guide to geophysical equations

297

Upload: juan-specht

Post on 02-Mar-2018

213 views

Category:

Documents


1 download

TRANSCRIPT

  • 7/26/2019 Guide to Geophysical Equations

    1/296

  • 7/26/2019 Guide to Geophysical Equations

    2/296

    A Students Guide to Geophysical Equations

    The advent of accessible student computing packages has meant that geophysics students

    can now easily manipulate datasets and gain rst-hand modeling experience essential

    in developing an intuitive understanding of the physics of the Earth. Yet to gain a more

    in-depth understanding of the physical theory, and to be able to develop new models and

    solutions, it is necessary to be able to derive the relevant equations from rst principles.

    This compact, handy book lls a gap left by most modern geophysics textbooks,

    which generally do not have space to derive all of the important formulae, showing the

    intermediate steps. This guide presents full derivations for the classical equations of

    gravitation, gravity, tides, Earth rotation, heat, geomagnetism, and foundational seismol-

    ogy, illustrated with simple schematic diagrams. It supports students through the suc-

    cessive steps and explains the logical sequence of a derivation facilitating self-study

    and helping students to tackle homework exercises and prepare for exams.

    w i l l i a m l o w r i e was born in Hawick, Scotland, and attended the University of

    Edinburgh, where he graduated in 1960 with rst-class honors in physics. He achieved a

    masters degree in geophysics at the University of Toronto and, in 1967, a doctorate at the

    University of Pittsburgh. After two years in the research laboratory of Gulf Oil Company

    he became a researcher at the Lamont-Doherty Geological Observatory of Columbia

    University. In 1974 he was elected professor of geophysics at the ETH Zrich (Swiss

    Federal Institute of Technology in Zurich), Switzerland, where he taught and researched

    until retirement in 2004. His research in rock magnetism and paleomagnetism consisted

    of deducing the Earths magnetic eld in the geological past from the magnetizations of

    dated rocks. The results were applied to the solution of geologic-tectonic problems, and

    to analysis of the polarity history of the geomagnetic eld. Professor Lowrie has authored

    135 scientic articles and a second edition of his acclaimed 1997 textbookFundamentals

    of Geophysicswas published in 2007. He has been President of the European Union of

    Geosciences (19879) and Section President and Council member of the American

    Geophysical Union (20002). He is a Fellow of the American Geophysical Union and

    a Member of the Academia Europaea.

  • 7/26/2019 Guide to Geophysical Equations

    3/296

  • 7/26/2019 Guide to Geophysical Equations

    4/296

    A Students Guide

    to Geophysical Equations

    W I L L I A M L O W R I E

    Institute of Geophysics

    Swiss Federal Institute of Technology

    Zurich, Switzerland

  • 7/26/2019 Guide to Geophysical Equations

    5/296

    c a m b r i d g e u n i v e r s i t y p r e s s

    Cambridge, New York, Melbourne, Madrid, Cape Town,

    Singapore, So Paulo, Delhi, Tokyo, Mexico City

    Cambridge University Press

    The Edinburgh Building, Cambridge CB2 8RU, UK

    Published in the United States of America by Cambridge University Press, New York

    www.cambridge.org

    Information on this title: www.cambridge.org/9781107005846

    William Lowrie 2011

    This publication is in copyright. Subject to statutory exception

    and to the provisions of relevant collective licensing agreements,

    no reproduction of any part may take place without the written

    permission of Cambridge University Press.

    First published 2011

    Printed in the United Kingdom at the University Press, Cambridge

    A catalogue record for this publication is available from the British Library

    Library of Congress Cataloguing in Publication data

    Lowrie, William, 1939

    A students guide to geophysical equations / William Lowrie.

    p. cm.

    Includes bibliographical references and index.

    ISBN 978-1-107-00584-6 (hardback)1. Geophysics Mathematics Handbooks, manuals, etc.

    2. Physics Formulae Handbooks, manuals, etc. 3. Earth Handbooks, manuals, etc.

    I. Title.

    QC809.M37L69 2011

    550.1051525dc22

    2011007352

    ISBN 978 1 107 00584 6 Hardback

    ISBN 978 0 521 18377 2 Paperback

    Cambridge University Press has no responsibility for the persistence or

    accuracy of URLs for external or third-party internet websites referred to

    in this publication, and does not guarantee that any content on such

    websites is, or will remain, accurate or appropriate.

    http://www.cambridge.org/http://www.cambridge.org/9781107005846http://www.cambridge.org/9781107005846http://www.cambridge.org/
  • 7/26/2019 Guide to Geophysical Equations

    6/296

    This book is dedicated to Marcia

  • 7/26/2019 Guide to Geophysical Equations

    7/296

  • 7/26/2019 Guide to Geophysical Equations

    8/296

    Contents

    Preface pagexi

    Acknowledgments xiii

    1 Mathematical background 1

    1.1 Cartesian and spherical coordinates 1

    1.2 Complex numbers 1

    1.3 Vector relationships 4

    1.4 Matrices and tensors 8

    1.5 Conservative force,eld, and potential 17

    1.6 The divergence theorem (Gausss theorem) 18

    1.7 The curl theorem (Stokestheorem) 20

    1.8 Poissons equation 23

    1.9 Laplaces equation 26

    1.10 Power series 28

    1.11 Leibnizs rule 32

    1.12 Legendre polynomials 32

    1.13 The Legendre differential equation 34

    1.14 Rodriguesformula 41

    1.15 Associated Legendre polynomials 43

    1.16 Spherical harmonic functions 491.17 Fourier series, Fourier integrals, and Fourier transforms 52

    Further reading 58

    2 Gravitation 59

    2.1 Gravitational acceleration and potential 59

    2.2 Keplers laws of planetary motion 60

    2.3 Gravitational acceleration and the potential of a solid

    sphere 66

    2.4 Laplaces equation in spherical polar coordinates 69

    2.5 MacCullaghs formula for the gravitational potential 74

    Further reading 85

    vii

  • 7/26/2019 Guide to Geophysical Equations

    9/296

    3 Gravity 86

    3.1 The ellipticity of the Earths gure 86

    3.2 The geopotential 88

    3.3 The equipotential surface of gravity 91

    3.4 Gravity on the reference spheroid 96

    3.5 Geocentric and geographic latitude 102

    3.6 The geoid 106

    Further reading 115

    4 The tides 116

    4.1 Origin of the lunar tide-raising forces 116

    4.2 Tidal potential of the Moon 119

    4.3 Loves numbers and the tidal deformation 1244.4 Tidal friction and deceleration of terrestrial and lunar

    rotations 130

    Further reading 136

    5 Earths rotation 137

    5.1 Motion in a rotating coordinate system 138

    5.2 The Coriolis and Etvs effects 140

    5.3 Precession and forced nutation of Earths rotation axis 142

    5.4 The free, Eulerian nutation of a rigid Earth 1555.5 The Chandler wobble 157

    Further reading 169

    6 Earths heat 170

    6.1 Energy and entropy 171

    6.2 Thermodynamic potentials and Maxwells relations 172

    6.3 The melting-temperature gradient in the core 176

    6.4 The adiabatic temperature gradient in the core 178

    6.5 The Grneisen parameter 179

    6.6 Heatow 182

    Further reading 197

    7 Geomagnetism 198

    7.1 The dipole magnetic eld and potential 198

    7.2 Potential of the geomagnetic eld 200

    7.3 The Earths dipole magnetic eld 205

    7.4 Secular variation 213

    7.5 Power spectrum of the internal eld 214

    7.6 The origin of the internal eld 217

    Further reading 225

    8 Foundations of seismology 227

    8.1 Elastic deformation 227

    viii Contents

  • 7/26/2019 Guide to Geophysical Equations

    10/296

    8.2 Stress 228

    8.3 Strain 233

    8.4 Perfectly elastic stressstrain relationships 239

    8.5 The seismic wave equation 244

    8.6 Solutions of the wave equation 252

    8.7 Three-dimensional propagation of plane P- and S-waves 254

    Further reading 258

    Appendix A Magnetic poles, the dipoleeld, and current loops 259

    Appendix B Maxwells equations of electromagnetism 265

    References 276

    Index 278

    Contents ix

  • 7/26/2019 Guide to Geophysical Equations

    11/296

  • 7/26/2019 Guide to Geophysical Equations

    12/296

    Preface

    This work was written as a supplementary text to help students understand the

    mathematical steps in deriving important equations in classical geophysics. It is

    not intended to be a primary textbook, nor is it intended to be an introduction to

    modern research in any of the topics it covers. It originated in a set of handouts, a

    kindof do-it-yourself manual, that accompanied a course I taught on theoretical

    geophysics. The lecture aids were necessary for two reasons. First, my lectures

    were given in German and there were no comprehensive up-to-date texts in the

    language; the recommended texts were in English, so the students frequentlyneeded clarication. Secondly, it was often necessary to explain classical theory

    in more detail than one nds in a multi-topic advanced textbook. To keep such a

    book as succinct as possible, the intermediate steps in the mathematical derivation

    of a formula must often be omitted. Sometimes the unassisted student cannotll

    in the missing steps without individual tutorial assistance, which is usually in

    short supply at most universities, especially at large institutions. To help my

    students in these situations, the do-it-yourself text that accompanied my lec-

    tures explained missing details in the derivations. This is the background against

    which I prepared the present guide to geophysical equations, in the hope that it

    might be helpful to other students at this level of study.

    The classes that I taught to senior grades were largely related to potential

    theory and primarily covered topics other than seismology, since this was the

    domain of my colleagues and better taught by a true seismologist than by a

    paleomagnetist! Theoretical seismology is a large topic that merits its own

    treatment at an advanced level, and there are several textbooks of classical

    and modern vintage that deal with this. However, a short chapter on the

    relationship of stress, strain, and the propagation of seismic waves is included

    here as an introduction to the topic.

    Computer technology is an essential ingredient of progress in modern geo-

    physics, but a well-trained aspiring geophysicist must be able to do more than

    xi

  • 7/26/2019 Guide to Geophysical Equations

    13/296

    apply advanced software packages. A fundamental mathematical understanding

    is needed in order to formulate a geophysical problem, and numerical computa-

    tional skills are needed to solve it. The techniques that enabled scientists to

    understand much about the Earth in the pre-computer era also underlie much of

    modern methodology. For this reason, a university training in geophysics still

    requires the student to work through basic theory. This guide is intended as a

    companion in that process.

