Optimal Spacecraft Trajectories John E Prussing Full Chapter PDF

You might also like

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

Optimal Spacecraft Trajectories John E.

Prussing
Visit to download the full and correct content document:
https://ebookmass.com/product/optimal-spacecraft-trajectories-john-e-prussing/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

John Owen : trajectories in Reformed Orthodox theology


Mcgraw

https://ebookmass.com/product/john-owen-trajectories-in-reformed-
orthodox-theology-mcgraw/

Subway John E. Morris [E. Morris

https://ebookmass.com/product/subway-john-e-morris-e-morris/

Spacecraft Systems Engineering 4th Edition – Ebook PDF


Version

https://ebookmass.com/product/spacecraft-systems-engineering-4th-
edition-ebook-pdf-version/

Fault-Tolerant Attitude Control of Spacecraft Qinglei


Hu

https://ebookmass.com/product/fault-tolerant-attitude-control-of-
spacecraft-qinglei-hu/
IoT and Spacecraft Informatics 1st Edition K.L. Yung

https://ebookmass.com/product/iot-and-spacecraft-informatics-1st-
edition-k-l-yung/

Guyton E Hall Tratado De Fisiologia Médica John E. Hall

https://ebookmass.com/product/guyton-e-hall-tratado-de-
fisiologia-medica-john-e-hall/

Sarcopenia 2nd Edition John E. Morley

https://ebookmass.com/product/sarcopenia-2nd-edition-john-e-
morley/

Chemistry 7th Edition John E. Mcmurry

https://ebookmass.com/product/chemistry-7th-edition-john-e-
mcmurry/

They Knew Lincoln John E Washington

https://ebookmass.com/product/they-knew-lincoln-john-e-
washington/
O P T I M A L S PA C E C R A F T T R A J E C T O R I E S
Optimal Spacecraft
Trajectories

john e. prussing
Professor Emeritus
Department of Aerospace Engineering
University of Illinois at Urbana-Champaign
Urbana, Illinois, USA

3
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© John E. Prussing 2018
The moral rights of the author have been asserted
First Edition published in 2018
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2017959440
ISBN 978–0–19–881108–4 (hbk.)
ISBN 978–0–19–881111–4 (pbk.)
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Preface

T his text takes its title from an elective course at the University of Illinois at Urbana-
Champaign that has been taught to graduate students for the past 29 years. The
book assumes the reader has a background of undergraduate engineering (aerospace,
mechanical, or related), physics, or applied mathematics. Parts of the book rely heav-
ily on Optimal Control Theory, but an entire chapter on that topic is provided for those
unfamiliar with the subject.
Books on the field of optimal spacecraft trajectories are very few in number and date
back to the 1960s, and it has been nearly 40 years since a comprehensive new book has
appeared. Hence this volume.
Some classical results are presented using modern formulations and both impulsive-
thrust and continuous-thrust trajectories are treated. Also included are topics not
normally covered, such as cooperative rendezvous and second-order conditions. An
unusually large number of appendices (7) is provided to supplement the main text.
The book is suitable for a one-semester graduate-level university course. It is also a
scholarly reference book. Problems are included at the ends of the chapters and some of
the appendices. They either illustrate the subject matter covered or extend it. Solving
them will enhance the reader’s understanding of the material, especially if they are
assigned by your instructor as homework!
I am indebted to all the students who have taken my graduate course and to the many
colleagues I have interacted with over the years.
I have greatly benefitted from interacting with some of the giants in optimal trajec-
tories: Derek F. Lawden, Theodore N. Edelbaum, John V. Breakwell, and Richard H.
Battin. These fine gentlemen have passed on, but I hope this book helps convey their
contributions, both directly and indirectly.
J.E.P.
Urbana. Illinois, USA
February, 2017
Introduction

Performance index for an optimal trajectory

T he performance index for an optimal spacecraft trajectory is typically minimum


propellant consumption, not because of the monetary cost of the propellant, but
because for each kilogram of propellant saved an additional kilogram of payload is
delivered to the final destination. The term “minimum fuel” is often used in place of
“minimum propellant”, even for chemical rockets in which the propellant is composed
of both fuel and oxidizer. As is often the case, the term “fuel” is used rather than “pro-
pellant” because it’s shorter – having one syllable instead of three. In some applications
other performance indices are used, such as minimum time, maximum range, etc.
In optimizing spacecraft trajectories the two major branches of optimization theory
are used: parameter optimization and optimal control theory. In parameter optimization
the parameters are constants and we minimize a function of a finite number of param-
eters. An example is to represent the propellant consumed as the sum of a finite number
of velocity changes caused by rocket thrusts. By contrast, optimal control is a problem of
infinite dimensions where the variables are functions of time. An example of this is the
fuel consumed by continuously varying the thrust magnitude for a finite duration rocket
thrust.

Types of orbital maneuvers


The three types of orbital maneuvers are orbital interception, orbital rendezvous, and
orbit transfer. In orbital interception at the final time tf the position vector of the space-
craft r(tf ) and the position vector of a target body r# (tf ) are equal: r(tf ) = r# (tf ). A word
of explanation of the term “target” would be helpful here. The term is more general (and
more benign) than describing a military target to be eliminated. For a planetary flyby the
target body is the flyby planet, such as Jupiter during the Cassini Mission to Saturn. At
the flyby time the spacecraft position vector and the flyby planet position vector are equal
and the spacecraft has intercepted the flyby planet.

Optimal Spacecraft Trajectories. John E. Prussing.


© John E. Prussing 2018. Published 2018 by Oxford University Press.
2 | Introduction

By contrast, in an orbital rendezvous both the position vector and the velocity vector
of the spacecraft at the final time are equal to those of the target body: r(tf ) = r# (tf ) and
v(tf ) = v# (tf ). Putting a lander on Mars is an example of a rendezvous of a spacecraft with
a planet.
Lastly, in an orbit transfer there is no target body. The desired final condition is a
specific orbit, characterized by values of semimajor axis, eccentricity, inclination, etc.
Any one of these maneuvers can be either time-fixed (constrained, closed, finite or
fixed horizon) or time-open (unconstrained, open, infinite horizon). Note that in a well-
posed time-open optimization problem the value of tf will be optimized as part of the
solution.

Notation and preliminaries


A bold-faced symbol denotes a vector (column matrix) x having components (elements)
xi . An example is the n-vector
⎡ ⎤
x1
⎢x ⎥
⎢ 2⎥
⎢ ⎥
x=⎢ · ⎥ (1)
⎢ ⎥
⎣ · ⎦
xn

A symbol x without the bold face is the magnitude of the vector x:


1 1
x = | x | = (x · x) 2 = (xT x) 2 (2)

We will encounter a scalar function of a vector φ(x), e.g., a gravitational potential


function, and our convention is that the gradient ∂φ
∂x
is a row (1 × n) vector:

∂φ ∂φ ∂φ ∂φ
= ... (3)
∂x ∂x1 ∂x2 ∂xn

One advantage of this convention is that

∂φ
dφ = dx (4)
∂x

with no transpose symbol being needed. The gradient appearing in Eq. (4) corresponds
T
to the gradient vector g = ∂φ∂x
, which is the vector normal to the surface φ = constant.
We will also encounter a Jacobian matrix, which is the derivative of a vector F with
respect to another vector x. If F is a p-vector and x is an n-vector the Jacobian is an p × n
matrix with ij element equal to ∂F
∂xj
i
, where i is the row index and j is the column index:
Notation and preliminaries | 3
⎡ ∂F ∂F ∂F1 ⎤
1 1
·
⎢ ∂x1 ∂x2 ∂xn ⎥
⎢ ⎥
⎢ ∂F ∂F ∂F2 ⎥
⎢ 2 2
· ⎥
∂F ⎢⎢ ∂x ∂x ∂x


= 1 2 n
∂x ⎢ ⎥ (5)
⎢ · · · · ⎥
⎢ ⎥
⎢ ⎥
⎣ ∂F ∂F ∂Fp ⎦
p p
·
∂x1 ∂x2 ∂xn

Finally, we distinguish between a function and a functional. A function maps numbers


into numbers. An example is f (x) = x2 which for f (2) = 4. By contrast a functional maps
1
functions into numbers. An example is J = f (x)dx, which for f (x) = x2 yields J = 1/3.
0
1 Parameter Optimization

1.1 Unconstrained parameter optimization

P arameter optimization utilizes the theory of ordinary maxima and minima. In our
analysis, we will use the notation in Chapter 1 of Ref. [1.1]. The unconstrained prob-
lem is to determine the value of the m-vector u of independent parameters (decision
variables) to minimize the function

L(u) (1.1)

where the scalar L is termed a performance index or cost function. We note that a max-
imum value of L can be determined by minimizing –L, so we will treat only minimization
conditions.
If the ui are independent (no constraints) and the first and second partial derivatives of
L are continuous, then a stationary solution u∗ satisfies the necessary conditions (NC) that

∂L
= 0T (1.2a)
∂u
where 0 is the zero vector (every element equal to zero) and that the Hessian matrix

∂ 2L
≥0 (1.2b)
∂u2 u∗

which is a shorthand notation that the m × m matrix having elements ∂u∂ i ∂u


2
L
j
evaluated

at u must be positive semidefinite (all eigenvalues zero or positive). Equation (1.2a)
represents m equations that determine the values of the m variables u∗i .
Sufficient conditions (SC) for a local minimum are the stationary condition (1.2b) and
that the Hessian matrix of (1.2b) be positive definite (all eigenvalues positive). (The

Optimal Spacecraft Trajectories. John E. Prussing.


© John E. Prussing 2018. Published 2018 by Oxford University Press.
Unconstrained parameter optimization | 5

semidefinite NC is simply a statement that for the stationary point to be a minimum,


it is necessary that it is not a maximum!)
In the general case of m parameters we can write the first variation of our function as

∂L ∂L
δL = δu1 + · · · + δum = 0 (1.3)
∂u1 ∂um

and conclude, using the conditions of Eq. (1.2a), that at a stationary point δL = 0. In a
series expansion about the stationary point:

L(u) = L(u∗ ) + δL + δ 2 L + · · · (1.4)

the term δ 2 L is the second variation. A zero value for the first variation and a positive
value for the second variation are the SC for a local minimum.
As an example consider u to be a 2-vector. The NC (1.2a) for a minimum provides
two equations that determine the values of u∗1 and u∗2 .
The SC is
⎡ 2 ⎤
∂ L ∂ 2L
2  ⎢ ∂u ∂u ∂u ∂u ⎥ 
T ∂ L ⎢ 1 1 1 2 ⎥ δu1
δ L = δu
2
δu = [δu1 δu2 ] ⎢ ⎥ >0 (1.5)
∂u2 u∗ ⎣ ∂ 2L ∂ L
2 ⎦ δu2
∂u2 ∂u1 ∂u2 ∂u2 u∗

where δu1 ≡ u1 – u∗1 and δu2 ≡ u2 – u∗2 are arbitrary infinitesimal variations away from
their stationary values. If the strict inequality in Eq. (1.5) is satisfied for all nonzero δu1
and δu2 the matrix is positive definite. A simple test is that a matrix is positive definite if
all its leading principal minors are positive. For the 2 × 2 matrix in Eq. (1.5) the required
conditions are that the first principal minor is positive:

∂ 2L
>0 (1.6a)
∂u1 ∂u1

and that the second principal minor (the determinant of the matrix in this case) is
positive:
2
∂ 2L ∂ 2L ∂ 2L
– >0 (1.6b)
∂u1 ∂u1 ∂u2 ∂u2 ∂u1 ∂u2

As a very simple example consider

L(u) = 2u21 + u22 + u1 u2 (1.7)

The reader can verify that the NC (1.2a) for this example has the unique solution
u1 = u2 = 0 and that Eqs (1.6a) and (1.6b) are also satisfied because the values of the
leading principal minors are 4 and 7. So the function has a minimum at u = 0.
6 | Parameter Optimization

1.2 Parameter optimization with equality constraints


If there are one or more constraints the values of the parameters are not independent, as
we assumed previously. As an example, if a point (x1 , x2 ) is constrained to lie on a circle
of radius R centered at the origin we have the constraint

x21 + x22 – R2 = 0 (1.8)

The constrained minimum problem is to determine the values of m-decision parameters u


that minimize the scalar cost L(x, u) of n + m parameters, where the n-state parameters x
are determined by the decision parameters through a set of n equality constraints:

f(x, u) = 0 (1.9)

where f is an n-vector. Note that the number of state parameters n is by definition equal to
the number of constraints. In general, n > m so there are n – m independent variables.
(If m = n the problem is overconstrained and there is no optimization possible; there is
either only one solution or no solution.)

