð Þ 2 � f 2 tð Þ dx2ð Þ2 where c2 is a constant. The metric tensor and conjugated metric tensor are given, respectively, by gij ¼ c2 0 0 �f 2 tð Þ " # gij ¼ c �2 0 0 �f�2 tð Þ " # then g ¼ c2f 2i2 ) ffiffiffigp ¼ cf i where i2 ¼ �1 is the imaginary number and with expression (5.6.8) R1212 ¼ � 1 2 ffiffiffi g p ∂ ∂x1 1ffiffiffi g p ∂g22 ∂x1 � � þ ∂ ∂x2 1ffiffiffi g p ∂g11 ∂x2 � � � it follows that R1212 ¼ � 1 2cf i ∂ ∂x1 1 cf i ∂g22 ∂x1 � � ¼ � 1 2cf i ∂ ∂x1 1 cf i � �2f _f� � � ¼ 1 2cf i ∂ ∂x1 2 _f ci � � ¼ €f c2f i2 ¼ � €f c2f For the components of the Ricci tensor it follows that Rij ¼ gpkRipkj R11 ¼ g22R1212 ¼ � 1 f 2 � €f c2f � � ¼ €f c2f 3 R22 ¼ g11R1212 ¼ 1 c2 � €f c2f � � ¼ � €f c4f R12 ¼ R21 ¼ g12R1212 ¼ 0 and for the Gauß curvature it results in K ¼ R1212 g ¼ � €fc2f c2f 2i2 ¼ €f c4f 3 270 5 Riemann Spaces 5.6.4 Einstein Tensor For the particular case in which the metric, the metric tensor, and its conjugated tensor are given, respectively, by ds2 ¼ h x1; x2� � dx1� �2 þ h x1; x2� � dx2� �2 gij ¼ h 0 0 h " # gij ¼ 1 h 0 0 1 h 2664 3775 where h x1; x2ð Þ > 0 is a function of the coordinates, g ¼ detgij ¼ h2, and the Ricci tensor is expressed by Rij ¼ gpkRipkj ¼ g11Ri11j þ g12Ri12j þ g21Ri21j þ g22Ri22j then Rij ¼ 1 h Ri11j þ Ri22j � � Developing this expression and with the symmetry of tensor Ripkj it follows that R11 ¼ 1 h R1111 þ R1221ð Þ ¼ 1 h R1221 R22 ¼ 1 h R2112 þ R2222ð Þ ¼ 1 h R2112 R12 ¼ 1 h R1112 þ R1222ð Þ ¼ 0 R21 ¼ 1 h R2111 þ R2221ð Þ ¼ 0 Let the scalar curvature R ¼ gijRij ¼ g11R11 þ g12R12 þ g21R21 þ g22R22 ¼ g11R11 þ 0þ 0þ g22R22 ¼ g11R11 þ g22R22 and with the components of the Ricci tensor as a function of the components of tensor Ripkj it follows that R ¼ 1 h 1 h R1221 þ 1 h 1 h R2112 As Ripkj ¼ Rpijk it results for the scalar curvature 5.6 Particular Cases of Riemann Spaces 271 R ¼ 1 h2 R1221 þ R1221ð Þ ¼ 2 h2 R1221 then R1221 ¼ h 2 2 R and with the substitution of this expression in the expressions of the components of the Ricci tensor it follows that R11 ¼ 1 h h2 2 R ¼ h 2 R ¼ R 2 g11 R22 1 h h2 2 R ¼ h 2 R ¼ R 2 g22 R12 ¼ R21 ¼ 0 These expressions allow relating the Ricci tensor with the scalar curvature and with the metric tensor, thus Rij ¼ R 2 gij ð5:6:8Þ and with the definition of the scalar curvature given by expression (5.4.6) and with the previous expression it follows R ¼ gijRij ¼ gijgij R 2 ¼ δ ii R 2 ¼ NR 2 or R 1� N 2 � � ¼ 0 ð5:6:9Þ then for the Riemann space E2 it is verified that Rij ¼ R ¼ 0. Consider the Einstein tensor given by its covariant components Gij ¼ Rij � 1 2 gijR ¼ �Kgij � 1 2 gijR ¼ �Kgij � 1 2 gij �2Kð Þ ¼ 0 then the tensor Gij is null for the Riemann space E2. 272 5 Riemann Spaces Free ebooks ==> www.Ebook777.com 5.6.5 Riemann Space with Constant Curvature The Riemann curvature in point xi2EN , in general, depends on this point in which it is defined and the vectors u and v that establish the plane π with respect to which it is calculated. It is admitted that this dependency does not exist, i.e., the space is isotropic, then the relation of the isotropy of the space with the Riemann curvature is established by the following theorem. Schur Theorem If all the points of a neighborhood in the Riemann space EN, being N > 2, are isotropic, then the curvature K is constant in all this neighborhood. To prove the validity of this theorem, let expression (5.3.13) be rewritten as Rijk‘ ¼ Gijk‘K ð5:6:10Þ with Gijk‘ ¼ gikgj‘ � gi‘gjk � � 6¼ 0 valid in the neighborhood of point xm of Riemann space EN. The covariant derivative of expression (5.6.11) with respect to variable xm is given by ∂mRijk‘ ¼ Gijk‘∂mK ð5:6:11Þ with ∂mGijk‘ ¼ 0, because, in general, ∂gij∂xm ¼ 0. With the permutation of indexes in the expression (5.6.12) ∂kRij‘m ¼ Gij‘m∂kK ð5:6:12Þ ∂‘Rijmk ¼ Gijmk∂‘K ð5:6:13Þ The sum of expressions (5.6.12)–(5.6.14) provides ∂mRijk‘ þ ∂kRij‘m þ ∂‘Rijmk ¼ Gijk‘∂mK þ Gij‘m∂kK þ Gijmk∂‘K but the left side of expression is the second Bianchi identity thus Gijk‘∂mK þ Gij‘m∂kK þ Gijmk∂‘K ¼ 0 and multiplying the terms of this expression by gikgj‘ it follows 5.6 Particular Cases of Riemann Spaces 273 www.Ebook777.com http://www.ebook777.com gikgj‘Gijk‘∂mK ¼ gikgj‘ gikgj‘ � gi‘gjk � � ¼ δ kk δ ‘‘ � δ k‘ δ ‘k ¼ N2 � N gikgj‘Gij‘m∂kK ¼ gikgj‘ gi‘gjm � gimgj‘ � � ¼ δ k‘ δ ‘m � δ kmδ ‘‘ ¼ δ km � Nδ km gikgj‘Gijmk∂‘K ¼ gikgj‘ gimgjk � gikgjm � � ¼ δ kmδ ‘k � δ kk δ ‘m ¼ δ ‘m � Nδ ‘m The sum of these three terms provides N2 � N� �∂mK þ δ km � Nδ km� �∂kK þ δ ‘m � Nδ ‘m� �∂‘K ¼ 0 it follows that N2 � N� �∂mK þ 1� Nð Þ∂mK þ 1� Nð Þ∂mK ¼ 0 whereby N2 � N� �þ 2 1� Nð Þ� �∂mK ¼ 0 ð5:6:14Þ For N > 2 this expression is null only if ∂mK ¼ 0, and as xm is an arbitrary coordinate it is concluded that K is constant in the neighborhood of this point in the Riemann space EN, which proves the Schur theorem. Expression (5.3.13), where K is a constant is the necessary and sufficient condition so that the curvature of the Riemann space EN is independent of the orientation considered. 5.6.6 Isotropy Another characteristic of this type of space is related with a scalar curvature. Let expression (5.3.13) be rewritten as Rijk‘ ¼ K gikgj‘ � gi‘gjk � � and multiplied by g‘i Rjk ¼ g‘iRijk‘ ¼ Kg‘i gikgj‘ � gi‘gjk � � ¼ K δ ‘k gj‘ � δ ‘‘gjk � � ¼ K gjk � Ngjk � � then Rjk ¼ K 1� Nð Þgjk ð5:6:15Þ 274 5 Riemann Spaces For the scalar curvature it follows that Rkk ¼ gkjRjk ¼ gkjK 1� Nð Þgjk ¼ K 1� Nð Þδ kk whereby R ¼ K 1� Nð ÞN ð5:6:16Þ This formulation shows that in the Riemann space E2 the tensor Rijk‘ leads to the Gauß curvature K, which is the reason for adopting the denomination curvature tensor by extension of this particular case for Riemann spaces of N dimensions. For the Riemann space EN, where N > 2, in which the Ricci tensor results from the substitution of expression (5.6.17) in expression (5.6.16), thus Rij ¼ R N gij ð5:6:17Þ where the ratio RN defines a scalar. The space in which the Ricci tensor is pro- portional to the metric tensor is called the Einstein space. The scalar curvature of the Einstein space is given by gpiRij ¼ K N gpigij following for the Ricci tensor with variance (1, 1) Rpj ¼ K N δ pi The covariant derivative of this expression with respect to variable xp is given by ∂pR p j ¼ K N ∂δpj ∂xj ¼ 0 and with expression (5.4.11) ∂pR p j ¼ 1 2 ∂R ∂xj ¼ 0 whereby ∂R ∂xj ¼ 0 ð5:6:18Þ then the Einstein space has constant curvature, i.e., is isotropic. The multiplying of expression (5.6.18) by vector uj allows researching the eigenvalues of the Ricci tensor, thus 5.6 Particular Cases of Riemann Spaces 275 Riju j ¼ R N giju j ¼ R N ui ) Rij � R N δij � � uj ¼ 0 where the scalar curvature is constant then the eigenvalues are equal to RN. In this case the eigenvectors of tensor Rij are undetermined. Exercise 5.5 Calculate the components of the curvature tensor Rijk‘, of the Ricci tensor Rij, the scalar curvature and the Gauß curvature K for the bidimensional spherical space which metric is given by ds2 ¼ r2 dφ2 þ sin 2φdθ2� � The metric tensor, the determinant g, and the conjugated tensor of gij are given, respectively, by gij ¼ r2 0 0 r2 sin 2φ " # g ¼ r4 sin 2φ gij ¼ 1 r2 0 0 1 r2 sin 2φ 2664 3775 For the partial derivatives of the metric tensor the result is g11, 1 ¼ g22, 2 ¼ 0, following for the Christoffel symbols Γ111 ¼ Γ222 ¼ Γ211 ¼ Γ212 ¼ Γ221 ¼ 0 Γ112 ¼ Γ121 ¼ g11g11, 2 2 ¼ 1 2r2 sin 2φ ∂ r2 sin 2φð Þ ∂φ ¼ � cos φ sin φ Γ122 ¼ � g11g22, 1 2 ¼ 1 2r2 ∂ r2 sin 2φð Þ ∂φ ¼ � sin φ � cos φ thus Rijk‘ ¼ ∂Γ ij‘ ∂xk � ∂Γ i jk ∂x‘ þ Γmj‘Γ imk � ΓmjkΓ im‘ R1212 ¼ g1mRm212