    Historically, most geophysicists came from the eld of physics, for which

    geophysics was an applied science. They generally had a sound training in

    mathematics. The modern geophysics student is more likely to have begun

    studies in an Earth science discipline, the mathematical background might be

    heavily oriented to the use of tailor-made packaged software, and some studentsmay be less able to handle advanced mathematical topics without help or

    tutoring. To ll these needs, the opening chapter of this book provides a

    summary of the mathematical background for topics handled in subsequent

    chapters.

    xii Preface

  • 7/26/2019 Guide to Geophysical Equations

    14/296

    Acknowledgments

    In writing this book I have beneted from the help and support of various

    people. At an early stage, anonymous proposal reviewers gave me useful

    suggestions, not all of which have been acted on, but all of which were

    appreciated. Each chapter was read and checked by an obliging colleague.

    I wish to thank Dave Chapman, Rob Coe, Ramon Egli, Chris Finlay, Valentin

    Gischig, Klaus Holliger, Edi Kissling, Emile Klingel, Alexei Kuvshinov,

    Germn Rubino, Rolf Sidler, and Doug Smylie for their corrections and sug-

    gestions for improvement. The responsibility for any errors that escaped scru-tiny is, of course, mine. I am very grateful to Derrick Hasterok and Dave

    Chapman for providing me with an unpublished gure from Derricks Ph.D.

    thesis. Dr. Susan Francis, Senior Commissioning Editor at Cambridge

    University Press, gave me constant support and friendly encouragement

    throughout the many months of writing, for which I am sincerely grateful.

    Above all, I thank my wife Marcia for her generous tolerance of the intrusion

    of this project into our retirement activities.

    xiii

  • 7/26/2019 Guide to Geophysical Equations

    15/296

  • 7/26/2019 Guide to Geophysical Equations

    16/296

    1

    Mathematical background

    1.1 Cartesian and spherical coordinates

    Two systems of orthogonal coordinates are used in this book, sometimes

    interchangeably. Cartesian coordinates (x, y, z) are used for a system with

    rectangular geometry, and spherical polar coordinates (r, , ) are used for

    spherical geometry. The relationship between these reference systems is shown

    inFig. 1.1(a). The convention used here for spherical geometry is dened as

    follows: the radial distance from the origin of the coordinates is denoted r, the

    polar angle (geographic equivalent: the co-latitude) lies between the radius

    and thez-axis (geographic equivalent: Earths rotation axis), and the azimuthal

    angle in the xy plane is measured from the x-axis (geographic equivalent:

    longitude). Position on the surface of a sphere (constantr) is described by the

    two anglesand. The Cartesian and spherical polar coordinates are linked as

    illustrated inFig. 1.1(b)by the relationships

    xr sin cos yr sin sin zr cos

    (1:1)

    1.2 Complex numbers

    The numbers we most commonly use in daily life are realnumbers. Some of

    them are also rationalnumbers. This means that they can be expressed as the

    quotient of two integers, with the condition that the denominator of the quotient

    must not equal zero. When the denominator is 1, the real number is an integer.

    Thus 4, 4/5, 123/456 are all rational numbers. A real number can also be

    irrational, which means it cannot be expressed as the quotient of two integers.

    1

  • 7/26/2019 Guide to Geophysical Equations

    17/296

    Familiar examples are , e (the base of natural logarithms), and some square

    roots, such as2,3,5, etc. The irrational numbers are real numbers that do

    not terminate or repeat when expressed as decimals.

    In certain analyses, such as determining the roots of an equation, it is

    necessary to nd the square root of a negative real number, e.g. (y2), where

    yis real. The result is an imaginarynumber. The negative real number can be

    written as (1)y2, and its square root is then(1)y. The quantity(1) is written

    iand is known as the imaginary unit, so that(y2) becomes iy.

    A complex number comprises a real part and an imaginary part. For example,

    z = x+ iy, in whichx and y are both real numbers, is a complex number with a

    real part xand animaginary part y. The composition of a complex number can

    be illustrated graphically with the aid of thecomplex plane(Fig. 1.2). The real

    part is plotted on the horizontal axis, and the imaginary part on the vertical axis.

    The two independent parts are orthogonal on the plot and the complex numberz

    r

    z

    xy

    (a)

    r

    y= rsinsin

    x= rsincos

    z= rcos

    rsin

    (b)

    Fig. 1.1. (a) Cartesian and spherical polar reference systems. (b) Relationshipsbetween the Cartesian and spherical polar coordinates.

    +x

    +y z= x+ iy

    Realaxis

    Imaginaryaxis

    r

    rcos

    rsin

    Fig. 1.2. Representation of a complex number on an Argand diagram.

    2 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    18/296

    is represented by their vector sum, dening a point on the plane. The distancer

    of the point from the origin is given by

    r ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffix2 y2p (1:2)The line joining the point to the origin makes an angle with the real (x-)axis,

    and so rhas real and imaginary components rcos and rsin , respectively. The

    complex numberzcan be written in polar form as

    zr cos isin (1:3)It is often useful to write a complex number in the exponential form introduced

    by Leonhard Euler in the late eighteenth century. To illustrate this we make useof innite power series; this topic is described inSection 1.10. The exponential

    function, exp(x), of a variablexcan be expressed as a power series as in (1.135).

    On substitutingx = i, the power series becomes

    exp i 1 i i 2

    2! i

    3

    3! i

    4

    4! i

    5

    5! i

    6

    6!

    1

    i 2

    2! i 4

    4! i 6

    6! i

    i 3

    3! i 5

    5! 1

    2

    2!

    4

    4!

    6

    6!

    i

    3

    3!

    5

    5!

    (1:4)

    Comparison with (1.135) shows that the rst bracketed expression on the

    right is the power series for cos ; the second is the power series for sin .

    Therefore

    exp i cos isin (1:5)On inserting (1.5) into (1.3), the complex numberzcan be written in exponential

    form as

    zr exp i (1:6)The quantityris themodulusof the complex number and is itsphase.

    Conversely, using (1.5) the cosine and sine functions can be dened as the

    sum or difference of the complex exponentials exp(i) and exp(i):

    cos exp i expi 2

    sin exp i expi 2i

    (1:7)

    1.2 Complex numbers 3

  • 7/26/2019 Guide to Geophysical Equations

    19/296

    1.3 Vector relationships

    A scalar quantity is characterized only by its magnitude; a vector has bothmagnitude and direction; a unit vector has unit magnitude and the direction of

    the quantity it represents. In this overview the unit vectors for Cartesian

    coordinates (x, y, z) are written (ex, ey, ez); unit vectors in spherical polar

    coordinates (r,,) are denoted (er,e,e). The unit vector normal to a surface

    is simply denotedn.

    1.3.1 Scalar and vector products

    The scalar product of two vectors a and b is dened as the product of their

    magnitudes and the cosine of the angle between the vectors:

    a bab cos (1:8)If the vectors are orthogonal, the cosine of the angle is zero and

    a b0 (1:9)The vector product of two vectors is another vector, whose direction is perpen-

    dicular to both vectors, such that a right-handed rule is observed. The magnitudeof the vector product is the product of the individual vector magnitudes and the

    sine of the angle between the vectors:

    a bj j ab sin (1:10)Ifa and b are parallel, the sine of the angle between them is zero and

    a b0 (1:11)

    Applying these rules to the unit vectors (ex, ey, ez), which are normal to each

    other and have unit magnitude, it follows that their scalar products are

    ex ey ey ez ez ex0ex ex ey ey ez ez1

    (1:12)

    The vector products of the unit vectors are

    ex

    ey

    ez

    ey ez exez ex eyex ex ey ey ez ez0

    (1:13)

    4 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    20/296

    A vectorawith components (ax,ay,az) is expressed in terms of the unit vectors

    (ex,ey,ez) as

    aaxex ayey azez (1:14)The scalar product of the vectorsaand bis found by applying the relationships

    in (1.12):

    a b axex ayey azez

    bxex byey bzez

    axbx ayby azbz (1:15)The vector product of the vectorsa and b is found by using (1.13):

    a b axex ayey azez bxex byey bzez aybz azby

    ex azbx axbz ey axby aybx

    ez (1:16)

    This result leads to a convenient way of evaluating the vector product of two

    vectors, by writing their components as the elements of a determinant, as

    follows:

    a

    b

    ex ey ez

    ax ay az

    bx by bz (1:17)

    The following relationships may be established, in a similar manner to the

    above, for combinations of scalar and vector products of the vectors a,b, andc:

    a b c b c a c a b (1:18)

    a b c b c a c a b (1:19)

    a

    b

    c

    b c a

    a b c

    (1:20)

    1.3.2 Vector differential operations

    The vector differential operator is dened relative to Cartesian axes (x,y,z)

    as

    r ex x

    ey y

    ez z

    (1:21)

    The vector operator determines the gradient of a scalar function, which may

    be understood as the rate of change of the function in the direction of each of the

    reference axes. For example, the gradient of the scalar function with respect to

    Cartesian axes is the vector

    1.3 Vector relationships 5

  • 7/26/2019 Guide to Geophysical Equations

    21/296

    r ex x

    ey y

    ez z

    (1:22)

    The vector operator can operate on either a scalar quantity or a vector. The

    scalar product of with a vector is called thedivergenceof the vector. Applied

    to the vectora it is equal to

    r a ex x

    ey y

    ez z

    axex ayey azez

    axx

    ayy

    azz

    (1:23)

    If the vectora is dened as the gradient of a scalar potential , as in(1.22), we

    can substitute potential gradients for the vector components (ax, ay, az). This

    gives

    r r x

    x

    y

    y

    z

    z

    (1:24)

    By convention the scalar product ( ) on the left is written 2. The resulting

    identity is very important in potential theory and is encountered frequently. In

    Cartesian coordinates it is

    r2 2

    x2

    2

    y2

    2

    z2 (1:25)

    The vector product of with a vector is called the curl of the vector. The

    curl of the vector a may be obtained using a determinant similar to (1.17):

    r aex ey ez

    =x =y =zax ay az

    (1:26)In expanded format, this becomes

    r a azy

    ayz

    ex ax

    z az

    x

    ey ay

    x ax

    y

    ez (1:27)

    The curl is sometimes called the rotation of a vector, because of its physical

    interpretation (Box 1.1). Some commonly encountered divergence and curl

    operations on combinations of the scalar quantity and the vectors a and b

    are listed below:

    r a r a r a (1:28)

    6 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    22/296

    r a b b r a a r b (1:29)

    r a r a r a (1:30)

    r a b ar b br a a r b b r a (1:31)

    r r 0 (1:32)

    Box 1.1. The curl of a vector

    The curl of a vector at a given point is related to the circulation of the vector

    about that point. This interpretation is best illustrated by an example, in

    which a uid is rotating about a point with constant angular velocity .