1.2.1 Lagrange multipliers

In the constrained problem the parameters can be treated as independent if we introduce


n additional variables λ1 , λ2 , . . . , λn , called Lagrange multipliers. We use these Lagrange
multipliers to adjoin the constraints to the cost function by forming an augmented
function H

H(x, u, λ) = L(x, u) + λT f(x, u) (1.10)

where λT ≡ [λ1 λ2 · · · λn ]. From Eq. (1.10) we see that if the constraints are satisfied,
f = 0, H is equal to L, and for first-order NC we can treat the problem as the unconstrained
optimization δH = 0. The necessary conditions are

∂H
= 0T (1.11a)
∂x
∂H
= 0T (1.11b)
∂u
and the constraint
∂H
= fT = 0T (1.11c)
∂λ
Equations (1.11a–c) are 2n + m equations in 2n + m unknowns (x, λ, and u).
Introducing the n additional variables λ seems to be counterproductive, because we
now have more unknowns to solve for. Instead of introducing the Lagrange multi-
pliers we could, in principle, solve the constraints f = 0 for the n state variables x in
Parameter optimization with equality constraints | 7

terms of the decision variables u, leaving only m equations to solve for the m decision
variables. However, those equations, while fewer, are usually more complicated than
Eqs (1.11a–c), so it is easier to solve the simpler 2n + m equations (see Problem 1.2).
The Lagrange multiplier λ has an interesting and useful interpretation, which makes it
worth solving for. It provides the sensitivity of the stationary cost to small changes in the
constants in the constraint equations. So, we can obtain the change in the stationary cost
due to small changes in these constants without resolving the problem. This can be very
useful in a complicated problem that is solved numerically.
This interpretation is discussed in Appendix A. To summarize, write the constraints as

f(x, u, c) = 0 (1.12)

where c is a q-dimensional vector of constants in the constraint equations. Due to changes


in these constants, dc, the change in the stationary cost is given by

dL∗ = λT fc dc (1.13)

∂f
where fc = ∂c
is an n × q matrix.
∂2L
The SC analogous to the Hessian matrix condition ∂u2 u∗
> 0 for the unconstrained
problem is given in Section 1.3 of Ref. [1.1] as

∂ 2L
x fu – fu fx Hxu + fu fx Hxx fx fu > 0
T –T T –T
= Huu – Hux f–1 –1
(1.14)
∂u2 f=0

where the symbol f–Tx is shorthand for the transpose of the inverse, which is equal to the
inverse of the transpose (see Problem 1.3).

Example 1.1 Minimization using a Lagrange multiplier

Determine the rectangle of maximum perimeter that can be inscribed in a circle of


radius R. In this example, there are two variables x and y and a single constraint—
the equation of the circle. So n = m = 1 and we arbitrarily assign y to be the decision
variable u. Let the center of the circle lie at (x, u) = (0, 0) and denote the corner of
the rectangle in the first quadrant as (x, u). To maximize the perimeter, we minimize
its negative and form

L(x, u) = –4(x + u) (1.15)

subject to the constraint

f (x, u) = x2 + u2 – R2 = 0 (1.16)

where x and u are scalars.


8 | Parameter Optimization

We construct the augmented function H as

H(x, u, λ) = L(x, u) + λf (x, u)


(1.17)
= –4(x + u) + λ(x2 + u2 – R2 )

Applying the NC of Eqs. (1.11a–c):


∂H
Hx = = –4 + 2λx = 0 (1.18a)
∂x
∂H
Hu = = –4 + 2λu = 0 (1.18b)
∂u
and the constraint
∂H
Hλ = = x2 + u2 – R2 = 0 (1.18c)
∂λ
where we use subscript notation for partial derivatives. Equations (1.18a–c) are easily
solved to yield

∗ ∗ 2
x =u = R (1.19)
2

2 2
λ= (1.20)
R
where Eq. (1.19) indicates that the rectangle of maximum perimeter is a square. To
verify that our stationary solution is indeed a minimum we use Eq. (1.14):

∂ 2L
= Huu – Hux fx–1 fu – fu fx–1 Hxu + fu fx–1 Hxx fx–1 fu
∂u2 f =0
√ √ √ (1.21)
4 2 4 2 8 2
= –0–0+ = >0
R R R
and this constrained second derivative√being positive satisfies the SC for a minimum,
resulting in a maximum perimeter of 4 2R = 5.657R.
Applying Eq. (1.13) with the single constant c being R:

2 2 √
dL∗ = λfR dR = (–2R)dR = – 4 2dR (1.22)
R
Consider a nominal value R = 1 and a change in the radius
√ of dR = 0.1. Using Eq.
(1.22) the change in the minimum cost is equal to –0.4 2 = –0.566, resulting in a
maximum perimeter
√ of 5.657 + 0.566 = 6.223. This compares very favorably with the
exact value of 4 2(1.1) = 6.223
Parameter optimization with an inequality constraint | 9

For this simple example using the interpretation of the Lagrange multiplier is not very
profound, but in a complicated problem, especially if the solution is obtained numerically
rather than analytically, a good approximation to the cost of a neighboring stationary
solution can be obtained without completely resolving the problem.
Appendix B provides a more complicated constrained minimum example: the
Hohmann transfer as a minimum v solution.

1.3 Parameter optimization with an inequality


constraint
Consider the simple case of a scalar cost L(y) subject to a constraint of the form f (y) ≤ 0,
where y is p dimensional and f is a scalar. We will consider only the simple case of
a single constraint, because that is all we need for the single application that we treat
in Section 7.2. Consideration of the general case results in the Karush–Kuhn–Tucker
conditions and is discussed in Section 1.7 of Ref. [1.1] (as the Kuhn–Tucker conditions).
As in the equality constraint case in Section 1.1 a Lagrange multiplier λ is introduced
and a scalar H is defined as:

H(y, λ) = L(y) + λf (y) (1.23)

Note that there is a total of p + 1 unknowns, namely the elements of y and λ. As in the
equality constraint case an NC can be written as

∂H
= 0T (1.24)
∂y

but, unlike the equality constraint case, the inequality constraint is either active or inactive,
resulting in the NC:

λ≥0 if f (y) = 0 (active) (1.25)

λ=0 if f (y) < 0 (inactive) (1.26)

Note that in either case the product λf = 0, resulting in H being equal to L.


Expanding Eq. (1.24):

∂H ∂L ∂f
= +λ = 0T (1.27)
∂y ∂y ∂y
or
∂L ∂f
= –λ (1.28)
∂y ∂y

which provides the geometrical interpretation that when the constraint is active the gra-
∂f
dients ∂L
∂y
and ∂y are oppositely directed (see Fig. 1.1 for p = 2). From that figure we see
10 | Parameter Optimization

f<0
Constant L fy
y2 Ly
fy Active
Ly constraint
f=0
Ly
Linareacy fy f=0
y1
f<0
y2 Unconstrained
Min L Inactive
constraint
f<0

y1
¶L ¶f
=–λ = 0T because
¶y ¶y λ=0

Figure 1.1 An active constraint and an inactive constraint

that at any point on the constraint not at the constrained minimum (where the gradients
are oppositely directed) we can slide along the constraint and decrease the cost until we
reach the constrained minimum.
This is even more evident when we note that the SC for the constraint being active is
λ > 0, indicating that the only way to decrease the cost is to violate the constraint.
If the constraint is inactive the constrained minimum is the unconstrained minimum.
In that case ∂L
∂y
= 0T as seen in Eq. (1.28) with λ = 0.

Example 1.2 Simple inequality constraint


In order to show how the NC are applied, consider the very simple example to
minimize
1
L(y) = y2 (1.29)
2
subject to the constraint

f (y) = y – a ≤ 0 (1.30)

where y and a are scalars. Use Eq. (1.23) to form H:


1
H(y, λ) = y2 + λ(y – a) (1.31)
2
Equation (1.24) provides the NC:
∂H
=y+λ=0 (1.32)
∂y
Problems | 11

where from Eqs. (1.25) and (1.26)

λ≥0 if y = a (active) (1.33)


λ=0 if y < a (inactive) (1.34)

We next assume the constraint is active or inactive and investigate whether the NC
are satisfied. Consider two separate cases, a = 1 and a = –1.
Case I a = 1
Assume the constraint is active. Then y = 1 and, from Eq. (1.32), λ = –1, which
violates NC (1.33). Assuming the constraint is inactive yields λ = 0 and y = 0, which
satisfy the NC. So, the constrained minimum is at y = 0 and is the same as the uncon-
strained minimum with a cost L = 0. Because the constraint is inactive the SC that
∂ 2 L/∂y2 = 1 > 0 is satisfied.
Case II a = –1
Assume the constraint is inactive. Then λ = 0 and y = 0, which violates the con-
straint y < –1. Assuming the constraint is active yields y = –1 and λ = 1, which satisfy
the NC. So the constrained minimum occurs at y = –1 with a cost L = 1/2. Note also
that the SC λ > 0 is satisfied.
Section 7.2 provides a more complicated example of an inequality constraint
minimization, namely, the terminal maneuver for an optimal cooperative impulsive
rendezvous.

Problems

1.1 a) Determine the value of the constant parameter k that “best” approximates a
given function f (x) on the interval a ≤ x ≤ b. Use as a cost function to be
minimized
b
1
J(k) = [k – f (x)]2 dx
2
a

This is the integral-squared error, which is the continuous version of the familiar
least-squares approximation in the finite-dimensional case.
Verify that a global minimum is achieved for your solution.
b) For a = 0, b = 1, and f (x) = x3 evaluate the cost J for (i) the optimal value of k
and (ii) another value of k that you choose.
1.2 Solve the problem in Example 1.1 without using a Lagrange multiplier.
12 | Parameter Optimization

1.3 Show that for a nonsingular square matrix the inverse of its transpose is equal to the
transpose of its inverse.
1.4* Minimize L(y) = 12 (y21 + y22 ) subject to the constraints y1 + y2 = 1 and y1 ≥ 3/4.
a) Attempt to satisfy the NC by assuming the inequality constraint is inactive.
b) Attempt to satisfy the NC by assuming the inequality constraint is active.
c) Determine whether the solution to the NC satisfies the SC.

Reference
[1.1] Bryson, A.E., Jr., and Ho, Y-C, Applied Optimal Control. Hemisphere Publishing, 1975.
2 Rocket Trajectories

2.1 Equations of motion

T he equation of motion of a spacecraft which is thrusting in a gravitational field can


be expressed in terms of the orbital radius vector r as:
r̈ = g(r) + ;  = u (2.1)

The variable  is the thrust acceleration vector. The scalar  is the magnitude of the
thrust acceleration defined as the thrust (force) T divided by the mass of the vehicle m.
The variable u is a unit vector in the thrust direction, and g(r) is the gravitational accel-
eration vector. Equation (2.1) is somewhat deceptive and looks like a simple statement
of “F equals ma” with the thrust term appearing on the right-hand side as if it were an
external force like gravity. In actuality, F does not equal ma (but it does equal the time
derivative of the linear momentum mv), because the mass is changing due to the gen-
eration of thrust and because the thrust is an internal force in the system defined as the
combination of the vehicle and the exhausted particles. A careful derivation of Eq. (2.1)
requires deriving the so-called rocket equation by equating the net external force (such as
gravity) to the time rate of change of the linear momentum of the vehicle-exhaust particle
system (see Sections 6.1–6.4 of Ref. [2.1]).
An additional equation expresses the change in mass of the spacecraft due to the
generation of thrust:
ṁ = –b; b≥0 (2.2)

In Eq. (2.2) b is the (nonnegative) mass flow rate. The thrust magnitude T is given by
T = bc, where c is the effective exhaust velocity of the engine. The word “effective” applies
in the case of high-thrust chemical engines where the exhaust gases may not be fully
expanded at the nozzle exit. In this case an additional contribution to the thrust exists
and the effective exhaust velocity is
Ae
c = ca + (pe – p∞ ) (2.3)
b

Optimal Spacecraft Trajectories. John E. Prussing.


© John E. Prussing 2018. Published 2018 by Oxford University Press.
14 | Rocket Trajectories

In Eq. (2.3) the subscript e refers to conditions at the nozzle exit, ca is the actual (as
opposed to effective) exhaust velocity, and p∞ is the ambient pressure. If the gases are
exhausted into the vacuum of space, p∞ = 0.
An alternative to specifying the effective exhaust velocity is to describe the engine in
terms of its specific impulse, defined to be:

(bc)t c
Isp = = (2.4)
(bt)g g

where g is the gravitational attraction at the earth surface, equal to 9.80665 m/s2 . The spe-
cific impulse is obtained by dividing the mechanical impulse delivered to the vehicle by
the weight (on the surface of the earth) of propellant consumed. The mechanical impulse
provided by the thrust force over a time t is simply (bc)t, and, in the absence of other
forces acting on the vehicle is equal to the change in its linear momentum. The weight
(measured on the surface of the earth) of propellant consumed during that same time
interval is (bt)g, as shown in Eq. (2.4). Note that if instead one divided by the mass
of the propellant (which is the fundamental measure of the amount of substance), the
specific impulse would be identical to the exhaust velocity. However, the definition in
Eq. (2.4) is in standard use with the value typically expressed in units of seconds.

