You are on page 1of 53

Turbulence an introduction for

scientists and engineers Davidson


Visit to download the full and correct content document:
https://textbookfull.com/product/turbulence-an-introduction-for-scientists-and-engineer
s-davidson/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Physics for scientists and engineers an interactive


approach 2nd Edition Hawkes

https://textbookfull.com/product/physics-for-scientists-and-
engineers-an-interactive-approach-2nd-edition-hawkes/

Introduction to Probability and Statistics for


Engineers and Scientists 6th Edition Sheldon M. Ross

https://textbookfull.com/product/introduction-to-probability-and-
statistics-for-engineers-and-scientists-6th-edition-sheldon-m-
ross/

Statistics for Engineers and Scientists William Navidi

https://textbookfull.com/product/statistics-for-engineers-and-
scientists-william-navidi/

Essential MATLAB for Engineers and Scientists Hahn

https://textbookfull.com/product/essential-matlab-for-engineers-
and-scientists-hahn/
Essential MATLAB for engineers and scientists Hahn

https://textbookfull.com/product/essential-matlab-for-engineers-
and-scientists-hahn-2/

Essential MATLAB for engineers and scientists Hahn

https://textbookfull.com/product/essential-matlab-for-engineers-
and-scientists-hahn-3/

Physics for scientists and engineers an interactive


approach 2nd Edition Peter J. Williams

https://textbookfull.com/product/physics-for-scientists-and-
engineers-an-interactive-approach-2nd-edition-peter-j-williams/

Probability statistics for engineers scientists Walpole

https://textbookfull.com/product/probability-statistics-for-
engineers-scientists-walpole/

Crystallography and surface structure an introduction


for surface scientists and nanoscientists Hermann

https://textbookfull.com/product/crystallography-and-surface-
structure-an-introduction-for-surface-scientists-and-
nanoscientists-hermann/
Turbulence
TURBULENCE
An Introduction for Scientists and Engineers

SECOND EDITION

p. a. davidson
University of Cambridge

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
© P. A. Davidson 2015
The moral rights of the author have been asserted
First Edition published in 2004
Second Edition published in 2015
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: 2014954276
ISBN 978–0–19–872258–8 (hbk.)
ISBN 978–0–19–872259–5 (pbk.)
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
For Henri and Mars
PREFACE TO THE SECOND EDITION

R oughly 10 years have passed since the publication of the first edition of this book, and
perhaps it is timely to reflect on the ambitions and structure of the text in the light of
recent advances in the subject. In the first edition three recurring themes were: (i) a pref-
erence for working in real space rather than Fourier space, wherever possible; (ii) the importance
of giving equal weight to the large and small scales in the discussion of homogeneous turbulence;
and (iii) a note of caution as to what may be achieved through the use of direct numerical simu-
lations (DNS) of turbulence. As to the first two themes, developments in the intervening years
have, by and large, tended to reinforce my particular preferences. However, I have been proved
overly pessimistic on the third point; the relentless rise in computing power has allowed DNS to
become the preferred means of attack for the theoretician interested in the fundamental structure
of turbulence, though of course DNS remains of limited value to the engineer or applied scientist
interested in practical problems. In any event, the text has been modified to reflect this growing
importance of DNS in fundamental studies.
The motivation for a second edition has been the desire to simplify the text where possible,
while updating those sections where the subject has moved on. As a consequence, significant parts
of Sections 5.1.3, 5.3.6, 6.3, 6.6, 9.2, 9.5, and Chapter 10 have been rewritten, and a new section,
6.2.10, has been introduced on passive scalar mixing. Nevertheless the spirit of the book remains,
I hope, unchanged; a desire to bridge the gap between the elementary coverage of turbulence
found in undergraduate texts and the advanced (if often daunting) discussion in monographs on
the subject. Above all, I have tried at all times to combine the maximum of physical insight with
the minimum of mathematical detail.
Once again it is a pleasure to acknowledge the help of friends and colleagues, particularly
the many stimulating discussions I continue to have with Julian Hunt, Yukio Kaneda, Per-Åge
Krogstad, and many colleagues in Cambridge. Oxford University Press were, as always, a pleasure
to work with.

P. A. Davidson
Cambridge, 2014

preface | vii
PREFACE TO THE FIRST EDITION

T
urbulence is all around us. The air flowing in and out of our lungs is turbulent, as is the
natural convection in the room in which you sit. Glance outside; the wind which gusts
down the street is turbulent, and it is turbulence which disperses the pollutants that belch
from the rear of motor cars, saving us from asphyxiation. Turbulence controls the drag on cars,
aeroplanes, and bridges, and it dictates the weather through its influence on large-scale atmos-
pheric and oceanic flows. The liquid core of the earth is turbulent, and it is this turbulence which
maintains the terrestrial magnetic field against the natural forces of decay. Even solar flares are
a manifestation of turbulence, since they are triggered by vigorous motion on the surface of the
sun. It is hard not to be intrigued by a subject which pervades so many aspects of our lives.
Yet curiosity can so readily give way to despair when the budding enthusiast embarks on ser-
ious study. The mathematical description of turbulence is complex and forbidding, reflecting the
profound difficulties inherent in describing three-dimensional, chaotic processes.
This is a textbook and not a research monograph. Our principal aim is to bridge the gap be-
tween the elementary, heuristic accounts of turbulence to be found in undergraduate texts, and
the more rigorous, if daunting, accounts given in the many excellent monographs on the sub-
ject. Throughout we seek to combine the maximum of physical insight with the minimum of
mathematical detail.
Turbulence holds a unique place in the field of classical mechanics. Despite the fact that the
governing equations have been known since 1845, there is still surprisingly little we can predict
with relative certainty. The situation is reminiscent of the state of electromagnetism before it was
transformed by Faraday and Maxwell. A myriad tentative theories have been assembled, often
centred around particular experiments, but there is not much in the way of a coherent theoretical
framework.1 The subject tends to consist of an uneasy mix of semi-empirical laws and deter-
ministic but highly simplified cartoons, bolstered by the occasional rigorous theoretical result.
Of course, such a situation tends to encourage the formation of distinct camps, each with its own
doctrines and beliefs. Engineers, mathematicians, and physicists tend to view turbulence in rather
different ways, and even within each discipline there are many disparate groups. Occasionally re-
ligious wars break out between the different camps. Some groups emphasize the role of coherent
vortices, while others downplay the importance of such structures and advocate the use of purely