    At distancerfrom the point the linear velocity of the uidv is equal to r.

    Taking the curl ofv, and applying the identity (1.31) with constant,

    r v r w r wr r w r r (1)To evaluate the rst term on the right, we use rectangular coordinates (x,y,z):

    wr r w ex

    x ey

    y ez

    z

    xexyey zez w ex ex ey ey ez ez

    3w (2)The second term is

    w r r x x

    y y

    z z

    xexyey zez

    xex

    yey

    zez

    w (3)

    Combining the results gives

    r v2w (4)

    w12r v (5)

    Because of this relationship between the angular velocity and the linear

    velocity of a

    uid, the curl operation is often interpreted as the rotationofthe uid. When v = 0 everywhere, there is no rotation. A vector that

    satises this condition is said to beirrotational.

    1.3 Vector relationships 7

  • 7/26/2019 Guide to Geophysical Equations

    23/296

    r r a 0 (1:33)

    r r a r r a r2a (1:34)

    It is a worthwhile exercise to establish these identities from basic principles,especially (1.19) and (1.31)(1.34), which will be used in later chapters.

    1.4 Matrices and tensors

    1.4.1 The rotation matrix

    Consider two sets of orthogonal Cartesian coordinate axes (x,y,z) and (x0,y0,

    z0) that are inclined to each other as in Fig. 1.3. Thex0-axis makes angles (1,1,

    1) with each of the (x,y,z) axes in turn. Similar sets of angles (2,2,2) and

    (3,3,3) are dened by the orientations of they0- andz0-axes, respectively, to

    the (x,y,z) axes. Let the unit vectors along the (x,y,z) a n d (x0,y0,z0)axesbe(n1,

    n2,n3) and (m1,m2,m3), respectively. The vectorr can be expressed in either

    system, i.e.,r = r(x,y,z) =r(x0,y0,z0), or, in terms of the unit vectors,

    rxn1yn2 zn3x0m1y0m2 z0m3 (1:35)We can write the scalar product (r m1) as

    r m1xn1 m1yn2 m1 zn3 m1x0 (1:36)The scalar product (n1 m1) = cos 1=11denes11as thedirection cosineof

    thex0-axis with respect to thex-axis (Box 1.2). Similarly, (n2 m1) = c o s1 = 12

    z0z

    x0x

    y0

    y

    1

    1

    1

    2

    3

    2

    2

    3

    3

    m

    1 n1

    m2

    n2

    m3n3

    Fig. 1.3. Two sets of Cartesian coordinate axes, (x, y, z) and (x0, y0, z0), with

    corresponding unit vectors (n1, n2, n3)and(m1, m2, m3), rotated relative to each other.

    8 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    24/296

    and (n3 m1) = cos 1= 13dene 12and 13as the direction cosines of the

    x0-axis with respect to they- andz-axes, respectively. Thus, (1.36) is equivalent to

    x011x 12y 13z (1:37)On treating they0- andz0-axes in the same way, we get their relationships to the

    (x,y,z) axes:

    y021x 22y 23zz031x 32y 33z

    (1:38)

    The three equations can be written as a single matrix equation

    x0y0z0

    24 35 11 12 1321 22 2331 32 33

    24 35 xyz

    24 35M xyz

    24 35 (1:39)The coefcients nm(n = 1, 2, 3; m = 1, 2, 3) are the cosines of the interaxial

    angles. By denition,12=21,23=32, and31=13, so the square matrixM

    is symmetric. It transforms the components of the vector in the (x, y, z)

    coordinate system to corresponding values in the (x0,y0,z0) coordinate system.

    It is thus equivalent to a rotation of the reference axes.

    Because of the orthogonality of the reference axes, useful relationships existbetween the direction cosines, as shown inBox 1.2. For example,

    11 2 12 2 13 2 cos21 cos21 cos21 1

    r2 x2 y2 z2 1

    (1:40)

    and

    1121 1222 1323cos 1cos 2 cos1cos2 cos 1cos 20

    (1:41)The last summation is zero because it is the cosine of the right angle between the

    x0-axis and they0-axis.

    These two results can be summarized as

    X3k1

    mknk 1; mn0; m6n

    (1:42)

    1.4.2 Eigenvalues and eigenvectors

    The transpose of a matrix Xwith elements nm is a matrix with elements mn(i.e., the elements in the rows are interchanged with corresponding elements in

    1.4 Matrices and tensors 9

  • 7/26/2019 Guide to Geophysical Equations

    25/296

    Box 1.2. Direction cosines

    The vectorr is inclined at angles,, and, respectively, to orthogonal

    reference axes (x,y,z) with corresponding unit vectors (ex,ey,ez), as inFig.

    B1.2. The vectorr can be written

    rxexyey zez (1)

    where (x,y,z) are the components ofrwith respect to these axes. The scalar

    products ofr withex,ey, andezare

    z

    x

    r

    y

    ex

    ey

    ez

    Fig. B1.2. Angles,, and dene the tilt of a vectorr relative to orthogonal

    reference axes (x, y, z), respectively. The unit vectors (ex, ey, ez) dene thecoordinate system.

    r exxr cos r eyyr cosr ezzr cos

    (2)

    Therefore, the vectorr in (1) is equivalent to

    r r cos ex r cos ey r cos ez (3)

    The unit vectoru in the direction ofr has the same direction as r but its

    magnitude is unity:

    u rr cos ex cos ey cos ezlex mey nez (4)

    where (l,m,n) are the cosines of the angles that the vectorr makes with

    the reference axes, and are called the direction cosines ofr. They are

    useful for describing the orientations of lines and vectors.

    10 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    26/296

    the columns). The transpose of a (3 1) column matrix is a (1 3) row matrix.

    For example, ifXis a column matrix given by

    Xx

    yz

    24 35 (1:43)

    then its transpose is the row matrix XT, where

    XT x y z (1:44)The matrix equation XTMX = K, where K is a constant, denes a quadric

    surface:

    XTMX x y z 11 12 1321 22 2331 32 33

    24 35 xyz

    24 35K (1:45)The symmetry of the matrix leads to the equation of this surface:

    The scalar product of two unit vectors is the cosine of the angle they form.

    Letu1and u2be unit vectors representing straight lines with direction

    cosines (l1,m1,n1) and (l2,m2,n2), respectively, and letbe the anglebetween the vectors. The scalar product of the vectors is

    u1 u2cos l1ex m1ey n1ez

    l2ex m2ey n2ez

    (5)

    Therefore,

    cos l1l2 m1m2 n1n2 (6)

    The square of a unit vector is the scalar product of the vector with itself and isequal to 1:

    u u r rr2

    1 (7)

    On writing the unit vectoru as in (4), and applying the orthogonality

    conditions from (2), we nd that the sum of the squares of the direction

    cosines of a line is unity:

    lex mey nez lex mey nez l2 m2 n2 1 (8)

    1.4 Matrices and tensors 11

  • 7/26/2019 Guide to Geophysical Equations

    27/296

    fx;y; z 11x2 22y2 33z2 212xy 223yz 231zxK (1:46)When the coefcientsnmare all positive real numbers, the geometric expres-

    sion of the quadratic equation is an ellipsoid. The normal direction n to the

    surface of the ellipsoid at the point P(x,y,z) is the gradient of the surface. Using

    the relationships between (x,y,z) and (x0,y0,z0) in(1.39) and the symmetry of

    the rotation matrix,nm= mnforn m, the normal direction has components

    f

    x2 11x 12y 13z 2x0

    f

    y2 21x 22y 23z 2y0

    f

    z2 31x 32y 33z 2z0 (1:47)

    and we write

    nx;y; z rf ex fx

    ey fx

    ez fx

    (1:48)

    nx;y; z 2x0exy0ey z0ez 2rx0;y0; z0 (1:49)The normalnto the surface at P(x,y,z) in the original coordinates is parallel to

    the vectorr at the point (x0,y0,z0) in the rotated coordinates (Fig. 1.4).

    The transformation matrixMhas the effect of rotating the reference axes from

    one orientation to another. A particular matrix exists that will cause the direc-

    tions of the (x0, y0, z0) axes to coincide with the (x, y, z) axes. In this case the

    normal to the surface of the ellipsoid is one of the three principal axes of the

    x y

    z P

    tangentplane

    (x0,y0,z0)

    (x,y,z)r

    n

    er

    Fig. 1.4. Location of a point (x, y, z) on an ellipsoid, where the normal n to the

    surface is parallel to the radius vector at the point (x0,y0,z0).

    12 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    28/296

  • 7/26/2019 Guide to Geophysical Equations

    29/296

    1.4.3 Tensor notation

    Equations describing vector relationships can become cumbersome when writ-

    ten in full or symbolic form. Tensor notation provides a succinct alternative wayof writing the equations. Instead of the alphabetic indices used in theprevious

    section, tensor notation uses numerical indices that allow summations to be

    expressed in a compact form.