2.2 High-thrust and low-thrust engines


A distinction between high- and low-thrust engines can be made based on the value of
the nondimensional ratio max /g. For high-thrust devices this ratio is greater than unity
and thus these engines can be used to launch vehicles from the surface of the earth. This
ratio extends as high as perhaps 100. The corresponding range of specific impulse val-
ues is between 200 and approximately 850 sec., with the lower values corresponding to
chemical rockets, both solid and liquid, and the higher values corresponding to nuclear
thermal rockets.
For low thrust devices the ratio max /g is quite small, ranging from approximately
10–2 down to 10–5 . These values are typical of electric rocket engines such as magneto-
hydrodynamic (MHD), plasma arc and ion devices, and solar sails. The electric engines
typically require separate power generators such as a nuclear radioisotope generator or
solar cells. The ratio for solar sails is of the order of 10–5 .

2.3 Constant-specific-impulse (CSI)


and variable-specific-impulse (VSI) engines
Two basic types of engines exist: CSI and VSI, also called power-limited engines. The
CSI category includes both high- and low- thrust devices. The mass flow rate b in some
cases can be continuously varied but is limited by a maximum value bmax . For this reason,
this type of engine is also described as a thrust-limited engine, with 0 ≤  ≤  max .
Constant-specific-impulse (CSI) and variable-specific-impulse | 15

The VSI category includes those low thrust engines which need a separate power
source to run the engine, such as an ion engine. For these engines, the power is limited
by a maximum value Pmax , but the specific impulse can be varied over a range of values.
The fuel expenditure for the CSI and VSI categories is handled separately.
The equation of motion (1) can be expressed as:

cb cb
v̇ = u + g(r); ≡  (2.5)
m m

For the CSI case Eq. (2.5) is solved using the fact that c is constant as follows:

cb
dv = u dt + g(r) dt (2.6)
m

Using Eq. (2.2),

dm
dv = –c u + g(r) dt (2.7)
m

This can be integrated (assuming constant u) to yield:

tf
v = v(tf ) – v(to ) = –cu(ln mf – ln mo ) + g(r) dt (2.8)
to

  tf
mo
v = c u ln + g(r) dt (2.9)
mf
to

which correctly indicates that, in the absence of gravity, the velocity change would be in
the thrust direction u. The actual velocity change achieved also depends on the gravita-
tional acceleration g(r) which is acting during the thrust period. The term in Eq. (2.9)
involving the gravitational acceleration g(r) is called the gravity loss. Note there is no
gravity loss due to an impulsive thrust.
If one ignores the gravity loss term for the time being, a cost functional representing
propellant consumed can be formulated. As will be seen, minimizing this cost functional
is equivalent to maximizing the final mass of the vehicle. Utilizing the fact that the thrust
is equal to the product of the mass flow rate b and the exhaust velocity c, one can write:

–m
ṁ = –b = (2.10)
c
dm 
= – dt (2.11)
m c
16 | Rocket Trajectories

For the CSI case the exhaust velocity c is constant and Eq. (2.11) can be integrated to
yield

 tf
mf 1
ln =–  dt (2.12)
mo c
to

or

 tf
mo
c ln =  dt ≡ JCSI (2.13)
mf
to

JCSI is referred to as the characteristic velocity of the maneuver or the V (pronounced


“delta vee”) and it is clear from Eq. (2.13) that minimizing JCSI is equivalent to maximiz-
ing the final mass mf .
In the impulsive thrust approximation for the unbounded thrust case (max → ∞)
the vector thrust acceleration is represented by


n
(t) = vk δ(t – tk ) (2.14)
k=1

with to ≤ t1 < t2 < · · · < tn ≤ tf representing the times of the n thrust impulses. (See
Sections 6.1–3 of Ref. [2.1].) Using the definition of a unit impulse,

tk
+

δ(t – tk ) dt = 1 (2.15)
tk–

and tk± ≡ lim ε → 0(tk ± ε); ε > 0.


Using Eq. (2.14) in Eq. (2.13):

tf 
n
JCSI = dt = vk (2.16)
to k=1

and the total propellant cost is given by the sum of the magnitudes of the velocity changes.
The corresponding cost functional for the VSI case is obtained differently. The exhaust
power (stream or beam power) is half of the product of the thrust and the exhaust
velocity:

1 1 1
P = Tc = mc = bc2 (2.17)
2 2 2
Constant-specific-impulse (CSI) and variable-specific-impulse | 17

Using this along with



b –ṁ d 1
= = (2.18)
m2 m2 dt m

results in

d 1 2
= (2.19)
dt m 2P

which integrates to

tf
1 1 1 2
– = dt (2.20)
mf mo 2 P
to

Maximizing mf for a given value of mo regardless of whether  is optimal or not is


obtained by running the engine at maximum power P = Pmax . This is not as obvious as it
looks in Eq. (2.20), because the value of  might be different for different values of P. To
see that the engine should be run at maximum power note that for a specified trajectory
r(t) the required vector thrust acceleration is given by Eq. (2.1) as

(t) = r̈(t) – g[r(t)] (2.21)

Thus, for a given trajectory r(t) (optimal or not), the final mass in Eq. (2.20) is
maximized by running the engine at maximum power.
For this reason the VSI cost functional can be taken to be

tf
1
JVSI =  2 dt (2.22)
2
to

To summarize, the cost functionals representing minimum propellant expenditure are


given by

tf
JCSI =  dt (2.23)
to

and

tf
1
JVSI =  2 dt (2.24)
2
to
18 | Rocket Trajectories

As seen in Eqs. (2.23) and (2.24) the minimum-propellant cost can be written in terms
of the control (t) rather than introducing the mass as an additional state variable whose
final value is to be maximized.

Problems

2.1 For an impulsive thrust consider first-order changes in J = v and in the final
mass mf . Obtain an expression for the marginal (fractional) change δmf /mf in terms
of the marginal change δJ/J. Assume the values of mo and c are specified and there is
no gravity loss.
2.2 Consider a constant thrust T = To in a CSI propulsion system.
a) Determine an expression for ˙ in terms of (only) the exhaust velocity c and the
thrust acceleration magnitude .
b) Solve the ODE in part (a) for (t). Express your answer in terms of To divided
by a time-varying mass.
2.3 Show how Eq. (2.19) follows from Eq. (2.18).

Reference
[2.1] Prussing, J.E. and Conway, B.A., Orbital Mechanics, Oxford University Press, 2nd Edition,
2012.
3 Optimal Control Theory

3.1 Equation of motion and cost functional

C onsider a dynamic system that operates between a specified initial time to and a
final time tf , which may be specified or unspecified. The equation of motion of the
system can be written as

ẋ = f(x, u, t); x(to ) = xo (specified) (3.1)

where x(t) is an n × 1 state vector and u(t) is an m × 1 control vector. Given an initial
state x(to ), the state x(t) is determined by the control u(t) through the equation of
motion (3.1).
The first-order form of Eq. (3.1) is completely general because any higher-order
equation can be transformed into first-order form. As an example consider Eq. (2.1):

r̈ = g(r) +  (2.1)

To transform to first-order form let

ṙ = v (3.2a)
v̇ = g +  (3.2b)

Then
 
ṙ v
ẋ = = (3.3)
v̇ g+

and a second-order equation for a 3-vector r has been transformed into a first-order
equation for a 6-vector x.

Optimal Spacecraft Trajectories. John E. Prussing.


© John E. Prussing 2018. Published 2018 by Oxford University Press.
20 | Optimal Control Theory

To formulate an optimal control problem we define a cost functional

tf
J = φ[x(tf ), tf ] + L[x(t), u(t), t] dt (3.4)
to

to be minimized by choice of a control u(t). The first term in Eq. (3.4) is sometimes
called the “terminal cost” and the second term the “path cost”.
As a simple example let φ = tf and L = 0. Then the cost is J = tf and we have a
minimum time problem.
As noted elsewhere, if we wish to maximize a performance index J̃, we minimize J = –J̃.
There are three basic forms of the cost functional, named after pioneers in the Calculus
of Variations:

Mayer form: J = φ[x(tf ), tf ]


tf
Lagrange form: J = L(x, u, t) dt
to
tf
Bolza form: J = φ[x(tf ), tf ] + L(x, u, t) dt [Eq. (3.4)]
to

There is also a special case of the Mayer form: J = xn (tf ), i.e., the cost is the final value of
the last component of the state vector.
Although the Bolza form appears to be the most general, all the forms have equal gen-
erality. To see this we will transform the Bolza form into the special case of the Mayer
form.

Example 3.1 Cost transformation


Given that x is an n-vector, define ẋn+1 = L with xn+1 (to ) = 0. Then

tf
L dt = xn+1 (tf ) – xn+1 (to ) = xn+1 (tf ).
to

Then,

J = φ[x(tf ), tf ] + xn+1 (tf ) ≡ xn+2 (tf ) (3.5)


General problem | 21

3.2 General problem


Over the time interval to ≤ t ≤ tf , where the final time can be specified or unspecified
(fixed or free, closed or open) there is the dynamic constraint of the equation of motion:

ẋ = f(x, u, t); x(to ) = xo (specified) (3.1)

and a cost to be minimized

tf
J = φ[x(tf ), tf ] + L[x(t), u(t), t] dt (3.4)
to

The ODE constraint of Eq. (3.1) is what makes Optimal Control Theory a generalization
of the Calculus of Variations.
There may be terminal constraints ψ[x(tf ), tf ] = 0, for example, r(tf ) – r# (tf ) = 0,
where r is the spacecraft position vector and r# is a target body position vector. This con-
straint then represents an orbital interception. And there may be control constraints such
as a bounded control | u(t) | ≤ M.
First consider the simplest optimal control problem: fixed final time, no terminal or
control constraints. We will derive first-order necessary conditions (NC) that must be
satisfied to determine the control u * (t) that minimizes the cost J. This is treated in more
detail in Section 3.3 of Ref. [3.1] and Section 2.3 of Ref. [3.2].
Before we begin our derivation of the NC, a fundamental point needs to be made. We
will be assuming that an optimal control u * (t) exists and deriving conditions that must
be satisfied. If no optimal control exists the NC are meaningless.
Peron’s Paradox illustrates this point and goes as follows: Let N = the largest integer.
We will prove by contradiction that N = 1.
If N > 1, N 2 > N, which is a contradiction because N is the largest integer. Therefore,
N = 1, a nonsensical result because there is no largest integer.
Assuming an optimal control u * (t) exists for the general problem, let’s proceed. First,
we adjoin the equation of motion constraint to the cost to form

tf
J̄ = φ[x(tf ), tf ] + L(x, u, t) + λT [f(x, u, t) – ẋ] dt (3.6)
to

Where λ(t) is an n × 1 vector variously called the “adjoint” vector, the “costate” vector,
or a “time-varying Lagrange multiplier” vector. Note that if the constraint equation of
motion (3.1) is satisfied, we have added zero to the cost, in the same spirit as the use of a
Lagrange multiplier in Section 2.2.1.
We will form the variation in the cost δ J̄ and derive NC that make δ J̄ = 0 due to a
small variation in the control about the optimal control δu(t) = u(t) – u∗ (t). Note that
22 | Optimal Control Theory

the variation is equal to the difference at a common value of time t and for this reason it
is sometimes called a contemporaneous variation, for which δt ≡ 0.
For convenience define the Hamiltonian

H(x, u, λ, t) ≡ L(x, u, t) + λT (t)f(x, u, t) (3.7)

(The terminology comes from classical mechanics, in which H = pT q̇ – L, where L is the


Lagrangian, p are the generalized momenta, and q̇ are the generalized velocities.)
Using the definition of the Hamiltonian, J̄ in Eq. (3.6) is equal to J̄ = φ +
tf
[H – λT ẋ] dt.
to
tf
Before forming the variation δ J̄ we will deal with the last term in J̄, namely – λT ẋ dt.
to
First, we derive a vector integration by parts formula by using d(λT x) = λT dx + dλT x.
T
Then d(λT x) = λT x = λT dx + dλT x. Using dx = ẋdt and dλT = λ̇ dt we obtain
T
λT ẋdt = λT x – λ̇ xdt.
tf
Applying this to the definite integral term – λT ẋ dt.
to

tf tf
T
– λT ẋdt = –λT (tf )x(tf ) + λT (to )x(to ) + λ̇ (t)x(t) dt (3.8)
to to

Using this result we rewrite Eq. (3.6) as

tf
T
J̄ = φ + λT (to )x(to ) – λT (tf )x(tf ) + [H + λ̇ x]dt (3.9)
to

Forming the cost variation δ J̄ due to a control variation δu(t):