1
One difference between turbulence and nineteenth-century electromagnetism is that the latter was eventually
refined into a coherent theory, whereas it is unlikely that this will ever occur in turbulence.

viii | preface
statistical methods of attack. Some believe in the formalism of fractals or chaos theory, others do
not. Some follow the suggestion of von Neumann and try to unlock the mysteries of turbulence
through massive computer simulations, others believe that this is not possible. Many engineers
promote the use of semi-empirical models of turbulence; most mathematicians find that this is
not to their taste. The debate is often vigorous and exciting and has exercised some of the finest
twentieth-century minds, such as L.D. Landau and G.I. Taylor. Any would-be author embarking
on a turbulence book must carefully pick his way through this minefield, resigned to the fact that
not everyone will be content with the outcome. But this is no excuse for not trying; turbulence is
of immense importance in physics and engineering, and despite the enormous difficulties of the
subject, significant advances have been made.
Roughly speaking, texts on turbulence fall into one of two categories. There are those which
focus on the turbulence itself and address such questions as, where does turbulence come from?
What are its universal features? To what extent is it deterministic? On the other hand, we have
texts whose primary concern is the influence of turbulence on practical processes, such as drag,
mixing, heat transfer, and combustion. Here the main objective is to parameterize the influence
of turbulence on these processes. The word modelling appears frequently in such texts. Applied
mathematicians and physicists tend to be concerned with the former category, while engineers
are necessarily interested in the latter. Both are important, challenging subjects.
On balance, this text leans slightly towards the first of these categories. The intention is to
provide some insight into the physics of turbulence and to introduce the mathematical appar-
atus which is commonly used to dissect turbulent phenomena. Practical applications, alas, take a
back seat. Evidently such a strategy will not be to everyone’s taste. Nevertheless, it seems natural
when confronted with such a difficult subject, whose pioneers adopted both rigorous and heuris-
tic means of attack, to step back from the practical applications and try and describe, as simply as
possible, those aspects of the subject which are now thought to be reasonably well understood.
Our choice of material has been guided by the observation that the history of turbulence has,
on occasions, been one of heroic initiatives which promised much yet delivered little. So we have
applied the filter of time and chosen to emphasize those theories, both rigorous and heuristic,
which look like they might be a permanent feature of the turbulence landscape. There is little
attempt to document the latest controversies, or those findings whose significance is still un-
clear. We begin, in Chapters 1 to 5, with a fairly traditional introduction to the subject. The
topics covered include: the origins of turbulence, boundary layers, the log-law for heat and mo-
mentum, free-shear flows, turbulent heat transfer, grid turbulence, Richardson’s energy cascade,
Kolmogorov’s theory of the small scales, turbulent diffusion, the closure problem, simple closure
models, and so on. Mathematics is kept to a minimum and we presuppose only an elementary
knowledge of fluid mechanics and statistics. (Those statistical ideas which are required are intro-
duced as and when they are needed in the text.) Chapters 1 to 5 may be appropriate as background
material for an advanced undergraduate or introductory postgraduate course on turbulence.
Next, in Chapters 6 to 8, we tackle the somewhat refined, yet fundamental, problem of homo-
geneous turbulence. That is, we imagine a fluid which is vigorously stirred and then left to itself.
What can we say about the evolution of such a complex system? Our discussion of homogeneous
turbulence differs from that given in most texts in that we work mostly in real space (rather than
Fourier space) and we pay as much attention to the behaviour of the large, energy-containing
eddies as we do to the small-scale structures.
Perhaps it is worth explaining why we have taken an unconventional approach to homoge-
neous turbulence, starting with our slight reluctance to embrace Fourier space. The Fourier trans-
form is conventionally used in turbulence because it makes certain mathematical manipulations

preface | ix
easier and because it provides a simple (though crude) means of differentiating between large
and small-scale processes. However, it is important to bear in mind that the introduction of the
Fourier transform produces no new information; it simply represents a transfer of information
from real space to Fourier space. Moreover, there are other ways of differentiating between large
and small scales, methods which do not involve the complexities of Fourier space. Given that
turbulence consists of eddies (blobs of vorticity) and not waves, it is natural to ask why we must
invoke the Fourier transform at all. Consider, for example, grid turbulence. We might picture this
as an evolving vorticity field in which vorticity is stripped off the bars of the grid and then mixed
to form a seething tangle of vortex tubes and sheets. It is hard to picture a Fourier mode being
stripped off the bars of the grid! It is the view of this author that, by and large, it is preferable
to work in real space, where the relationship between mathematical representation and physical
reality is, perhaps, a little clearer.
The second distinguishing feature of Chapters 6–8 is that equal emphasis is given to both large
and small scales. This is a deliberate attempt to redress the current bias towards small scales in
monographs on homogeneous turbulence. Of course, it is easy to see how such an imbalance
developed. The spectacular success of Kolmogorov’s theory of the small eddies has spurred a vast
literature devoted to verifying (or picking holes in) this theory. Certainly it cannot be denied that
Kolmogorov’s laws represent one of the milestones of turbulence theory. However, there have
been other success stories too. In particular, the work of Landau, Batchelor, and Saffman on the
large-scale structure of homogeneous turbulence stands out as a shinning example of what can
be achieved through careful, physically motivated analysis. So perhaps it is time to redress the
balance, and it is with this in mind that we devote part of Chapter 6 to the dynamics of the large-
scale eddies. Chapters 6 to 8 may be suitable as background material for an advanced postgraduate
course on turbulence, or can act as a reference source for professional researchers.
The final section of the book, Chapters 9 and 10, covers certain special topics rarely discussed
in introductory texts. The motivation here is the observation that many geophysical and astro-
physical flows are dominated by the effects of body forces, such as buoyancy, Coriolis, and Lorentz
forces. Moreover, certain large-scale flows are approximately two dimensional and this has led to
a concerted investigation of two-dimensional turbulence over the last few years. We touch on the
influence of body forces in Chapter 9 and two-dimensional turbulence in Chapter 10.
There is no royal route to turbulence. Our understanding of it is limited and what little we
do know is achieved through detailed and difficult calculation. Nevertheless, it is hoped that this
book provides an introduction which is not too arduous and which allows the reader to retain at
least some of that initial sense of enthusiasm and wonder.
It is a pleasure to acknowledge the assistance of many friends and colleagues. Alan Bailey,
Kate Graham, and Teresa Cronin all helped in the preparation of the manuscript, Jean Delery
of ONERA supplied copies of Henri Werle’s beautiful photographs, while the drawing of the
cigarette plume and the copy of Leonardo’s sketch are the work of Fiona Davidson. I am grateful
to Julian Hunt, Marcel Lesieur, Keith Moffatt, and Tim Nickels for many interesting discussions
on turbulence. In addition, several useful suggestions were made by Ferit Boysan, Jack Herring,
Jon Morrison, Mike Proctor, Mark Saville, Christos Vassilicos, and John Young. Finally, I would
like to thank Stephen Davidson who painstakingly read the entire manuscript, exposing the many
inconsistencies in the original text.

