Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Orbital Mechanics for Engineering

Students Fourth Edition. Edition


Howard D. Curtis
Visit to download the full and correct content document:
https://ebookmass.com/product/orbital-mechanics-for-engineering-students-fourth-edi
tion-edition-howard-d-curtis/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Aerodynamics for Engineering Students 7th Edition E.L.


Houghton

https://ebookmass.com/product/aerodynamics-for-engineering-
students-7th-edition-e-l-houghton/

Statistical Mechanics: Fourth Edition R.K. Pathria

https://ebookmass.com/product/statistical-mechanics-fourth-
edition-r-k-pathria/

Chemistry for Engineering Students 4th Edition Lawrence


S. Brown

https://ebookmass.com/product/chemistry-for-engineering-
students-4th-edition-lawrence-s-brown/

Process Control Instrumentation Technology 8th Edition


Curtis D. Johnson

https://ebookmass.com/product/process-control-instrumentation-
technology-8th-edition-curtis-d-johnson/
Graduate Research, Fourth Edition: A Guide for Students
in the Sciences Densmore

https://ebookmass.com/product/graduate-research-fourth-edition-a-
guide-for-students-in-the-sciences-densmore/

Wiley Abridged Print Companion for Engineering


Mechanics: Statics | 9th Edition J.L. Meriam

https://ebookmass.com/product/wiley-abridged-print-companion-for-
engineering-mechanics-statics-9th-edition-j-l-meriam/

Engineering Mechanics 4 edition Edition S. Timoshenko

https://ebookmass.com/product/engineering-mechanics-4-edition-
edition-s-timoshenko/

Statistical Mechanics: Fourth Edition. Instructor's


Manual R.K. Pathria

https://ebookmass.com/product/statistical-mechanics-fourth-
edition-instructors-manual-r-k-pathria/

System Dynamics for Engineering Students: Concepts and


Applications 2nd Edition Nicolae Lobontiu

https://ebookmass.com/product/system-dynamics-for-engineering-
students-concepts-and-applications-2nd-edition-nicolae-lobontiu/
Orbital Mechanics for
Engineering Students
Orbital Mechanics for
Engineering Students
Fourth Edition

Howard D. Curtis
Professor Emeritus, Aerospace Engineering
Embry-Riddle Aeronautical University
Daytona Beach, Florida
Butterworth-Heinemann is an imprint of Elsevier
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom
50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States
# 2020 Elsevier Ltd. All rights reserved.
No part of this publication may be reproduced or transmitted in any form or by any means, electronic or
mechanical, including photocopying, recording, or any information storage and retrieval system, without
permission in writing from the publisher. Details on how to seek permission, further information about the
Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance
Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the Publisher
(other than as may be noted herein).
MATLAB® is a trademark of The MathWorks, Inc. and is used with permission. The MathWorks does not
warrant the accuracy of the text or exercises in this book. This book’s use or discussion of MATLAB®
software or related products does not constitute endorsement or sponsorship by The MathWorks of a
particular pedagogical approach or particular use of the MATLAB® software.
Notices
Knowledge and best practice in this field are constantly changing. As new research and experience broaden
our understanding, changes in research methods, professional practices, or medical treatment may become
necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using
any information, methods, compounds, or experiments described herein. In using such information or methods
they should be mindful of their own safety and the safety of others, including parties for whom they have a
professional responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability
for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise,
or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the Library of Congress
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library
ISBN: 978-0-08-102133-0

For information on all Butterworth-Heinemann publications


visit our website at https://www.elsevier.com/books-and-journals

Publisher: Katey Birtcher


Acquisition Editor: Steve Merken
Editorial Project Manager: Nathaniel McFadden
Production Project Manager: Maria Bernard
Cover Designer: Victoria Pearson
Typeset by SPi Global, India
To my beloved wife, Mary
For her patience, encouragement, and love
Preface
The purpose of this book is to provide an introduction to space mechanics for undergraduate engineer-
ing students. It is not directed toward graduate students, researchers, and experienced practitioners,
who may nevertheless find useful review material within the book’s contents. The intended readers
are those who are studying the subject for the first time and have completed courses in physics, dy-
namics, and mathematics through differential equations and applied linear algebra. I have tried my best
to make the text readable and understandable to that audience. In pursuit of that objective I have in-
cluded a large number of example problems that are explained and solved in detail. Their purpose is not
to overwhelm but to elucidate. I find that students like the “teach by example” method. I always assume
that the material is being seen for the first time and, wherever possible, I provide solution details so as to
leave little to the reader’s imagination. The numerous figures throughout the book are also intended to
aid comprehension. All of the more labor-intensive computational procedures are accompanied by
MATLAB® code.
For this, the fourth edition, I have retained the content and style of the previous editions and
corrected all the errors discovered by me or reported to me by readers. Except for the new
Chapter 9 on basic lunar trajectories and an expanded discussion of quaternions in Chapter 11 the book
remains essentially the same. Adding the new chapter required the following reshuffling:

Topic This edition Previous edition

Lunar trajectories Chapter 9 Absent


Introduction to orbital perturbations Chapter 10 Chapter 12
Rigid body dynamics Chapter 11 Chapter 9
Satellite attitude dynamics Chapter 12 Chapter 10
Rocket vehicle dynamics Chapter 13 Chapter 11

The organization of the book remains the same as that of the third edition. Chapter 1 is a review of
vector kinematics in three dimensions and of Newton’s laws of motion and gravitation. It also focuses
on the issue of relative motion, crucial to the topics of rendezvous and satellite attitude dynamics. The
material on ordinary differential equation solvers will be useful for students who are expected to code
numerical simulations in MATLAB or other programming languages. Chapter 2 presents the vector-
based solution of the classical two-body problem, resulting in a host of practical formulas for the anal-
ysis of orbits and trajectories of elliptical, parabolic, and hyperbolic shape. The restricted three-body
problem is covered to introduce the notion of Lagrange points and to present the numerical solution of a
lunar trajectory problem. Chapter 3 derives Kepler’s equations, which relate position to time for the
different kinds of orbits. The universal variable formulation is also presented. Chapter 4 is devoted
to describing orbits in three dimensions. Coordinate transformations and the Euler elementary rotation
sequences are defined. Procedures for transforming back and forth between the state vector and the
classical orbital elements are addressed. The effect of the earth’s oblateness on the motion of an orbit’s
ascending node and eccentricity vector is described, pending a more detailed explanation in Chapter 10.
Chapter 5 is an introduction to preliminary orbit determination, including Gibbs’ and Gauss’ methods
and the solution of Lambert’s problem. Auxiliary topics include topocentric coordinate systems, Julian

xiii
xiv Preface

day numbering, and sidereal time. Chapter 6 presents the common means of transferring from one orbit
to another by impulsive delta-v maneuvers, including Hohmann transfers, phasing orbits, and plane
changes. Chapter 7 is a brief introduction to relative motion in general and to the two-impulse rendez-
vous problem in particular. The latter is analyzed using the Clohessy-Wiltshire equations, which are
derived in this chapter. Chapter 8 is an introduction to interplanetary mission design using patched
conics. Chapter 9 extends the patched conic method and the restricted three-body approach to lunar
trajectory analysis. Chapter 10 is an introduction to common orbital perturbations: drag, nonspherical
gravitational field, solar radiation pressure, and lunar and solar gravity. Chapter 11 presents those el-
ements of rigid body dynamics required to characterize the attitude of a space vehicle. Euler’s equa-
tions of rotational motion are derived and applied in a number of example problems. Euler angles, yaw-
pitch-roll angles, and quaternions are presented as ways to describe the attitude of rigid body.
Chapter 12 describes the methods of controlling, changing, and stabilizing the attitude of spacecraft
by means of thrusters, gyros, and other devices. Chapter 13 is a brief introduction to the characteristics
and design of multistage launch vehicles.
Chapters 1 through 4 form the core of a first orbital mechanics course. The time devoted to
Chapter 1 depends on the background of the student. It might be surveyed briefly and used thereafter
simply as a reference. What follows Chapter 4 depends on the objectives of the course.
Chapters 5 through 10 carry on with the subject of orbital mechanics. Chapter 6 on orbital maneu-
vers should be included in any case. Coverage of Chapters 5, 7, 8, and 9 is optional. However, if
Chapters 8 and 9 on interplanetary and lunar missions is to form a part of the course, then the solution
of Lambert’s problem (Section 5.3) must be studied beforehand.
Chapter 10 is appropriate for a course devoted exclusively to orbital mechanics with an introduction
to perturbations, which is a whole topic unto itself.
Chapters 11 and 12 must be covered if the course objectives include an introduction to spacecraft
dynamics. In that case Chapters 5, 7, 8, and 9 would probably not be studied in depth.
Chapter 13 is optional if the engineering curriculum requires a separate course in propulsion includ-
ing rocket dynamics.
The important topic of spacecraft control systems is omitted. However, the material in this book and
a course in control theory provide the basis for the study of spacecraft attitude control.
To understand the material and to solve problems requires using a lot of undergraduate mathemat-
ics. Mathematics, of course, is the language of engineering. Students must not forget that the English
mathematician and physicist Sir Isaac Newton (1642–1727) had to invent calculus so he could solve
orbital mechanics problems in more than just a heuristic way. Newton’s 1687 publication Mathemat-
ical Principles of Natural Philosophy (“the Principia”) is one of the most influential scientific works of
all time. It must be noted that his contemporary, the German mathematician Gottfried Wilhelm von
Leibnitz (1646–1716) is credited with inventing infinitesimal calculus independently of Newton in
the 1670s.
In addition to honing their math skills, students are urged to take advantage of computers (which,
incidentally, use the binary numeral system developed by Leibnitz). There are many commercially
available mathematics software packages for personal computers. Wherever possible they should be
used to relieve the burden of repetitive and tedious calculations. Computer-programming skills can
and should be put to good use in the study of orbital mechanics. The elementary MATLAB programs
referred to in Appendix D of this book illustrate how many of the procedures developed in the text can
Preface xv

be implemented in software. All the scripts were developed and tested using MATLAB version 9.2
(release 2017a). Information about MATLAB, which is a registered trademark of The MathWorks,
Inc., may be obtained from
The MathWorks, Inc.
3 Apple Hill Drive
Natick, MA 01760-2089, USA
www.mathworks.com
Appendix A presents some tables of physical data and conversion factors. Appendix B is a road map
through the first three chapters, showing how the most fundamental equations of orbital mechanics are
related. Appendix C shows how to set up the n-body equations of motion and program them in
MATLAB. Appendix D contains listings of all the MATLAB algorithms and example problems
presented in the text. Appendix E shows that the gravitational field of a spherically symmetric body
is the same as if the mass were concentrated at its center. Appendix F explains how to deal with a
computational issue that arises in some perturbation analyses.

SUPPLEMENTS TO THE TEXT


For purchasers of the book, copies of the MATLAB M-files listed in Appendix D can be freely down-
loaded from this book’s companion website. Also available on the companion website are a set of an-
imations that accompany the text. To access these files, please visit https://www.elsevier.com/books-
and-journals/book-companion/9780081021330.
For instructors using this book for a course, please visit www.textbooks.elsevier.com to register for
access to the solutions manual, PowerPoint lecture slides, and other resources.

ACKNOWLEDGEMENTS
Since the publication of the first three editions and during the preparation of this one, I have received
helpful criticism, suggestions, and advice from many sources locally and worldwide. I thank them all
and regret that time and space limitations prohibited the inclusion of some recommended additional
topics that would have enhanced the book.
It has been a pleasure to work with the people at Elsevier, in particular Joseph P. Hayton, Publisher;
Steve Merken, Senior Acquisitions Editor; and Nate McFadden, Senior Developmental Editor.
I appreciate their enthusiasm for the book, their confidence in me, and all the work they did to move
this project to completion.
Finally and most importantly, I must acknowledge the patience and support of my wife, Mary, who
was a continuous source of optimism and encouragement throughout the revision effort.

Howard D. Curtis
Daytona Beach, FL, United States
CHAPTER

DYNAMICS OF POINT MASSES


1
1.1 INTRODUCTION
This chapter serves as a self-contained reference on the kinematics and dynamics of point masses as
well as some basic vector operations and numerical integration methods. The notation and concepts
summarized here will be used in the following chapters. Those familiar with the vector-based dynamics
of particles can simply page through the chapter and then refer back to it later as necessary. Those who
need a bit more in the way of review will find that the chapter contains all the material they need to
follow the development of orbital mechanics topics in the upcoming chapters.
We begin with a review of vectors and some vector operations, after which we proceed to the prob-
lem of describing the curvilinear motion of particles in three dimensions. The concepts of force and
mass are considered next, along with Newton’s inverse-square law of gravitation. This is followed
by a presentation of Newton’s second law of motion (“force equals mass times acceleration”) and
the important concept of angular momentum.
As a prelude to describing motion relative to moving frames of reference, we develop formulas for
calculating the time derivatives of moving vectors. These are applied to the computation of relative
velocity and acceleration. Example problems illustrate the use of these results, as does a detailed con-
sideration of how the earth’s rotation and curvature influence our measurements of velocity and accel-
eration. This brings in the curious concept of Coriolis force. Embedded in exercises at the end of the
chapter is practice in verifying several fundamental vector identities that will be employed frequently
throughout the book.
The chapter concludes with an introduction to numerical methods, which can be called upon to
solve the equations of motion when an analytical solution is not possible.

1.2 VECTORS
A vector is an object that is specified by both a magnitude and a direction. We represent a vector graph-
ically by a directed line segment (i.e., an arrow pointing in the direction of the vector). The end opposite
the arrow is called the tail. The length of the arrow is proportional to the magnitude of the vector. Ve-
locity is a good example of a vector. We say that a car is traveling eastward at 80 km/h. The direction is
east and the magnitude, or speed, is 80 km/h. We will use boldface type to represent vector quantities
and plain type to denote scalars. Thus, whereas B is a scalar, B is a vector.
Orbital Mechanics for Engineering Students. https://doi.org/10.1016/B978-0-08-102133-0.00001-5
# 2020 Elsevier Ltd. All rights reserved.
1
2 CHAPTER 1 DYNAMICS OF POINT MASSES

FIG. 1.1
All of these vectors may be denoted A, since their magnitudes and directions are the same.

FIG. 1.2
Parallelogram rule of vector addition. A + B ¼ C.