    Let the Cartesian coordinates (x,y,z) be replaced by coordinates (x1,x2, x3)

    and let the corresponding unit vectors be (e1, e2, e3). The vectora in (1.14)

    becomes

    aa1e1 a2e2 a3e3 Xi1;2;3

    aiei (1:56)

    A convention introduced by Einstein drops the summation sign and tacitly

    assumes that repetition of an index implies summation over all values of the

    index, in this case from 1 to 3. The vectora is then written explicitly

    aaiei (1:57)Alternatively, the unit vectors can be implied and the expression ai is under-

    stood to represent the vectora. Using the summation convention, (1.15) for the

    scalar product of two vectorsa and b is

    a ba1b1 a2b2 a3b3aibi (1:58)Suppose that two vectorsa and b are related, so that each component ofa is a

    linear combination of the components ofb. The relationship can be expressed in

    tensor notation as

    aiTijbj (1:59)

    The indicesiandjidentify components of the vectorsaandb; each index takeseach of the values 1, 2, and 3 in turn. The quantity Tij is a second-order (or

    second-rank) tensor, representing the array of nine coefcients (i.e., 32). A

    vector has three components (i.e., 31) a n d i s a rst-order tensor; a scalar property

    has a single (i.e., 30) value, its magnitude, and is a zeroth-order tensor.

    To write the cross product of two vectors we need to dene a new quantity,

    the Levi-Civitapermutation tensorijk. It has the value +1 when a permutation

    of the indices is even (i.e., 123 = 231 = 312= 1) and the value 1 when a

    permutation of the indices is odd (i.e., 132 = 213 = 321 = 1). If any pair of

    indices is equal,ijk= 0. This enables us to write the cross product of two vectors

    in tensor notation. Letu be the cross product of vectors a and b:

    u a b a2b3 a3b2 e1 a3b1 a1b3 e2 a1b2 a2b1 e3 (1:60)

    14 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    30/296

    In tensor notation this is written

    uiijkajbk (1:61)This can be veried readily for each component ofu. For example,

    u1123a2b3 132a3b2a2b3 a3b2 (1:62)The tensor equivalent to the unit matrix dened in (1.55) is known as

    Kroneckers symbol,ij, or alternatively theKronecker delta.It has the values

    ij 1; if ij0; if i6j

    (1:63)

    Kroneckers symbol is convenient for selecting a particular component of a

    tensor equation. For example, (1.54) can be written in tensor form using the

    Kronecker symbol:

    ijij

    xj0 (1:64)This represents the set of simultaneous equations in (1.50). Likewise, the

    relationship between direction cosines in (1.42) simplies to

    mknkmn (1:65)in which a summation over the repeated index is implied.

    1.4.4 Rotation of coordinate axes

    Letvkbe a vector related to the coordinates xlby the tensorTkl

    vkTklxl (1:66)

    A second set of coordinates xn is rotated relative to the axes xl so that the

    direction cosines of the angles between corresponding axes are the elements of

    the tensornl:

    x0nnlxl (1:67)Let the same vector be related to the rotated coordinate axes xn by the tensor

    Tkn:

    v0kT0knx0n (1:68)vk and vk are the same vector, expressed relative to different sets of axes.

    Therefore,

    v0kknvnknTnlxl (1:69)

    1.4 Matrices and tensors 15

  • 7/26/2019 Guide to Geophysical Equations

    31/296

  • 7/26/2019 Guide to Geophysical Equations

    32/296

    r e1 x1

    e2 x2

    e3 x3

    ei xi

    (1:78)

    Several shorthand forms of this equation are in common use; for example,

    xi r i;ii (1:79)

    Thedivergenceof the vectora is written in tensor notation as

    r a a1x1

    a2x2

    a3x3

    aixi

    iai (1:80)

    Thecurl(orrotation) of the vectora becomes

    r a e1 a3x2

    a2x3

    e2 a1

    x3 a3

    x1

    e3 a2

    x1 a1

    x2

    (1:81)

    r a iijkak

    xjijkjak (1:82)

    1.5 Conservative force,eld, and potential

    If the work done in moving an object from one point to another against a force is

    independent of the path between the points, the force is said to be conservative.

    No work is done if the end-point of the motion is the same as the starting point;

    this condition is called the closed-path test of whether a force is conservative. In

    a real situation, energy may be lost, for example to heat or friction, but in an

    ideal case the total energyEis constant. The workdWdone against the force F is

    converted into a gain dEP in the potential energy of the displaced object. Thechange in the total energydEis zero:

    dEdEP dW0 (1:83)The change in potential energy when a force with components (Fx, Fy, Fz)

    parallel to the respective Cartesian coordinate axes (x,y,z) experiences elemen-

    tary displacements (dx,dy,dz) is

    dEP dW Fx dx Fy dy Fzdz (1:84)The value of a physical force may vary in the space around its source. For

    example, gravitational and electrical forces decrease with distance from a

    source mass or electrical charge, respectively. The region in which a physical

    1.5 Conservative force,eld, and potential 17

  • 7/26/2019 Guide to Geophysical Equations

    33/296

  • 7/26/2019 Guide to Geophysical Equations

    34/296

  • 7/26/2019 Guide to Geophysical Equations

    35/296

  • 7/26/2019 Guide to Geophysical Equations

    36/296

  • 7/26/2019 Guide to Geophysical Equations

    37/296

  • 7/26/2019 Guide to Geophysical Equations

    38/296

  • 7/26/2019 Guide to Geophysical Equations

    39/296

  • 7/26/2019 Guide to Geophysical Equations

    40/296

  • 7/26/2019 Guide to Geophysical Equations

    41/296

    is distributed in the volume with mean density , a volume integral can replace

    the enclosed mass:

    ZZZV

    r aG dV 4GZZZ

    V

    dV (1:114)

    ZZZV

    r aG 4G dV0 (1:115)

    For this to be generally true the integrand must be zero. Consequently,

    r aG

    4G (1:116)

    The gravitational acceleration is the gradient of the gravitational potential UGas

    in (1.88):

    r rUG 4G (1:117)

    r2UG4G (1:118)Equation (1.118) is known as Poissons equation, after Simon-Denis Poisson

    (17811840), a French mathematician and physicist. It describes the gravita-

    tional potential of a mass distribution at a point that is within the mass distri-

    bution. For example, it may be used to compute the gravitational potential at a

    point inside the Earth.

    1.9 Laplaces equation

    Another interesting case is the potential at a point outside the mass distribution.

    Let S be a closed surface outside the point mass m. The radius vectorrfrom the

    point mass mnow intersects the surface S at two points A and B, where it forms

    angles1and2with the respective unit vectorsn1andn2normal to the surface

    (Fig. 1.10). Leterbe a unit vector in the radial direction. Note that the outward

    normaln1forms an obtuse angle with the radius vector at A. The gravitational

    acceleration at A is a1and its ux through the surface area dS1is

    dN1

    a1 n1dS1

    Gm

    r2

    1 r n1

    dS1 (1:119)

    dN1 Gmr21

    cos 1 dS1Gm cos 1dS1

    r21(1:120)

    26 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    42/296

    dN1Gm d (1:121)The gravitational acceleration at B is a2 and its ux through the surface area dS2 is

    dN2 a2 n2dS2 Gm cos 2dS2r22

    (1:122)

    dN2 Gm d (1:123)The total contribution of both surfaces to the gravitational ux is

    dNdN1 dN20 (1:124)Thus, the total ux of the gravitational acceleration aGthrough a surface S that

    does not include the point massm is zero. By invoking the divergence theorem

    we have for this situationZV

    r aG dVZS

    aG n dS0 (1:125)

    For this result to be valid for any volume, the integrand must be zero:

    r aG r rUG 0 (1:126)

    r2UG 0 (1:127)Equation (1.127) is Laplaces equation, named after Pierre Simon, Marquis de

    Laplace (17491827), a French mathematician and physicist. It describes the

    gravitational potential at a point outside a mass distribution. For example, it is

    applicable to the computation of the gravitational potential of the Earth at an

    external point or on its surface.

    In Cartesian coordinates, which are rectilinear, Laplaces equation has the

    simple form

    S

    m a1

    n2

    d

    dS1

    1

    2

    dS2n1

    a2A B

    er

    Fig. 1.10. Representation of the gravitational ux through a closed surface S that

    does not enclose the source of the ux (the point massm).

    1.9 Laplaces equation 27

  • 7/26/2019 Guide to Geophysical Equations

    43/296

  • 7/26/2019 Guide to Geophysical Equations

    44/296

    f0 a0; dfdx

    x0

    a1

    d2

    fdx2

    x0

    2a2; d3

    fdx3

    x0

    3 2 a3dnf

    dxn

    x0

    nn 1n 2 . . . 3 2 1 ann!an

    (1:133)

    On inserting these values for the coefcients into (1.130) we get the power

    series for(x):

    f

    x

    f

    0

    x

    df

    dx

    x0

    x2

    2!

    d2f

    dx

    2 x0

    x3

    3!

    d3f

    dx

    3 x0

    xn

    n!

    dnf

    dxn

    x0

    (1:134)

    This is the MacLaurin series for(x) about the origin,x = 0. It was derived in the

    eighteenth century by the Scottish mathematician Colin MacLaurin (1698

    1746) as a special case of a Taylor series.

    The MacLaurin series is a convenient way to derive series expressions for

    several important functions. In particular,

    sin xx x3

    3! x

    5

    5! x

    7

    7! 1 n1 x

    2n1

    2n 1 !

    cos x1 x2

    2! x

    4

    4! x

    6

    6! 1 n1 x

    2n2

    2n 2 !

    expx ex 1 x x2

    2! x

    3

    3! x

    n1

    n 1 !

    ln 1

    x

    loge 1

    x

    x

    x2

    2

    x3

    3

    x4

    4 1

    n1 xn

    n

    (1:135)

    1.10.2 Taylor series

    We can write the power series in (1.134) for(x) centered on any new origin, for

    examplex = x0. To do this we substitute (x x0) forx in the above derivation.

    The power series becomes

    fx fx0 x x0 dfdx xx0

    x x0 2

    2!

    d2f

    dx2 xx0 x x0

    3

    3!

    d3f

    dx3

    xx0

    x x0 n

    n!

    dnf

    dxn

    xx0

    (1:136)

    1.10 Power series 29

  • 7/26/2019 Guide to Geophysical Equations

    45/296

    This is called a Taylor series, after an English mathematician, Brooks Taylor

    (16851731), who described its properties in 1712.

    The MacLaurin and Taylor series are both approximations to the function

    (x). The remainder between the true function and its power series is a measure

    of how well the function is expressed by the series.