P. A. Davidson
Cambridge, 2003

x | preface
CONTENTS

Part 1 The classical picture of turbulence


1. The ubiquitous nature of turbulence 3
1.1 The experiments of Taylor and Bénard 4
1.2 Flow over a cylinder 7
1.3 Reynolds’ experiment 8
1.4 Common themes 9
1.5 The ubiquitous nature of turbulence 13
1.6 Different scales in a turbulent flow: a glimpse at the energy cascade of
Kolmogorov and Richardson 18
1.7 The closure problem of turbulence 27
1.8 Is there a ‘theory of turbulence’? 29
1.9 The interaction of theory, computation, and experiment 30
2. The equations of fluid mechanics 34
2.1 The Navier–Stokes equation 35
2.1.1 Newton’s second law applied to a fluid 35
2.1.2 The convective derivative 38
2.1.3 Integral versions of the momentum equation 40
2.1.4 The rate of dissipation of energy in a viscous fluid 41
2.2 Relating pressure to velocity 43
2.3 Vorticity dynamics 44
2.3.1 Vorticity and angular momentum 44
2.3.2 The vorticity equation 48
2.3.3 Kelvin’s theorem 52
2.3.4 Tracking vorticity distributions 55
2.4 A definition of turbulence 57
3. The origins and nature of turbulence 61
3.1 The nature of chaos 62
3.1.1 From non-linearity to chaos 63
3.1.2 More on bifurcations 66
3.1.3 The arrow of time 69

contents | xi
3.2 Some elementary properties of freely evolving turbulence 71
3.2.1 Various stages of development 73
3.2.2 The rate of destruction of energy in fully developed turbulence 77
3.2.3 How much does the turbulence remember? 81
3.2.4 The need for a statistical approach and different methods of taking averages 85
3.2.5 Velocity correlations, structure functions, and the energy spectrum 88
3.2.6 Is the asymptotic state universal? Kolmogorov’s theory 94
3.2.7 The probability distribution of the velocity field 97
4. Turbulent shear flows and simple closure models 105
4.1 The exchange of energy between the mean flow and the turbulence 107
4.1.1 Reynolds stresses and the closure problem of turbulence 108
4.1.2 The eddy viscosity theories of Boussinesq and Prandtl 111
4.1.3 The transfer of energy from the mean flow to the turbulence 114
4.1.4 A glimpse at the k-ε model 119
4.2 Wall-bounded shear flows and the log-law of the wall 122
4.2.1 Turbulent flow in a channel and the log-law of the wall 122
4.2.2 Inactive motion—a problem for the log-law? 127
4.2.3 Turbulence profiles in channel flow 129
4.2.4 The log-law for a rough wall 131
4.2.5 The structure of a turbulent boundary layer 131
4.2.6 Coherent structures 134
4.2.7 Spectra and structure functions near the wall 139
4.3 Free shear flows 142
4.3.1 Planar jets and wakes 142
4.3.2 The round jet 148
4.4 Homogeneous shear flow 152
4.4.1 The governing equations 152
4.4.2 The asymptotic state 156
4.5 Heat transfer in wall-bounded shear flows—the log-law revisited 157
4.5.1 Turbulent heat transfer near a surface and the log-law for temperature 157
4.5.2 The effect of stratification on the log-law—the atmospheric boundary layer 163
4.6 More on one-point closure models 169
4.6.1 A second look at the k-ε model 169
4.6.2 The Reynolds stress model 176
4.6.3 Large eddy simulation: a rival for one-point closures? 180
5. The phenomenology of Taylor, Richardson, and Kolmogorov 188
5.1 Richardson revisited 191
5.1.1 Time and length scales in turbulence 191
5.1.2 The energy cascade pictured as the stretching of turbulent eddies 195
5.1.3 The dynamic properties of turbulent eddies: linear and angular impulse 202
5.2 Kolmogorov revisited 211
5.2.1 Dynamics of the small scales 211
5.2.2 Turbulence-induced fluctuations of a passive scalar 221
5.3 The intensification of vorticity and the stretching of material lines 228
5.3.1 Enstrophy production and the skewness factor 228

xii | contents
5.3.2 Sheets or tubes? 231
5.3.3 Examples of concentrated vortex sheets and tubes 234
5.3.4 Are there singularities in the vorticity field? 236
5.3.5 The stretching of material line elements 240
5.3.6 The interplay of the strain and vorticity fields 243
5.4 Turbulent diffusion by continuous movements 252
5.4.1 Taylor diffusion of a single particle 254
5.4.2 Richardson’s law for the relative diffusion of two particles 257
5.4.3 The influence of mean shear on turbulent dispersion 260
5.5 Why turbulence is never Gaussian 263
5.5.1 The experimental evidence and its interpretation 263
5.5.2 A glimpse at closure schemes which assume near-Gaussian statistics 267
5.6 Closure 268