Observe that a vector is specified solely by its magnitude and direction. If A is a vector, then all
vectors having the same physical dimensions, the same length, and pointing in the same direction as A
are denoted A, regardless of their line of action, as illustrated in Fig. 1.1. Shifting a vector parallel to
itself does not mathematically change the vector. However, the parallel shift of a vector might produce
a different physical effect. For example, an upward 5-kN load (force vector) applied to the tip of an
airplane wing gives rise to quite a different stress and deflection pattern in the wing than the same load
acting at the wing’s midspan.
The magnitude of a vector A is denoted kAk, or, simply A.
Multiplying a vector B by the reciprocal of its magnitude produces a vector that points in the di-
rection of B, but it is dimensionless and has a magnitude of one. Vectors having dimensionless mag-
nitude are called unit vectors. We put a hat (^) over the letter representing a unit vector. Then we can tell
simply by inspection that, for example, u ^ is a unit vector, as are B^ and ^e.
^A . As pointed out above,
It is convenient to denote the unit vector in the direction of the vector A as u
we obtain this vector from A as follows:

A
^A ¼
u (1.1)
A

^C ¼ C=C, u
Likewise, u ^F ¼ F=F, etc.
The sum or resultant of two vectors is defined by the parallelogram rule (Fig. 1.2). Let C be the sum
of the two vectors A and B. To form that sum using the parallelogram rule, the vectors A and B are
1.2 VECTORS 3

FIG. 1.3
Three-dimensional, right-handed Cartesian coordinate system.

shifted parallel to themselves (leaving them unaltered) until the tail of A touches the tail of B. Drawing
dotted lines through the head of each vector parallel to the other completes a parallelogram. The
diagonal from the tails of A and B to the opposite corner is the resultant C. By construction, vector
addition is commutative; that is,
A+B¼B+A (1.2)

A Cartesian coordinate system in three dimensions consists of three axes, labeled x, y, and z,
which intersect at the origin O. We will always use a right-handed Cartesian coordinate system,
which means if you wrap the fingers of your right hand around the z axis, with the thumb
pointing in the positive z direction, your fingers will be directed from the x axis toward the y axis.
Fig. 1.3 illustrates such a system. Note that the unit vectors along the x, y, and z axes are, respectively,
^i, ^j, and k.
^
In terms of its Cartesian components, and in accordance with the above summation rule, a vector A
is written in terms of its components Ax, Ay, and Az as
A ¼ Ax^i + Ay^j + Az k
^ (1.3)

The projection of A on the xy plane is a vector denoted Axy . It follows that


Axy ¼ Ax^i + Ay^j

According to the Pythagorean theorem, the magnitude of A in terms of its Cartesian components is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
A¼ A2x + A2y + A2z (1.4)

From Eqs. (1.1) and (1.3), the unit vector in the direction of A is
^A ¼ cos θx^i + cos θy^j + cos θz k
u ^ (1.5)
4 CHAPTER 1 DYNAMICS OF POINT MASSES

FIG. 1.4
Direction angles in three dimensions.

where
Ax Ay Az
cos θx ¼ cos θy ¼ cos θz ¼ (1.6)
A A A
The direction angles θx, θy, and θz are illustrated in Fig. 1.4, and they are measured between the vector
and the positive coordinate axes. Note carefully that the sum of θx, θy, and θz is not in general known a
priori and cannot be assumed to be, say, 180 degrees.

EXAMPLE 1.1
Calculate the direction angles of the vector A ¼ ^i  4^j + 8k.
^

Solution
First, compute the magnitude of A by means of Eq. (1.4),
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
A ¼ 12 + ð4Þ2 + 82 ¼ 9
Then Eq. (1.6) yields
   
Ax 1
θx ¼ cos 1 ¼ cos 1 ) θx ¼ 83:62 degrees
A 9
   
Ay 4
θy ¼ cos 1 ¼ cos 1 ) θy ¼ 116:4 degrees
A 9
   
Az 8
θz ¼ cos 1 ¼ cos 1 ) θz ¼ 27:27 degrees
A 9
Observe that θx + θy + θz ¼ 227.3 degrees.

Multiplication and division of two vectors are undefined operations. There are no rules for com-
puting the product AB and the ratio A/B. However, there are two well-known binary operations on
1.2 VECTORS 5

vectors: the dot product and the cross product. The dot product of two vectors is a scalar defined as
follows:
A  B ¼ AB cos θ (1.7)
where θ is the angle between the heads of the two vectors, as shown in Fig. 1.5. Clearly,
AB¼BA (1.8)
If two vectors are perpendicular to each other, then the angle between them is 90 degrees. It follows
from Eq. (1.7) that their dot product is zero. Since the unit vectors ^i, ^j, and k
^ of a Cartesian coordinate
system are mutually orthogonal and of magnitude 1, Eq. (1.7) implies that
^i  ^i ¼ ^j  ^j ¼ k
^k ^¼1
(1.9)
^i  ^j ¼ ^i  k^ ¼ ^j  k
^¼0

Using these properties, it is easy to show that the dot product of the vectors A and B may be found in
terms of their Cartesian components as
A  B ¼ Ax Bx + Ay By + Az Bz (1.10)

If we set B ¼ A, then it follows from Eqs. (1.4) and (1.10) that


pffiffiffiffiffiffiffiffiffiffi
A¼ AA (1.11)
The dot product operation is used to project one vector onto the line of action of another. We can
imagine bringing the vectors tail to tail for this operation, as illustrated in Fig. 1.6. If we drop a per-
pendicular line from the tip of B onto the direction of A, then the line segment BA is the orthogonal
projection of B onto the line of action of A. BA stands for the scalar projection of B onto A. From trig-
onometry, it is obvious from the figure that
BA ¼ B cos θ

FIG. 1.5
The angle between two vectors brought tail to tail by parallel shift.

FIG. 1.6
Projecting the vector B onto the direction of A.
6 CHAPTER 1 DYNAMICS OF POINT MASSES

^A be the unit vector in the direction of A. Then,


Let u
¼1
zffl}|ffl{
^ A ¼ kBk ku
Bu ^A k cos θ ¼ B cos θ
Comparing this expression with the preceding one leads to the conclusion that
A
^A ¼ B 
BA ¼ B  u (1.12)
A
^A is given by Eq. (1.1). Likewise, the projection of A onto B is given by
where u
B
AB ¼ A 
B
Observe that AB ¼ BA only if A and B have the same magnitude.

EXAMPLE 1.2
Let A ¼ ^i + 6^j + 18k
^ and B ¼ 42^i  69^j + 98k:
^ Calculate
(a) the angle between A and B;
(b) the projection of B in the direction of A;
(c) the projection of A in the direction of B.
Solution
First, we make the following individual calculations.
A  B ¼ ð1Þð42Þ + ð6Þð69Þ + ð18Þð98Þ ¼ 1392 (a)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
A ¼ ð1Þ2 + ð6Þ2 + ð18Þ2 ¼ 19 (b)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
B ¼ ð42Þ2 + ð69Þ2 + ð98Þ2 ¼ 127 (c)
(a) According to Eq. (1.7), the angle between A and B is
 
AB
θ ¼ cos 1
AB
Substituting Eqs. (a), (b), and (c) yields
 
1392
θ ¼ cos 1 ¼ 54:77 degrees
19  127
(b) From Eq. (1.12), we find the projection of B onto A.
A AB
BA ¼ B  ¼
A A
Substituting Eqs. (a) and (b) we get
1392
BA ¼ ¼ 73:26
19
(c) The projection of A onto B is
B AB
AB ¼ A  ¼
B B
1.2 VECTORS 7

Substituting Eqs. (a) and (c) we obtain


1392
AB ¼ ¼ 10:96
127

The cross product of two vectors yields another vector, which is computed as follows:
A  B ¼ ðAB sinθÞ^
nAB (1.13)

where θ is the angle between the heads of A and B, and n^AB is the unit vector normal to the plane defined
by the two vectors. The direction of n^AB is determined by the right-hand rule. That is, curl the fingers of
the right hand from the first vector (A) toward the second vector (B), and the thumb shows the direction
^AB (Fig. 1.7). If we use Eq. (1.13) to compute B  A, then n
of n ^AB points in the opposite direction,
which means
B  A ¼ ðA  BÞ (1.14)

Therefore, unlike the dot product, the cross product is not commutative.
The cross product is obtained analytically by resolving the vectors into Cartesian components.
   
A  B ¼ Ax^i + Ay^j + Az k
^  Bx^i + By^j + Bz k
^ (1.15)

Since the set ^i^j k


^ is a mutually perpendicular triad of unit vectors, Eq. (1.13) implies that
^i ^i ¼ 0 ^j  ^j ¼ 0 ^k
k ^¼0
^i  ^j ¼ k
^ ^j  k^ ¼ ^i ^ ^i ¼ ^j (1.16)
k

Expanding the right-hand side of Eq. (1.15), substituting Eq. (1.16), and making use of Eq. (1.14)
leads to
 
A  B ¼ Ay Bz  Az By ^i  ðAx Bz  Az Bx Þ^j + Ax By  Ay Bx k
^ (1.17)

It may be seen that the right-hand side is the determinant of the matrix
2 3
^i ^j ^
k
4 Ax Ay Az 5
Bx By Bz

FIG. 1.7
^AB is normal to both A and B and defines the direction of the cross product A  B.
n
8 CHAPTER 1 DYNAMICS OF POINT MASSES

Thus, Eq. (1.17), can be written as


^i ^j ^
k
A  B ¼ Ax Ay Az (1.18)
Bx By Bz

where the two vertical bars stand for the determinant. Obviously, the rule for computing the cross prod-
uct, though straightforward, is a bit lengthier than that for the dot product. Remember that the dot prod-
uct yields a scalar whereas the cross product yields a vector.
The cross product provides an easy way to compute the normal to a plane. Let A and B be any two
vectors lying in the plane, or, let any two vectors be brought tail to tail to define a plane, as shown in
Fig. 1.7. The vector C ¼ A  B is normal to the plane of A and B. Therefore, n ^ AB ¼ C=C, or
AB
^AB ¼
n (1.19)
kA  Bk

EXAMPLE 1.3
Let A5  3^i + 7^j + 9k
^ and B56^i  5^j + 8k.
^ Find a unit vector that lies in the plane of A and B and is perpendicular to A.

Solution
The plane of vectors A and B is determined by parallel-shifting the vectors so that they meet tail to tail. Calculate the vector
D ¼ A  B.
^i ^j k
^
D ¼ 3 7 9 ¼ 101^i + 78^j  27k
^
6 5 8
Note that A and B are both normal to D. We next calculate the vector C ¼ D  A.
^i ^j ^
k
C ¼ 101 78 27 ¼ 891^i  828^j + 941k
^
3 7 9
C is normal to D as well as to A. A, B, and C are all perpendicular to D. Therefore, they are coplanar. Thus, C is not only
perpendicular to A, but it also lies in the plane of A and B. Therefore, the unit vector we are seeking is the unit vector in the
direction of C. That is

C 891^i  828^j + 941k ^


^C ¼
u ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
C 2
891 + ð828Þ + 9412
2

^C ¼ 0:5794^i  0:5384^j + 0:6119k


u ^

In the chapters to follow, we will often encounter the vector triple product, A  (B  C). By resolving
A, B, and C into their Cartesian components, it can easily be shown that the vector triple product can be
expressed in terms of just the dot products of these vectors as follows:
A  ðB  CÞ ¼ BðA  CÞ  CðA  BÞ (1.20)
Because of the appearance of the letters on the right-hand side, this is often referred to as the “bac–cab
rule.”
1.3 KINEMATICS 9

EXAMPLE 1.4
If F ¼ E  {D  [A  (B  C)]}, use the bac–cab rule to reduce this expression to one involving only dot products.
Solution
First, we invoke the bac–cab rule to obtain
8 baccab rule
9
>
< zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{>
=
F ¼ E  D  ½BðA  CÞ  CðA  BÞ
>
: >
;

Expanding and collecting terms leads to


F ¼ ðA  CÞ½E  ðD  BÞ  ðA  BÞ½E  ðD  CÞ
We next apply the bac–cab rule twice on the right-hand side.
baccab rule baccab rule
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
F ¼ ðA  CÞ ½DðE  BÞ  BðE  DÞ ðA  BÞ ½DðE  CÞ  CðE  DÞ
Expanding and collecting terms yields the sought-for result.

F ¼ ½ðA  CÞðE  BÞ  ðA  BÞðE  CÞD  ðA  CÞðE  DÞB + ðA  BÞðE  DÞC

Another useful vector identity is the “interchange of the dot and the cross”:
A  ðB  C Þ ¼ ðA  B Þ  C (1.21)
It is so-named because interchanging the operations in the expression A  B  C yields A  B  C.
The parentheses in Eq. (1.21) are required to show which operation must be carried out first, according
to the rules of vector algebra. (For example, (A  B)  C, the cross product of a scalar and a vector,
is undefined.) It is easy to verify Eq. (1.21) by substituting A ¼ Ax^i + Ay^j + Az k,
^ B ¼ Bx^i + By^j + Bz k,
^
^ ^ ^
and C ¼ Cx i + Cy j + Cz k and observing that both sides of the equal sign reduce to the same expression.

1.3 KINEMATICS
To track the motion of a particle P through Euclidean space, we need a frame of reference, consisting of
a clock and a nonrotating Cartesian coordinate system. The clock keeps track of time t, and the xyz axes
of the Cartesian coordinate system are used to locate the spatial position of the particle. In nonrelativ-
istic mechanics, a single “universal” clock serves for all possible Cartesian coordinate systems. So when
we refer to a frame of reference, we need to think only of the mutually orthogonal axes themselves.
The unit of time used throughout this book is the second (s). The unit of length is the meter (m), but
the kilometer (km) will be the length unit of choice when large distances and velocities are involved.
Conversion factors between kilometers, miles, and nautical miles are listed in Table A.3.
Given a frame of reference, the position of the particle P at a time t is defined by the position vector
r(t) extending from the origin O of the frame out to P itself, as illustrated in Fig. 1.8. The components of
r(t) are just the x, y, and z coordinates,
rðtÞ ¼ xðtÞ^i + yðtÞ^j + zðtÞk
^
10 CHAPTER 1 DYNAMICS OF POINT MASSES

FIG. 1.8
Position, velocity, and acceleration vectors.

The distance of P from the origin is the magnitude or length of r, denoted k r k or just r,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
krk ¼ r ¼ x2 + y 2 + z2

As in Eq. (1.11), the magnitude of r can also be computed by means of the dot product operation,
pffiffiffiffiffiffiffiffi
r¼ rr

The velocity v and acceleration a of the particle are the first and second time derivatives of the position
vector,
dxðtÞ^ dyðtÞ^ dyðtÞ ^
vðtÞ ¼ i+ j+ k ¼ vx ðtÞ^i + vy ðtÞ^j + vz ðtÞk
^
dt dt dt
dvx ðtÞ^ dvy ðtÞ^ dvz ðtÞ ^
aðtÞ ¼ i+ j+ k ¼ ax ðtÞ^i + ay ðtÞ^j + az ðtÞk
^
dt dt dt
The derivatives of ^i, ^j, and k
^ are zero since axes of the Cartesian frame have fixed directions. It is con-
venient to represent the time derivative by means of an overhead dot. In this shorthand notation, if ( ) is
any quantity, then
 dð Þ  d2 ð Þ ⋯ d3 ð Þ
ð Þ¼ ð Þ¼ ð Þ¼ etc:
dt dt2 dt3
Thus, for example,
v ¼ r_
a ¼ v_ ¼ €r
vx ¼ x_ vy ¼ y_ vz ¼ z_
ax ¼ v_ x ¼ x€ ay ¼ v_ y ¼ y€ az ¼ v_ z ¼ z€

The locus of points that a particle occupies as it moves through space is called its path or trajectory.
If the path is a straight line, then the motion is rectilinear. Otherwise, the path is curved, and the motion
1.3 KINEMATICS 11

^t is the unit vector tangent to the


is called curvilinear. The velocity vector v is tangent to the path. If u
trajectory, then
v ¼ v^
ut (1.22)

where the speed v is the magnitude of the velocity v. The distance ds that P travels along its path in the
time interval dt is obtained from the speed by
ds ¼ vdt

In other words,
v ¼ s_

The distance s, measured along the path from some starting point, is what the odometers in our auto-
mobiles record. Of course, s, _ our speed along the road, is indicated by the dial of the speedometer.
Note carefully that v 6¼ r_ (i.e., the magnitude of the derivative of r does not equal the derivative of
the magnitude of r).