δ J̄ = δφ + δ[λT (to )x(to )] – δ[λT (tf )x(tf )]

tf
T
+δ [H + λ̇ x] dt
to

∂φ
= δx(tf ) + λT (to )δx(to ) – λT (tf )δx(tf )
∂x(tf )
tf 
∂H ∂H T
+ δx + δu + λ̇ δx dt (3.10)
∂x ∂u
to
General problem | 23

Combining coefficients:

∂φ
δ J̄ = – λT (tf ) δx(tf ) + λT (to )δx(to )
∂x(tf )
tf   (3.11)
∂H T ∂H
+ + λ̇ δx + δu dt
∂x ∂u
to

Because δx(t) and δu(t) are arbitrary, for δ J̄ = 0 the NC are obtained from Eq. (3.11) by
setting individual coefficients equal to zero:

∂H
T ∂L ∂f
λ̇ = – =– – λT (3.12)
∂x ∂x ∂x
∂φ
λT (tf ) = (3.13)
∂x(tf )
∂H
= 0T (3.14)
∂u(t)
either δxk (to ) = 0 or λk (to ) = 0, k = 1, 2, . . . , n (3.15)

where δxk (to ) will be equal to 0 if the component xk (to ) is specified, but if xk (to ) is
unspecified, λk (to ) = 0. This represents a generalization of the initial condition x(to ) =
xo to allow some of the initial states to be unspecified. An additional NC is that the
constraint Eq. (3.1) must be satisfied: ẋ = f with initial conditions given by Eq. (3.15).
Equation (3.13) has a simple geometrical interpretation as a gradient vector. The final
value of the adjoint vector is orthogonal to the line or surface φ = constant.
Note that there are 2n + m unknowns: x(t), λ(t), and u(t). Equation (3.1) with initial
conditions given by (3.15) provide n equations; Eqs. (3.12) and (3.13) provide n equa-
tions; and Eq. (3.14) provides m equations, for a total of 2n + m NC equations in that
same number of unknowns. So, in principle, the optimal control problem is solved! But
there are several details to be worked out and observations to be made.
First note that even if the system equation ẋ = f is nonlinear the adjoint equation
(3.12) for λ(t) is linear because L and f are not functions of λ. And if the cost is of the
Mayer form (no L), the adjoint equation is homogeneous. But also note that the bound-
ary condition (3.13) is at the final time in contrast to the conditions in Eq. (3.15) at the
initial time. So, we have a split Two-Point Boundary Value Problem (2PBVP), which
indicates a fundamental difficulty in solving the NC for an optimal control problem,
because not all boundary conditions are given at the same value of time.
Note that the adjoint equation (3.12) and the system equation (3.1) are differential
equations, but what is often called the optimality condition (3.14) is an algebraic equation
that applies at all values of to ≤ t ≤ tf . It requires that the optimal control result in a
24 | Optimal Control Theory

stationary value of H at each point along the trajectory. Later we will learn the stronger
result that the optimal control must minimize the value of H.
A simple argument shows why Eq. (3.14) is an NC. Consider a scalar control u and
assume ∂H/∂u = 0 on a small finite interval t. Because u is arbitrary, if ∂H/∂u > 0,
choose δu < 0 and if ∂H/∂u < 0, choose δu > 0. In either case ∂H/∂u δu < 0, result-
ing in δJ < 0 and a decrease in the cost. Therefore, it is necessary that ∂H/∂u = 0 for J
to be a minimum.
Hidden in Eq. (3.11) is an interpretation of the vector λ. If all the other NC are
satisfied

δ J̄ = δJ = λ(to )T δx(to ) (3.16)

which provides the interpretation

∂J
λ(to )T = (3.17)
∂x(to )

That is, each component λk (to ) is the sensitivity of the minimum cost to a small change
in the corresponding component of the initial state. This means that the change in cost
due to a small change δxk (to ) can be calculated using Eq. (3.16) without resolving the
problem.
Note that this is consistent with Eq. (3.15) that λk (to ) = 0 if xk (to ) is unspecified. The
best value of xk (to ) to minimize J is the value for which the cost is insensitive to small
changes.
Finally, note that Eq. (3.17) can be generalized because any time t along an optimal
trajectory can be considered to be the initial time for the remainder of the trajectory (as
in today is the first day of the rest of your life). So, we can replace to by the time t and state
that
∂J
λ(t)T = (3.18)
∂x(t)
This is an application of Bellman’s Principle of Optimality. Stated simply: Any part of an
optimal trajectory is itself an optimal trajectory.

Example 3.2 Constant Hamiltonian


Form the total time derivative of H(x, u, λ, t):

dH ∂H ∂H ∂H ∂H
= ẋ + u̇ + λ̇ + (3.19)
dt ∂x ∂u ∂λ ∂t
Using ẋ = f and ∂H/∂λ = fT

dH ∂H T ∂H ∂H
= + λ̇ f + u̇ + (3.20)
dt ∂x ∂u ∂t
General problem | 25

Applying the NC results in

dH ∂H
= (3.21)
dt ∂t

So if ∂H/∂t = 0, i.e., H is not an explicit function of t, then H is a constant. This


occurs if neither L nor f is an explicit function of t, i.e., L = L(x, u) and f = f(x, u).
That property for f occurs when the system is time invariant (autonomous). In clas-
sical mechanics the constant H is known as the Jacobi Integral and often represents
conservation of total mechanical energy.
What good is a constant Hamiltonian? It can be used to evaluate the accuracy of a
numerically integrated solution by monitoring the value of H as the solution is gener-
ated. Also, as with conservation of energy, if the value of H is known at one value of
time, the unknown values of some variables at another time can be calculated.

Example 3.3 Rigid body control


Consider the single axis rotation of a rigid body for 0 ≤ t ≤ T, where the final time T
is specified. In terms of the angular velocity ω and the moment of inertia I the angular
momentum is given by h = Iω. With an external moment Mext acting about the axis
the equation of motion is ḣ = I ω̇ = Mext . In state and control variable notation, let
x = ω and u = Mext /I. Then the equation of motion is that of a simple integrator ẋ = u
with x(0) = xo , a specified initial angular velocity.
To formulate an optimal control problem define a cost

T
1 1
J = kx2 (T) + u2 (t)dt (3.22)
2 2
0

(The half factors are included to avoid factors of 2 when we differentiate.) The
squared terms ensure that positive contributions to the cost occur regardless of the
algebraic sign of the variables. The cost includes a terminal cost term to penalize the
final state (angular velocity) and a path cost to penalize control (external moment).
So, the effect is to slow down (for k > 0) the rotation without using an excessive
amount of external moment, which presumably incurs some sort of cost such as pro-
pellant. The constant k is a weighting coefficient that determines the relative weight
of each term in the cost. If we allow k < 0 we can represent a spin-up maneuver in
which a large final angular velocity is desired. From the form of terminal cost term,
we can expect that in the limit as k → ∞ x(T) → 0 and we stop the rotation. So,
this example has the essential aspects of a simple optimal control problem for which
n = m = 1.
26 | Optimal Control Theory

To apply the NC form the Hamiltonian

1
H = u2 + λu (3.23)
2

Then the adjoint equation is

∂H
λ̇ = – =0 (3.24)
∂x

so λ is constant and

∂φ
λ(t) = λ(T) = = kx(T) (3.25)
∂x(T)

The optimality condition is

∂H
=u+λ=0 (3.26)
∂u

so that

u = –λ = constant = –kx(T) (3.27)

which indicates two important facts: (i) The adjoint variable λ provides important
information to determine the optimal control, and (ii) of all the possible u(t) for this
problem the optimal control is a constant. (Note that, in addition to a stationary value
of H, it is minimized because Huu = 1 > 0.)
To solve the NC substitute for the control,

ẋ = –kx(T) (3.28)

which yields

x∗ (t) = –kx(T)t + xo (3.29)

Evaluating at t = T and solving:

xo
x∗ (T) = (3.30)
1 + kT

and a larger k > 0 results in a smaller x(T) because more weight is given to the
terminal cost term.
Terminal constraints and unspecified final time | 27

The rest of the solution is given by

–kxo
u∗ (t) = (3.31)
1 + kT
and

∗ 1 + k(T – t)
x (t) = xo (3.32)
1 + kT

where x(t) is conveniently expressed in terms of the “time-to-go” T – t and the state
changes linearly in time from its initial value to its final value.
Some limiting cases are instructive: As k → ∞, x∗ (T) → 0 and u∗ (t) → –xo /T
indicating that as infinite weight is put on the terminal cost the control drives the
initial state to zero in the allotted time.
As k → 0, x∗ (T) → xo , and u∗ (t) → 0 indicating that when all the weight is put
on the control cost no control is used and the state remains unchanged.
The optimal cost is calculated to be

kx2o
J∗ = (3.33)
2(1 + kT)

and we can explicitly verify the interpretation of λ(0) in Eq. (3.17) as ∂J ∗ /∂xo =
kxo /(1 + kT) = –u(0) = λ(0).
Finally, since ∂H/∂t = 0, H = constant = H(T) and

1 k2 x2o
H(T) = – <0 (3.34)
2 (1 + kT)2

And we note anecdotally that ∂J ∗ /∂T = H(T) (we will see why this is true later).
This means that there is no optimal value of T; the larger the value of T the smaller
the value of J*.
With that observation we seem to have extracted all we can from this example!

3.3 Terminal constraints and unspecified final time


In addition to the general problem considered in Section 3.2 we now include an unspeci-
fied (open) final time tf and terminal constraints of the form ψ[x(tf ), tf ] = 0, where ψ is
a q × 1 vector and 0 ≤ q ≤ n. Our augmented cost is now

tf
J̄ = φ + ν ψ +
T
L + λT [f – ẋ]dt (3.35)
to
28 | Optimal Control Theory

which is Eq. (3.6) with an additional terminal cost term and an unspecified upper limit
on the integral. The vector ν is a constant Lagrange multiplier q-vector.
For convenience define

[x(tf ), tf ] ≡ φ + ν T ψ (3.36)

where the vector ν has a useful interpretation (see Problem 3.4).


Consider the perturbation dJ̄ due to δu(t), where the symbol “d” includes the effect of
a differential change dtf , for example, the noncontemporaneous variation dx(tf ) ≡ x(tf +
dtf ) – x∗ (tf ). The result is a modified version of Eq. (3.11):


∂ ∂
dJ̄ = dx(tf ) + + Lf dtf – λT (tf )δx(tf ) + λT (to )δx(to )
∂x(tf ) ∂tf
tf  
∂H T ∂H
+ + λ̇ δx + δu dt (3.37)
∂x ∂u
to

where Lf ≡ L[(x(tf )), u(tf ), tf ].


Next, we observe that dx(tf ) = x(tf + dtf ) – x∗ (tf ) is to first order equal to x(tf ) +
ẋ(tf )dtf – x∗ (tf ). Using δx(tf ) = x(tf ) – x∗ (tf ) we can write dx(tf ) = δx(tf ) + ẋ(tf )dtf ,
where the last term is the contribution due to the change dt f . However,

ẋ(tf )dtf = [ẋ ∗ (tf ) + δẋ(tf )]dtf


(3.38)
= ẋ ∗ (tf ) + dtf + second order term

So to first order we have the “d – δ” rule:

dx(tf ) = δx(tf ) + ẋ ∗ (tf )dtf (3.39)

This is shown in Figure 3.1.


Next use δx(tf ) = dx(tf ) – ẋ(tf )dtf to write Eq. (3.37) as

 
∂ ∂
dJ̄ = – λT (tf ) dx(tf ) + + Lf + λT (tf )ẋ(tf ) dtf + λT (to )δx(to )
∂x(tf ) ∂tf

tf  
∂H ∂H
+ + λ̇T δx + δu dt (3.40)
∂x ∂u
to
Pontryagin minimum principle | 29

.
x(tf)dtf
dx(tf)
δx(tf)

x* (t) Nominal

Perturbed
x(t)

t
tf tf + dtf

Figure 3.1 Relationship between dx(tf ), ∀x(tf ), and dtf

3.4 Pontryagin minimum principle


The statement that the optimal control must provide a stationary value of the
Hamiltonian, as expressed by the NC ∂H/∂u = 0T can be strengthened to the Minimum
Principle, the NC that the optimal control must minimize the Hamiltonian, specifically,

H[x∗ , u∗ , λ, t] ≤ H[x∗ , u, λ, t] (3.41)

This has great generality and applies even in the case that the control is an element of a
closed set: u(t) ∈ U. The standard reference for the Minimum Principle is Ref. [3.3] and
Ref. [3.4] provides a short introduction to the subject.
The proof of this principle is notoriously difficult, but applying it is easy. A very sim-
ple example is for the bounded control: –1 ≤ u(t) ≤ 1 and the Hamiltonian H = 3u.
Application of the principle determines u∗ (t) = –1. In fact, for this example there is no
value of u that provides a stationary value of H, so that weaker condition does not provide
the optimal control.