Part 2 Freely decaying, homogeneous turbulence


6. Isotropic turbulence (in real space) 275
6.1 Introduction: exploring isotropic turbulence in real space 275
6.1.1 Deterministic cartoons versus statistical phenomenology 276
6.1.2 The strengths and weaknesses of Fourier space 281
6.1.3 An overview of this chapter 283
6.2 The governing equations of isotropic turbulence 293
6.2.1 Some kinematics: velocity correlation functions and structure functions 293
6.2.2 More kinematics: the simplifications of isotropy and the vorticity
correlation function 300
6.2.3 A summary of the kinematic relationships 304
6.2.4 Dynamics at last: the Karman–Howarth equation 308
6.2.5 Kolmogorov’s four-fifths law 310
6.2.6 The skewness factor and enstrophy production (reprise) 312
6.2.7 The dynamical equation for the third-order correlations
and the problem of closure 313
6.2.8 Closure of the dynamical equations in the equilibrium range 314
6.2.9 Quasi-normal-type closure schemes (part 1) 316
6.2.10 Passive scalar mixing in isotropic turbulence and Yaglom’s
four-thirds law 318
6.3 The dynamics of the large scales 320
6.3.1 The classical view: Loitsyansky’s integral and Kolmogorov’s decay laws 322
6.3.2 Landau’s angular momentum 323
6.3.3 Batchelor’s pressure forces 328
6.3.4 Saffman’s spectrum 333
6.3.5 A consistent theory of the large scales in Batchelor turbulence 342
6.3.6 A summary of the dynamics of the large scales 346
6.4 The characteristic signature of eddies of different shape 348
6.4.1 Townsend’s model eddy and its relatives 349
6.4.2 Turbulence composed of Townsend’s model eddies of different sizes 353
6.4.3 Other model eddies 356

contents | xiii
6.5 Intermittency in the inertial-range eddies 357
6.5.1 A problem for Kolmogorov’s theory? 358
6.5.2 The log-normal model of intermittency 360
6.5.3 The β̂ model of intermittency 363
6.6 Measuring the distribution of energy and enstrophy across the different
eddy sizes 366
6.6.1 A real-space function which represents, approximately, the variation
of energy with scale 366
6.6.2 Relating energy distributions in real and Fourier space 374
6.6.3 Cascade dynamics in real space 379
7. The role of numerical simulations 394
7.1 What is DNS or LES? 394
7.1.1 Direct numerical simulations (DNS) 394
7.1.2 Large eddy simulations (LES) 398
7.2 On the dangers of periodicity 404
7.3 Structure in chaos 406
7.3.1 Tubes, sheets, and cascades 407
7.3.2 On the taxonomy of worms and clusters of worms 412
8. Isotropic turbulence (in spectral space) 419
8.1 Kinematics in spectral space 420
8.1.1 The Fourier transform and its properties 421
8.1.2 The Fourier transform as a filter 425
8.1.3 The autocorrelation function and power spectrum 427
8.1.4 The transform of the correlation tensor and the three-dimensional
energy spectrum 431
8.1.5 One-dimensional energy spectra in three-dimensional turbulence 434
8.1.6 Relating the energy spectrum to the second-order structure function 438
8.1.7 A footnote: singularities in the spectrum due to anisotropy 440
8.1.8 Another footnote: the transform of the velocity field 441
8.1.9 Definitely the last footnote: what do E(k) and E1 (k) really represent? 442
8.2 Dynamics in spectral space 445
8.2.1 An evolution equation for E(k) 445
8.2.2 Closure in spectral space 448
8.2.3 Quasi-normal type closure schemes (part 2) 454

Part 3 Special topics


9. The influence of rotation, stratification, and magnetic fields on turbulence 467
9.1 The importance of body forces in geophysics and astrophysics 467
9.2 The influence of rapid rotation and stable stratification 470
9.2.1 The Coriolis force 470
9.2.2 The Taylor–Proudman theorem 473
9.2.3 Properties of inertial waves 474
9.2.4 Turbulence in rapidly rotating systems 479
9.2.5 From rotation to stratification (or from cigars to pancakes) 485

xiv | contents
9.3 The influence of magnetic fields I: the MHD equations 488
9.3.1 The interaction of moving conductors and magnetic
fields: a qualitative overview 489
9.3.2 From Maxwell’s equations to the governing equations of MHD 494
9.3.3 Simplifying features of low magnetic Reynolds number MHD 498
9.3.4 Simple properties of high magnetic Reynolds number MHD 499
9.4 The influence of magnetic fields II: MHD turbulence 503
9.4.1 The growth of anisotropy in MHD turbulence 505
9.4.2 The evolution of eddies at low magnetic Reynolds number 507
9.4.3 The Landau invariant for homogeneous MHD turbulence 513
9.4.4 Decay laws at low magnetic Reynolds number 514
9.4.5 Turbulence at high magnetic Reynolds number 516
9.4.6 The shaping of eddies by combined Coriolis and Lorentz forces 520
9.5 Turbulence in the core of the earth 522
9.5.1 An introduction to planetary dynamo theory 523
9.5.2 Numerical simulations of the geodynamo 529
9.5.3 Various cartoons of the geodynamo 532
9.5.4 An α2 model of the geodynamo based on inertial wave packets 538
9.6 Turbulence near the surface of the sun 553
10. Two-dimensional turbulence 559
10.1 The classical picture of 2D turbulence: Batchelor’s self-similar
spectrum, the inverse energy cascade, and the E(k)~k–3 enstrophy
flux law 560
10.1.1 What is two-dimensional turbulence? 560
10.1.2 What does the turbulence remember? 565
10.1.3 Batchelor’s self-similar spectrum 565
10.1.4 The inverse energy cascade of Batchelor and Kraichnan 567
10.1.5 Different scales in 2D turbulence 570
10.1.6 The shape of the energy spectrum and the k–3 law 571
10.1.7 Problems with the k–3 law 574
10.2 Coherent vortices: a problem for the classical theories 577
10.2.1 The evidence 577
10.2.2 The significance 579
10.3 The governing equations in statistical form 581
10.3.1 Correlation functions, structure functions, and the energy spectrum 582
10.3.2 The two-dimensional Karman–Howarth equation 586
10.3.3 Four consequences of the Karman–Howarth equation 586
10.3.4 The two-dimensional Karman–Howarth equation in spectral space 588
10.4 Variational principles for predicting the final state in confined domains 591
10.4.1 Minimum enstrophy 592
10.4.2 Maximum entropy 594
10.5 Quasi-two-dimensional turbulence: bridging the gap with reality 594
10.5.1 The governing equations for shallow-water, rapidly rotating flow 595
10.5.2 The Karman–Howarth equation for shallow-water, rapidly
rotating turbulence 597