EXAMPLE 1.5
The position vector in meters is given as a function of time in seconds as
  
r ¼ 8t2 + 7t + 6 ^i + 5t3 + 4 ^j + 0:3t4 + 2t2 + 1 k
^ ðmÞ (a)
At t ¼ 10 s, calculate (a) v (the magnitude of the derivative of r) and (b) r_ (the derivative of the magnitude of r).
Solution
(a) The velocity v is found by differentiating the given position vector with respect to time,
dr 
v¼ ¼ ð16t + 7Þ^i + 15t2^j + 1:2t3 + 4t k
^
dt
The magnitude of this vector is the square root of the sum of the squares of its components,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
v ¼ 1:44t6 + 234:6t4 + 272t2 + 224t + 49
Evaluating this at t ¼ 10 s, we get
v ¼ 1953:3m=s
(b) Calculating the magnitude of r in Eq. (a) leads to
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r ¼ 0:09t8 + 26:2t6 + 68:6t4 + 152t3 + 149t2 + 84t + 53
The time derivative of this expression is
dr 0:36t7 + 78:6t5 + 137:2t3 + 228t2 + 149t + 42
r_ ¼ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dt 0:09t8 + 26:2t6 + 68:6t4 + 152t3 + 149t2 + 84t + 53
Substituting t ¼ 10 s yields

r_ ¼ 1935:5 m=s

^t in the Cartesian coordinate


If v is given, then we can find the components of the unit tangent u
frame of reference by means of Eq. (1.22):
v vx vy vz ^  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
^t ¼ ¼ ^i + ^j + k
u v ¼ v2x + v2y + v2z (1.23)
v v v v
12 CHAPTER 1 DYNAMICS OF POINT MASSES

The acceleration may be written as


^t + an u
a ¼ at u ^n (1.24)
where at and an are the tangential and normal components of acceleration, given by
v2
at ¼ v_ ð¼ s€Þ an ¼ (1.25)
ρ
where ρ is the radius of curvature, which is the distance from the particle P to the center of curvature of
^n is perpendicular to u
the path at that point. The unit principal normal u ^t and points toward the center of
curvature C, as shown in Fig. 1.9. Therefore, the position of C relative to P, denoted rC/P, is
rC=P ¼ ρ^
un (1.26)

The orthogonal unit vectors u ^t and u


^n form a plane called the osculating plane. The unit normal to the
^b , the binormal, and it is obtained from u
osculating plane is u ^t and u
^n by taking their cross product:
u ^t  u
^b ¼ u ^n (1.27)
From Eqs. (1.22), (1.24), and (1.27), we have
v  a ¼ v^ ^ t + an u
ut  ðat u ^ n Þ ¼ van ðu ^n Þ ¼ van u
^t  u ^b ¼ kv  ak^
ub
That is, an alternative to Eq. (1.27) for calculating the binormal vector is
va
^b ¼
u (1.28)
kv  ak
^t , u
Note that u ^n , and u
^b form a right-handed triad of orthogonal unit vectors. That is
^b  u
u ^t ¼ u
^n ^t  u
u ^n ¼ u
^b ^n  u
u ^b ¼ u
^t (1.29)

FIG. 1.9
Orthogonal triad of unit vectors associated with the moving point P.
1.3 KINEMATICS 13

The center of curvature lies in the osculating plane. When the particle P moves an incremental distance
ds, the radial from the center of curvature to the path sweeps out a small angle, dϕ, measured in the
osculating plane. The relationship between this angle and ds is
ds ¼ ρdϕ
_ or
so that s_ ¼ ρϕ,
v
ϕ_ ¼ (1.30)
ρ

EXAMPLE 1.6
Relative to a Cartesian coordinate system, the position, velocity, and acceleration of a particle P at a given instant are

r ¼ 250^i + 630^j + 430k


^ ðmÞ (a)

v ¼ 90^i + 125^j + 170k


^ ðm=sÞ (b)

a ¼ 16^i + 125^j + 30k
^ m=s2 (c)
Find the coordinates of the center of curvature at that instant.
Solution
The coordinates of the center of curvature C are the components of its position vector rC. Consulting Fig. 1.9, we observe
that
rC ¼ r + ρ^
un (d)
^n is the unit principal normal vector. The
where r is the position vector of the point P, ρ is the radius of curvature, and u
position vector r is given in Eq. (a), but ρ and u^n are unknowns at this point. We must use the geometry of Fig. 1.9 to
find them.
We begin by seeking the value of u ^n , using the first of Eqs. (1.29),
u ^b  u
^n ¼ u ^t (e)
^t is found at once from the velocity vector in Eq. (b) by means of Eq. 1.23,
The unit tangent vector u
v
^t ¼
u
v
where
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
v¼ 902 + 1252 + 1702 ¼ 229:4 m=s (f)
Thus,

90^i + 125^j + 170k


^
^t ¼
u ¼ 0:39233^i + 0:54490^j + 0:74106k
^ (g)
229:4
^b we insert the given velocity and acceleration vectors into Eq. (1.28),
To find the binormal u
^i ^j ^
k
90 125 170
va 16 125 30 17,500^i + 20^j + 9250k ^
(h)
^b ¼
u ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kv  ak kv  ak 2
ð17, 500Þ + 20 + 92502 2

¼ 0:88409^i + 0:0010104^i + 0:46731k


^
14 CHAPTER 1 DYNAMICS OF POINT MASSES

Substituting Eqs. (g) and (h) back into Eq. (e) finally yields the unit principal normal

^i ^j ^
k
^n ¼ 0:88409 0:0010104 0:46731 ¼ 0:25389^i + 0:8385^j  0:48214k
u ^ (i)
0:39233 0:5449 0:74106
The only unknown remaining in Eq. (d) is ρ, for which we appeal to Eq. (1.25),

v2
ρ¼ (j)
an
^n ,
The normal acceleration an is calculated by projecting the acceleration vector a onto the direction of the unit normal u
   
^n ¼ 16^i + 125^j + 30k
an ¼ a  u ^  0:25389^i + 0:8385^j  0:48214k ^ ¼ 86:287m=s2 (k)
Putting the values of v and an from Eqs. (f) and (k) into Eq. ( j) yields the radius of curvature,

229:42
ρ¼ ¼ 609:89m (l)
86:287
Upon substituting Eqs. (a), (i), and (l) into Eq. (d), we obtain the position vector of the center of curvature C,
   
rC ¼ 250^i + 630^j + 430k ^ + 609:89 0:25389^i + 0:8385^j  0:48214k ^

¼ 95:159^i + 1141:4^j + 135:95k


^ ðmÞ
Therefore, the coordinates of C are
x ¼ 95:16 m y ¼ 1141 m z ¼ 136:0 m

1.4 MASS, FORCE, AND NEWTON’S LAW OF GRAVITATION


Mass, like length and time, is a primitive physical concept: it cannot be defined in terms of any other
physical concept. Mass is simply the quantity of matter. More practically, mass is a measure of the
inertia of a body. Inertia is an object’s resistance to changing its state of motion. The larger its inertia
(the greater its mass), the more difficult it is to set a body into motion or bring it to rest. The unit of mass
is the kilogram (kg).
Force is the action of one physical body on another, either through direct contact or through a dis-
tance. Gravity is an example of force acting through a distance, as are magnetism and the force between
charged particles. The gravitational force Fg between two masses m1 and m2 having a distance r be-
tween their centers is
m1 m2
Fg ¼ G (1.31)
r2
This is Newton’s law of gravity, in which G, the universal gravitational constant, has the value
G ¼ 6.6742(1011)m3/(kg  s2). Due to the inverse-square dependence on distance, the force of gravity
rapidly diminishes with the amount of separation between the two masses. In any case, the force of
gravity is minuscule unless at least one of the masses is extremely big.
The force of a large mass (such as the earth) on a mass many orders of magnitude smaller (such as a
person) is called weight, W. If the mass of the large object is M and that of the relatively tiny one is m,
then the weight of the small body is
 
Mm GM
W ¼G ¼ m
r2 r2
1.4 MASS, FORCE, AND NEWTON’S LAW OF GRAVITATION 15

or
W ¼ mg (1.32)

where
GM
g¼ (1.33)
r2
g has units of acceleration (m/s2) and is called the acceleration of gravity. If planetary gravity is the only
force acting on a body, then the body is said to be in free fall. The force of gravity draws a freely falling
object toward the center of attraction (e.g., center of the earth) with an acceleration g. Under ordinary
conditions, we sense our own weight by feeling contact forces acting on us in opposition to the force of
gravity. In free fall, there are, by definition, no contact forces, so there can be no sense of weight. Even
though the weight is not zero, a person in free fall experiences weightlessness, or the absence of gravity.
Let us evaluate Eq. (1.33) at the surface of the earth, whose radius according to Table A.1 is
6378 km. Letting g0 represent the standard sea level value of g, we get
GM
g0 ¼ (1.34)
R2E

In SI units,
g0 ¼ 9:807m=s2 (1.35)

Substituting Eq. (1.34) into Eq. (1.33) and letting z represent the distance above the earth’s surface, so
that r ¼ RE + z, we obtain
R2E g0
g ¼ g0 ¼ (1.36)
ðRE + zÞ2 ð1 + z=RE Þ2

FIG. 1.10
Variation of the acceleration of gravity with altitude.
16 CHAPTER 1 DYNAMICS OF POINT MASSES

Commercial airliners cruise at altitudes on the order of 10 km (6 miles). At that height, Eq. (1.36)
reveals that g (and hence weight) is only three-tenths of a percent less than its sea level value.
Thus, under ordinary conditions, we ignore the variation of g with altitude. A plot of Eq. (1.36) out
to a height of 2000 km (the upper limit of low earth orbit operations) is shown in Fig. 1.10. The var-
iation of g over that range is significant. Even so, at space station altitude (400 km), weight is only
about 10% less than it is on the earth’s surface. The astronauts experience weightlessness, but they
clearly are not weightless.

EXAMPLE 1.7
Show that in the absence of an atmosphere, the shape of a low-altitude ballistic trajectory is a parabola. Assume the ac-
celeration of gravity g is constant and neglect the earth’s curvature.
Solution
Fig. 1.11 shows a projectile launched at t ¼ 0 s with a speed v0 at a flight path angle γ 0 from the point with coordinates
(x0, y0).
Since the projectile is in free fall after launch, its only acceleration is that of gravity in the negative y direction:
x€¼ 0
y€¼ g
Integrating with respect to time and applying the initial conditions leads to
x ¼ x0 + ðv0 cos γ 0 Þt (a)
1
y ¼ y0 + ðv0 sin γ 0 Þt  gt2 (b)
2
Solving Eq. (a) for t and substituting the result into Eq. (b) yields
1 g
y ¼ y0 + ðx  x0 Þtan γ 0  ðx  x0 Þ2 (c)
2 v20 cos 2 γ 0
This is the equation of a second-degree curve, a parabola, as sketched in Fig. 1.11.

FIG. 1.11
Flight of a low-altitude projectile in free fall (no atmosphere).
1.5 NEWTON’S LAW OF MOTION 17

EXAMPLE 1.8
An airplane flies a parabolic trajectory like that in Fig. 1.11 so that the passengers will experience free fall (weightlessness).
What is the required variation of the flight path angle γ with speed v? Ignore the curvature of the earth.
Solution
Fig. 1.12 reveals that for a “flat” earth, dγ ¼  dϕ. That is,

γ_ ¼ ϕ_
It follows from Eq. (1.30) that
ρ_γ ¼ v (1.37)
The normal acceleration an is just the component of the gravitational acceleration g in the direction of the unit principal
normal to the curve (from P toward C). From Fig. 1.12, then,
an ¼ g cos γ (a)
Substituting the second of Eqs. (1.25) into Eq. (a) and solving for the radius of curvature yields
v2
ρ¼ (b)
gcos γ
Combining Eqs. (1.37) and (b), we find the time rate of change of the flight path angle,
g cos γ
γ_ ¼ 
v

FIG. 1.12
Relationship between dγ and dϕ for a “flat” earth.

1.5 NEWTON’S LAW OF MOTION


Force is not a primitive concept like mass because it is intimately connected with the concepts of mo-
tion and inertia. In fact, the only way to alter the motion of a body is to exert a force on it. The degree to
which the motion is altered is a measure of the force. Newton’s second law of motion quantifies this. If
the resultant or net force on a body of mass m is Fnet, then
Fnet ¼ ma (1.38)
18 CHAPTER 1 DYNAMICS OF POINT MASSES

In this equation, a is the absolute acceleration of the center of mass. The absolute acceleration is mea-
sured in a frame of reference that itself has neither translational nor rotational acceleration relative to
the fixed stars. Such a reference is called an absolute or inertial frame of reference.
Force is related to the primitive concepts of mass, length, and time by Newton’s second law. The
unit of force, appropriately, is the Newton, which is the force required to impart an acceleration of 1 m/
s2 to a mass of 1 kg. A mass of 1 kg therefore weighs 9.807 N at the earth’s surface. The kilogram is not
a unit of force.
Confusion can arise when mass is expressed in units of force, as frequently occurs in US engineer-
ing practice. In common parlance either the pound or the ton (2000 lb) is more likely to be used to
express the mass. The pound of mass is officially defined precisely in terms of the kilogram, as shown
in Table A.3. Since 1 lb of mass weighs 1 lb of force where the standard sea level acceleration of gravity
(Eq. 1.35) exists, we can use Newton’s second law to relate the pound of force to the Newton:
11b ðforceÞ ¼ 0:4536kg  9:807m=s2 ¼ 4:448N

The slug is the quantity of matter accelerated at 1 ft/s2 by a force of 1 lb. We can again use Newton’s
second law to relate the slug to the kilogram. Noting the relationship between feet and meters in
Table A.3, we find
1lb 4:448N kg  m=s2
1 slug ¼ 2
¼ 2
¼ 14:59 ¼ 14:59kg
1ft=s 0:3048m=s m=s2

EXAMPLE 1.9
On a NASA mission, the space shuttle Atlantis orbiter was reported to weigh 239,255 lb just prior to liftoff. On orbit 18 at an
altitude of about 350 km, the orbiter’s weight was reported to be 236,900 lb. (a) What was the mass, in kilograms, of
Atlantis on the launchpad and in orbit? (b) If no mass was lost between launch and orbit 18, what would have been the
weight of Atlantis, in pounds?
Solution
(a) The given data illustrate the common use of weight in pounds as a measure of mass. The “weights” given are actually
the mass in pounds of mass. Therefore, prior to launch
0:4536kg
mlaunchpad ¼ 239,255lb ðmassÞ  ¼ 108,500kg
1lbðmassÞ
In orbit,
0:4536kg
morbit 18 ¼ 236,900lb ðmassÞ  ¼ 107,500kg
1lbðmassÞ
The decrease in mass is the propellant expended by the orbital maneuvering and reaction control rockets on the orbiter.
(b) Since the space shuttle launchpad at the Kennedy Space Center is essentially at sea level, the launchpad weight of
Atlantis in pounds (force) was numerically equal to its mass in pounds (mass). With no change in mass, the force of
gravity at 350 km would be, according to Eq. (1.36),
0 12
B 1 C
W ¼ 239,255lbðforceÞ  @ A ¼ 215,000lbðforceÞ
350
1+
6378
1.5 NEWTON’S LAW OF MOTION 19

The integral of a force F over a time interval is called the impulse of the force,
ð t2
I¼ Fdt (1.39)
t1

Impulse is a vector quantity. From Eq. (1.38) it is apparent that if the mass is constant, then
ð t2
dv
I net ¼ m dt ¼ mv2  mv1 (1.40)
t1 dt
That is, the net impulse on a body yields a change mΔv in its linear momentum, so that
I net
Δv ¼ (1.41)
m
If Fnet is constant, then I net ¼ Fnet Δt, in which case Eq. (1.41) becomes
Fnet
Δv ¼ Δt ðif Fnet is constantÞ (1.42)
m
Let us conclude this section by introducing the concept of angular momentum. The moment of the
net force about O in Fig. 1.13 is
MO Þnet ¼ r  Fnet
Substituting Eq. (1.38) yields
dv
MO Þnet ¼ r  ma ¼ r  m (1.43)
dt
But, keeping in mind that the mass is constant,
 
dv d dr d
rm ¼ ðr  mvÞ   mv ¼ ðr  mvÞ  ðv  mvÞ
dt dt dt dt
Since v  mv ¼ m(v  v) ¼ 0, it follows that Eq. (1.43) can be written
dHO
MO Þnet ¼ (1.44)
dt

FIG. 1.13
The absolute acceleration of a particle is in the direction of the net force.
20 CHAPTER 1 DYNAMICS OF POINT MASSES

where HO is the angular momentum about O,


HO ¼ r  mv (1.45)

Thus, just as the net force on a particle changes its linear momentum mv, the moment of that force about
a fixed point changes the moment of its linear momentum about that point. Integrating Eq. (1.44) with
respect to time yields
ð t2
MO Þnet ¼ HO Þ2  HO Þ1 (1.46)
t1

The integral on the left is the net angular impulse. This angular impulse-momentum equation is the
rotational analog of the linear impulse-momentum relation given above in Eq. (1.40).