Example 3.4 Minimum principle application

Consider a minimum time problem for a vehicle traveling at a constant scalar speed vo .
The control vector is a steering command in the form of a unit vector u in the
direction of the velocity. The equation of motion is for 0 ≤ t ≤ T:

ṙ = vo u

with r(0) = ro = 0 (specified). The terminal constraint is ψ = r(T) = 0. The initial


time is zero and the cost is T.
Considering φ = T and L = 0 the Hamiltonian is H = L + λT f = vo λT u.
30 | Optimal Control Theory

Because H is linear in u it has no stationary value, so we choose u to minimize H


by aligning u opposite to λ (to minimize the dot product) and enforce the fact that u
is a unit vector. The result is u = –λ/λ.
T
Another NC is λ̇ = –∂H/∂r = 0T indicating that λ(t) is constant and hence u(t)
is constant indicating a constant steering direction.
From Eq. (3.36)

= φ + ν T ψ = T + ν T r(T)

and the boundary condition is simply λT (T) = ∂ /∂r(T) = ν T . Using Eq. (3.47)
we calculate = 1 + vo λT u = 1 – vo λ = 0 which yields the value of λ = 1/vo .
From the equation of motion ṙ = vo u, and because u is constant, we have r(t) –
r(0) = vo tu. Evaluating at the final time and applying the terminal constraint yields
r(T) = ro + vo Tu = 0, from which we solve for u = –ro /Tvo . Requiring u to be a unit
vector yields the result that the minimum time is T = ro /vo and the optimal control is
u = –ro /ro . Note that ∂J/∂ro = λo = 1/vo .
Finally note that this minimum-time example is actually a minimum-distance
example made to sound more exciting as a minimum-time vehicle motion problem.
The result is the well-known straight-line solution.

Problems

3.1 Based on the discussion at the end of the Preface, is the cost in Mayer form a function
or a functional? Explain your answer.
 T
3.2 Show that Eqs. (3.1) and (3.12) can be written as the pair of equations ẋ = ∂H ∂λ
 T
and λ̇ = – ∂H∂x
, which are known as Hamilton’s Equations in classical mechanics.
3.3 In Eq. (3.10) why are there no δλ terms?
3.4 The terminal constraint (3.45) can be generalized to ψ[x(tf ), tf , c] = 0, where c is a
vector of constants in the constraints. Show that the vector ν provides the interpret-
ation of a change in the cost due to a small change in c, namely, ∂J/∂c = ν T ψ c . Also
q
show that in component form we have ∂J/∂ci = νj ∂ψj /∂ci .
j=1
∂f ∂f
3.5 Consider the state variational equation δẋ = ∂x δx + ∂u δu. Show for a cost in the
Mayer form that along a solution satisfying the NC the product λT (t)δx(t) is
constant.
3.6 In Example 3.3 for k < 0:
a) Determine the range of values of k for which the solution makes sense.
b) Determine the maximum value of x * (T) that can be achieved.
References | 31

3.7 a) In Example 3.4 consider r = [x y z] with specified initial position components


xo and yo , but an unspecified value zo . Determine the optimal solution including
the optimal value of zo using the NC.
b) Determine the values of H(T) and .
c)* Determine the optimal solution including the optimal value of α for the ter-
minal constraint r(T) = αa, where a is a specified vector for which ro × a = 0
and –∞ < α < ∞ (sketch this).

References
[3.1] Longuski, J.M., Guzman, J.J., and Prussing, J.E., Optimal Control with Aerospace Applications,
Springer, New York, 2014.
[3.2] Bryson, A.E. Jr and Ho, Y-C, Applied Optimal Control, Hemisphere Publishing, Washington,
D.C., 1975.
[3.3] Pontryagin, L.S., Boltyanski, V.G., Gamkrelidze, R.V, and Mishchenko, E.F., The
Mathematical Theory of Optimal Processes. Wiley, New York, 1962.
[3.4] Ross, I.M., A Primer on Pontryagin’s Principle in Optimal Control, Collegiate Publishers,
Carmel CA, 2009.
4 Optimal Trajectories

4.1 Optimal constant-specific-impulse trajectory

C onstant-specific-impulse and variable-specific-impulse rocket engines are discussed


in Section 2.3, with the cost functional for a CSI engine given by Eq. (2.23):

tf
JCSI =  dt (4.1)
to

The equation of motion in first-order form is given by Eq. (3.2):


   
ṙ v
ẋ = = (4.2)
v̇ g+

The first-order necessary conditions for an optimal CSI trajectory were first derived by
Lawden in Ref. [4.1] using classical Calculus of Variations. In the derivation that follows,
an Optimal Control Theory formulation is used, but the derivation is similar to that of
Lawden. One difference is that the mass is not defined as a state variable, but is kept track
of indirectly.
In order to minimize the cost in Eq. (4.1) one forms the Hamiltonian using Eqs. (3.7)
and (3.2) as

H =  + λTr v + λTv [g(r) + u] (4.3)

where we have written the thrust acceleration vector  as the product u, where  is the
scalar thrust acceleration magnitude and u is a unit vector in the thrust direction.
The adjoint equations (3.41) are then
T ∂H
λ̇r = – = –λTv G(r) (4.4)
∂r

Optimal Spacecraft Trajectories. John E. Prussing.


© John E. Prussing 2018. Published 2018 by Oxford University Press.
Optimal constant-specific-impulse trajectory | 33

T ∂H
λ̇v = – = –λTr (4.5)
∂v
where
∂g(r)
G(r) ≡ (4.6a)
∂r
is the symmetric 3 × 3 gravity gradient matrix. (See Problem 4.1.)

Example 4.1 Derivation of a gravity gradient matrix


For the inverse-square gravitational field: g(r) = – rμ2 rr = – rμ3 r show that the gravity
gradient matrix G(r) of Eq. (4.6) is equal to G(r) = – rμ5 (3rrT – r2 I3 ), where I3 is the
3 × 3 identity matrix.
For g = – μr
r3
we have:
  
∂g ∂r ∂r
G= = r3 –μ + μr 3r2 /r6 (4.6b)
∂r ∂r ∂r

Using ∂r/∂r = I3 and differentiating r2 = rT r yields 2r(∂r/∂r) = 2rT , so ∂r/∂r =


rT /r and we obtain the final result:

3μ(r2 rrT /r – μr3 I3 ) μ


G= = 5 (3rrT – r2 I3 ) (4.6c)
r6 r

For terminal constraints of the form


 
ψ r(tf ), v(tf ), tf = 0 (4.7)

which may describe an orbital intercept, rendezvous, etc., the boundary conditions on
Eqs. (4.4–4.5) are given in terms of

 ≡ ν T ψ[r(tf ), v(tf ), tf ] (4.8)

by Eq. (3.42):

∂ ∂ψ
λTr (tf ) = = νT (4.9)
∂r(tf ) ∂r(tf )
∂ ∂ψ
λTv (tf ) = = νT (4.10)
∂v(tf ) ∂v(tf )

There are two control variables, the thrust direction u and the thrust acceleration magni-
tude , that must be chosen to satisfy the Minimum Principle discussed in Section 3.4,
34 | Optimal Trajectories

i.e., to minimize the instantaneous value of the Hamiltonian H. By inspection, the


Hamiltonian of Eq. (4.3) is minimized over the choice of thrust direction by aligning
the unit vector u(t) opposite to the adjoint vector λv (t). Because of the significance of
the vector –λv (t) Lawden termed it the primer vector p(t):

p(t) ≡ –λv (t) (4.11)

The optimal thrust unit vector is then in the direction of the primer vector, specifically:

p(t)
u(t) = (4.12)
p(t)

and

λTv u = –λv = –p (4.13)

in the Hamiltonian of Eq. (4.3).


From Eqs. (4.5) and (4.11) it is evident that

λr (t) = ṗ(t) (4.14)

Equations (4.4, 4.5, 4.11, 4.14) combine to yield the primer vector equation

p̈ = G(r)p (4.15)

The boundary conditions on the solution to Eq. (4.15) are obtained from
Eqs. (4.9–4.10):

∂ψ
p(tf ) = –ν T (4.16)
∂v(tf )
∂ψ
ṗ(tf ) = ν T (4.17)
∂r(tf )

Note that in Eq. (4.16) the final value of the primer vector for an optimal intercept is the
zero vector, because the terminal constraint ψ does not depend on v(tf ).
Using Eqs. (4.11–4.15) the Hamiltonian of Eq. (4.3) can be rewritten as

H = –(p – 1) + ṗT v – pT g (4.18)

To minimize the Hamiltonian over the choice of the thrust acceleration magnitude ,
one notes that the Hamiltonian is a linear function of  and thus the minimizing
Optimal constant-specific-impulse trajectory | 35

Γmax

S(t)

MT MT MT
0
NT NT t

Figure 4.1 Three-burn CSI switching function and thrust profile.

value for 0 ≤  ≤ max will depend on the algebraic sign of the coefficient of  in Eq.
(4.18). It is convenient to define the switching function

S(t) ≡ p – 1 (4.19)

The choice of the thrust acceleration magnitude  that minimizes H is then given by the
“bang-bang” control law:

max for S > 0 (p > 1)
= (4.20)
0 for S < 0 (p < 1)

That is, the thrust magnitude switches between its limiting values of 0 (an NT null-thrust
arc) and Tmax (an MT maximum-thrust arc) each time S(t) passes through 0 (p(t) passes
through 1) according to Eq. (4.20). Figure 4.1 shows an example switching function for
a three-burn trajectory.
The possibility also exists that S(t) ≡ 0 (p(t) ≡ 1) on an interval of finite duration.
From Eq. (4.18) it is evident that in this case the thrust acceleration magnitude is not
determined by the Minimum Principle and may take on intermediate values between 0
and max . This IT “intermediate thrust arc” in Ref. [4.1] is referred to as a singular arc in
optimal control (Ref. [4.2]).
Lawden explained the origin of the term primer vector in a personal letter in 1990:
In regard to the term ‘primer vector’, you are quite correct in your supposition. I served in
the artillery during the war [World War II] and became familiar with the initiation of the
burning of cordite by means of a primer charge. Thus, p = 1 is the signal for the rocket motor
to be ignited.
For a CSI engine running at constant maximum thrust Tmax the mass flow rate bmax
will be constant. If the engine is on for a total of t time units,

max (t) = Tmax /(mo – bmax t) (4.21)

Other necessary conditions are that the variables p and ṗ must be continuous every-
where. Equation (4.19) then indicates that the switching function S(t) is also continuous
everywhere.
36 | Optimal Trajectories

Even though the gravitational field is time-invariant, the Hamiltonian in this formula-
tion does not provide a first integral (constant of the motion) on an MT arc, because 
is an explicit function of time as shown in Eq. (4.21). From Eq. (4.18)

H = –S + ṗT v – pT g (4.22)

Note that the Hamiltonian is continuous everywhere because S = 0 at the discontinuities


in the thrust acceleration magnitude.

4.2 Optimal impulsive trajectory


For a high-thrust CSI engine the thrust durations are very small compared with the
times between thrusts. Because of this one can approximate each MT arc as an impulse
(Dirac delta function) having unbounded magnitude (max → ∞) and zero duration.
The primer vector then determines both the optimal times and directions of the thrust
impulses with p ≤ 1 corresponding to S ≤ 0. The impulses can occur only at those
instants at which S = 0 (p = 1). These impulses are separated by NT arcs along which
S < 0 (p < 1). At the impulse times the primer vector is then a unit vector in the optimal
thrust direction.
The necessary conditions (NC) for an optimal impulsive trajectory first derived by
Lawden are in Ref. [4.1].
For a linear system these NC are also sufficient conditions (SC) for an optimal trajec-
tory, Appendix C and (Ref. [4.3]). Also in Appendix C, an upper bound on the number
of impulses required for an optimal solution is given.
Figure 4.2 shows a trajectory (at top) and a primer vector magnitude (at bottom) for
an optimal three-impulse solution. The orbital motion is counterclockwise and canonical
units are used. The canonical time unit is the orbital period of the circular orbit that has a
radius of one canonical length unit. The initial orbit is a unit radius circular orbit shown
as the topmost orbit going counterclockwise from the symbol ⊕ at (1, 0) to (–1, 0).
The transfer time is 0.5 original (initial) orbit periods (OOP). The target is in a coplanar
circular orbit of radius 2, with an initial lead angle (ila) of 270◦ and shown by the symbol
 at (0, –2). The spacecraft departs  and intercepts  at approximately (1.8, –0.8) as
shown. The + signs at the initial and final points indicate thrust impulses and the + sign
on the transfer orbit very near (0, 0) indicates the location of the midcourse impulse. The
magnitudes of the three vs are shown at the left, with the total v equal to 1.3681 in
units of circular orbit speed in the initial orbit.
The bottom graph in Fig. 4.2 displays the time history of the primer vector magnitude.
Note that it satisfies the necessary conditions of Table 4.1 for an optimal transfer.
The examples shown in this chapter are coplanar, but the theory and applications
apply to three-dimensional trajectories as well, e.g., Prussing and Chiu, Ref. [4.4].
Optimal variable-specific-impulse trajectory | 37

tf = 0.5 OOP, ∆V = 1.3681


0.5585
0.5 0.14239
0.66717
Canonical distance

−0.5

−1

−1.5
−2
−4 −3 −2 −1 0 1 2 3 4
ila = 270°
Canonical distance
1.2

1
Primer magnitude

0.8

0.6

0.4

0.2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time [OOP]

Figure 4.2 Optimal three-impulse trajectory and primer magnitude. Figure 4.2 was
generated using the MATLAB computer code written by S.L. Sandrik for Ref. [4.9].