contents | xv
Epilogue 601
Appendix 1: Vector identities and an introduction to tensor notation 603
A1.1 Vector identities and theorems 603
A1.2 An introduction to tensor notation 606
Appendix 2: The properties of isolated vortices: invariants, far-field properties,
and long-range interactions 611
A2.1 The far-field velocity induced by an isolated eddy 611
A2.2 The pressure distribution in the far field 613
A2.3 Integral invariants of an isolated eddy: linear and angular impulse 614
A2.4 Long-range interactions between eddies 617
Appendix 3: Hankel transforms and hypergeometric functions 620
A3.1 Hankel transforms 620
A3.2 Hypergeometric functions 621
Appendix 4: The kinematics of homogeneous, axisymmetric turbulence 623

Subject index 625

xvi | contents
PA R T 1

The classical picture of turbulence

Cross section through a turbulent jet carrying a passive scalar. The colours indicate the sca-
lar concentration. Note that the edge of the jet is highly convoluted. (See the discussion in
Section 4.3.1.)
Image courtesy of P.E. Dimotakis, California Institute of Technology.
CHAPTER 1

The ubiquitous nature


of turbulence
Vether it’s worth goin’ through so much, to learn so little, as the charity-boy said ven he got to the
end of the alphabet, is a matter o’ taste.
Charles Dickens, The Pickwick Papers

T
he study of turbulence is not easy, requiring a firm grasp of applied mathematics and
considerable physical insight into the dynamics of fluids. Worse still, even after the various
theoretical hypotheses have been absorbed, there are relatively few situations in which we
can make definite predictions!
For example, perhaps the simplest (and oldest) problem in turbulence concerns the decay of
energy in a cloud of turbulence. That is, we stir up a fluid and then leave it to itself. The turbulence
decays because of viscous dissipation and a natural question to ask is, how fast does the kinetic
energy decline? Theoretical physicists and applied mathematicians have been trying to answer
this question for over half a century and still they cannot agree. At times, one is tempted to side
with Weller the elder in The Pickwick Papers.
Nevertheless, there are certain predictions that can be made based on a variety of physical
arguments. This is important because turbulent motion is the natural state of most fluids. As you
read this book the air flowing up and down your larynx is turbulent, as is the natural convection
in the room in which you sit. Glance outside: if you are lucky you will see leaves rustling in the
turbulent wind. If you are unlucky you will see pollutants belching out from the rear of motor cars,
and it is turbulent convection in the street that disperses the pollutants, saving the pedestrians
from an unfortunate fate.
Engineers need to know how to calculate the aerodynamic drag on planes, cars, and buildings.
In all three cases the flow will certainly be turbulent. At a larger scale, motion in the oceans and
in the atmosphere is turbulent, so weather forecasters and oceanographers study turbulence. Tur-
bulence is also important in geophysics, since it is turbulent convection in the core of the Earth
that maintains the Earth’s magnetic field despite the natural forces of decay. Even astrophysicists

Turbulence: An Introduction for Scientists and Engineers. Second Edition. P. A. Davidson.


© P. A. Davidson 2015. Published in 2015 by Oxford University Press.
study turbulence, since it controls phenomena such as solar flares, the 22-year solar cycle, and the
rate of flow of mass through accretion discs onto the surfaces of young and dying stars.
It is the ubiquitous nature of turbulence, from the eruption of solar flares to the rustling of
leaves, that makes the subject both important and intriguing. The purpose of this chapter is to
give some indication as to just how widespread turbulence really is, and to introduce some of its
elementary properties. We start by describing some simple laboratory experiments.

1.1 The experiments of Taylor and Bénard


It is an empirical observation that the motion of a very viscous or slow-moving fluid tends to be
smooth and regular. We call this laminar flow. However, if the fluid viscosity is not too high, or
the characteristic speed is moderate to large, then the movement of the fluid becomes irregular
and chaotic, i.e. turbulent. The transition from laminar to turbulent motion is nicely illustrated
in a number of simple experiments, of which the most famous are probably those of Reynolds,
Bénard, and Taylor.
In 1923 Taylor described a remarkably simple, yet thought-provoking, experiment. Suppose
that we have two concentric cylinders and that the annular gap between the cylinders is filled
with a liquid, say water. The inner cylinder is made to rotate while the outer one remains sta-
tionary. At low rotation rates the fluid within the gap does what you would expect: it also rotates,
being dragged around by the inner cylinder. At higher rotation rates, however, something unex-
pected happens. At a certain critical speed, toroidal vortices suddenly appear, superimposed on
the primary circular motion. (This is shown schematically in Figure 1.1(a).) These axisymmetric
structures are called Taylor vortices, for an obvious reason, and they arise because of an instability
of the basic rotary flow. The net motion of a fluid particle is now helical, confined to a toroidal
surface.
The reason for the sudden appearance of Taylor vortices is related to the centrifugal force. This
tends to drive the rotating fluid radially outward. Below the critical speed this force is balanced by
a radial pressure gradient and there is no radial motion. However, the centrifugal force is greatest

(a) z (b) (c)


r

R d

Taylor vortices Wavy Taylor vortices Turbulent Taylor vortices

Figure 1.1 Flow between concentric cylinders. As the rotation rate of the inner cylinder increases,
the flow becomes progressively more complex until eventually turbulence sets in.