    1.10.3 Binomial series

    Finite series

    An important series is the expansion of the function (x) = (a + x)n. Ifn is a

    positive integer, the expansion of(x) is a nite series, terminating after (n + 1)

    terms. Evaluating the series for some low values ofn gives the following:

    n0 : a x 0 1n1 : a x 1 a xn2 : a x 2 a2 2ax x2n3 : a x 3 a3 3a2x 3ax2 x3n4 : a x 4 a4 4a3x 6a2x2 4ax3 x4

    (1:137)

    The general expansion of(x) is therefore

    a x n an nan1x n n 1 1 2 a

    n2x2

    n n 1 . . . n k 1 k!

    ankxk xn(1:138)

    The coefcient of the general kth term is equivalent to

    n n 1 . . . n k 1 k!

    n!

    k! n k ! (1:139)

    This is called thebinomial coefcient.

    When the constanta is equal to 1 and n is a positive integer, we have the

    useful series expansion

    1 x nXnk0

    n!

    k! n k !xk (1:140)

    Innite series

    If the exponent in (1.140) is not a positive integer, the series does not terminate,

    but is an innite series. The series for(x) = (1 +x)p, in which the exponentpis

    not a positive integer, may be derived as a MacLaurin series:

    30 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    46/296

  • 7/26/2019 Guide to Geophysical Equations

    47/296

    1.11 Leibnizs rule

    Assume thatu(x) and v(x) are differentiable functions ofx. The derivative oftheir product is

    d

    dx u x v x u x dv x

    dx v x du x

    dx (1:144)

    If we dene the operatorD = d/dx, we obtain a shorthand form of this equation:

    D uv u Dv v Du (1:145)We can differentiate the product (uv) a second time by parts,

    D2 uv D D uv u D2v Du Dv Dv Du v D2uu D2v 2 Du Dv v D2u (1:146)

    and, continuing in this way,

    D3 uv u D3v 3 Du D2v 3 D2u Dv v D3uD4 uv u D4v 4 Du D3v

    6 D2u

    D2v

    4 D2u

    Dv v D4u

    (1:147)

    The coefcients in these equations are the binomial coefcients, as dened in

    (1.139). Thus aftern differentiations we have

    Dn uv Xnk0

    n!

    k! n k ! Dku

    Dnkv

    (1:148)

    This relationship is known as Leibnizs rule, after Gottfried Wilhelm Leibniz

    (16461716), who invented innitesimal calculus contemporaneously with

    Isaac Newton (16421727); each evidently did so independently of the other.

    1.12 Legendre polynomials

    LetrandRbe the sides of a triangle that enclose an angle and letube the side

    opposite this angle (Fig. 1.11). The angle and sides are related by the cosine rule

    u2 r2 R2 2rR cos (1:149)Inverting this expression and taking the square root gives

    1

    u 1

    R

    "1 2

    r

    R

    !cos

    r

    R

    !2#1=2(1:150)

    32 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    48/296

  • 7/26/2019 Guide to Geophysical Equations

    49/296

    Equation (1.156) is known as the generating function for the polynomials Pn(x).

    Using this result, and substituting h = r/R and x = cos , we nd that (1.151)

    becomes

    1

    u 1

    R

    X1n0

    r

    R

    !nPn cos (1:157)

    The polynomialsPn(x) orPn(cos ) are called Legendre polynomials, after the

    French mathematician Adrien-Marie Legendre (17521833). The dening

    equation (1.157) is called the reciprocal-distance formula. An alternative for-

    mulation is given inBox 1.4.

    1.13 The Legendre differential equation

    The Legendre polynomials satisfy an important second-order partial differential

    equation, which is called the Legendre differential equation. To derive this

    equation we will carry out a sequence of differentiations, starting with the

    generating function in the form

    1 2xh h2 1=2 (1:158)Differentiating this function once with respect toh gives

    h x h 1 2xh h2 3=2 x h 3 (1:159)

    Differentiating twice with respect to x gives

    xh 1

    2xh

    h2 3=2 h

    3

    3 1

    h

    x

    (1:160)

    2

    x2 3h2

    x3h25

    5 1

    3h2

    2

    x2

    (1:161)

    Next we perform successive differentiations of the product (h) with respect to

    h. The rst gives

    h h h

    h h x h 3 (1:162)

    34 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    50/296

    On repeating the differentiation, and taking (1.159) into account, we get

    2

    h2 h

    h h x h 32

    h x 2h 3

    x h 3 3h x h 25 x 2h 3 2x 3h 3 3h x h 25 (1:163)

    Now substitute for3, from (1.160), and 5, from (1.161), giving

    Box 1.4. Alternative form of the reciprocal-distance formula

    The sides and enclosed angle of the triangle in Fig. 1.11are related by the

    cosine rule

    u2 r2 R2 2rR cos (1)

    Instead of takingRoutside the brackets as in (1.150), we can moveroutside

    and write the expression foru as

    1

    u

    1

    r

    1

    2

    R

    r cos

    R

    r

    2

    " #1=2

    (2)

    Following the same treatment as inSection 1.12, but now withhR=randx = cos , we get

    1

    u1

    r 1 2xh h2 1=2 1

    r 1 t 1=2 (3)

    wheret= 2xh h2. The function (1 t)1/2 is expanded as a binomial series,

    which again gives an innite series inh, in which the coefcient ofhn

    isPn(x). The dening equation is as before:

    x; h 1 2xh h2 1=2 X1n0

    hnPn x (4)

    On substitutinghR=rand x = cos , we nd an alternative form for thegenerating equation for the Legendre polynomials:

    1u1

    r

    X1n0

    Rr

    nPn cos (5)

    1.13 The Legendre differential equation 35

  • 7/26/2019 Guide to Geophysical Equations

    51/296

    2

    h2 h 2x 3h 1

    h

    x

    3h x h 2 1

    3h2

    2

    x2

    (1:164)

    Multiply throughout byh:

    h

    2

    h2 h 2x 3h

    x

    x h 2

    2

    x2

    2x x

    3h

    x

    x h 2

    2

    x2

    (1:165)

    The second term on the right can be replaced as follows, again using (1.160) and

    (1.161):

    3h

    x3h23 1

    2

    2

    x2

    1 2xh h2 2x2

    (1:166)

    On substituting into (1.165) and gathering terms, we get

    h 2

    h2

    h

    x

    h

    2

    1

    2xh

    h2 h i

    2

    x2 2x

    x (1:167)

    h

    2

    h2 h x2 1 2

    x2

    2x

    x

    (1:168)

    The Legendre polynomialsPn(x) are dened in (1.156) as the coefcients ofhn

    in the expansion of as a power series. On multiplying both sides of (1.156) by

    h, we get

    h X

    1

    n0 hn

    1

    Pnx (1:169)We differentiate this expression twice and multiply by h to get a result that can

    be inserted on the left-hand side of (1.168):

    h h

    X1n0

    n 1 hnPnx (1:170)

    h

    2

    h2 h X

    1

    n0n n

    1

    hnP

    nx

    (1:171)

    Using (1.156), we can now eliminate and convert (1.168) into a second-order

    differential equation involving the Legendre polynomials Pn(x),

    36 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    52/296

  • 7/26/2019 Guide to Geophysical Equations

    53/296

  • 7/26/2019 Guide to Geophysical Equations

    54/296

    Thus

    d

    dx 1

    x2 PlP0

    nP0

    l

    Pn n n 1 l l 1 PlPn0 (1:186)Now integrate each term in this equation with respect to x over the range

    1x 1. We get

    1 x2 PlP0n P0lPn 1x1 n n 1 l l 1 Z1

    x1PlPn dx0

    (1:187)

    The rst term is zero on evaluation of (1 x2) atx = 1; thus the second termmust also be zero. Forn l the condition for orthogonality of the Legendre

    polynomials is Z 1x1

    PnxPlxdx0 (1:188)

    1.13.2 Normalization of the Legendre polynomials

    A function is said to be normalized if the integral of the square of the function overits range is equal to 1. Thus we must evaluate the integral

    R1x1 Pn x 2 dx. We

    begin by recalling the generating function for the Legendre polynomials given in

    (1.156), which we rewrite forPn(x) andPl(x) individually:X1n0

    hnPn x 1 2xh h2 1=2

    (1:189)

    X1l0 h

    l

    Pl x 1 2xh h2 1=2 (1:190)

    Multiplying these equations together gives

    X1l0

    X1n0

    hnlPn x Pl x 1 2xh h2 1

    (1:191)

    Now let l= n and integrate both sides with respect to x, taking into account

    (1.188):

    X1n0

    h2nZ1

    x1Pn x 2 dx

    Z1x1

    dx

    1 h2 2xh (1:192)

    1.13 The Legendre differential equation 39

  • 7/26/2019 Guide to Geophysical Equations

    55/296

    The right-hand side of this equation is a standard integration that results in a

    natural logarithm:

    Z dxa bx

    1b

    ln a bx (1:193)

    The right-hand side of (1.192) therefore leads to

    Z1x1

    dx

    1 h2 2xh 1

    2h ln 1 h2 2xh

    1

    x1

    1

    2h ln 1 h2

    2h ln 1 h2 2h (1:194)and

    Z1x1

    dx

    1 h2 2xh 1

    h 1

    2ln 1 h 2 1

    2ln 1 h 2

    1h

    ln 1 h ln 1 h (1:195)

    Using the MacLaurin series for the natural logarithms as in (1.135), we get

    ln 1 h h h2

    2 h

    3

    3 h

    4

    4 1 n1h

    n

    n (1:196)

    ln 1 h h h2

    2 h

    3

    3 h

    4

    4 1 n1h

    n

    n (1:197)

    Subtracting the second equation from the rst gives

    Z1x1

    dx

    1 h2 2xh2

    h h h

    3

    3 h

    5

    5

    2

    h

    X1n0

    h2n1

    2n 1 (1:198)

    Inserting this result into (1.192) gives

    X1n0

    h2nZ1

    x1Pn x 2 dx 2

    h

    X1n0

    h2n1

    2n 1 (1:199)

    X1n0

    h2nZ1

    x1Pn x 2 dx 2

    2n 1

    0@

    1A0 (1:200)

    40 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    56/296

  • 7/26/2019 Guide to Geophysical Equations

    57/296

    x2 1 Dn2f 2x n 1 Dn1f 2 n 1n12

    Dnf2nx Dn1f

    2n n

    1

    Dnf

    (1:207)

    On gathering terms and bringing all to the left-hand side, we have

    x2 1 Dn2f 2x Dn1f nn 1Dnf0 (1:208)Now we deney(x) such that

    y x

    Dnf

    dn

    dxn

    x2

    1 n (1:209)

    and we have

    x2 1 d2ydx2

    2x dydx

    nn 1y0 (1:210)

    On comparing with (1.174), we see that this is the Legendre equation. The

    Legendre polynomials must therefore be proportional to y(x), so we can write

    Pnx cndn

    dxn x

    2

    1 n (1:211)The quantitycnis a calibration constant. To determine cnwe rst write

    dn

    dxn x2 1 n dn

    dxn x 1 n x 1 n (1:212)

    then we apply Leibnizs rule to the product on the right-hand side of the

    equation:

    dndxn

    x2 1 n Xnm0

    n!m! n m !

    dmdxm

    x 1 ndnmdxnm

    x 1 n (1:213)

    The successive differentiations of (x 1)n give

    d

    dx x 1 n n x 1 n1

    d2

    dx2 x 1 n n n 1 x 1 n2

    dn1dxn1

    x 1 n n n 1 n 2 . . . 321 x 1 n! x 1 dn

    dxn x 1 n n!