Table 4.1 Impulsive necessary conditions

1. The primer vector and its first derivative are continuous everywhere.
2. The magnitude of the primer vector satisfies p(t) ≤ 1 with the impulses occurring at those
instants at which p = 1.
3. At the impulse times the primer vector is a unit vector in the optimal thrust direction.
4. As a consequence of the above conditions, dp/dt = ṗ = ṗT p = 0 at an intermediate impulse
(not at the initial or final time).

4.3 Optimal variable-specific-impulse trajectory


A variable specific impulse (VSI) engine is also known as a power-limited (PL) engine,
because the power source is separate from the engine itself, e.g., solar panels, radioiso-
tope thermoelectric generator (RTG), etc. The power delivered to the engine is bounded
between 0 and a maximum value Pmax , with the optimal value being constant and equal
to the maximum, as discussed in Section 2.3. The cost functional representing minimum
propellant consumption for the VSI case is given by Eq. (2.22) as
38 | Optimal Trajectories

tf
1
JVSI =  2 (t) dt (4.23)
2
to

Writing  2 as  T  the corresponding Hamiltonian function can be written as

1
H =  T  + λTr v + λTv [g(r) + ] (4.24)
2

For the VSI case there is no need to consider the thrust acceleration magnitude and
direction separately, so the vector  is used in place of the term u that appears
in Eq. (4.3).
Because H is a nonlinear function of  the Minimum Principle is applied by setting

∂H
=  T + λTv = 0T (4.25)
∂
or

(t) = –λv (t) = p(t) (4.26)

using the definition of the primer vector in Eq. (4.11). Thus, for a VSI engine the optimal
thrust acceleration vector is equal to the primer vector: (t) = p(t).
Because of this Eq. (4.2), written as r̈ = g(r) + , can be combined with Eq. (4.15), as
in Ref. [4.5] to yield a fourth-order differential equation in r:

riv – Ġṙ + G(g – 2r̈) = 0 (4.27)

Every solution to Eq. (4.27) is an optimal VSI trajectory through the gravity field g(r).
But desired boundary conditions, such as specified position and velocity vectors at the
initial and final times, must be satisfied.
From (Eq. 4.25) we get:
T
∂ 2H ∂ ∂H
= = I3 (4.28)
∂ 2 ∂ ∂

where I3 is the 3 × 3 identity matrix. Eq. (4.28) shows that the (Hessian) matrix of
second partial derivatives is positive definite, verifying that H is minimized.
Because the VSI thrust acceleration of Eq. (4.26) is continuous, a procedure described
in Ref. [4.6] to test whether second-order NC and SC are satisfied can be applied.
Equation (4.28) shows that a NC for minimum cost (Hessian matrix positive semi-
definite) and part of the SC (Hessian matrix positive definite) are satisfied. The other
condition that is both a NC and SC is the Jacobi no-conjugate-point condition. Reference
[4.6] details the test for that.
Solution to the primer vector equation | 39

4.4 Solution to the primer vector equation


The primer vector equation Eq. (4.15) can be written in first-order form as the linear
system:
    
d p O3 I3 p
= (4.29)
dt ṗ G O3 ṗ

where O3 is the 3 × 3 zero matrix.


Equation (4.29) is of the form ẏ = F(t) y and its solution can be written in terms of a
transition matrix (t, to ) as

y(t) = (t, to )y(to ) (4.30)

for a specified initial condition y(to ).


Glandorf in Ref. [4.7] and Battin in Ref. [4.8] present forms of the transition matrix
for an inverse-square gravitational field. [In Ref. [4.7] the missing Eq. (2.33) is (t, to ) =
P(t)P–1 (to ).]
Note that on an NT (no-thrust or coast) arc the variational (linearized) state equation
is, from Eq. (4.2)
    
δṙ O3 I3 δr
δẋ = = (4.31)
δv̇ G O3 δv

which is the same equation as Eq. (4.29). So, the transition matrix in Eq. (4.30) is also
the transition matrix for the state variation, i.e., the state transition matrix in Ref. [4.8].
Also shown in Ref. [4.8] is that this state transition matrix has the usual properties
from linear system theory (Appendix D) and is also symplectic, the definition of which is
that

T (t, to )J(t, to ) = J (4.32)