4 | the ubiquitous nature of turbulence


at points where the rotation rate is highest (near the inner cylinder) and so there is always a
tendency for fluid at the inner surface to move outward, displacing the outer layers of fluid. This
tendency is held in check by pressure and viscous forces. At the critical rotation rate, however, the
viscous forces are no longer able to suppress radial disturbances and the flow becomes unstable
to the slightest perturbation. Of course, all of the rapidly rotating fluid cannot move uniformly
outward because the outer fluid is in the way. Thus the flow breaks up into bands (or cells) as
shown in Figure 1.1(a). This is known as the Rayleigh instability.1
Now suppose we increase the rotation rate a little more, say by 25%. The Taylor vortices them-
selves become unstable and so-called wavy Taylor vortices appear. These have the appearance
of non-axisymmetric toroidal vortices which migrate around the inner cylinder (Figure 1.1(b)).
Note that, although complex, this flow is still laminar (non-chaotic).
Suppose we now increase the speed of the inner cylinder yet further. More complex, unsteady
structures start to emerge (so-called modulated wavy Taylor vortices) until eventually, when
the speed of the inner cylinder is sufficiently high, the flow becomes fully turbulent. In this
final state the time-averaged flow pattern resembles that of the steady Taylor vortices shown in
Figure 1.1(a), although the cells are a little larger. However, superimposed on this mean flow we
find a chaotic component of motion, so that individual fluid particles are no longer confined to
toroidal surfaces. Rather, as they are swept around by the mean flow, there is a constant jostling
for position, almost as if the particles are in a state of Brownian motion (Figure 1.1(c)). A typical
measurement of, say, uz (t) would look something like that shown in Figure 1.2. (Here ū is the
time-averaged value of uz .)
In fact, it is not just the rotation rate, , that determines which regime prevails. The kinematic
viscosity of the fluid, v, the annular gap, d, the radius of the inner cylinder, R, and the length of
the apparatus, L, are also important. We can construct three independent dimensionless groups
from , v, d, R, and L, say,
2 d3 R
Ta = , d/R , L/R.
v2
The first of these groups, Ta, is called the Taylor number. (Different authors use slightly different
definitions of Ta.) When the apparatus is very long, L  R, and the gap very narrow, d  R, it
happens to be the value of Ta, and only Ta, which determines the onset of Taylor vortices. In fact,
the critical value of Ta turns out to be 1.70 × 103 and the axial wavenumber of the vortices is
k = 2π/λ = 3.12/d, λ being the wavelength.
Let us now consider a quite different experiment, often called Bénard convection or Rayleigh–
Bénard convection. Suppose that a fluid is held between two large, flat, parallel plates, as shown

Figure 1.2 Variation of the axial component of velocity with


u time at some typical location in the annulus. When the flow
is turbulent there is a mean component of motion plus a
t random component.

1
Rayleigh identified the instability mechanism and produced a stability criterion for inviscid, rotating flows.
Taylor later extended the theory to viscous flows of the type shown in Figure 1.1(a).

the experiments of taylor and b énard | 5


(a) (b) (c)
COLD (T = T0)

No motion d
g

HOT (T = T0 + ΔT)
Low ΔT Bénard cells Turbulent convection
(High ΔT)

Figure 1.3 Fluid is held between two flat, parallel plates. The lower plate is heated. At low values
of T the fluid is quiescent. As T is increased natural convection sets in, first in the form of
regular convection cells and then in the form of turbulent flow.

in Figure 1.3(a). The lower plate is maintained at temperature T = T0 + T and the upper plate
at temperature T 0 . At low values of T the fluid remains stagnant and heat passes from one plate
to the other by molecular conduction. Of course, there is an upward buoyancy force which is
greatest near the lower plate and tends to drive the fluid upward. However, at low values of T
this force is exactly balanced by a vertical pressure gradient. We now slowly increase T. At a
critical value of T the fluid suddenly starts to convect as shown in Figure 1.3(b).
The flow consists of hot rising fluid and cold falling regions. This takes the form of regular
cells, called Bénard cells, which are reminiscent of Taylor vortices.2 The rising fluid near the top
of the layer loses its heat to the upper plate by thermal conduction, whereupon it starts to fall.
As this cold fluid approaches the lower plate it starts to heat up and sooner or later it reverses
direction and starts to rise again. Thus we have a cycle in which potential energy is continuously
released as light fluid rises and dense fluid falls. Under steady conditions the rate of working of
the buoyancy force is exactly balanced by viscous dissipation within the fluid.
The transition from the quiescent state to an array of steady convection cells is, of course,
triggered by an instability associated with the buoyancy force. If we perturb the quiescent state,
allowing hot fluid to rise and cold fluid to fall, then potential energy will be released and, if the
viscous forces are not too excessive, the fluid will accelerate, giving rise to an instability. On the
other hand, if the viscous forces are large then viscous dissipation can destroy all of the potential
energy released by the perturbation. In this case the quiescent configuration is stable. Thus we
expect the quiescent state to be stable if T is small and v large, and unstable if T is large and v
is small.
It is possible to set up a variational problem where one looks for the form of perturbation that
maximizes the rate of release of potential energy and minimizes the viscous dissipation. On the
assumption that the observed convection cells have just such a shape this yields the critical values
of T and v at which instability first sets in. It turns out that the instability criterion is

gβTd3
Ra = ≥ 1.70 × 103 ,

2
In plan view the convection cells can take a variety of forms, depending on the value of (T)/(T)CRIT and on
the shape of the container. Two-dimensional rolls and hexagons are both common.