EXAMPLE 1.10
A particle of mass m is attached to point O by an inextensible string of length l, as illustrated in Fig. 1.14. Initially, the string
is slack when m is moving to the left with a speed v0 in the position shown. Calculate (a) the speed of m just after the string
becomes taut and (b) the average force in the string over the small time interval Δt required to change the direction of the
particle’s motion.
Solution
(a) Initially, the position and velocity of the particle are

r1 ¼ c^i + d^j v1 ¼ v0^i


The angular momentum about O is
^i ^j ^
k
H1 ¼ r1  mv1 ¼ c d ^
0 ¼ mvo d k (a)
mvo 0 0
Just after the string becomes taut,
pffiffiffiffiffiffiffiffiffiffiffiffiffi
r2 ¼  l2  d 2^i + d^j v2 ¼ vx^i + vy^j (b)
and the angular momentum is
^i ^j ^
k  pffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffi
H2 ¼ r2  mv2 ¼  l2  d2 d 0 ¼ mvx d  mvy l2  d2 k ^ (c)
mvx mvy 0
Initially, the force exerted on m by the slack string is zero. When the string becomes taut, the force exerted on m passes
through O. Therefore, the moment of the net force on m about O remains zero. According to Eq. (1.46),
H 2 ¼ H1

FIG. 1.14
Particle attached to O by an inextensible string.
1.6 TIME DERIVATIVES OF MOVING VECTORS 21

Substituting Eqs. (a) and (c) yields


pffiffiffiffiffiffiffiffiffiffiffiffiffi
vx d + l2  d 2 vy ¼ vo d (d)
The string is inextensible, so the component of the velocity of m along the string must be zero:
v2  r2 ¼ 0
Substituting v2 and r2 from Eq. (b) and solving for vy, we get
rffiffiffiffiffiffiffiffiffiffiffiffi
l2
vy ¼ vx 1 (e)
d2
Solving Eqs. (d) and (e) for vx and vy leads to
rffiffiffiffiffiffiffiffiffiffiffiffi
d2 d2 d
vx ¼  2 vo vy ¼  1  2 vo (f)
l l I
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Thus, the speed, v ¼ vx 2 + vy 2 , after the string becomes taut is
d
v ¼ vo
l
(b) From Eq. (1.40), the impulse on m during the time it takes the string to become taut is
" rffiffiffiffiffiffiffiffiffiffiffiffi ! #   rffiffiffiffiffiffiffiffiffiffiffiffi
d2 ^ d2 d ^   d2 d2 d
I ¼ mðv2  v1 Þ ¼ m  2 vo i  1  2 vo j  vo i ¼ 1  2 mvo i  1  2 mvo^j
^ ^
l l l l l l

The magnitude of this impulse, which is directed along the string, is


rffiffiffiffiffiffiffiffiffiffiffiffi
d2
I ¼ kI k ¼ 1  2 mvo
l
Hence, the average force in the string during the small time interval Δt required to change the direction of the velocity
vector turns out to be
sffiffiffiffiffiffiffiffiffiffiffiffi
I d2 mvo
Favg ¼ ¼ 1  2
Δt l Δt

1.6 TIME DERIVATIVES OF MOVING VECTORS


Fig. 1.15(a) shows a vector A inscribed in a rigid body B that is in motion relative to an inertial frame of
reference (a rigid, Cartesian coordinate system, which is fixed relative to the fixed stars). The magni-
tude of A is fixed. The body B is shown at two times, separated by the differential time interval dt. At
time t + dt, the orientation of vector A differs slightly from that at time t, but its magnitude is the same.
According to one of the many theorems of the prolific 18th-century Swiss mathematician Leonhard
Euler (1707–1783), there is a unique axis of rotation about which B, and therefore A, rotates during
the differential time interval. If we shift the two vectors A(t) and A(t + dt) to the same point on the
axis of rotation, so that they are tail to tail, as shown in Fig. 1.15(b), we can assess the difference
dA between them caused by the infinitesimal rotation. Remember that shifting a vector to a parallel
line does not change the vector. The rotation of the body B is measured in the plane perpendicular
to the instantaneous axis of rotation. The amount of rotation is the angle dθ through which a line
element normal to the rotation axis turns in the time interval dt. In Fig. 1.15(b) that line element is
22 CHAPTER 1 DYNAMICS OF POINT MASSES

FIG. 1.15
Displacement of a rigid body. (a) Change in orientation of an embedded vector A. (b) Differential rotation of A
about the instantaneous rotation axis.

the component of A normal to the axis of rotation. We can express the difference dA between A(t) and
A(t + dt) as
magnitude of dA
zfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflffl{
dA ¼ ½ðkAk  sinϕÞdθ n ^ (1.47)
where n^ is the unit normal to the plane defined by A and the axis of rotation, and it points in the direction
of the rotation. The angle ϕ is the inclination of A to the rotation axis. By definition,
dθ ¼ kωkdt (1.48)
where ω is the angular velocity vector, which points along the instantaneous axis of rotation, and its
direction is given by the right-hand rule. That is, wrapping the right hand around the axis of rotation,
with the fingers pointing in the direction of dθ, results in the thumb defining the direction of ω. This is
evident in Fig. 1.15(b). It should be pointed out that the time derivative of ω is the angular acceleration,
usually given the symbol α. Thus,

α¼ (1.49)
dt
Substituting Eq. (1.48) into Eq. (1.47), we get
^ ¼ ðkωk  kAk  sinϕÞ^
dA ¼ kAk  sinϕ  kωkdt  n n dt (1.50)
By definition of the cross product, ω  A is the product of the magnitude of ω, the magnitude of A, the
sine of the angle between ω and A, and the unit vector normal to the plane of ω and A, in the rotation
direction. That is,
^
ω  A ¼ kωk  kAk  sinϕ  n (1.51)
Substituting Eq. (1.51) into Eq. (1.50) yields
dA ¼ ω  Adt
1.6 TIME DERIVATIVES OF MOVING VECTORS 23

Dividing through by dt, we finally obtain


 
dA d
¼ωA if kAk ¼ 0 (1.52)
dt dt

Eq. (1.52) is a formula we can use to compute the time derivative of any vector of constant magnitude.

EXAMPLE 1.11
Calculate the second time derivative of a vector A of constant magnitude, expressing the result in terms of ω and its de-
rivatives and A.
Solution
Differentiating Eq. (1.52) with respect to time, we get
d2 A d dA d dω dA
¼ ¼ ðω  AÞ ¼ A+ω
dt2 dt dt dt dt dt
Using Eqs. (1.49) and (1.52), this can be written

d2 A
¼ α  A + ω  ðω  AÞ (1.53)
dt2

EXAMPLE 1.12
Calculate the third derivative of a vector A of constant magnitude, expressing the result in terms of ω and its derivatives
and A.
Solution
d3 A d d2 A d
¼ ¼ ½α  A + ω  ðω  AÞ
dt3 dt dt2 dt
d d
¼ ðα  AÞ + ½ω  ðω  AÞ
dt
 dt 
dα dA dω d
¼ A+α +  ðω  AÞ + ω  ðω  AÞ
dt dt dt dt
 
dα dω dA
¼  A + α  ðω  AÞ + α  ðω  AÞ + ω  A+ω
dt dt dt

¼  A + α  ðω  AÞ + fα  ðω  AÞ + ω  ½α  A + ω  ðω  AÞg
dt

¼  A + α  ðω  AÞ + α  ðω  AÞ + ω  ðα  AÞ + ω  ½ω  ðω  AÞ
dt

¼  A + 2α  ðω  AÞ + ω  ðα  AÞ + ω  ½ω  ðω  AÞ
dt
d3 A dα
¼  A + 2α  ðω  AÞ + ω  ½α  A + ω  ðω  AÞ
dt3 dt

Let XYZ be a rigid inertial frame of reference and xyz a rigid moving frame of reference, as shown in
Fig. 1.16. The moving frame can be moving (translating and rotating) freely on its own accord, or it can
24 CHAPTER 1 DYNAMICS OF POINT MASSES

FIG. 1.16
Fixed (inertial) and moving rigid frames of reference.

be attached to a physical object, such as a car, an airplane, or a spacecraft. Kinematic quantities mea-
sured relative to the fixed inertial frame will be called absolute (e.g., absolute acceleration), and those
measured relative to the moving system will be called relative (e.g., relative acceleration). The unit
vectors along the inertial XYZ system are ^I, ^ ^ whereas those of the moving xyz system are
J, and K,
^i, ^j, and k.
^ The motion of the moving frame is arbitrary, and its absolute angular velocity is Ω. If, how-
ever, the moving frame is rigidly attached to an object, so that it not only translates but also rotates with
it, then the frame is called a body frame and the axes are referred to as body axes. A body frame clearly
has the same angular velocity as the body to which it is bound.
Let B be any time-dependent vector. Resolved into components along the inertial frame of refer-
ence, it is expressed analytically as
B ¼ BX^I + BY ^ ^
J + BZ K

where BX, BY, and BZ are functions of time. Since ^I, ^J, and K
^ are fixed, the time derivative of B is simply
dB dBX ^ dBY ^ dBZ ^
¼ I+ J+ K
dt dt dt dt
dBX/dt, dBY/dt, and dBZ/dt are the components of the absolute time derivative of B.
B may also be resolved into components along the moving xyz frame, so that, at any instant,
B ¼ Bx^i + By^j + Bz k
^ (1.54)

Using this expression to calculate the time derivative of B yields


dB dBx ^ dBy ^ dBz ^ d^i d^j ^
dk
¼ i+ j+ k + Bx + By + Bz (1.55)
dt dt dt dt dt dt dt
The orthogonal unit vectors ^i, ^j, and k ^ are not fixed in space but are continuously changing
direction; therefore, their time derivatives are not zero. They obviously have a constant magnitude
1.6 TIME DERIVATIVES OF MOVING VECTORS 25

(unity) and, being attached to the xyz frame, they all have the angular velocity Ω. It follows from
Eq. (1.52) that
d^i d^j ^
dk
¼ Ω ^i ¼ Ω  ^j ^
¼ Ωk
dt dt dt
Substituting these on the right-hand side of Eq. (1.55) yields
dB dBx ^ dBy ^ dBz ^     
¼ i+ j+ k + Bx Ω ^i + By Ω  ^j + Bz Ω  k^
dt dt dt dt
dBx ^ dBy ^ dBz ^     
¼ i+ j+ k + Ω  Bx^i + Ω  By^j + Ω  Bz k ^
dt dt dt
dBx ^ dBy ^ dBz ^  
¼ i+ j+ k + Ω  Bx^i + By^j + Bz k
^
dt dt dt
In view of Eq. (1.54), this can be written as

dB dB
¼ + ΩB (1.56)
dt dt rel
where

dB dBx ^ dBy ^ dBz ^
¼ i+ j+ k (1.57)
dt rel dt dt dt
dB/dt)rel is the time derivative of B relative to the moving frame. Eq. (1.56) shows how the absolute
time derivative is obtained from the relative time derivative. Clearly, dB/dt ¼ dB/dt)rel only when the
moving frame is in pure translation (Ω ¼ 0).
Eq. (1.56) can be used recursively to compute higher order time derivatives. Thus, differentiating
Eq. (1.56) with respect to t, we get

d2 B d dB dΩ dB
¼ + B+Ω
dt2 dt dt rel dt dt
Using Eq. (1.56) in the last term yields
 
d2 B d dB dΩ dB
¼ + B+Ω + ΩB (1.58)
dt2 dt dt rel dt dt rel
Eq. (1.56) also implies that
  
d dB d2 B dB
¼ +Ω (1.59)
dt dt rel dt2 rel dt rel

where

d2 B d2 Bx ^ d2 By ^ d2 Bz ^
¼ i+ 2 j+ 2 k
dt2 rel dt2 dt dt
Substituting Eq. (1.59) into Eq. (1.58) yields
  
d2 B d2 B dB dΩ dB
¼ + Ω  +  B + Ω  +ΩB
dt2 dt2 rel dt rel dt dt rel
26 CHAPTER 1 DYNAMICS OF POINT MASSES

Collecting terms, this becomes


 
d2 B d2 B _  B + Ω  ðΩ  BÞ + 2Ω  dB
¼ 2 +Ω (1.60)
dt2 dt rel dt rel
_  dΩ=dt is the absolute angular acceleration of the xyz frame.
where Ω
Formulas for higher order time derivatives are found in a similar fashion.

1.7 RELATIVE MOTION


Let P be a particle in arbitrary motion. The absolute position vector of P is r and the position of P
relative to the moving frame is rrel. If rO is the absolute position of the origin of the moving frame,
then it is clear from Fig. 1.17 that
r ¼ rO + rrel (1.61)
Since rrel is measured in the moving frame,
rrel ¼ x^i + y^j + zk
^ (1.62)
where x, y, and z are the coordinates of P relative to the moving reference.
The absolute velocity v of P is dr/dt, so that from Eq. (1.61) we have
drrel
v ¼ vO + (1.63)
dt
where vO ¼ drO/dt is the (absolute) velocity of the origin of the xyz frame. From Eq. (1.56), we can write
drrel
¼ vrel + Ω  rrel (1.64)
dt
where vrel is the velocity of P relative to the xyz frame (so that ^i, ^j, and k
^ are held fixed):

drrel dx^ dy^ dz ^
vrel ¼ ¼ i+ j+ k (1.65)
dt rel dt dt dt

FIG. 1.17
Absolute and relative position vectors.
1.7 RELATIVE MOTION 27

Substituting Eq. (1.64) into Eq. (1.63) yields


v ¼ vO + Ω  rrel + vrel (1.66)

The absolute acceleration a of P is dv/dt, so that from Eq. (1.63) we have


d2 rrel
a ¼ aO + (1.67)
dt2
where aO ¼ dvO/dt is the absolute acceleration of the origin of the xyz frame. We evaluate the second
term on the right using Eq. (1.60).
 
d2 rrel d2 rrel _  rrel + Ω  ðΩ  rrel Þ + 2Ω  drrel
¼ 2 +Ω (1.68)
dt2 dt rel dt rel

Since vrel ¼ drrel/dt)rel and arel ¼ d2rrel/dt2)rel, this can be written


d2 rrel _  rrel + Ω  ðΩ  rrel Þ + 2Ω  vrel
¼ arel + Ω (1.69)
dt2
Upon substituting this result into Eq. (1.67), we find
_  rrel + Ω  ðΩ  rrel Þ + 2Ω  vrel + arel
a ¼ aO + Ω (1.70)

The cross product 2Ω  vrel is called the Coriolis acceleration after Gustave Gaspard de Coriolis
(1792–1843), the French mathematician who introduced this term (Coriolis, 1835). Because of the
number of terms on the right, Eq. (1.70) is sometimes referred to as the five-term acceleration formula.