    (1:214)

    42 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    58/296

  • 7/26/2019 Guide to Geophysical Equations

    59/296

    1 x2 ddx

    P00n 22x d

    dxP0n n n 1 12

    d

    dxPn0 (1:222)

    Next, we differentiate this expression again, observing the same rules andgathering terms,

    1x2 d2dx2

    P00n 2x d

    dxP00n

    4 d

    dxP0n 4x

    d2

    dx2P0n

    n n 1 2 d

    2

    dx2Pn 0

    (1:223)

    1 x2

    d2

    dx2P00n 2x

    d2

    dx2P0n 4x

    d2

    dx2P0n n n 1 2 4

    d2

    dx2Pn0

    (1:224)

    1 x2 d2dx2

    P00n 6xd2

    dx2P0n n n 1 6

    d2

    dx2Pn0 (1:225)

    which, as we did with (1.222), we can write in the form

    1 x2 d2dx2

    P00n 23xd2

    dx2P0n n n 1 23

    d2

    dx2Pn0 (1:226)

    On following step-by-step in the same manner, we get after the third

    differentiation

    1 x2 d3dx3

    P00n 2 4 xd3

    dx3P0n n n 1 3 4

    d3

    dx3Pn0 (1:227)

    Equations (1.222), (1.226), and (1.227) all have the same form. The higher-

    order differentiation is accompanied by systematically different constants. By

    extension, differentiating (1.219)m times (wherem n) yields the differential

    equation

    1 x2 dmdxm

    P00n 2 m 1 x dm

    dxmP0n n n 1 m m 1

    dm

    dxmPn0(1:228)

    Now let themth-order differentiation ofPnbe written as

    dm

    dxmPnx Qx

    1

    x2

    m=2

    (1:229)

    Substitution of this expression into (1.228) gives a new differential equation

    involving Q(x). We need to determine bothdm=dxmP0n anddm=dxmP00n , sorst we differentiate (1.229) with respect tox:

    44 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    60/296

  • 7/26/2019 Guide to Geophysical Equations

    61/296

    The functionsQ(x) involve two parameters, the degreenand orderm, and are

    writtenPn,m(x). Thus

    1 x2 d2dx2

    Pn;m x 2x ddx

    Pn;m x n n 1 m2

    1 x2

    Pn;m x 0

    (1:237)

    This is the associated Legendre equation. The solutions Pn,m(x) orPn,m(cos ),

    wherex = cos , are calledassociated Legendre polynomials, and are obtained

    from the ordinary Legendre polynomials using the denition ofQ in (1.229):

    Pn;mx 1 x2 m=2 dm

    dxmPnx (1:238)

    Substituting Rodriguesformula (1.218) forPn(x) into this equation gives

    Pn;mx 1 x2 m=2

    2nn!

    dnm

    dxnm x2 1 n (1:239)

    The highest power ofx in the function (x2 1)n isx2n. After 2ndifferentiations

    the result will be a constant, and a further differentiation will give zero.Therefore n + m 2n, and possible values ofm are limited to the range 0 m n.

    1.15.1 Orthogonality of associated Legendre polynomials

    For succinctness we again writePn,m(x) as simplyPn,m. The dening equations

    for the associated Legendre polynomialsPn,mand Pl,mare

    1 x2 Pn;m 00 2x Pn;m 0 n n 1 m21 x2 Pn;m0 (1:240)1 x2 Pl;m 00 2x Pl;m 0 l l 1 m2

    1 x2

    Pl;m0 (1:241)

    As for the ordinary Legendre polynomials, we multiply (1.240) by Pl,m and

    (1.241) byPn,m:

    1 x2

    Pn;m 00Pl;m 2x Pn;m 0Pl;m n n 1 m

    2

    1 x2 Pn;mPl;m0(1:242)

    46 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    62/296

  • 7/26/2019 Guide to Geophysical Equations

    63/296

  • 7/26/2019 Guide to Geophysical Equations

    64/296

    1.16 Spherical harmonic functions

    Several geophysical potential

    elds

    for example, gravitation and geomagnet-ism satisfy the Laplace equation. Spherical polar coordinates are best suited

    for describing a global geophysical potential. The potential can vary with

    distance r from the Earths center and with polar angular distance and

    azimuth (equivalent to co-latitude and longitude in geographic terms)

    on any concentric spherical surface. The solution of Laplaces equation in

    spherical polar coordinates for a potential U may be written (see Section

    2.4.5,(2.104))

    UX1n0

    Xnm0

    Anrn Bn

    rn1

    amn cosm bmn sinm

    Pmn cos (1:252)

    HereAn,Bn,amn, and b

    mn are constants that apply to a particular situation. On the

    surface of the Earth, or an arbitrary sphere, the radial part of the potential of a

    point source at the center of the sphere has a constant value and the variation

    over the surface of the sphere is described by the functions in and . We are

    primarily interested in solutions outside the Earth, for which An is zero. Also we

    can set the constantBnequal toRn+1

    , whereRis the Earths mean radius. Thepotential is then given by

    UX1n0

    Xnm0

    R

    r

    n1amn cosm bmn sinm

    Pmn cos (1:253)

    Let thespherical harmonic functionsCmn ; and Smn ; be dened as

    C

    m

    n ; cosm Pm

    n cos Smn ; sinm Pmn cos

    (1:254)

    The variation of the potential over the surface of a sphere may be described by

    these functions, or a more general spherical harmonic function Ymn ; thatcombines the sine and cosine variations:

    Ymn ; Pmn cos cosm

    sin

    m

    (1:255)

    Like their constituent parts the sine, cosine, and associated Legendre

    functions spherical harmonic functions are orthogonal and can be

    normalized.

    1.16 Spherical harmonic functions 49

  • 7/26/2019 Guide to Geophysical Equations

    65/296

    1.16.1 Normalization of spherical harmonic functions

    Normalization of the functions Cmn ; and Smn ; requires integrating thesquared value of each function over the surface of a unit sphere. The element ofsurface area on a unit sphere is d = sin dd (Box 1.3) and the limits of

    integration are 0 and 0 2. The integral is

    ZZS

    Cmn ; 2

    d Z

    0

    Z20

    Cmn ; 2

    sin dd

    Z

    0

    Z20

    cosmPmn cos 2 sin dd (1:256)

    Letx = cos in the associated Legendre polynomial, so thatdx = sin dand

    the limits of integration are 1 x 1. The integration becomes

    Z1

    x1 Z2

    0

    cos2md

    8>:

    9>=>; P

    mn x

    2dx Z

    1

    x1

    Pmn x 2

    dx (1:257)

    Normalization of the associated Legendre polynomials gives the result in

    (1.248), thus

    ZZS

    Cmn ; 2

    d 22n 1

    n m!n m ! (1:258)

    The normalization of the function Smn ; by this method delivers the sameresult.

    Spherical harmonic functions make it possible to express the variation of a

    physical property (e.g., gravity anomalies,g(, )) on the surface of the Earth as

    an innite series, such as

    g ; X1n0

    Xnm0

    amnCmn ; bmnSmn ;

    (1:259)

    The coefcients amn and bmn may be obtained by multiplying the functiong ;

    by Cmn ; or Smn ; , respectively, and integrating the product over thesurface of the unit sphere. The normalization properties give

    50 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    66/296

  • 7/26/2019 Guide to Geophysical Equations

    67/296

    nodal lines of latitude and the longitudinal lines separate sectors in which the

    potential is greater or less than the uniform value. An example of a sectorial

    spherical harmonic is Y5

    5

    ;

    , shown inFig. 1.12(b).

    In the general case (m 0, n m) the potential varies with both latitude and

    longitude. There aren m nodal lines of latitude andm nodal great circles (2m

    meridians) of longitude. The appearance of the spherical harmonic resembles a

    patchwork of alternating regions in which the potential is greater or less than the

    uniform value. An example of atesseralspherical harmonic isY45 ; , whichis shown inFig. 1.12(c).