6 | the ubiquitous nature of turbulence


where Ra (a dimensionless parameter) is known as the Rayleigh number, β is the expansion co-
efficient of the fluid, d is the height of the gap, and α the thermal diffusivity. The wavenumber
(2π/wavelength) of the convection cells is k = 3.12/d.
Of course, we have seen these numbers before. The critical value of the Taylor number for
narrow annuli is
2 d3 R
Ta = = 1.70 × 103
v2
and the wavenumber of the Taylor cells at the onset of instability is k = 3.12/d. At first sight this
coincidence seems remarkable. However, it turns out that there is an analogy between axisym-
metric flow with swirl (in the narrow-gap approximation) and natural convection. The analogue
of the buoyancy force is the centrifugal force and the angular momentum, = ruθ , in a rotating
axisymmetric flow is convected and diffused just like the temperature, T, in Bénard’s experiment.3
In fact, Rayleigh first derived his famous criterion for the instability of a rotating, inviscid fluid
through a consideration of the analogy between the buoyancy and centrifugal forces.
Now suppose Ra is slowly increased above its critical value. Then there is a point at which the
Bénard cells themselves become unstable, and more complex, unsteady structures start to appear.
Eventually, when Ra is large enough, we reach a state of turbulent convection.
It seems that the general picture that emerges from both Taylor’s and Bénard’s experiments is
the following. For high viscosities the basic configuration is stable. As the viscosity is reduced we
soon reach a point where the basic equilibrium is unstable to infinitesimal perturbations and the
system bifurcates (changes) to a more complex state, consisting of a steady laminar flow in the
form of regular cells. As the viscosity is reduced even further the new flow itself becomes unstable
and we arrive at a more complex motion. Subsequent reductions in viscosity give rise to pro-
gressively more complex flows until eventually, for sufficiently small v, the flow is fully turbulent.
At this point the flow field consists of a mean (time-averaged) component plus a random, chaotic
motion. This sequence of steps is called the transition to turbulence.

1.2 Flow over a cylinder


This kind of behaviour, in which a fluid passes through a sequence of flow regimes of increasing
complexity as v is reduced, is also seen in external flows. Consider, for example, flow past a cy-
linder, as shown in Figure 1.4. An inverse measure of v is given by the (dimensionless) Reynolds
number, Re = ud/v, where u is the upstream speed of the fluid and d the diameter of the cylin-
der. At high values of v we have a steady, symmetric flow pattern. As Re approaches unity the
upstream–downstream symmetry is lost and in the range Re ~ 5–40 we find steady vortices at-
tached to the rear of the cylinder. When Re reaches a value of around 40 an instability is observed
in the form of an oscillation of the wake, and by the time we reach Re ~ 100 the vortices start to
peel off from the rear of the cylinder in a regular, periodic manner. The flow is still laminar but
we have now lost the top–bottom symmetry. This is, of course, the famous Karman vortex street.
The laminar Karman street persists up to Re ~ 200 at which point three-dimensional instabilities
develop. By Re ~ 400 low levels of turbulence start to appear within the vortices. Nevertheless,

3
Actually it turns out that there is a 0.7% difference in the critical values of Ta and Ra, so the analogy is very
nearly, but not exactly, perfect.

flow over a cylinder | 7


(a)

Re « 1

(b)
Re ~ 10

(c)

Re ~ 100

(d)

Re ~ 104

(e)
Figure 1.4 Flow behind a cylinder:
(a) Re < 1; (b) 5 < Re < 40; Re ~ 106
(c) 100 < Re < 200; (d) Re ~ 104 ;
and (e) Re ~ 106 .

the periodicity of the vortex shedding remains quite robust. Eventually, for high values of Re,
say 105 , the turbulence spreads out of the vortices and we obtain a fully turbulent wake. Within
this wake, however, one can still detect coherent vortex shedding. Thus we have the same sort of
pattern of behaviour as seen in Taylor’s experiment. As v is reduced the flow becomes more and
more complex until, eventually, turbulence sets in.
This example is important since it illustrates the key role played by the Reynolds number
in determining the state of a flow. In general, the Reynolds number, Re = u /v, represents the
ratio of inertial to viscous forces in a fluid, provided that is chosen appropriately.4 Thus, when
Re is large, the viscous forces, and hence viscous dissipation, are small. Such flows are prone to
instabilities and turbulence, as evidenced by our example of flow over a cylinder.

1.3 Reynolds’ experiment


The transition from laminar to turbulent flow, and the important role played by Re in this tran-
sition, was first pointed out by Reynolds in 1883. Reynolds was concerned with flow along a
straight, smooth pipe which, despite having a particularly simple geometry, turns out to be rather
more subtle and complex than either Taylor’s or Bénard’s experiments.
In his now famous paper, Reynolds clearly distinguished between the two possible flow re-
gimes (laminar and turbulent) and argued that the parameter controlling the transition from one

4
The inertial forces are of order u2 /lu where lu is a length scale typical of the streamline pattern, say the curvature
of the streamlines, while the viscous forces are of order vu/lv2 , where lv is a length typical of cross-stream gradients
in velocity (see Chapter 2).

8 | the ubiquitous nature of turbulence


initiation formation of turbulent slug laminar region turbulent slug

Figure 1.5 Turbulent slugs near the inlet of a pipe.

regime to another had to be Re = ud/v, where d is the pipe diameter and u the mean flow down
the pipe. He also noted that the critical value of Re at which turbulence first appears is very sen-
sitive to disturbances at the entrance to the pipe. Indeed, he suggested that the instability that
initiates the turbulence might require a perturbation of a certain magnitude, for a given value of
Re, for the unstable motion to take root and turbulence to set in. For example, when the inlet
disturbances were minimized, he found the laminar flow to be stable up to Re ~ 13,000, whereas
turbulence typically appears at Re ~ 2000 if no particular effort is taken to minimize the disturb-
ances. We now know that Reynolds was right. The current view is that fully developed laminar
pipe flow is stable to infinitesimal disturbances for all values of Re, no matter how large, and in-
deed recent experiments (with very special inlet conditions) have achieved laminar flow for values
of Re up to 90,000. It is the size of the disturbance, and the type of inlet, that matters.
Reynolds also examined what happens when turbulence is artificially created in the pipe. He
was particularly interested in whether or not there is a value of Re below which the turbulence
dies out. It turns out that there is, and that this corresponds to Re ~ 2000.
The modern view is the following. The inlet conditions are very important. When we have
a simple, straight inlet, as shown in Figure 1.5, and Re exceeds 104 , turbulence tends to appear
first in the annular boundary layer near the inlet. The turbulence initially takes the form of small,
localized patches of chaotic motion. These ‘turbulent spots’ then spread and merge until a slug
of turbulence fills the pipe (Figure 1.5). For lower values of Re, on the other hand, the boundary
layer near the inlet is thought to be stable to small disturbances. Thus the perturbations which
initiate transition in the range 2000–104 must be present at the pipe inlet, or else represent a
finite-amplitude instability of the boundary layer near the inlet. In any event, whatever the origin
of the turbulence, it appears that transition starts with a series of intermittent turbulent slugs
passing down the pipe (Figure 1.5). Provided Re exceeds ~ 2000, these slugs tend to grow in
length at the expense of the non-turbulent fluid between them, eventually merging to form fully
developed turbulence. When Re is less than ~ 2000, on the other hand, any turbulent slugs that
are generated near the inlet simply decay.