where
 
O3 I3
J= (4.33)
–I3 O3

Note that J2 = –I6 indicating that J is a matrix analog of the imaginary number i = –1.
As shown in Problem 4.7 the inverse of the  matrix can be simply obtained by matrix
multiplication. This is especially useful when the state transition matrix is determined
numerically because the inverse matrix –1 (t, to ) = (to , t) can be computed without
explicitly inverting a 6 × 6 matrix. As seen from Eq. (4.30) this inverse matrix yields the
variable y(to ) in terms of y(t): y(to ) = (to , t)y(t).
Another random document with
no related content on Scribd:
in the day. It was a day too merry for work, and very disturbing to the
heart of any man.
In the noontime of that bright day it happened that a pedlar of
summer stuffs came through the countryside, and he carried on his
shoulder a great heap of his stuffs, of every hue and shade, and
some were flowered, and as he went he called, “Cloth—good cloth
for sale!”
When he came to this house where the man and the woman and the
old mother and the little children sat in the shade of their willow tree
and ate their noon meal he halted and cried, “Shall I stay, goodwife,
and show you my stuffs?”
But the mother called back, “We have no money to buy, unless it be
a foot of some common cheap stuff for this new son of mine. We be
but poor farmer folk and not able to buy new clothes nor much of any
stuffs except such as must be had to keep us from bareness!”
And the old woman, who must always put in her bit, cried in her little
old shrill voice, “Aye, it is true what my daughter-in-law says, and the
stuffs be very poor these days and washed to shreds in a time or
two, and I mind when I was young I wore my grandmother’s coat and
it was good till I was married and needing something new but still
only for pride’s sake, for the coat was good enough still, but here I be
in my second shroud and nearly ready for a third, the stuffs be so
poor and weak these days—”
Then the pedlar came near, scenting sale, and he was a man with
very pleasant and courteous coaxing ways such as pedlars have,
and he humored the mother and had a good kindly word for the old
woman, too, and he said to her, “Old mother, here I have a bit of
cloth as good as any the ancients had and good enough even for
that new grandson you have—goodwife, it is a bit left from a large
piece that a rich lady bought in a great village I went through today,
and she bought it for her only son. Of her I did ask the honest price
seeing she cut from a whole piece, but since there is only this bit left,
I will all but give it to you, goodwife, in honor of the fine new son you
have there at your breast.”
So saying these words smoothly and as though all in one flowing
breath, the pedlar drew from out his pack a very pretty end of cloth,
and it was as he said, flowered with great red peonies upon a grass-
green ground.
The old woman cried out with pleasure because her dim eyes could
see its hues so clear and bright, and the mother loved it when she
saw it. She looked down then at the babe upon her breast, naked
except for a bit of old rag about its belly, and it was true he was a fat
and handsome child, the prettiest of her three, and like the father,
and he would look most beautiful in that bit of flowery stuff. So it
seemed to the mother and she felt her heart grow weak in her and
she said unwillingly, “How much is that bit then? But still I cannot buy
it for we have scarcely enough to feed these children and this old
soul and pay the landlord too. We cannot buy such stuffs as rich
women put upon their only sons.”
The old woman looked very doleful at this, and the little girl had
slipped from her place and went to peer at the bright cloth, putting
her dim eyes near to see it. Only the elder lad ate on, caring nothing,
and the man sat idly, singing a little, careless of this bit of stuff for no
one but a child.
Then the pedlar dropped his voice low and coaxing and he held the
cloth near the child, but not too near either, careful lest some soil
come on it if it were not bought, and he said half whispering, “Such
cloth—such strength—such color—I have had many a piece pass
through my hands, but never such a piece as this. If I had a son of
my own I would have saved it out for him, but I have only a poor
barren wife who gives me no child at all, and why should the cloth be
wasted on such as she?”
The old woman listened to this tale and when she heard him say his
wife was barren she was vastly diverted and she cried out, “A pity,
too, and you so good a man! And why do you not take a little wife,
good man, and try again and see what you can do? I ever say a man
must try three women before he knows the fault is his—”
But the mother did not hear. She sat musing and unsure, and her
heart grew weaker still, for she looked down at her child and he was
so beautiful with this fine new stuff against his soft golden skin and
his red cheeks that she yielded and said, “What is your least price,
then, for more I cannot pay?”
Then the pedlar named a sum, and it was not too high and not as
high as she had feared, and her heart leaped secretly. But she shook
her head and looked grave and named half the sum, as the custom
was in bargaining in those parts. This was so little that the pedlar
took the cloth back quickly and put it in its place and made to go
away again, and then the mother, remembering her fair child, called
out a sum a little more, and so haggling back and forth and after
many false starts away the pedlar made, he threw down his pack
again and pulled forth the bit, agreeing at last to somewhat less than
he had asked, and so the mother rose to fetch the money from the
cranny in the earthen wall where it was kept.
Now all this time the man sat idly by, singing, and his high voice
made soft and small and stopping sometimes to sup down his hot
water that he drank always after he had eaten, and he took no part in
this bargain. But the pedlar being a very clever fellow and eager to
turn to his account every passing moment, took care to spread out
seemingly in carelessness a piece of grass cloth that he had, and it
was that cloth made of wild flax which cools the flesh upon a hot day
in summer, and in color it was like the sky, as clear, as blue. Then
the pedlar glanced secretly at the man to see if the man saw it, and
he said half laughing, “Have you bought a robe for yourself yet this
summer? For if you have not, I have it for you here, and at a price I
swear is cheaper than it can be bought in any shop in town.”
But the man shook his head and a dark look came down upon his
idle, pretty face, and he said with bitterness, “I have nothing
wherewith to buy myself anything in this house. Work I have and
nothing else, and all I gain for it is more to feed, the more I work.”
Now the pedlar had passed through many a town and countryside
and it was his trade to know men’s faces and he saw at a glance that
this man was one who loved his pleasure, and that he was like a lad
held down to life he was not ready for, and so he said in seeming
kindliness and pity, “It is true that I can see you have a very hard life
and little gain, and from your fine looks I see it is too hard a life. But if
you buy yourself a new robe you will find it like a very potent new
medicine to put pleasure in your heart. There is nothing like a new
summer robe to put joy in a man, and with that ring upon your finger
shined and cleaned and your hair smoothed with a bit of oil and this
new robe upon you, I swear I could not see a prettier man even in a
town.”
Now the man heard this and it pleased him and he laughed aloud
half sheepishly, and then he remembered himself and said, “And
why should I not for once have a new robe for myself? There is
nothing ahead but one after another of these young ones, and am I
forever to wear my old rags?” And he stooped swiftly and fingered
the good stuff in his fingers and while he looked at it the old mother
was excited by the thought and she cried, “It is a very fair piece, my
son, and if you must have a robe then this is as pretty a blue robe as
ever I did see, and I remember once your father had such a robe—
was it when we were wed? But no, I was wed in winter, yes, in
winter, for I sneezed so at the wedding and the men laughed to see
a bride sneezing so—”
But the man asked suddenly and roughly, “How much will it be for a
robe?”
Now when the pedlar said the price, at that moment the mother
came forth with the money in her hand counted and exact to the last
penny and she cried out alarmed, “We can spend no more!”
At that cry of hers some desire hardened in the man and he said
wilfully, “But I will have myself a robe cut from this piece and I like it
very well so that I will have it for the once! There are those three
silver pieces I know we have.”
Now those three coins were of good value and coins the mother had
brought with her when she came to be wed, and her own mother had
handed them to her for her own when she left her home. They were
her precious possession and she had never found the hour when
she could spend them. Even when she had bought the coffin for the
old mother when they thought her dying, she had pinched and
borrowed and would not spend her own, and often the thought of
those three silver pieces was in her mind for safe riches, and they
were there if ever times grew too hard, some war or hardship that
might come at any hour and lose them the fruits of their land. With
those three coins in the wall she knew they could not starve for a
while. So now she cried, “That silver we cannot spend!”
But the man leaped up as swift as a swallow and darted past her in a
fury and he went to that cranny and searched in it and seized the
silver. Yet the woman was after him, too, and she caught him and
held him and hung to him as he ran. But she was not quick enough
and never quick enough for his litheness. He threw her aside so that
she fell upon the earthen floor, and the child still in her arms, and he
ran out shouting as he ran, “Cut me off twelve feet of it and the foot
and more to spare that is the custom!”
This the pedlar made haste to do, and he took the silver coins
quickly, although indeed they were somewhat less than he had
asked, but he was anxious to be away and yet have his stuff sold,
too. When the mother came out at last the pedlar was gone and the
man stood in the green shade of the tree, the blue stuff bright and
new in his two hands, and her silver gone. The old woman sat afraid
and when she saw the mother come she began in haste to speak of
this or that in a loud creaky voice, “A very pretty blue, my son, and
not dear, and a long summer since you had a grass cloth—”
But the man looked blackly at the woman, his face dark and red, and
he roared at her, still bold with his anger, “Will you make it, then, or
shall I take it to some woman and pay her to make it and tell her my
wife will not?”
But the mother said nothing. She sat down again upon her little stool
and she sat silent at first, pale and shaken with her fall, and the child
she held still screamed in fright. But she paid no heed to him. She
set him on the ground to scream, and twisted up afresh the knot of
her loosened hair. She panted for a while and swallowed once or
twice and at last she said, not looking at the man, “Give it to me
then. I will make it.”
She was ashamed to have another do it and know the quarrel more
than they did now, watching from their doors when they heard the
angry cries.
But from that day on the woman harbored this hour against the man.
Even while she cut the cloth and shaped it, and she did it well and
the best she knew to do, for it was good stuff and worth good care,
still she took no pleasure in the work and while she made the robe
she stayed hard and silent with the man, and she said no small and
easy thing about the day or what had happened in the street or any
little thing such as contented women say about a house. And
because she was hard with him in these small ways the man was
sullen and he did not sing and as soon as he had eaten he went
away to the wayside inn and he sat there among the men and drank
his tea and gambled far into the night, so that he must needs sleep
late the next day. When he did so in usual times she would scold him
and keep him miserable until he gave over for peace’s sake, but now
she let him sleep and she went alone to the fields, hard and silent
against him whatever he might do, though her heart was dreary, too,
while she kept it hard.
Even when the robe was done at last, and she was long in making it
because there was the rice to be set and planted, even when it was
done she said nothing of how it looked upon him. She gave it to him
and he put it on and he shined his ring with bits of broken stone and
he smoothed his hair with oil he poured from the kitchen bottle and
he went swaggering down the street.
Yet even when this one and that cried out to him how fine he was
and how fine his robe, he took no full sweet pleasure in himself as he
might have done. She had said no word to him. No, when he had
lingered at the door an instant she went on with her task, bending to
the short-handled broom and sweeping about the house and never
looking up to ask if the robe fitted him or if his body was suited to its
shape, as she was wont to do if she had made him even so much as
a pair of new shoes. At last he had even said, half shy, “It seems to
me you have sewed this robe better than any robe I ever had, and it
fits me as a townsman’s does.”
But still she would not look up. She set the broom in its corner and
went and fetched a roll of cotton wool and set herself to spinning it to
thread, since she had used her store in the making of the blue robe.
At last she answered bitterly, “At the cost it was to me it should look
like an emperor’s robe.”
But she would not look at him, no, not even when he flung himself
down the street. She would not even look at him secretly when his
back was turned because she was so bitter against him, although
her heart knew the blue robe suited him well.
V
THROUGH that day long the mother watched for the man to come
home. It was a day when the fields could be left to their own growing,
for the rice was planted in its pools, and in the shallow water and in
the warm sunlight the green young plants waved their newly forming
heads in the slight winds. There was no need to go out to the land
that day.
So the mother sat under the willow tree spinning and the old woman
came to sit beside her, glad of one to listen to what she said, and
while she talked she unfastened her coat and stretched her thin old
withered arms in the hot sun and felt the good heat in her bones, and
the children ran naked in the sunshine too. But the mother sat
silently on, twisting the spindle with a sure movement between her
thumb and the finger she wet on her tongue, and the thread came
out close spun and white, and when she had made a length of it she
wound it about a bit of bamboo polished smooth to make a spool.
She spun as she did all things, firmly and well, and the thread was
strong and hard.
Slowly the sun climbed to noon and she put her spinning down and
rose.
“He will be coming home soon and hungry for all his blue robe,” she
said dryly, and the old woman answered, cackling with her ready,
feeble laughter, “Oh, aye, what is on a man’s belly is not the same as
what is in it—”
The mother went then and dipped rice with a gourd from the basket
where they kept it stored, and she leveled the gourd with her other
hand so not a grain was spilled, and she poured the rice into a
basket made of finely split bamboo and went along the path to the
pond’s edge, and as she went she looked down the street. But she
saw no glimpse of new blue. She stepped carefully down the bank
and began to wash the rice, dipping the basket into the water and
scrubbing the grain with her brown strong hands, dipping it again
and again until the rice shone clean and white as wet pearls. On her
way back she stooped to pull a head of cabbage where it grew, and
threw a handful of grass to the water buffalo tethered under a tree,
and so she came again to the house. Now the elder boy came home
from the street leading his sister by the hand, and the mother asked
him quietly, “Saw you your father on the street or in the inn or at
anyone’s door?”
“He sat a while at the inn drinking tea this morning,” the boy replied,
wondering. “And I saw his robe, new and blue, and it was pretty and
our cousin when he knew how much it cost said it had cost my father
very dear.”
“Aye, it cost him dear, I swear!” said the mother, suddenly, her voice
hard.
And the girl piped up, echoing her brother, “Yes, his robe was blue—
even I could see that it was blue.”
But the mother said no more. The babe began to weep where he lay
sleeping in a winnowing basket and she went and picked him up and
opened her coat and held him to her breast, and she suckled him as
she went to cook the meal. But first she called to the old woman,
“Turn yourself where you sit, old mother, and watch and tell me if you
see the new blue of his robe, and I will put the meal on the table.”
“I will, then, daughter,” called the old dame cheerfully.
Yet when the rice was cooked and flaked, white and dry as the man
loved it, still he did not come. When the cabbage was tender and the
woman had even made a bit of sweet and sour sauce to pour upon
its heart, as he loved it, he did not come.
They waited a while and the old woman grew hungry and faint with
the smell of the food in her nostrils and she cried out, in a sudden
small anger, being so hungry, “Wait no more for that son of mine!
The water is leaking out of my mouth and my belly is as empty as a
drum and still he is not here!”
So the mother gave the old woman her bowl then and she fed the
children too and even let them eat of the cabbage, only she saved
the heart of it for him. She ate also after this, but sparingly for she
seemed less zestful in her hunger today, somehow, so there was still
much rice left and a good bowlful of the cabbage and this she put
carefully away where the wind would catch it and keep it fresh. It
would be as good at night as it was now if she heated it again. Then
she gave suck to the babe, and he drank his fill and slept, a round,
fat, sturdy child, sleeping in the strong sun and brown and red with
its heat, and the two children stretched in the shade of the willow
tree and slept and the old woman nodded on her bench, and over
the whole small hamlet the peace of sleep and the silence of the
heat of noonday fell, so that even the beasts stood with drooping,
drowsy heads.
Only the mother did not sleep. She took up her spindle and she sat
herself in the shade of the willow tree that cast its shadow on the
western part of the threshing-floor and she twisted the thread and
wound it. But after a while she could not work. Through the morning
she had worked steadily and smoothly, twisting and turning and
spinning, but now she could not be still. It was as though some
strange anxiety gathered like a power in her body. She had never
known the man not to come home for his food. She murmured to
herself, “It must be he has gone into the town to game or for
something or other.”
This she had not thought of, but the more she thought upon it the
more it seemed true that so he had done. And after a while her
cousin-neighbor came out to go to his fields and after a while his wife
awoke from where she had sat sleeping by a tree, and she called,
“Has your man gone for the day somewhere?”
The mother answered easily, “Aye, he has gone to the town on some
business of his own,” and the cousin searching slowly among his
hoes and spades for what he wanted called in his thin voice, “Aye, I
saw him gay in his new blue robe and set for town!”
“Aye,” said the woman.
Now her heart eased itself somewhat, and she fell to spinning again
with more zeal, since the cousin had seen him set for town. He had
gone for a day’s pleasure, doubtless, flinging himself off for the day
to revenge himself on her. It was what he would do with his new
gown and that brass ring of his scrubbed bright and clean and his
hair covered with oil. She nursed her anger somewhat at the
thought. But her anger was dead, and she could not make it live
again, because it was mingled with some strange anxiety still, for all
the cousin’s words.
The afternoon wore on long and hot. The old woman woke and cried
that her mouth was dry as bark and the mother rose and fetched her
tea to drink, and the children woke and rolled in the dust a while and
rose at last to play, and the babe woke and lay merry in his basket,
happy with his sleep.
Still the mother could not rest. If she could have slept she would
have, and on any common day she could have dropped easily into
sleep even as she worked, since she was so sound and robust that
sleep came on her deep and sweet and without her seeking it. But
there was some gnawing in her heart today that held her wide awake
and as though she listened for some sound that must come.
She rose at last impatient with her waiting and weary of the empty
street that was empty for her so long as she did not see the one she
sought, and she took up the babe and set him on her thigh and she
took her hoe and went to the field, and she called to the old woman,
“I go to weed the corn on the south hillside.” And as she went she
thought to herself that it would be easier if she were not at the
house, and the hours would pass more quickly if she pushed her
body to some hard labor.
So through the afternoon she worked in the corn field, her face
shielded from the sun’s heat with a blue cotton kerchief, and up and
down she moved her hoe unceasingly among the green young corn.
It was but a small, ragged field, for all of their land which could bear
it they put into rice, terracing even the hillsides as high as water
could be forced, because rice is a more dainty food than corn and
sells for higher price.
The sun poured down upon the shadeless hill and beat upon her and
soon her coat was wet and dark with her sweat. But she would not
rest at all except sometimes to suckle the babe when he cried, and
then she sat flat on the earth and suckled him and wiped her hot
face and stared across the brilliant summer land, seeing nothing.
When he was satisfied she put him down again to work once more
and she worked until her body ached and her mind was numb and
she thought of nothing now except of those weeds falling under the
point of her hoe and withering in the dry hot sunshine. At last the sun
rested on the edge of the land and the valley fell into sudden
shadow. Then she straightened herself and wiped her wet face with
her coat and she muttered aloud, “Surely he will be home waiting—I
must go to make his food.” And picking up the child from the bed of
soft earth where she had laid him she went home.
But he was not there. When she turned the corner of the house he
was not there. The old woman was peering anxiously toward the
field, and the two children sat upon the doorstep waiting and weary
and they cried out when they saw her and she said bewildered,
“Your father—is he not come yet?”
“He has not come and we are hungry,” cried the boy, and the girl
echoed in her broken, childish way, “Not come, and we are hungry!”
and sat with her eyes fast shut against the piercing last golden rays
of the sun. And the old mother rose and hobbled to the edge of the
threshing-floor and called out shrilly to the cousin coming home,
“Saw you my son anywhere?”
But the mother cried out in sudden impatience, “Let be, old mother!
Do not tell all he is not come!”
“Well, but he does not,” said the old woman, peering, troubled.
But the mother said no more. She fetched cold rice for the children
and heated a little water and poured it over the rice for the old
woman and found a morsel of some old food for the dog, and while
they ate she went down the street, the babe upon her arm, to the
wayside inn. There were but few guests there now, and only a
scattered one or two on his way home to some near village, for it
was the hour when men are in their homes and the day’s work done.
If he were there, she thought, he would be sitting at a table nearest
the street where he could hear and see whatever passed, or at a
table with a guest, for he would not be alone if he could help it, or if
there was a game going on, he would be in the middle of it. But
although she stared as she came there was no glint of a new blue
robe and no clatter of gambling upon a table. She went and looked
within the door then, but he was not there. Only the innkeeper stood
resting himself after the evening meal and he leaned against the wall
by his stove, his face black with the smoke and grease of many
days, for in such a blackish trade as his it seemed to him but little
use to wash himself, seeing he was black again so soon.
“Have you seen the father of my children?” the mother called.
But the innkeeper picked at his teeth with his black fingernail and
sucked and called back idly, “He sat here a while in that new blue
robe of his this morning and then he went townward for the day.” And
smelling some new gossip he cried afresh, “What—has aught
happened, goodwife?”
“Nothing—nothing—” replied the mother in haste. “He had business
in the town and it kept him late, I dare swear, and it may be he will
spend the night somewhere and come home tomorrow.”
“And what business?” asked the innkeeper suddenly curious.
“How can I know, being but a woman?” she answered and turned
away.
But on the way home while her lips called answer back to those who
called to her as she passed, she thought of something. When she
reached the house she went in and went to that cranny and felt in it.
It was empty. Well she knew there had been a precious small store
of copper coins there, and a small silver bit, too, because he had
sold the rice straw for a good price a day or two ago, being clever at
such things, and he brought a good part of the money back. She had
taken it from him and counted it and put it into the cranny and there it
should be. But it was not there.
Then she knew indeed that he was gone. It came over her in a daze
that he was truly gone. She sat down suddenly there in the earthen
house upon the earthen floor and holding the babe in her arms she
rocked herself back and forth slowly and in silence. Well, he was
gone! Here was she with the three children and the old woman, and
he gone!
The babe began to fret suddenly and without knowing what she did
she opened her bosom to him. The two children came in, the girl
whimpering and rubbing her eyes, and the old woman came in
leaning on her staff and saying over and over, “I do wonder where is
my son. Daughter, did my son say where he was? A very strange
thing where my son is gone—”
Then the mother rose and said, “He will be back tomorrow,
doubtless, old mother. Lie you down now and sleep. He will be back
tomorrow.”
The old mother listened and echoed, comforted, “Oh, aye, back
tomorrow doubtless,” and went to her pallet, feeling through the dim
room.
Then the mother led the two children into the dooryard and washed
them as her wont was on a summer’s night before they slept, and
she poured a gourdful of water over each of them, rubbing their
smooth brown flesh clean with her palm as she poured. But she did
not hear what they said, nor did she heed the girl’s moaning of her
eyes. Only when they went to the bed and the boy cried, astonished
that his father was not come, “And where does my father sleep,
then?”—only then did the mother answer out of her daze, “Doubtless
in the town, for he will come home tomorrow or in a day or so,” and
she added in sudden anger, “Doubtless when that bit of money is
gone he will be home again,” and she added again and most bitterly,
“And that new blue robe will be filthy and ready for me to wash
already, doubtless!”
And she was somehow glad she could be angry at him, and she held
her anger, clinging to it, because it made him seem more near, and
she clung to it while she led in the beast and barred the door against
the night and she muttered, “I dare swear I shall be just asleep when
he comes pounding at the door, even tonight!”
But in the dark night, in the still, hot night, in the silence of the closed
room, her anger went out of her and she was afraid. If he did not
come back what would she do, a lone woman and young?... The bed
was enormous, empty. She need take no care tonight, she might
spread her arms and legs out as she would. He was gone. Suddenly
there fell upon her the hottest longing for that man of hers. These six
years she had lain against him. Angry she might be with him in the
day, but at night she was near to him again and she forgot his idle
ways and his childishness. She remembered now how good and fair
he was to look upon, not coarse in the mouth and foul of breath as
most men are, but a very fair young man to see, and his teeth as
white as rice. So she lay longing for him, and all her anger was gone
out of her and only longing left.
When the morning came she rose weary with her sleeplessness, and
again she could be hard. When she rose and he did not come and
she had turned the beasts out and fed the children and the old
woman, she hardened herself and over and over she muttered half
aloud, “He will come when his money is gone—very well I know he
will come then!”
When the boy stared at the emptiness of the bed and when he asked
astonished, “Where is my father still?” she replied sharply and in a
sudden loud voice, “I say he is away a day or two, and if any asks
you on the street you are to say he is away a day or so.”
Nevertheless on that day when the children were off to play here and
there she did not go to the fields. No, she set her stool so that she
could see through the short single street of the hamlet if any came
that way, and while she made answer somehow to the old mother’s
prattling she thought to herself that the blue robe was so clear a blue
she could see it a long way off and she set herself to spinning, and
with every twist she gave to the spindle she looked secretly down the
road. And she counted over in her mind the money he had taken and
how many days it might last, and it seemed to her it could not last
more than six or seven days, except he had those nimble lucky
fingers of his to game with and so he might make more and stay a
little longer, too, before he must come back. Times there were as the
morning wore on when she thought she could not bear the old
mother’s prattling voice any more, but she bore it still for the hope of
seeing the man come home perhaps.
When the children wandered home at noon hungry and the boy
spied the cabbage bowl set aside for his father and asked for some,
she would not let him have it. She cuffed him soundly when he
asked again and she answered loudly, “No, it is for your father. If he
comes home tonight he will be hungry and want it all for himself.”
The long still summer’s afternoon wore on, and he did not come, and
the sun set in its old way, heavy and full of golden light, and the
valley was filled with the light for a little while, and the night came
and it was deep and dark and now she refused no more. She set the
bowl before the children and she said, “Eat what you will, for it will
spoil if it is left until another day, and who knows—” and she dipped
up some of the sweet and sour sauce and gave it to the old woman
saying, “Eat it, and I will make fresh if he comes tomorrow.”
“Will he come tomorrow then?” the old woman asked, and the
mother answered sombrely, “Aye, tomorrow perhaps.”
That night she laid herself down most sorrowful and afraid upon her
bed and this night she said openly to her own heart that none knew if
he would ever come back again.