    1.17 Fourier series, Fourier integrals, andFourier transforms

    1.17.1 Fourier series

    Analogously to the representation of a continuous function by a power series

    (Section 1.10), it is possible to represent a periodic function by an innite sum

    of terms consisting of the sines and cosines of harmonics of a fundamental

    frequency. Consider a periodic function(t) with periodthat is dened in the

    interval 0 t , so that (a)(t) is

    nite within the interval; (b) (t) is periodicoutside the interval, i.e., (t + ) = (t); and (c) (t) is single-valued in the

    interval except at a nite number of points, and is continuous between these

    points. Conditions (a)(c) are known as the Dirichlet conditions. If they are

    satised,(t) can be represented as

    ft a02

    X1n1

    ancos nt bnsin nt (1:262)

    where = 2/ and the factor 1

    2 in the rst term is included for reasons ofsymmetry. This representation of (t) is known as a Fourier series. The

    orthogonal properties of sine and cosine functions allow us to nd the coef-

    cientsan andbn of thenth term in the series by multiplying (1.262) by sin(nt)

    or cos(nt) and integrating over a full period:

    an2

    Z=2t

    =2

    ftcos nt dt

    bn2

    Z=2t=2

    ftsin nt dt(1:263)

    52 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    68/296

    Instead of using trigonometric functions, we can replace the sine and

    cosine terms with complex exponentials using the denitions in (1.7), i.e., we

    write

    cos nt exp int expint 2

    sin nt exp int expint 2i

    (1:264)

    Using these relationships in (1.262) yields

    ft X1n0

    an2

    exp int expint bn2i

    exp int expint

    X1n0

    an ibn2

    exp int

    X1n0

    an ibn2

    expint

    (1:265)

    The summation indices are dummy variables, so in the second sum we can

    replacen by n, and extend the limits of the sum to n = ; thus

    ft X1n0

    an ibn2

    exp int

    X0

    n1

    an ibn2

    exp int

    X1

    n1

    an an i bn bn 2

    exp int (1:266)

    If we denecnas the complex number

    cn an

    a

    n

    i bn

    b

    n

    2 (1:267)the Fourier series (1.262) can be written in complex exponential form as

    ft X1

    n1cnexp int (1:268)

    In this case the harmonic coefcientscnare given by

    cn1

    Z=2t=2

    ftexpint dt (1:269)

    1.17 Fourier series, integrals, and transforms 53

  • 7/26/2019 Guide to Geophysical Equations

    69/296

  • 7/26/2019 Guide to Geophysical Equations

    70/296

    Box 1.5. Transition from Fourier series to Fourier integral

    The complex exponential Fourier series for a function(t) is

    ft X1

    n1cnexp int (1)

    where the complex coefcientscnare given by

    cn1

    Z=2

    t=2

    ftexpint dt (2)

    In these expressions = 2/is the fundamental frequency and is the

    fundamental period. From one value ofn to the next, the harmonic frequency

    changes by = 2/, so the factor preceding the second equation can be

    replaced by 1/= /(2). To avoid confusion when we insert (2) into (1), we

    change the dummy variable of the integration tou, giving

    cn

    2Z=2

    u=2fuexpinu du (3)

    After insertion, (1) becomes

    ft X1

    n1

    2

    Z=2u=2

    fuexpinu du

    0B@

    1CA

    exp int

    X1

    n1

    2

    Z=2u=2

    fuexp in t u du0B@

    1CA (4)

    We now dene the function within the integral as

    F Z=2

    u=2

    fuexp in t u du (5)

    The initial Fourier series becomes

    1.17 Fourier series, integrals, and transforms 55

  • 7/26/2019 Guide to Geophysical Equations

    71/296

    f t 12 X

    1

    1

    F (6)

    We now let the incremental frequency become very small, tending in the

    limit to zero; this is equivalent to letting the periodbecome innite. The

    indexnis dropped becauseis now a continuous variable; the discrete sum

    becomes an integral and the functionf(t) is

    f t 12

    Z1

    1

    F d (7)

    while the functionF() from (5) becomes

    F Z1

    u1fuexp i t u du (8)

    On insertingF() into (7) we get

    f t 12

    Z11

    Z1u1

    fuexp i t u du24 35d

    12

    Z11

    Z1u1

    fuexpiu du24

    35 exp it d (9)

    The quantity in square brackets, on changing the variable from uback tot, is

    g 12

    Z1t1

    ftexpit dt (10)

    and the original expression can now be written

    f t Z1

    1g exp it d (11)

    The equivalence of these two equations is known as the Fourier integral

    theorem.

    56 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    72/296

    Fourier series that represent odd or even functions consist of sums of sines or

    cosines, respectively. In the same way, there are sine and cosine Fourier integrals

    that represent odd and even functions, respectively. Suppose that the function (t)

    iseven, and let us replace the complex exponential in (1.272) using(1.5):

    g 12

    Z1t1

    ftcos t isin t dt (1:273)

    The sine function is odd, so, if(t) is even, the product(t)sin(t) is odd, and the

    integral of the second term is zero. The product(t)cos(t) is even, and we can

    convert the limits of integration to the positive interval:

    g 12

    Z1t1

    ftcos t dt

    1

    Z1t0

    ftcos t dt (1:274)

    Thus, if(t) is even, then g() is also even. Similarly, one nds that, if(t) is

    odd,g() is also odd.

    Now we expand the exponential in (1.271) and apply the same conditions of

    evenness and oddness to the products:

    ft Z1

    1g cos t isin t d

    2Z1

    0

    g cos t d (1:275)

    If we were to substitute (1.275) back into (1.274), the integration would be

    preceded by a constant 2/, the product of the two constants in these equations.

    Equations (1.275) and (1.274) form a Fourier-transform pair, and it does not

    matter how the factor 2/ is divided between them. We will associate it here

    entirely with the second equation, so that we have the pair of equations

    ft Z1

    0

    g cos t d

    g 2

    Z1t0

    ftcos t dt(1:276)

    1.17 Fourier series, integrals, and transforms 57

  • 7/26/2019 Guide to Geophysical Equations

    73/296

    The even functions(t) andg() areFourier cosine transformsof each other.

    A similar treatment for a function (t) that is oddleads to a similar pair of

    equations in which the Fourier transform g() is alsooddand

    ft Z1

    0g sin t d

    g 2

    Z1t0

    ftsin t dt(1:277)

    The odd functions(t) andg() areFourier sine transformsof each other.

    f u r t h e r r e a d i n g

    Boas, M. L. (2006).Mathematical Methods in the Physical Sciences, 3rd edn. Hoboken,

    NJ: Wiley, 839 pp.

    James, J. F. (2004). A Students Guide to Fourier transforms, 2nd edn. Cambridge:

    Cambridge University Press, 135 pp.

    58 Mathematical background

  • 7/26/2019 Guide to Geophysical Equations

    74/296

    2

    Gravitation

    2.1 Gravitational acceleration and potential

    The Universal Law of Gravitation deduced by Isaac Newton in 1687 describes

    the force of gravitational attraction between two point masses m and Msepa-

    rated by a distancer. Let a spherical coordinate system (r,,) be centered on

    the point mass M. The force of attraction F exerted on the point mass m acts

    radially inwards towards M, and can be written

    F G mMr2

    er (2:1)

    In this expression, Gis the gravitational constant (6.674 21 1011 m3 kg1 s2),

    eris the unit radial vector in the direction of increasing r, and the negative sign

    indicates that the force acts inwardly, towards the attracting mass. The gravita-

    tional accelerationaGat distanceris the force on a unit mass at that point:

    aG GM

    r2 er (2:2)

    The acceleration aGmay also be written as the negative gradient of a gravita-

    tional potentialUG

    aG rUG (2:3)

    The gravitational acceleration for a point mass is radial, thus the potential

    gradient is given by

    UGr

    G Mr2

    (2:4)

    UG GM

    r (2:5)

    59

  • 7/26/2019 Guide to Geophysical Equations

    75/296

    In Newtons time the gravitational constant could not be veried in a laboratory

    experiment. The attraction between heavy masses of suitable dimensions is weak

    and the effects of friction and air resistance relatively large, so the rst successful

    measurement of the gravitational constant by Lord Cavendish was not made until

    more than a century later, in 1798. However, Newton was able to conrm the

    validity of the inverse-square law of gravitation in 1687 by using existing

    astronomic observations of the motions of the planets in the solar system.

    These had been summarized in three important laws by Johannes Kepler in

    1609 and 1619. The small sizes of the planets and the Sun, compared with the

    immense distances between them, enabled Newton to consider these as point

    masses and this allowed him to verify the inverse-square law of gravitation.

    2.2 Keplers laws of planetary motion

    Johannes Kepler (15711630), a German mathematician and scientist, formu-

    lated his laws on the basis of detailed observations of planetary positions by

    Tycho Brahe (15461601), a Danish astronomer. The observations were made

    in the late sixteenth century, without the aid of a telescope. Kepler found that the

    observations were consistent with the following three laws (Fig. 2.1).

    1. The orbit of each planet is an ellipse with the Sun at one focus.

    2. The radius from the Sun to a planet sweeps over equal areas in equal intervals

    of time.

    S

    Q

    Q*

    p

    a

    b

    r P(r,)

    P*

    aphelion perihelion

    Fig. 2.1. Illustration of Keplers laws of planetary motion. The orbit of each planetis an ellipse with the Sun at its focus (S); a,b, andpare the semi-major axis, semi-

    minor axis, and semi-latus rectum, respectively. The area swept by the radius to a

    planet in a given time is constant (i.e., area SPQ equals area SP*Q*); the square of

    the period is proportional to the cube of the semi-major axis. After Lowrie (2007).

    60 Gravitation

  • 7/26/2019 Guide to Geophysical Equations

    76/296

    3. The square of the period is proportional to the cube of the semi-major axis of

    the orbit.

    The fundamental assumption is that the planets move under the inuence of acentral, i.e., radially directed force. For a planet of mass mat distancerfrom the

    Sun the forceF can be written

    F md2r

    dt2 frer (2:6)

    The angular momentumh of the planet about the Sun is

    h

    r

    m

    dr

    dt (2:7)

    Differentiating with respect to time, the rate of change of angular momentum is

    dh

    dt m

    d

    dt r

    dr

    dt

    m

    dr

    dt

    dr

    dt

    m r

    d2r

    dt2

    (2:8)

    The rst term on the right-hand side is zero, because the vector product of a

    vector with itself (or with a vector parallel to itself) is zero. Thus

    dh

    dt r m

    d2r

    dt2 (2:9)

    On substituting from (2.6) and applying the same condition, we have

    dh

    dt r frer fr r er 0 (2:10)

    This equation means thath is a constant vector; the angular momentum of the

    system is conserved. On taking the scalar product ofh and r, we obtain

    r h r r mdr

    dt

    (2:11)

    Rotating the sequence of the vectors in the triple product gives

    r h mdr

    dt r r 0 (2:12)

    This result establishes that the vectorr describing the position of a planet isalways perpendicular to its constant angular momentum vectorh and therefore

    denes a plane. Every planetary orbit is therefore a plane that passes through the

    Sun. The orbit of the Earth denes theeclipticplane.