EXAMPLE 1.13
At a given instant, the absolute position, velocity, and acceleration of the origin O of a moving frame are
9
rO ¼ 100^I + 200^J + 300K
^ ðmÞ > >
=
vO ¼ 50^I + 30^J  10K^ ðm=sÞ ðgivenÞ (a)
 >
>
aO ¼ 15^I + 40J^ + 25K^ m=s2 ;

The angular velocity and acceleration of the moving frame are


)
Ω ¼ 1:0^I  0:4^J + 0:6K
^ ðrad=sÞ
 ðgivenÞ (b)
_ ^ ^
Ω ¼ 1:0I  0:3J  0:4K ^ rad=s2

The unit vectors of the moving frame are


9
^i ¼ 0:5571^I + 0:7428^J + 0:3714K
^ > >
=
^j ¼ 0:06331^I + 0:4839^J  0:8728K
^ ðgivenÞ (c)
>
>
^ ¼ 0:8280^I + 0:4627^J + 0:3166K
k ^ ;

The absolute position, velocity, and acceleration of P are


9
r ¼ 300^I  100^J + 150K ^ ðmÞ >
>
=
v ¼ 70^I + 25J^  20K^ ðm=sÞ ðgivenÞ (d)
 >
>
a ¼ 7:5^I  8:5^J + 6:0K
^ m=s2 ;

Find (a) the velocity vrel and (b) the acceleration arel of P relative to the moving frame.
28 CHAPTER 1 DYNAMICS OF POINT MASSES

Solution
^ in terms of ^i, ^j, and k
Let us first use Eq. (c) to solve for ^I, ^J, and K ^ (three equations in three unknowns):

^I ¼ 0:5571^i  0:06331^j  0:8280k^


^J ¼ 0:7428^i + 0:4839^j + 0:4627k^ (e)
K^ ¼ 0:3714^i  0:8728^j  0:3166k ^
(a) The relative position vector is
 
rrel ¼ r  rO ¼ 300^I  100^J + 150K
^  100^I + 200^J + 300K
^ ¼ 200^I  300^J  150K
^ ðmÞ (f)
From Eq. (1.66), the relative velocity vector is

vrel ¼ v  vO  Ω  rrel
^I J^ K^
 
¼ 70^I + 25^J  20K
^  50^I + 30^J  10K
^  1:0 0:4 0:6
200 300 150
  
¼ 70^I + 25J^  20K
^  50^I + 30J^  10K
^  240^I + 270^J  220K
^
or

vrel ¼ 120^I  275^J + 210K


^ ðm=sÞ (g)
To obtain the components of the relative velocity along the axes of the moving frame, substitute Eq. (e) into Eq. (g),
 
vrel ¼ 120 0:5571^i  0:06331^j  0:8280k ^
   
 275 0:7428^i + 0:4839^j + 0:4627k
^ + 210 0:3714^i  0:8728^j + 0:3166k
^

so that

vrel ¼ 193:1^i  308:8^j + 38:60k


^ ðm=sÞ (h)

^v in the direction of vrel,


Alternatively, in terms of the unit vector u
 
vrel ¼ 366:2^
uv ðm=sÞ u ^v ¼ 0:5272^i  0:8432^j + 0:1005k
^ (i)

(b) To find the relative acceleration, we use the five-term acceleration formula, Eq. (1.70):

_  rrel  Ω  ðΩ  rrel Þ  2ðΩ  vrel Þ


arel ¼ a  aO  Ω
^I J^ K^ ^I ^
K J^ ^I ^J ^
K
¼ a  aO  1:0 0:30:4  Ω  1:0 0:4 0:6  2 1:0 0:4 0:6
200 300 150 200 300 150 120 275 210
^I ^J ^
K
 
^  1:0 0:4 0:6  162^I  564^J  646K
¼ a  aO  165^I  230^J + 240K ^
240 270 220
  
¼ 7:5^I  8:5^J + 6K
^  15^I + 40^J + 25K
^  165^I  230^J + 240K
^
 
^ ^ ^ ^ ^
 74I + 364J + 366K  162I  564J  646K ^

arel ¼ 99:5^I + 381:5^J + 21:0K


^ ðm=sÞ2 (j)
The components of the relative acceleration along the axes of the moving frame are found by substituting Eq. (e) into
Eq. ( j):
 
arel ¼ 99:5 0:5571^i  0:06331^j  0:8282k ^
   
+ 381:5 0:7428^i + 0:4839^j + 0:4627k ^ + 21:0 0:3714^i  0:8728^j + 0:3166k
^

arel ¼ 346:6^i + 160:0^j + 100:8k


^ ðm=s2 Þ (k)
1.7 RELATIVE MOTION 29

Or, in terms of the unit vector ûa in the direction of arel,


 
arel ¼ 394:8^ua ðm=s2 Þ ^ ua ¼ 0:8778^i + 0:4052^j + 0:2553k
^ (l)

Fig. 1.18 shows the nonrotating inertial frame of reference XYZ with its origin at the center C of the
earth, which we shall assume to be a sphere. That assumption will be relaxed in Chapter 5. Embedded in
the earth and rotating with it is the orthogonal x0 y0 z0 frame, also centered at C, with the z0 axis parallel to
Z, the earth’s axis of rotation. The x0 axis intersects the equator at the prime meridian (0 degree
longitude), which passes through Greenwich in London, England. The angle between X and x0 is
θG, and the rate of increase of θG is just the angular velocity Ω of the earth. P is a particle (e.g., an
airplane or spacecraft), which is moving in an arbitrary fashion above the surface of the earth. rrel is
the position vector of P relative to C in the rotating x0 y0 z0 system. At a given instant, P is directly over
point O, which lies on the earth’s surface at longitude Λ and latitude ϕ. Point O coincides instanta-
neously with the origin of what is known as a topocentric-horizon coordinate system xyz. For our
purposes, x and y are measured positive eastward and northward along the local latitude and meridian,

FIG. 1.18
0 0 0
Earth-centered inertial frame (XYZ); earth-centered noninertial x y z frame embedded in and rotating with the
earth; and a noninertial, topocentric-horizon frame xyz attached to a point O on the earth’s surface.
30 CHAPTER 1 DYNAMICS OF POINT MASSES

respectively, through O. The tangent plane to the earth’s surface at O is the local horizon. The z axis is
the local vertical (straight up), and it is directed radially outward from the center of the earth. The unit
vectors of the xyz frame are ^i^jk,
^ as indicated in Fig. 1.18. Keep in mind that O remains directly below P,
so that as P moves, so do the xyz axes. Thus, the ^i^jk
^ triad, which comprises the unit vectors of a spherical
coordinate system, varies in direction as P changes location, thereby accounting for the curvature of
the earth.
Let us find the absolute velocity and acceleration of P. It is convenient to first obtain the velocity
and acceleration of P relative to the nonrotating earth, and then use Eqs. (1.66) and (1.70) to calculate
their inertial values.
The relative position vector can be written
^
rrel ¼ ðRE + zÞk (1.71)
where RE is the radius of the earth, and z is the height of P above the earth (i.e., its altitude). The time
derivative of rrel is the velocity vrel relative to the nonrotating earth,
^
vrel ¼
drrel ^ + ðRE + zÞ dk
¼ z_ k (1.72)
dt dt
^
To calculate dk=dt, we must use Eq. (1.52). The angular velocity ω of the xyz frame relative to the
nonrotating earth is found in terms of the rates of change of latitude ϕ and longitude Λ,
ω ¼ ϕ_ ^i + Λ_ cos ϕ^j + Λ_ sin ϕk
^ (1.73)
Thus,
^
dk ^ ¼ Λ_ cos ϕ^i + ϕ_ ^j
¼ ωk (1.74)
dt
Let us also record the following for future use:
d^j
¼ ω  ^j ¼ Λ_ sin ϕ^i  ϕ_ k
^ (1.75)
dt
d^i
¼ ω ^i ¼ Λ_ sin ϕ^j  Λ_ cosϕk
^ (1.76)
dt
Substituting Eq. (1.74) into Eq. (1.72) yields the velocity in the nonrotating frame resolved along the
topocentric-horizon axes,
vrel ¼ x_^i + y_^j + z_ k
^ (1.77a)
where
x_ ¼ ðRE + zÞΛ_ cosϕ y_ ¼ ðRE + zÞϕ_ (1.77b)
It is convenient to use these results to express the rates of change of latitude and longitude in terms of
the components of relative velocity over the earth’s surface,
y_ x_
ϕ_ ¼ Λ_ ¼ (1.78)
RE + z ðRE + zÞ cos ϕ
1.7 RELATIVE MOTION 31

The time derivatives of these two expressions are


ðRE + zÞy€ y_ z_ ðRE + zÞ x€cos ϕ  ðz_ cos ϕ  y_ sinϕÞx_
ϕ€ ¼ Λ€ ¼ (1.79)
ðRE + zÞ 2 ðRE + zÞ2 cos 2 ϕ
The acceleration of P relative to the nonrotating earth is found by taking the time derivative of vrel.
From Eqs. (1.77a) and (1.77b) we thereby obtain
d^i d^j ^
dk
arel ¼ x€^i + y€^j + z€k
^ + x_
+ y_ + z_
dt dt dt
 
¼ z_ Λ_ cosϕ + ðRE + zÞΛ€ cosϕ  ðRE + zÞϕ_ Λ_ sinϕ ^i + z_ ϕ_ + ðRE + zÞϕ€ ^j + z€k
^
    
+ ðRE + zÞΛ_ cos ϕ ω ^i + ðRE + zÞϕ_ ω  ^j + z_ ω  k ^

Substituting Eq. (1.74) through Eq. (1.76) together with Eqs. (1.78) and (1.79) into this expression
yields, upon simplification,
   
x_ ðz_  y_ tan ϕÞ ^ y_ z_ + x_ 2 tan ϕ ^ x_ 2 + y_ 2 ^
arel ¼ x€+ i + y€+ j + z€ k (1.80)
RE + z RE + z RE + z
Observe that the curvature of the earth’s surface is neglected by letting RE + z become infinitely large,
in which case
arel Þneglecting earth’s curvature ¼ x€^i + y€^j + z€k
^

That is, for a “flat earth,” the components of the relative acceleration vector are just the derivatives of
the components of the relative velocity vector.
For the absolute velocity we have, according to Eq. (1.66),
v ¼ vC + Ω  rrel + vrel (1.81)
^ ¼ cos ϕ^j + sin ϕk,
From Fig. 1.18, it can be seen that K ^ which means the angular velocity of the earth is
^ ¼ Ω cos ϕ^j + Ω sin ϕk
Ω ¼ ΩK ^ (1.82)
Substituting this, together with Eqs. (1.71) and (1.77a) and the fact that vC ¼ 0, into Eq. (1.81) yields
v ¼ ½x_ + ΩðRE + zÞ cosϕ^i + y_^j + z_ k
^ (1.83)
From Eq. (1.70) the absolute acceleration of P is
_  rrel + Ω  ðΩ  rrel Þ + 2Ω  vrel + arel
a ¼ aC + Ω
_ ¼ 0, we find, upon substituting Eqs. (1.71), (1.77a), (1.80), and (1.82), that
Since aC ¼ Ω
x_ ðz_  y_ tanϕÞ
a ¼ x€+ + 2Ωðz_ cos ϕ  y_ sinϕÞ ^i
 RE + z 
y_ z_ + x_ tan ϕ
2
+ y€+ + Ω sinϕ½ΩðRE + zÞcos ϕ + 2x_  ^j (1.84)
 RE + z 
x_ 2 + y_ 2 ^
+ z€  Ωcosϕ½ΩðRE + zÞ cos ϕ + 2x_  k
RE + z
32 CHAPTER 1 DYNAMICS OF POINT MASSES

Some special cases of Eqs. (1.83) and (1.84) follow.


Straight and level, unaccelerated flight: z_ ¼ z€¼ x€¼ y€¼ 0
v ¼ ½x_ + ΩðRE + zÞ cos ϕ^i + y_^j (1.85a)
 
x_ y_ tanϕ x_ tan ϕ 2
a¼ + 2Ωy_ sin ϕ ^i + + Ω sin ϕ½ΩðRE + zÞcos ϕ + 2x_  ^j
RE + z RE + z
 2 2  (1.85b)
x_ + y_ ^
 + Ω cos ϕ½ΩðRE + zÞ cos ϕ + 2x_  k
RE + z
Flight due north (y) at a constant speed and altitude: z_ ¼ z€¼ x_ ¼ x€¼ y€¼ 0
v ¼ ΩðRE + zÞ cos ϕ^i + y_^j (1.86a)
y_2
a ¼ 2Ωy_ sin ϕ^i + Ω2 ðRE + zÞ sin ϕ cos ϕ^j  ^
+ Ω2 ðRE + zÞ cos 2 ϕ k (1.86b)
RE + z
Flight due east (x) at a constant speed and altitude: z_ ¼ z€¼ x€¼ y_ ¼ y€¼ 0
v ¼ ½x_ + ΩðRE + zÞcos ϕ^i (1.87a)
 2 
x_ tan ϕ
a¼ + Ω sin ϕ½ΩðRE + zÞ cos ϕ + 2x_  ^j
RE +2 z  (1.87b)
x_ ^
 + Ω cos ϕ½ΩðRE + zÞ cos ϕ + 2x_  k
RE + z
Flight straight up (z): x_ ¼ x€¼ y_ ¼ y€¼ 0
v ¼ ΩðRE + zÞ cos ϕ^i + z_ k
^ (1.88a)

a ¼ 2Ωðz_ cos ϕÞ^i + Ω2 ðRE + zÞ sin ϕ cos ϕ^j + z€ Ω2 ðRE + zÞ cos 2 ϕ k^ (1.88b)

Stationary: x_ ¼ x€¼ y_ ¼ y€¼ z_ ¼ z€¼ 0


v ¼ ΩðRE + zÞcos ϕ^i (1.89a)

a ¼ Ω2 ðRE + zÞ sin ϕ cos ϕ^j  Ω2 ðRE + zÞ cos 2 ϕk


^ (1.89b)

EXAMPLE 1.14
An airplane of mass 70,000 kg is traveling due north at a latitude 30°N, at an altitude of 10 km (32,800 ft), with a speed of
300 m/s (671 mph). Calculate (a) the components of the absolute velocity and acceleration along the axes of the
topocentric-horizon reference frame and (b) the net force on the airplane. Assume the winds aloft are zero.