Nevertheless, there was the hope of the seven days when his money
might be gone. One by one the seven days came, and in each one it
seemed to her in the midst of her waiting as though the day was
come for his return. She had never been a woman to gad about the
little hamlet or chatter overmuch with the other women there. But
now one after another of these twenty or so came by to see and ask,
and they asked where her man was, and they cried, “We are all one
house in this hamlet and all somehow related to him and kin,” and at
last in her pride the mother made a tale of her own and she
answered boldly, from a sudden thought in her head, “He has a
friend in a far city, and the friend said there was a place there he
could work and the wage is good so that we need not wear
ourselves upon the land. If the work is not suited to him he will come
home soon, but if it be such work as he thinks fit to him, he will not
come home until his master gives him holiday.”
This she said as calmly as she ever spoke a truth, and the old
woman was astounded and she cried, “And why did you not tell me
so good a lucky thing, seeing I am his mother?”
And the mother made a further tale and she answered, “He told me
not to speak, old mother, because he said your tongue was as loose
in your mouth as any pebble and all the street would know more than
he did, and if he did not like it he would not have them know it.”
“Did he so, then!” cackled the old mother, leaning forward on her
staff to peer at her daughter’s face, her old empty jaws hanging, and
she said half hurt, “It is true I ever was a good talker, daughter, but
not so loose as any pebble!”
Again and again the mother told the tale and once told she added to
it now and then to make it seem more perfect in its truth.
Now there was one woman who came often past her house, a widow
woman who lived in an elder brother’s house, and she had not
overmuch to do, being widowed and childless, and she sat all day
making little silken flowers upon a shoe she made for herself, and
she could ponder long on any little curious thing she heard. So she
pondered on this strange thing of a man gone, and one day she
thought of something and she ran down the street as fast as she
could on her little feet and she cried shrewdly to the mother, “But
there has no letter come a long time to this hamlet and I have not
heard of any letter coming to that man of yours!”
She went secretly to the only man who knew how to read in the
hamlet, and he wrote such few letters as any needed to have written
and read such as came for any, and so added a little to his
livelihood. This man the widow asked secretly, “Did any letter come
for Li The First, who was son to Li The Third in the last generation?”
And when the man said no, the gossip cried out, “But there was a
letter, or so his wife says, and but a few days ago.”
Then the man grew jealous lest they had taken the letter to some
other village writer and he denied again and again, and he said,
“Very well I know there was no letter, nor any answering letter, nor
has anyone come to me to read or write or to buy a stamp to put on
any letter and I am the only one who has such stamps. And there
has not come so much as a letter carrier this way for twenty days or
more.”
Then the widow smelled some strange thing and she told
everywhere, whispering that the wife of Li The First lied and there
had been no letter and doubtless the husband had run away and left
his wife. Had there not been a great quarrel over the new robe, so
that the whole hamlet heard them cursing each other, and the man
had pushed her down and struck her even? Or so the children said.
But when the talk leaked through to the mother she answered stoutly
that what she said was true and that she had made the new blue
robe on purpose for the man to go to the far town, and that the
quarrel was for another thing. As for the letter, there was no letter but
the news had come by word of mouth from a traveling pedlar who
had come in from the coast.
Thus did the mother lie steadfastly and well, and the old woman
believed the tale heartily and cried out often of her son and how rich
he would be, and the mother kept her face calm and smooth and she
did not weep as women do when their men run away and shame
them. At last the tale seemed true to all, and even the gossip was
silenced somewhat and could only mutter darkly over her silken
flowers, “We will see—as time comes, we will see if there is money
sent or any letter written, or if he comes home ever and again.”
So the little stir in the hamlet died down and the minds of people
turned to other things and they forgot the mother and her tale.

Then did the mother set herself steadfastly to her life. The seven
days were long past and the man did not come and the rice ripened
through the days and hung heavy and yellow and ready for the
harvest and he did not come. The woman reaped it alone then
except for two days when the cousin came and helped her when his
own rice was cut and bound in sheaves. She was glad of his help
and yet she feared him too, for he was a man of few words, honest
and few, and his questions were simple and hard not to answer
truthfully. But he worked silently and asked her nothing and he said
nothing except the few necessary words he must until he went away,
and then he said, “If he is not come when the time is here to divide
the grain with the landlord, I will help you then, for the new agent is a
wily, clever man, and of a sort ill for a woman to do with alone.”
She thanked him quietly, glad of his help, for she knew the agent but
a little, since he was new in the last years to those parts, and a
townsman who had a false heartiness in all he did and said.
So day had passed into month, and day after day the woman had
risen before the dawn and she left the children and the old woman
sleeping, and she set their food ready for them to eat when they
woke, and taking the babe with her in one arm and in her other hand
the short curved sickle she must use in reaping she set out to the
fields. The babe was large now and he could sit alone and she set
him down upon the earth and let him play as he would, and he filled
his hands with earth and put it to his mouth and ate of it and spat it
out hating it and yet he forgot and ate of it again until he was
covered with the muddy spew. But whatever he did the mother could
not heed him. She must work for two and work she did, and if the
child cried he must cry until she was weary and could sit down to
rest and then she could put her breast to his earthy mouth and let
him drink and she was too weary to care for the stains he left upon
her.
Handful by handful she reaped the stiff yellow grain, bending to
every handful, and she heaped it into sheaves. When gleaners came
to her field to glean what she might drop, as beggars and gleaners
do at harvest time, she turned on them, her face dark with sweat and
earth, and drawn with the bitterness of labor, and she screamed
curses at them, and she cried, “Will you glean from a lone woman
who has no man to help her? I am poorer than you, you beggars,
and you cursed thieves!” And she cursed them so heartily and she
so cursed the mothers that bore them and the sons they had
themselves that at last they let her fields be, because they were
afraid of such powerful cursing.
Then sheaf by sheaf she carried the rice to the threshing-floor and
there she threshed it, yoking the buffalo to the rude stone roller they
had, and she drove the beast all through the hot still days of autumn,
and she drove herself, too. When the grain was threshed, she
gathered the empty straw and heaped it and tossed the grain up and
winnowed it in the winds that came sometimes.
Now she pressed the boy into labor too and if he lagged or longed to
play she cuffed him out of her sheer weariness and the despair of
her driven body. But she could not make the ricks. She could not
heap the sheaves into the ricks, for this the man had always done,
since it was a labor he hated less than some, and he did it always
neatly and well and plastered the tops smooth with mud. So she
asked the cousin to teach her this one year and she could do it
thenceforth with the boy if the man stayed longer than a year, and
the cousin came and showed her how and she bent her body to the
task and stretched and threw the grass to him as he sat on top of the
rick and spread it, and so the rice was harvested.
She was bone-thin now with her labor and with being too often
weary, and every ounce of flesh was gone from her, and her skin
was burnt a dark brown except the red of cheeks and lips. Only the
milk stayed in her breasts rich and full. Some women there are
whose food goes all to their own fat and none to child or food for
child, but this woman was made for children, and her motherhood
would rob her own body ruthlessly if there was any need for child.
Then came the day set for measuring out the landlord’s share of all
the harvest. Now this landlord of the hamlet and the fields about it
never came himself to fetch his share. He lived an idle rich man in
some far city or other, since the land was his from his fathers, and he
sent in his place his agent, and this year it was a new agent, for his
old agent had left him the last year, being rich enough after twenty
years to cease his labors. This new agent came now and he came to
every farmer in that hamlet, and the mother waited at her own door,

You might also like