    2.2 Keplers laws of planetary motion 61

  • 7/26/2019 Guide to Geophysical Equations

    77/296

    2.2.1 Keplers Second Law

    Let the position of a planet in its orbit be described by polar coordinates ( r,)

    with respect to the Sun. The coordinates are dened so that the angleis zero atthe closest approach of the planet to the Sun (perihelion). The angular momen-

    tum at an arbitrary point of the orbit has magnitude

    h mr2d

    dt (2:13)

    In a short interval of time t the radius vector from the Sun to the planet

    moves through a small angle and denes a small triangle. The area A of

    the triangle is

    DA 1

    2r2 D (2:14)

    The rate of change of the area swept over by the radius vector is

    dA

    dt lim

    Dt!0

    DA

    Dt

    limDt!0

    1

    2r2D

    Dt

    (2:15)

    dA

    dt

    1

    2 r

    2d

    dt (2:16)

    On inserting from (2.13) we get

    dA

    dt

    h

    2m (2:17)

    Thus the area swept over by the radius vector in a given time is constant. This is

    Keplers Second Law of planetary motion.

    2.2.2 Keplers First Law

    If just the gravitational attraction of the Sun acts on the planet (i.e., we ignore the

    interactions between the planets), the total energy Eof the planet is constant.

    The total energy E is composed of the planets orbital kinetic energy and its

    potential energy in the Suns gravitational eld:

    1

    2 m

    dr

    dt 2

    1

    2 mr2 d

    dt 2

    Gm

    S

    r E (2:18)

    The rst term here is the planets linear (radial) kinetic energy, the second term

    is its rotational kinetic energy (with mr2 being the planets moment of inertia

    62 Gravitation

  • 7/26/2019 Guide to Geophysical Equations

    78/296

    about the Sun), and the third term is the gravitational potential energy. On

    writing

    drdt

    drd

    ddt

    (2:19)

    and rearranging terms we get

    dr

    d

    2d

    dt

    2 r2

    d

    dt

    2 2G

    S

    r 2

    E

    m (2:20)

    Now, to simplify later steps, we make a change of variables, writing

    u 1r

    (2:21)

    Then

    dr

    d

    d

    d

    1

    u

    1

    u2du

    d

    r2

    du

    d

    (2:22)

    Substituting from (2.22) into (2.20) gives

    r2ddt

    2dud

    2 r2 ddt

    2 2G Sr 2Em (2

    :23)

    With the result of (2.13) we have

    r2d

    dt

    h

    m

    r d

    dt

    1

    r

    h

    m

    u

    h

    m

    (2:24)

    On replacing these expressions, ( 2.23) becomes

    h

    m

    2du

    d

    2 u2

    h

    m

    2 2uGS 2

    E

    m (2:25)

    du

    d

    2 u2 2uGS

    m2

    h2 2

    Em

    h2 (2:26)

    The rest of the evaluation is straightforward, if painstaking. First we add a

    constant to each side,

    du

    d

    2 u2 2uGS

    m2

    h2 GS

    m2

    h2

    2 2

    Em

    h2 GS

    m2

    h2

    2(2:27)

    2.2 Keplers laws of planetary motion 63

  • 7/26/2019 Guide to Geophysical Equations

    79/296

    du

    d

    2 u GS

    m2

    h2

    2 2

    Em

    h2 GS

    m2

    h2

    2(2:28)

    Next, we move the second term to the right-hand side of the equation, giving

    du

    d

    2 2

    Em

    h2 GS

    m2

    h2

    2 u GS

    m2

    h2

    2(2:29)

    du

    d

    2 GS

    m2

    h2

    21

    2Eh2

    G2S2m3

    u GS

    m2

    h2

    2(2:30)

    Now, we dene some combinations of these terms, as follows:

    u0 GSm2

    h2 (2:31)

    e2 1 2Eh2

    G2S2m3 (2:32)

    Using these dened terms, (2.30) simplies to a more manageable form:

    du

    d 2

    u2

    0e

    2

    u u0

    2

    (2:33)

    du

    d

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiu20e

    2 u u0 2

    q (2:34)

    The solution of this equation, which can be tested by substitution, is

    u u0 1 e cos (2:35)

    The angleis dened to be zero at perihelion. The negative square root in (2.34)

    is chosen because, asincreases,rincreases andu must decrease. Let

    p 1

    u0

    h2

    GSm2 (2:36)

    r p

    1 e cos (2:37)

    This is the polar equation of an ellipse referred to its focus, and is the proof of

    Keplers First Law of planetary motion. The quantityeis the eccentricity of the

    ellipse, whilep is thesemi-latus rectumof the ellipse, which is half the length of

    a chord passing through the focus and parallel to the minor axis (Fig. 2.1).

    These equations show that three types of trajectory around the Sun are

    possible, depending on the value of the total energyEin (2.18). If the kinetic

    64 Gravitation

  • 7/26/2019 Guide to Geophysical Equations

    80/296

    energy is greater than the potential energy, the value ofEin (2.32) is positive,

    andeis greater than 1; the path of the object is a hyperbola. If the kinetic energy

    and potential energy are equal, the total energy is zero and e is exactly 1; the path

    is a parabola. In each of these two cases the object can escape to innity, and the

    paths are called escape trajectories. If the kinetic energy is less than the potential

    energy, the total energy Eis negative and the eccentricity is less than 1. In this

    case (corresponding to a planet or asteroid) the object follows an elliptical orbit

    around the Sun.

    2.2.3 Keplers Third Law

    It is convenient to describe the elliptical orbit in Cartesian coordinates (x, y),

    centered on the mid-point of the ellipse, instead of on the Sun. Dene thex-axis

    parallel to the semi-major axis a of the ellipse and they-axis parallel to the semi-

    minor axisb. The equation of the ellipse in Fig. 2.1is

    x2

    a2

    y2

    b2 1 (2:38)

    The semi-minor axis is related to the semi-major axis by the eccentricity e, so

    that

    b2 a2 1 e2

    (2:39)

    The distance of the focus of the ellipse from its center is by denition ae. The

    lengthpof the semi-latus rectum is the value ofyfor a chord through the focus.

    On settingy = p and x = ae in (2.38), we obtain

    p2

    b2 1

    ae 2

    a2 1 e2

    (2:40)

    p2 a2 1 e2 2

    (2:41)

    Now consider the application of Keplers Second Law to an entire circuit of the

    elliptical orbit. The area of the ellipse isab, and the period of the orbit is T, so

    dA

    dt

    ab

    T (2:42)

    Using (2.17),

    h

    m

    2ab

    T (2:43)

    2.2 Keplers laws of planetary motion 65

  • 7/26/2019 Guide to Geophysical Equations

    81/296

    From (2.36) and (2.43) we get the value of the semi-latus rectum,

    p

    1

    GS

    h

    m 2

    1

    GS

    2ab

    T 2

    (2:44)

    Substituting from (2.41) gives

    a 1 e2

    42a2b2

    GST2

    42a4

    GST2 1 e2

    (2:45)

    After simplifying, we nally get

    T2

    a3

    42

    GS (2:46)

    The quantities on the right-hand side are constant, so the square of the period is

    proportional to the cube of the semi-major axis, which is Keplers Third Law.

    2.3 Gravitational acceleration and the potential

    of a solid sphere

    The gravitational potential and acceleration outside and inside a solid spheremay be calculated from the Poisson and Laplace equations, respectively.

    2.3.1 Outside a solid sphere, using Laplaces equation

    Outside a solid sphere the gravitational potentialUG satises Laplaces equation

    (Section 1.9). If the density is uniform, the potential does not vary with the

    polar angle or azimuth . Under these conditions, Laplaces equation in

    spherical polar coordinates ( 2.67) reduces to

    r r2

    UG

    r

    0 (2:47)

    This implies that the bracketed quantity that we are differentiating must be a

    constant,C,

    r2UG

    r C (2:48)

    UG

    r

    C

    r2 (2:49)

    The gravitational acceleration outside the sphere is therefore

    66 Gravitation

  • 7/26/2019 Guide to Geophysical Equations

    82/296

  • 7/26/2019 Guide to Geophysical Equations

    83/296

    UG

    r

    4

    3Gr (2:58)

    aG r5R UG

    r 4

    3Gr

    er (2:59)

    This shows that the gravitational acceleration inside a homogeneous solid

    sphere is proportional to the distance from its center. At the surface of the

    sphere,r= R, and the gravitational acceleration is

    aG R 4

    3GR er

    GM

    R2 er (2:60)

    where the massMof the sphere is

    M4

    3R3 (2:61)

    To obtain the potentialinside the solid sphere, we must integrate ( 2.58). This

    gives

    UG 23Gr2 C2 (2:62)

    The constant of integrationC2is obtained by noting that the potential must be

    continuous at the surface of the sphere. Otherwise a discontinuity would exist

    and the potential gradient (and force) would be innite. Equating (2.54) and (

    2.62) atr= R gives

    2

    3GR2

    C

    2

    GM

    R

    4

    3GR2 (2:63)

    C2 2GR2 (2:64)

    The gravitational potential inside the uniform solid sphere is therefore given by

    UG 2

    3Gr2 2GR2 (2:65)

    UG

    2

    3G r2 3R2 (2:66)

    A schematic graph of the variation of the gravitational potential inside and

    outside a solid sphere is shown in Fig. 2.2.

    68 Gravitation

  • 7/26/2019 Guide to Geophysical Equations

    84/296

    2.4 Laplaces equation in spherical polar coordinates

    In the above examples the sphere was assumed to have uniform density so that

    only the radial term in Laplaces equation had to be solved. This is also the case

    when density varies only with radius. In the Earth, however, lateral variations of

    the density distribution occur, and the gravitational potential UG is then a

    solution of the full Laplace equation

    1

    r2

    rr2

    UG

    r

    1

    r2 sin

    sin

    UG

    1

    r2 sin2

    2UG

    2 0 (2:67)

    This equation is solved using the method ofseparation of variables. This is a

    valuable mathematical technique, which allows the variables in a partial differ-

    ential equation to be separated so that only terms in one variable are on one side

    of the equation and terms in other variables are on the opposite side. A trial

    solution forUGis

    UG r; ; < r (2:68)

    Here , , and are all functions of a single variable only, namelyr,, and,

    respectively. Multiplying (2.67) byr2 and inserting (2.68) forUGgives

    rr2