Solution
(a) First, using the sidereal rotation period of the earth in Table A.1, we note that the earth’s angular velocity is
2π radians 2π radians 2π radians
Ω¼ ¼ ¼ ¼ 7:292  105 radians=s
sidereal day 23:93h 86,160 s
From Eq. (1.86a), the absolute velocity is
 
v ¼ ΩðRE + zÞcosϕ^i + y_^j ¼ 7:292  105  ð6378 + 10Þ  103 cos 30∘ ^i + 300^j
1.7 RELATIVE MOTION 33

FIG. 1.19
Components of the net force on the airplane.

or

v ¼ 403:4^i + 300^j ðm=sÞ


The 403.4 m/s (901 mph) component of velocity to the east (x direction) is due entirely to the earth’s rotation.
From Eq. (1.86b), the absolute acceleration is
y_ 2
a ¼ 2Ωy_ sin ϕ^i + Ω2 ðRE + zÞsin ϕ cos ϕ^j  ^
+ Ω2 ðRE + zÞcos 2 ϕ k
 RE +z
5 ∘^
¼ 2 7:292  10  300  sin 30 i

+ 7:292  105  ð6378 + 10Þ  103  sin 30∘  cos30∘^j
2

3002 

2
^
+ 7:292  105  ð6378 + 10Þ  103  cos 2 30∘ k
ð6378 + 10Þ  103
or

a ¼ 0:02187^i + 0:01471^j  0:03956k


^ ðm=s2 Þ

The westward (negative x) acceleration of 0.02187 m/s2 is the Coriolis acceleration.


(b) Since the acceleration in part (a) is the absolute acceleration, we can use it in Newton’s law to calculate the net force on
the airplane,
 
Fnet ¼ ma ¼ 70,000 0:02187^i + 0:01471^j  0:03956k ^

¼ 1531^i + 1029^j  2769k


^ ðNÞ

Fig. 1.19 shows the components of this relatively small force. The forward (y) and downward (negative z) forces are in
the directions of the airplane’s centripetal acceleration, caused by the earth’s rotation and, in the case of the downward
force, by the earth’s curvature as well. The westward force is in the direction of the Coriolis acceleration, which is due
to the combined effects of the earth’s rotation and the motion of the airplane. These net external forces must exist if the
airplane is to fly in the prescribed path.
In the vertical direction, the net force is that of the upward lift L of the wings plus the downward weight W of the
aircraft, so that
Fnet Þz ¼ L  W ¼ 2769 ) L ¼ W  2769N
34 CHAPTER 1 DYNAMICS OF POINT MASSES

Thus, the effect of the earth’s rotation and curvature is to apparently produce an outward centrifugal force, reducing the
weight of the airplane a bit, in this case by about 0.4%. The fictitious centrifugal force also increases the apparent drag
in the flight direction by 1029 N. That is, in the flight direction
Fnet Þy ¼ T  D ¼ 1029 N
where T is the thrust and D is the drag. Hence
T ¼ D + 1029 ðNÞ
The 1531-N force to the left, produced by crabbing the airplane very slightly in that direction, is required to balance
the fictitious Coriolis force, which would otherwise cause the airplane to deviate to the right of its flight path.

1.8 NUMERICAL INTEGRATION


Analysis of the motion of a spacecraft leads to ordinary differential equations with time as the inde-
pendent variable. It is often impractical if not impossible to solve them exactly. Therefore, the ability to
solve differential equations numerically is important. In this section, we will take a look at a few com-
mon numerical integration schemes and investigate their accuracy and stability by applying them to
some problems that do have an analytical solution.
Particle mechanics is based on Newton’s second law (Eq. 1.38), which may be written as
F
€r ¼ (1.90)
m

This is a second-order, ordinary differential equation for the position vector r as a function of time.
Depending on the complexity of the force function F, there may or may not be a closed-form, analytical
solution of Eq. (1.90). In the most trivial case, the force vector F and the mass m are constant, which
means we can use elementary calculus to integrate Eq. (1.90) twice to get
F 2
r¼ t + C1 t + C2 ðF and m constantÞ (1.91)
2m

C1 and C2 are the two vector constants of integration. Since each vector has three components, there are
a total of six scalar constants of integration. If the position and velocity are both specified at time t ¼ 0
to be r0 and r_ 0 , respectively, then we have an initial value problem. Applying the initial conditions to
Eq. (1.91), we find C1 ¼ r_ 0 and C2 ¼ r0, which means
F 2
r¼ t + r_ 0 t + r0 ðF and m constantÞ
2m

On the other hand, we may know the position r0 at t ¼ 0 and the velocity r_ f at a later time t ¼ tf. These
are boundary conditions and this is an example of a boundary value problem. Applying the boundary
conditions to Eq. (1.91) yields C1 ¼ r_ f  ðF=mÞtf and C2 ¼ r0, which means
 
F 2 F
r¼ t + r_ f  tf t + r0 ðF and m constantÞ
2m m

For the remainder of this section we will focus on the numerical solution of initial value problems only.
1.8 NUMERICAL INTEGRATION 35

In general, the function F in Eq. (1.90) is not constant but is instead a function of time t, position r,
and velocity r_ . That is, F ¼ Fðt, r, r_ Þ. Let us resolve the vector r and its derivatives as well as the force F
into their Cartesian components in three-dimensional space:
r ¼ x^i + y^j + zk
^ r_ ¼ x_^i + y_^j + z_ k
^ €r ¼ x€^i + y€^j + z€k
^ F ¼ Fx^i + Fy^j + Fz k
^

The three components of Eq. (1.90) are


Fx ðt, r, r_ Þ Fy ðt, r, r_ Þ Fz ðt, r, r_ Þ
x€¼ y€¼ z€¼ (1.92)
m m m
These are three second-order differential equations. For the purpose of numerical solution, they must be
reduced to six first-order differential equations. This is accomplished by introducing six auxiliary vari-
ables y1 through y6, defined as follows:
y1 ¼ x y2 ¼ y y3 ¼ z
(1.93)
y4 ¼ x_ y5 ¼ y_ y6 ¼ z_
In terms of these auxiliary variables, the position and velocity vectors are
r ¼ y1^i + y2^j + y3 k
^ r_ ¼ y4^i + y5^j + y6 k
^

Taking the derivative d/dt of each of the six expressions in Eq. (1.93) yields
dy1 =dt ¼ x_ dy2 =dt ¼ y_ dy3 =dt ¼ z_
dy4 =dt ¼ x€ dy5 =dt ¼ y€ dy6 =dt ¼ z€
Upon substituting Eqs. (1.92) and (1.93), we arrive at the six first-order differential equations
y_ 1 ¼ y4
y_ 2 ¼ y5
y_ 3 ¼ y6
Fx ðt, y1 , y2 , y3, y4, y5, y6 Þ
y_ 4 ¼ (1.94)
m
Fy ðt, y1 , y2 , y3, y4, y5, y6 Þ
y_ 5 ¼
m
Fz ðt, y1 , y2 , y3, y4, y5, y6 Þ
y_ 6 ¼
m
These equations are coupled because the right-hand side of each one contains variables that belong to
other equations as well. Eq. (1.94) can be written more compactly in vector notation as
y_ ¼ f ðt, yÞ (1.95)
_ and f are
where the column vectors y, y,
8 9 8 9 8 9
>
> y1 >
> >
> y_ 1 >
> >
> y4 >
>
>y >
> > >
> y_ > > >
> >
>
>
< 2> = >
< 2> = >
<
y5 >
=
y3 y_ 3 y6
y¼ y_ ¼ f¼ (1.96)
>
> y4 > > y_ 4 > > Fx ðt, yÞ=m >
>y >
> >
>
>
>
> _
>
>
>
>
>
>
>
>
>
>
: ;5 > > y
: ; 5 > : Fy ðt, yÞ=m >
> ;
y6 y_ 6 Fz ðt, yÞ=m
36 CHAPTER 1 DYNAMICS OF POINT MASSES

Note that in this case f(t, y) is shorthand for f(t, y1, y2, y3, y4, y5, y6). Any set of one or more ordinary
differential equations of any order can be cast in the form of Eq. (1.95).

EXAMPLE 1.15
Write the third-order nonlinear differential equation

x xx€+ x_ 2 ¼ 0 (a)
as three first-order differential equations.
Solution
Introducing the three auxiliary variables
y1 ¼ x y2 ¼ x_ y3 ¼ x€ (b)
we take the derivative of each one to get
dy1 =dt ¼ dx=dt ¼ x_
_
dy2 =dt ¼ dx=dt ¼ x€
From ðaÞ
… z}|{
€ ¼x
dy3 =dt ¼ dx=dt ¼ xx€ x_ 2
Substituting Eq. (b) on the right of these expressions yields
y_ 1 ¼ y2
y_ 2 ¼ y3 (c)
y_ 3 ¼ y1 y3  y22
This is a system of three first-order, coupled ordinary differential equations. It is an autonomous system, since time t does
not appear explicitly on the right-hand side. The three equations can therefore be written compactly as y_ ¼ f ðyÞ.

Before discussing some numerical integration schemes, it will be helpful to review the concept of
the Taylor series, named after the English mathematician Brook Taylor (1685–1731). Recall from cal-
culus that if we know the value of a function g(t) at time t and wish to approximate its value at a neigh-
boring time t + h, we can use the Taylor series to express g(t + h) as an infinite power series in h,

gðt + hÞ ¼ gðtÞ + c1 h + c2 h2 + c3 h3 + ⋯ + cn hn + O hn + 1 (1.97)

The coefficients cm are found by taking successively higher order derivatives of g(t) according to the
formula
1 dm gðtÞ
cm ¼ (1.98)
m! dtm
O(hn+1) (“order of h to the n + 1”) means that the remaining terms of this infinite series all have hn+1 as a
factor. In other words,
O ð hn + 1 Þ
lim ¼ cn + 1
h!0 hn + 1
O(hn+1) is the truncation error due to retaining only terms up to hn. The order of a Taylor series expan-
sion is the highest power of h retained. The more terms of the Taylor series that we keep, the more
accurate will be the representation of the function g(t + h) in the neighborhood of t. Reducing h lowers
the truncation error. For example, if we reduce h to h/2, then O(hn) goes down by a factor of (1/2)n.
1.8 NUMERICAL INTEGRATION 37

EXAMPLE 1.16
Expand the function sin(t + h) in a Taylor series about t ¼ 1. Plot the Taylor series of order 1, 2, 3, and 4 and compare them
with sin(1 + h) for 2 < h < 2.
Solution
The nth-order Taylor power series expansion of sin(t +h) is written
sin ðt + hÞ ¼ pn ðhÞ
where, according to Eqs. (1.97) and (1.98), the polynomial pn is given by
X
n
hm dm sin t
pn ðhÞ ¼
m¼0
m! dtm
Thus, the zeroth- through fourth-order Taylor series polynomials in h are
h0 d0 sin t
p0 ¼ ¼ sin t
0! dt0
h dsin t
p1 ¼ p0 + ¼ sin t + h cost
1! dt
2 2
h d sint h2
p2 ¼ p1 + ¼ sint + h cost  sint
2! dt2 2
h3 d3 sint h2 h3
p3 ¼ p2 + 3
¼ sint + h cost  sint  cost
3! dt 2 6
h4 d4 sint h2 h3 h4
p4 ¼ p3 + ¼ sint + h cost  sint  cost + sin t
4! dt4 2 6 24
For t ¼ 1, p1 through p4 as well as sin(t + h) are plotted in Fig. 1.20. As expected, we see that the higher degree Taylor
polynomials for sin(1+ h) lie closer to sin(1+ h) over a wider range of h.

FIG. 1.20
Plots of zeroth- to fourth-order Taylor series expansions of sin(1 + h).

The numerical integration schemes that we shall examine are designed to solve first-order ordinary
differential equations of the form shown in Eq. (1.95). To obtain a numerical solution of y_ ¼ f ðt, yÞ
over the time interval t0 to tf, we divide or “mesh” the interval into N discrete times t1, t2, t3, … ,
tN, where t1 ¼ t0 and tN ¼ tf. The step size h is the difference between two adjacent times on the mesh
38 CHAPTER 1 DYNAMICS OF POINT MASSES

(i.e., h ¼ ti+1  ti). h may be constant for all steps across the entire time span t0 to tf. Modern methods
have adaptive step size control in which h varies from step to step to provide better accuracy and
efficiency.
Let us denote the values of y and y_ at time ti as yi and fi, respectively, where fi ¼ f(ti, yi). In an initial
value problem, the values of all components of y at the initial time t0 together with Eq. (1.95) provide
the information needed to determine y at the subsequent discrete times.

1.8.1 RUNGE-KUTTA METHODS


The Runge-Kutta (RK) methods were originally developed by the German mathematicians Carl Runge
(1856–1927) and Martin Kutta (1867–1944). In the explicit, single-step RK methods, yi+1 at ti + h is
obtained from yi at ti by the formula
yi + 1 ¼ yi + hϕðti, yi , hÞ (1.99)
The increment function ϕ is an average of the derivative dy/dt over the time interval ti to ti + h. This
average is obtained by evaluating the derivative f(t, y) at several points or “stages” within the time
interval. The order of an RK method reflects the accuracy to which ϕ is computed, compared with
a Taylor series expansion. An RK method of order p is called an RKp method. An RKp method is
as accurate in computing yi from Eq. (1.99) as is the pth-order Taylor series
yðti + hÞ ¼ yi + c1 h + c2 h2 + ⋯cp hp (1.100)
An attractive feature of the RK schemes is that only the first derivative f(t, y) is required, and it is avail-
able from the differential equation itself (Eq. (1.95)). By contrast, the pth-order Taylor series expansion
in Eq. (1.100) requires computing all derivatives of y through order p.
The higher the RK order, the more stages there are and the more accurate is ϕ. The number of stages
equals the order of the RK method if the order is less than 5. If the number of stages is s, then there are s
times et within the interval ti to ti + h at which we evaluate the derivatives f(t, y). These times are given
by specifying numerical values of the nodes am in the expression
e
tm ¼ ti + am h m ¼ 1,2, …,s
At each of these times the value of e
y is obtained by providing numerical values for the coupling co-
efficients bmn in the formula
X
m1
e
ym ¼ yi + h bmnef n m ¼ 1,2, …,s (1.101)
n¼1

The vector of derivatives ef m is evaluated at stage m by substituting e


tm and e
ym into Eq. (1.95),

ef m ¼ f etm , e
ym m ¼ 1, 2,…, s (1.102)

The increment function ϕ is a weighted sum of the derivatives ef m over the s stages within the time
interval ti to ti + h,
X
s
ϕ¼ cmef m (1.103)
m¼1
Another random document with
no related content on Scribd:
with his wooden gun. But, above all, my attention was attracted by
his sister, a girl of eleven years, as lovely as Cupid, quiet, pensive,
pale, with large, musing eyes, slightly projecting out of their circles.
The other children had somehow offended her; for that reason, she
came into the very room where I sat, and, betaking herself into a
corner, was soon occupied with her doll. The guests looked with
great deference in the direction of her father, a wealthy proprietor,
and some one mentioned in a half-whisper that a dowry of three
hundred thousand rubles had already been laid aside for her.
I turned around to glance at those interested in this circumstance,
and my gaze fell upon Julian Mastakovich, who, having thrust his
hands behind him and inclined his head a trifle to the side, was
listening with a marked intentness to the chatter of these folk.
Afterward I could not help but feel astonished at the sageness of the
hosts in distributing the children’s gifts. The little girl who already had
a dowry of three hundred thousand rubles received the most
expensive doll. Then followed the other gifts, growing lower in value
in proportion to the lower standing of the parents of these happy
children. The last youngster, a boy of ten years, meagre, diminutive,
freckled, and red-haired, received only a small volume of tales
dealing with the bountifulness of nature, the joy of tears, and the like;
the book contained no pictures, not even a decoration. He was the
son of a poor widow, the governess of the host’s children, and had a
haunted, suppressed look. He was dressed in a wretched cotton
jacket. Having received his book, he hovered for a long time around
the toys. He had the most intense longing to play with the other
children, but dared not. It was evident that he already felt and
understood his position.
It is a favorite occupation of mine to observe children. It is highly
interesting to mark in them certain early and free inclinations of their
natures. I noted how the red-haired boy was tempted by the
expensive playthings of the other children—and especially by a toy
theatre, in which he showed a most eager desire to play some rôle—
to such a degree that he adopted an ingratiating manner to attain his
end. He smiled and joined the other children in their play, gave up
his apple to one puffed-up youngster who already had a whole
handkerchiefful of gifts tied to his body, and even offered to carry
another boy on his back, if only they would not drive him away from
the theatre. Soon, however, a bully in the party gave him a sound
drubbing. The boy did not dare to cry out. Presently the governess,
his mother, appeared, and ordered him not to interfere with the other
children’s play. The boy came into the room where the little girl was.
She permitted him to join her, and the two of them were at once
absorbed very earnestly in the rich doll.
I had been sitting in the ivy bower a half-hour and had almost dozed
off, while listening to the small chatter of the red-haired boy and the
beauty with three hundred thousand rubles’ dowry, solicitous over
the doll, when suddenly Julian Mastakovich walked into the room. He
took advantage of a particularly disgraceful quarrel among the
children to steal out of the reception-room. I had noticed that only a
few moments before he was discussing very fervently with the father
of the future rich bride, whose acquaintance he had only just made,
the preëminence of one kind of service over another. At this instant
he stood as if lost in thought, and seemed to be making a calculation
of some sort upon his fingers.
“Three hundred ... three hundred,” he whispered. “Eleven ... twelve
... thirteen ... sixteen ... five years! Say, at four per cent—five times
twelve equal sixty; at compound interest ... well, let us suppose in
five years it ought to reach four hundred. Yes, that’s it.... But the
rascal surely has it salted away at more than four per cent. Eight or
ten is more likely. Well, let’s say five hundred—five hundred
thousand at the very least; not counting a few extra for rags ... h’m
...”
Having ended his calculation, he sneezed vigorously and moved to
leave the room, when suddenly, his eye alighting upon the little girl,
he stopped. He did not see me behind the vases of flowers. He
seemed to me to be violently agitated. Either his calculation had
upset him, or something else; but he did not know what to do with his
hands, and was unable to remain on one spot. His agitation
increased— ne plus ultra—when he stopped and threw another
determined glance at the future bride. He was about to move
forward, but first looked around. Then he approached the child on his
tiptoes, as if conscious of guilt. Smiling, he bent over her and kissed
her head; while she, not expecting this onslaught, cried out from
fright.
“What are you doing here, sweet child?” he asked in a whisper,
glancing around him, and pinching the little girl’s cheek.
“We are playing....”
“Ah! With him?” Julian Mastakovich looked askew at the boy. “Go
into the next room, like a nice little boy,” he said to him.
The boy was silent and gazed at him with perturbed eyes. Julian
Mastakovich looked around once more and bent over the little girl.
“And what have you, sweet child, a doll?” he asked.
“Yes, a doll,” answered the little girl, frowning, and quailing visibly.
“A doll.... And do you know, sweet child, what the doll is made of?”
“I don’t know,” answered the little girl in a whisper, lowering her head.
“Of rags, my darling.... And you, my boy, you had better go into the
other room to your fellows,” said Julian Mastakovich, as he looked
severely at the youngster. The girl and the boy frowned and caught
hold of each other. They did not wish to part.
“And do you know why they gave you this doll?” asked Julian
Mastakovich, lowering his voice more and more.
“I don’t know.”
“Because you have been a lovely and well-behaved child the entire
week.”
At this juncture, Julian Mastakovich, agitated to the utmost, looked
round and, lowering his tone to a whisper, asked finally in an almost
inaudible voice, dying away more and more from agitation and
impatience:
“And will you love me, sweet girlie, when I shall come as a guest to
your papa and mamma?”
Having said this, Julian Mastakovich made one more effort to kiss
the lovely child; but the red-haired boy, quick to see that she was at
the point of tears, seized her hands and, out of deep sympathy for
her, began to whimper. Julian Mastakovich became quite angry.
“Begone, begone from here, begone!” he said to the boy. “Begone
into the other room! Begone to your own fellows!”
“No, don’t go! Don’t go! You had better go,” said the young girl, “but
leave him alone, leave him alone!” She was almost in tears.
Presently there was a commotion just within the door. Julian
Mastakovich immediately rose to his feet, somewhat frightened. The
red-haired boy was even more frightened. He left his companion and
stole out silently, with his hands brushing the wall, into the dining-
room. To hide his confusion, Julian Mastakovich followed him. He
was as red as a lobster, and when he looked in the glass he seemed
appalled as his own image. Perhaps he was annoyed at his rage and
impatience. Perhaps the calculation he made earlier on his fingers
had so affected him, tempting and inflaming him, that,
notwithstanding his position and dignity, he was impelled to act like a
young boy to attain his object, despite the fact that the object in any
case could be attained only five years hence. I followed the
esteemed gentleman into the dining-room and witnessed a strange
scene. Julian Mastakovich, his face all red from irritation and malice,
was pursuing the red-haired boy, who, retreating farther and farther
from him, did not know what to do with himself in his fright.
“Begone with you! What are you doing here? Begone, you good-for-
nothing! Begone! Stealing fruit, are you? Stealing fruit? Begone,
good-for-nothing! Begone, unclean one! Begone, begone to the likes
of yourself!”
The frightened boy, driven to desperate measures, tried to get under
the table. Then his pursuer, enraged to the last degree, drew out his
long batiste handkerchief and lashed it out at the cowering boy.
It is necessary to mention that Julian Mastakovich was a trifle fat. He
was a satiated, red-cheeked, stoutish person, large at the waist and
with fat legs; he was as round as a nut. He began to perspire, to
pant, and to grow fearfully red. His fury knew no bounds, so great
was his feeling of malice and—who knows?—perhaps jealousy. I
laughed out loud. Julian Mastakovich turned around, and in spite of
his importance was covered with most abject confusion. At this
instant the host entered by the opposite door. The boy climbed out
from under the table and wiped his knees and elbows. Julian
Mastakovich made haste to put his handkerchief, which he held by
one corner, to his nose.
The host, not without perplexity, surveyed the three of us; but, like a
man who understood life and looked at it with a serious eye, availed
himself of the opportunity to speak to his guest alone.
“This is the youngster,” said he, pointing at the red-haired boy,
“whom I had the pleasure of mentioning to you....”
“Ah?” answered Julian Mastakovich, not yet fully recovered from his
discomfiture.
“He is the son of the governess of my children,” continued the host in
an appealing voice. “She is a poor woman, a widow, the wife of an
honest official; and it is for this reason that ... Julian Mastakovich, is
it possible to....”
“Oh, no, no!” Julian Mastakovich made haste to exclaim. “No, Philip
Alekseievich; I am sorry, but it is utterly impossible. There is no
vacancy, and even if there were, there would be ten candidates for
the place, each having a greater right to it than he.... It is a great pity,
a great pity....”
“Yes, a pity,” repeated the host. “He is such a modest, quiet lad....”
“And quite a scamp, I should say,” added Julian Mastakovich, his
mouth hysterically athwart. “Begone, boy! Why are you standing
there? Go to your equals!”
At this point he could not restrain himself any longer, and looked at
me with one eye. I too could not resist, and laughed straight in his
face. Julian Mastakovich turned away immediately, and with
sufficient distinctness for me to hear asked the host the identity of
“that strange young man.” They exchanged whispers and left the
room. I observed afterward how Julian Mastakovich, listening to the
host, shook his head incredulously.
Having laughed to my heart’s content, I returned to the reception-
room. There the great man, surrounded by the fathers and the
mothers of families, the host and the hostess, was speaking with
great warmth to a lady to whom he had just been introduced. The
lady held by her hand the little girl with whom only ten minutes
before he had made the scene. Now he was lavish in his praises and
raptures over the beauty, talents, manners, and breeding of the
lovely child. He was plainly playing the wheedler before the mother.
She listened to him, almost with tears of joy in her eyes. The father’s
lips smiled. The prevailing spirit of good-will rejoiced the heart of the
host. Even all the guests lent a sympathetic hand, and made the
children stop their games in order not to interfere with the
conversation. The entire atmosphere was saturated with devotion. I
heard later how the mother of the interesting little girl, touched to the
very depths of her heart, begged Julian Mastakovich, in most
effusive language, to do her the great honor of conferring on the
house more often his precious presence; I heard with what
undisguised joy Julian Mastakovich accepted the invitation, and how
the guests, dispersing afterward in various directions as propriety
demanded, exchanged with one another complimentary salutations
regarding the host, the hostess, the little girl, and in particular Julian
Mastakovich.
“Is this gentleman married?” I asked almost aloud of an
acquaintance who stood nearest to Julian Mastakovich.
Julian Mastakovich threw at me a searching and malicious glance.
“No!” answered my acquaintance, mortified deeply at the
awkwardness which I committed purposely....

Not long ago I was passing the—— Church, and I was astonished at
the tremendous crowd that had gathered there. Every one talked
about a wedding. It was a bleak day in late autumn. I made my way
through the crowd and caught a glimpse of the bridegroom. He was
a round, satiated, pot-bellied little person, very much adorned. He
ran hither and thither, fussed, and gave orders. At last a murmur
went through the crowd, announcing the arrival of the bride. I
squeezed through the crowd and saw an astoundingly beautiful girl,
who had hardly experienced the first bloom of spring. But the
beautiful girl was pale and sad. She looked bewildered; and it
seemed to me that her eyes were red from newly-shed tears. The
classic rigidity of her features imparted to her beauty a kind of dignity
and strength. But through all this rigidity and dignity, through all this
sadness, there penetrated the first aspect of childhood’s innocence;
it suggested something naïve, fragile, and juvenile to the last degree;
and though the look bespoke resignation, it also seemed to utter a
silent prayer for mercy.
It was said in the crowd that she had just passed her sixteenth
birthday. An intent scrutiny of the bridegroom suddenly revealed him
to me as Julian Mastakovich, whom I had not seen for exactly five
years. I looked at her.... My God! I quickly made haste to leave the
church. In the crowd they were telling each other how rich the bride
was, that she had a dowry of five hundred thousand rubles ... and so
much besides in rags....
“At any rate, his calculation was a good one!” I reflected, as I jostled
my way into the street.
KOROLENKO THE EXILE

No intelligent outlander, I suppose, but marvels at the patience with


which the Russian people endure the exile system that has so long
brewed hell-broth for the nation to drink. When some violent offense
is answered by such punishment, we do not demur, but when
trivialities are magnified, and the police stupidly blunder, our blood
boils with protest.
So many times has Vladimir Korolenko been banished, that exile
must seem to him almost a normal condition, and freedom from
police surveillance a happy freak of fortune. And yet, more than any
other distinguished Russian writer, he is free from pessimism—his
writings are filled with passages of lyric sweetness.
Sixty years ago—in July, 1853—Vladimir was born at Jitomir, in the
government of Volynia. His father, of Cossack blood, was a district
judge in the cities of Dubno and Rovno, having previously served as
district attorney, and also as a minor judge. He was an honest man,
since he forbore to enrich himself with bribes, but made his modest
salary suffice. This course—eccentric in those days—left his wife in
straitened circumstances when he died. Vladimir was about fifteen at
the time, and still in the Gymnasium at Rovno, but his mother, the
daughter of a Polish landed proprietor, was enabled to keep him in
school and also maintain her other children, three boys and two girls.
The future author entered the Institute of Technology at St.
Petersburg in due course, and for two years fought off the extremes
of nakedness and hunger by coloring maps in the intervals of study,
for he had come to the great capital with only seventeen rubles in his
purse. The third year found him in Moscow, in the Petrovsk (St.
Peter’s) Agricultural Academy, and here, in the third year of his new
course (1875), he got his first taste of exile. His unforgivable crime
was to participate in a joint address of the students to the Faculty!
For this he was banished to the government of Vologda, but the
sentence was not completely carried out, for some one relented and
before he reached the place he was bidden to return to his home at
Kronstadt. Here for one year he was kept under police surveillance.
At the end of the year he was allowed to remove with his family to St.
Petersburg, where he worked in peace as a proof-reader, until
February, 1879. But he was soon to learn that Government never
forgets, for twice during that month was his home officially searched,
and at length he, together with his brother, his brother-in-law, and his
cousin, was banished to Glazof, in the government of Vyatka, and
presently still further north, to Vyshne Volotsk, where he was
confined in a political prison—and all without a trial, the reading of
charges, or any semblance of human justice.
The whole term of his exile was spent without a single gleam of light
to make clear his offense. But after his release in 1880, he learned
that his exile was due to his having attempted to break prison—an
offense which was alleged against him before he had ever been in
prison!
The circumstances of his release were fortunate. Prince Imeritinsky
had been deputized to investigate the condition of the political
prisons and to report on the causes of incarceration. Among other
prisons, he visited that at Vyshne Volotsk, and Korolenko was
already on the way to Yakutsk, Siberia, when the message came
ordering his release—probably as a result of the investigation.
Even then entire freedom was not granted him, for he was “allowed”
to settle at Perm; and here he began his active work as a writer,
though he had written successfully as early as 1879.
In 1881 Alexander III became Emperor of Russia, and all his
subjects were required to take oath of allegiance. But Korolenko
refused, because in addition the government officers demanded that
he betray his friends by giving details of any revolutionary
enterprises in which he knew them to be engaged. Rather than
become a party to such villainy, the young man chose further exile,
and for the succeeding three years lived miserably in Yakutsk, in
East Siberia. At length he returned to the ancient Tartar city of Nijni
Novgorod, on the Volga, where he now lives with his family.
All this period of maddening oppression was aggravated by the fact
that his mother needed his help. When in 1879 Korolenko began to
contribute literary sketches to such Russian periodicals as Russian
Thought, The Northern Messenger, and Annals of the Fatherland,
the meagre honorariums were indeed a blessing to his loved ones.

The thing that “goes without saying” often needs to be said just the
same. That a writer is likely to reproduce his life-experiences in his
writings is one of these truisms, yet it will always remain an
interesting occupation to trace connection between life and literary
product in the work of an author of individuality.
Korolenko came from “Little Russia,” and began to find his subjects
in the towns and villages of the west country in which he was born,
but naturally he turned at length to depicting the life of the extreme
Siberian east.
That Korolenko has been formed in opinion and moulded to iron
fortitude of heart by his severe experiences in exile is shown by his
remarkable story, “The Wondrous Maid,” in which the Nihilist is
depicted as a simple gendarme, whose manhood transfigures his
Nihilism and his work as an officer. Again, our author proved his
independence in a letter to the St. Petersburg Academy, in which, as
did Chekhov before him, he courteously declined membership
because the Academy had struck the name of Gorki from its list of
members.
It was in 1885, while in exile in Yakutsk, that he wrote his famous
“Makár’s Dream.” It is an odd fantasy, this story of the Yakut who,
having gotten half frozen in the wood, dreams that he is dragged
before the tribunal of the great Lord Toyon—a nondescript judge who
is neither of heaven nor of earth. Before a great scale, whose one
end is a small golden platter and whose other a huge wooden bowl,
the peasant is summoned to explain the acts of his life. At length,
when his cheatings and stealings are found to have outweighed all of
the deeds of service and faithfulness in his life, he suddenly breaks
into an unwonted eloquence of protest. He is unwilling to bear the
penalty of being turned into a beast of burden by becoming the horse
of a church official, not because the horse is badly treated, for it is
well fed—better fed, indeed, than he, the peasant, has ever been—
but he protests because the penalty is unjust. This appeal to justice
seems to move the great Toyon, and he ends by saying to the
dejected Makár, “Have patience, poor soul, thou art no longer on
earth: here will be found justice, even for thee!” And as he speaks
the scales begin to tremble, and the wooden bowl, filled with his evil
deeds, rises higher and higher, as though weighed down by his good
acts.
Surely, the great meed of injustice suffered by The Exile himself
gave inspiration for the message of mercy at the end of this fantastic
tale.
What may be called Korolenko’s Siberian era is further illustrated in
his sketches of a Siberian tourist, nine of which cover about one
hundred pages of ordinary size. All the sketches are remarkable for
local color and fine understanding of character. The one unfortunate
tendency is toward unfinished situation, for the sense of coming to
an adequate close is inseparable from good story-telling. It is but fair
to observe, however, that this trait of incompleteness is characteristic
of the sketch as a fictional form.
Throughout this series I have frequently asserted the obvious fact
that Russian themes have largely reflected the Russian
temperament, as is shown by the realistically direct and often terrible
pictures which fill the pages of their literature. Altogether apart from
our interest in the literate expressions of a great and alien people,
we must feel a sort of gruesome fascination as we are thrilled to the
point of horror in reading these simple yet titanic records of gloom.
All this raises the question of what is the difference between
fascination and charm—for charm, from the Anglo-American
viewpoint, is almost an unknown element in Russian literature.
Fascination they all possess; but charm is fascination plus delight. In
Korolenko we do have a writer of charm; and, besides, a charm that
is not the reflex of literatures other than his own—it evidently springs
from the sweetness of a spirit which all of the bitterness of
banishment could not defile. Here is a high and final test of native
fineness.
As compared with the stories of Garshin, with their “terrible,
incoherent cries of woe,” Korolenko’s tales are idyllic. A rhythmical,
lyrical measure beats enchantingly in his nature passages, whose
intimacy with the life of the woods inevitably recalls the French
Theuriet. “The Forest Whispers,” one of his longer short-stories, is
simply redolent of tree-fragrance. We feel the wandering airs of the
glades; we hear the never-ceasing swish of majestic boughs; we
stand rapt in the cathedral silences of the green-shadowy aisles. The
peasant tale is the thread on which these pearls are strung, but the
pearls hide the string.
Listen to this passage. What Loti has evoked from the inscrutable
sea, Korolenko has charmed from the forest with his enchanter’s
wand.
In the forest there was always a murmur, regular, continuous, like the
faint echo of a distant peal of bells; soft and indistinct, like a song
without words, or like the confused recollection of bygone days. The
murmur never ceased by day or night, for it was an old dense forest of
pines that had never been touched by woodman’s saw or axe. Lofty
pines, a hundred years old, with their red, sturdy trunks, stood in close
array, waving, in response to each breath of wind, their high-tufted
tops. Below, all was quiet; the air was filled with an odor of tar; through
the thick layer of pine-cones, with which the ground was strewn,
pushed gay ferns, in all the luxury of their rich fringes, and standing
motionless, their leaves unstirred by the breeze. In damp nooks green
grasses rose up on their high stalks; and the white clover bent its
heavy head, overcome, as it were, with dreamy lassitude. And above
flowed the murmur of the forest, the mingling sighs of the old pine-
wood.
Besides “The Forest Whispers,” two stories belong especially to
Korolenko’s Little-Russian group—“Iom-Kipour” (the Jewish Day of
Expiation) and “The Blind Musician.” The former relates how a Little-
Russian miller, good Christian though he is, narrowly escapes being
carried away by the Devil, in the place of the Jewish tavern-keeper
Iankiel, because, like him, he has tried to make money out of the
poor peasants—the same tendency to penetrate to the inner life
which we discover in other of Korolenko’s work, for he rose above
the realistic school, with its pathological records.
“The Blind Musician” is a remarkable psychological story—about
forty thousand words in length—in which all the sensations of the
blind are portrayed with sympathy and intelligence. The author has
not attempted to build up a meretricious interest by surrounding his
blind characters with the usual accompaniments to be found in fiction
—poverty and physical distress. Disallowing all such devices, he
wonderfully pictures the life of a child born blind in the home of a
wealthy family, his advance to boyhood, his love-life, and finally his
manhood’s experiences as a brilliant musician, “who attempts to
reproduce the sensations of sight by means of sounds.”
The following passage is typical:
The boy imaged to himself depth in the form of the soft murmur of the
stream as it flowed at the foot of the precipice, or of the frightened
splash of pebbles thrown from its top. Distance sounded in his ears
like the confused notes of a dying song. At times, in the sultry
noonday, when over the whole of nature there reigns a quiet so
profound that we can only divine the uninterrupted noiseless course of
life, the face of the blind boy would light up with a strange expression.
It seemed as if, under the influence of the silence that prevailed
around, there rose from the depth of his soul sounds audible only to
himself, to which he was listening with rapt attention. It was easy to
believe that at such moments a vague but productive train of thought
was awakening in his soul, like to the imperfectly caught melody of an
unknown song.

Two prose poems, of harmonious diction and fine human feeling, I


have space only to mention—“Easter Night,” and “The Old Bell-
Ringer,” which Korolenko calls “A Spring Idyl.” The latter is
reproduced herewith in a new translation for this series, and from it
the tone of the former may well be inferred.

Though not a great novelist—if he can be classed as a novelist at all


—Korolenko is the exponent of normality. He is more like Turgenev
than is any other living writer, though comparison with the Greatest
must not be taken to imply equality. The anarchistic, anti-Christian
Artsybashev, whose big-fisted novel, “Sanin,” forms an iconoclastic
type of its own, cannot approach Korolenko in lucid attractiveness.
Tolstoi, Korolenko followed, but at a distance, for he was of the
romantic school and little inclined to Tolstoi’s ultra-idealism,
particularly that of the last period.
One more refreshing characteristic of our author I venture to name—
human sympathy. True, he does not always temper his pity for the
“unfortunates” with the sound judgment of the moralist. Whether they
suffer deservedly or not, he does not deeply inquire—it is enough for
him that they suffer.
Well, I love him for that very trait of all-embracing sympathy. When a
man lets his heart go unleashed by the eternal judgment as to
whether the victim has sinned and may be suffering a righteous
punishment, he rises to utmost humanity—which is to say, the divine
spirit of the Great Master whose heart was Pity.
THE OLD BELL-RINGER
By Vladimir Korolenko
It had grown dark.
The tiny village, resting on the ledge of a remote stream, in a pine
forest, had become enveloped in that twilight which is peculiar to
starry spring nights, when the thin mist, rising from the earth,
deepens the shadows of the woods and fills the open spaces with a
silvery blue vapor.... How still was everything, and pensive and sad!
The village was quietly dreaming.
The dark outlines of the wretched huts were but vaguely visible; here
and there lights were aglimmer; now and then you could hear a gate
creak; a dog’s bark would start suddenly and die away; occasionally
out of the dark woods the figure of a pedestrian would emerge, or
that of a horseman; or a cart would pass by with a jolting noise.
These were the inhabitants of lone forest settlements, gathering to
their church to greet the great spring holiday.
The church stood on a little hill, in the very middle of the village. Its
windows were all alight. Its belfry—an old, tall, and dark structure—
pierced the blue sky.
The steps of the staircase creaked as the old bell-ringer ascended
the belfry, and soon his little lantern looked like a star suddenly
sprung into space.
It was hard for the old man to mount the steep staircase. His old legs
had already served their time, and his eyesight had grown dim.... It
was time an old man had rest, but God seemed slow in sending
deliverance. The old bell-ringer had buried sons and grandsons; he
had escorted both young and old to their final resting-place; but he
himself was still alive. It was hard!... So many times had he greeted
Easter that he had lost count—he could not even remember how
many times he had awaited here his last hour. And now once more
God had willed that he should be here.
Having reached the top, he leaned his elbow on the railing.
Below, around the church, he could discern the wretchedly kept
graves of the village burial-place; as if to protect, old crosses stood
over them with outstretched arms. Here and there a young birch-tree
inclined over them its branches, as yet leafless.... The aromatic odor
of young buds ascended from below towards Mikheyich, and with it
came a feeling of the sad tranquillity of eternal sleep.
And what would he be doing a year hence? Would he once more
climb this height, under this bronze bell, to arouse with a resounding
peal the lightly-slumbering night, or would he be resting ... down
there, in some dark corner of the graveyard, under a cross? God
knows!... He was ready, but in the meantime the Lord called him
once more to greet the holiday.
“All glory be to God!” whispered his lips, accustomed to the old
formula. Mikheyich raised his eyes towards the sky, dense with
millions of stars, and crossed himself.

“Mikheyich, Mikheyich!” a trembling voice, also that of an old man,


suddenly called him from below. The aged sexton looked up towards
the belfry, even fixed his palm over his blinking, tear-wet eyes, and
still could not see Mikheyich.
“What do you want? I am here,” answered the bell-ringer, leaning out
from the belfry. “Can’t you see me?”
“No, I can’t see. Isn’t it time to strike? What do you think?”
Both of them glanced at the stars. Thousands of God’s lights
twinkled on high. The fiery “Wagoner” was already far above the
horizon. Mikheyich pondered.
“No, not yet; wait just a little longer.... I know when to ...”

He knew. He had no need of a timepiece. God’s stars always told


him when the time came. The earth and the sky, the white cloud
floating silently across the expanse of blue, the indistinct murmur of
dark pines below, and the rippling of the stream concealed by the
dark—all were familiar to him, near to him.... Not in vain had he
spent his life here.
For the moment his entire long past unrolled before him.... He
recalled how he ascended the belfry with his father for the first
time.... Good Lord! how long ago it was!—and what a short time it
seemed!... He saw himself once more a fair-haired lad; his eyes
were kindled; the wind—not the sort that raises the dust of the street,
but rather a more rare wind, flapping, as it were, its noiseless wings
high above the earth—played with his hair.... There below, so far, so
far away, he saw some sort of little people; and the houses of the
village also seemed small, and the forest receded into the distance,
and the round-shaped meadow, upon which stood the village,
seemed immense, almost boundless.
“Well, here it is, all here!” smiled the old man, glancing at the small
spot of earth.
“So life, too, is like that,” he reflected. “When one is young, one sees
neither its end nor its edge.” ... And yet here it was, as if in the palm
of one’s hand, from the very beginning to the very grave he had just
been contemplating in the corner of the burial-ground.... What of
that? Glory be to the Lord!—It was time for rest. It was a hard road,
and he had traversed it an honest man; and the damp earth was his
mother.... Soon—if only soon!...
Well, the time had come. Mikheyich glanced once more at the stars,
removed his cap, crossed himself, and began to gather up the ropes
of the bells.... A few more moments, and the nocturnal air trembled
from the resounding stroke.... Another, a third, a fourth ... one after
the other, filling the lightly-slumbering pre-festal night with an
outpouring of powerful, lingering, resonant, singing tones.

The bell grew silent. The service in church had begun. It was the
habit of Mikheyich in former years to go down and to stop in a corner
near the door in order to pray and listen to the chanting. This time,
however, he remained in the tower. It was difficult for him; aside from
that, he felt intensely fatigued. He sat down on a little bench, and as
he listened to the dying tones of the agitated bronze he grew deeply
pensive. What were his thoughts? He himself could hardly have
answered the question.... The bell-tower was but dimly lighted by his
lantern. The still vibrating bells were lost in the darkness; faint
murmurs of the chant reached him occasionally from below, and the
nocturnal wind stirred the ropes fastened to the iron hearts of the
bells.
The old fellow let fall his gray head upon his breast. His mind was in
a state of delirious fancy. “Now they are singing a hymn,” he thought,
and he imagined himself among the others in church. He heard an
outpouring of children’s voices in a choir; he saw the figure of the
long-since-departed priest Nahum exhorting the congregation to
prayer; he saw hundreds of peasants’ heads, like ripe corn before
the wind, bend low and stand erect again.... The peasants were
crossing themselves.... Familiar faces, all of them, and all faces of
the dead. Here was the stern face of his father; here, beside his
father, his older brother, crossing himself and sighing. And he himself
stood here, in the bloom of health and strength and full of the
unconscious yearning for happiness and the joy of life.... Where, oh,
where, was this happiness?... The old man’s mind flared up for a
moment, like a dying flame, flashing with a bright, quick movement
and illuminating for the moment all the passages of his past life....
Hard work, sorrow, care.... Oh, where was this happiness? A hard
fate can bring furrows to a young face, give a stoop to a strong back,
and cause one to sigh like an older man.
There, on the left, among the women of the village, humbly inclining
her head, stood his sweetheart. A good woman, hers be the
Kingdom of God! How much had she not suffered, that fine soul!...
Constant need and labor and the inevitable womanly sorrow will
cause a handsome woman to wither; her eyes will lose their sparkle;
and the expression of perpetual, dull-like fright before each
unawaited blow of life will change the most superbly beautiful
creature.... Yes, and where was her happiness?... One son remained
to them, their one hope and joy, and he fell a victim to human
weakness.
And he too was here, his rich enemy, bending low time and again,
seeking to pray away the bitter tears of orphans he had wronged;
repeatedly he was performing upon himself the sign of the cross,
falling on his knees and touching the ground with his forehead....
And Mikheyich’s heart boiled over within him, while the dark faces of
the ikons looked down severely from their walls upon human sorrow
and human iniquity.
All that was past, all that behind him.... Now the entire world seemed
to him like a dark bell-tower, where the wind blew in the dusk, stirring
the bell-ropes.... “Let the Lord judge you!” whispered the old man,
shaking his gray head, while tears silently ran down his cheeks.

“Mikheyich! Mikheyich!... You haven’t fallen asleep?” someone


shouted up to him from below.
“Eh?” returned the old man, and quickly jumped to his feet. “Lord!
Have I in truth fallen asleep? That never happened before!”
With an accustomed hand, Mikheyich quickly caught the ropes.
Below him moved the peasant throng, a veritable ant-hill; the holy
banners aglimmer with gold brocade fluttered in the wind.... The
procession made a circuit of the church, and presently Mikheyich
heard the joyous cry, “Christ has risen from the dead!”
Coming like a mighty wave, the cry whelmed the old man’s heart....
And it seemed to Mikheyich that brighter flared the lights of the
waxen candles, and that stronger grew the agitation of the people;
the holy banners seemed to become more alive; and the suddenly
awakened wind caught up the waves of sound and with broad
sweeps lifted them high, where they became one with the loud
triumphant music of the bell.

Never before had old Mikheyich rung so well!


It was as if the old man’s brimming-over heart had passed into the
inanimate bronze; and it seemed as if the reverberations at the same
time sang and throbbed, laughed and wept, and, uniting in a rare
harmony, rose higher and higher unto the starry sky. The stars
themselves seemed to him to take on a new sparkle, to burst into
flame, while the sounds trembled and flowed, and again came down
to earth with a loving embrace.
A powerful bass loudly proclaimed: “Christ has risen!”
While two tenor voices, constantly atremble from the repeated blows
of the iron hearts, mingled with the bass joyously and resonantly:
“Christ has risen!”
And, again, two most slender soprano voices, seemingly in haste not
to be left behind, stole in among the more powerful ones, little
children, as it were, and sang in emulation: “Christ has risen!”
The entire belfry seemed to tremble and to shake; and the wind
blowing in the face of the bell-ringer appeared to flap its mighty
wings and to repeat: “Christ has risen!”

You might also like