1.4 Common themes


The examples above suggest that there are at least two types of transition to turbulence. There are
flows in which turbulent motion appears first in small patches. Provided Re is large enough, the
turbulent patches grow and merge until fully developed turbulence is established. This is typical
of transition to turbulence in boundary layers (Figure 1.6) and in pipes. The key point here is that,
in the transition region, the turbulence is somewhat intermittent, being interspersed by quiescent,
laminar regions.

common themes | 9
2D instability 3D instability turbulent spots

u Plate u

laminar transition to turbulence fully-developed


f low turbulence

Figure 1.6 Cartoon of transition to turbulence in a boundary layer in which the plate is flat and
smooth and the level of free-stream turbulence is low.

240
220

200
180
160
140 x
120
10
y 100
0
10 80
20
30 60
z 40
50 40

Figure 1.7 Computer simulation of turbulence created by a small rectangular object mounted on
a plate.
Figure courtesy of Neil Sandham.

Such flows also tend to be very sensitive to the level of the disturbances. For example, free-
stream turbulence or imperfections on the surface of the plate can trigger turbulence in an
otherwise stable boundary layer. This is illustrated in Figure 1.7 which shows a computer simula-
tion of turbulence created by a small rectangular object mounted on a plate. It is striking that such
a small object can create a long and persistent wedge of turbulence. In this case the height of the

10 | the ubiquitous nature of turbulence


object is comparable with the boundary-layer thickness, but much smaller surface imperfections
can have much the same effect.
Then there are flows in which, when a certain threshold is exceeded, chaos develops uniformly
throughout the fluid. This might start out as a simple instability of the mean flow, leading to a
more complicated laminar motion, which in turn becomes unstable and breaks up into even more
complex structures (see Figures 1.1 and 1.3). A sequence of such instabilities leads eventually to
random, chaotic motion, i.e. turbulence.
A familiar example of this second type of transition occurs in the buoyant plume from a cigar-
ette. As the plume rises the fluid accelerates and Re becomes larger. Sooner or later an instability
sets in and the laminar plume starts to exhibit a complex, three-dimensional structure. Shortly
thereafter the plume becomes fully turbulent (Figure 1.8). Again, the transition to turbulence in
such flows can be sensitive to the level of the disturbances.
The essential point is that fluid motion is almost always inherently unstable, and that incipient
instabilities are suppressed only if the viscous dissipation is high enough. However, virtually all
fluids have an extremely low viscosity. This is true of water, air, blood, the molten metal in the
core of the earth, and the atmosphere of the sun. The implication is that turbulence is the natural
state of things and indeed this is the case. Consider, for example, Bénard convection in a layer of

Figure 1.8 Sketch of a cigarette plume.


After Corrsin (1961). Drawing by F.C. Davidson (with
permission).

common themes | 11
water one inch deep. The initial instability sets in at a temperature difference of T ~ 0.01 ◦ C
and the final transition to turbulence occurs at T ~ 0.01 ◦ C, which is not a huge temperature
difference! Alternatively, we might consider Taylor’s experiment. If R = 10 cm, the annular gap is
1 cm, and the fluid is water, then the flow will become unstable when the peripheral speed of the
inner cylinder exceeds a mere 1 cm/s. Evidently, laminar flow is the exception and not the rule in
nature and technology.
So far we have carefully avoided giving a formal definition of turbulence. In fact, it turns out
to be rather difficult, and possibly not very useful, to provide one.5 It is better simply to note
that when v is made small enough all flows develop a random, chaotic component of motion.
We group these complex flows together, call them turbulent, and then note some of their common
characteristics. They are:

(i) the velocity field fluctuates randomly in time and is highly disordered in space, exhibiting
a wide range of length scales;
(ii) the velocity field is unpredictable in the sense that a minute change to the initial
conditions will produce a large change to the subsequent motion.

To illustrate the second point, consider the following experiment. Suppose we tow an initially
stationary cylinder through a quiescent fluid at a fixed speed creating a turbulent wake. We do
the experiment one hundred times and on each occasion we measure the velocity as a function of
time at a point a fixed distance downstream of the cylinder, say x0 in a frame of reference travelling
with the cylinder. Despite the nominally identical conditions we find that the function u(x0 , t) is
quite different on each occasion. This is because there will always be some minute difference in
the way the experiment is carried out and it is in the nature of turbulence that these differences
are amplified. It is striking that, although the governing equations of incompressible flow are
quite simple (essentially Newton’s second law applied to a continuum), the exact details of u(x, t)
appear to be, to all intents and purposes, random and unpredictable.
Now suppose that we do something different. We measure u(x0 , t) for some considerable
period of time and then calculate the time average of the signal, ū(x0 ). We do this a second time
and then a third. Each time we obtain the same value for ū(x0 ), as indicated in Figure 1.9. The
same thing happens if we calculate u2 . Evidently, although u(x, t) appears to be random and
unpredictable, its statistical properties are not. This is the first hint that any theory of turbulence
has to be a statistical one, and we shall return to this point time and again in the chapters which
follow.
In summary, then, the statistical properties of a turbulent flow are uniquely determined by the
boundary conditions and the initial conditions. If a sequence of nominally identical experiments
is carried out, then the statistical properties of the flow will not change from one experiment to
the next, even though the details of the individual realizations will be different. To be more pre-
cise: the inevitable minute variations in the different realizations of the experiments produce only
minute variations in the statistics. In principle, u(t) is also uniquely determined by the boundary
conditions and initial conditions. However, in any laboratory experiment the initial conditions
can be controlled only to within a certain level of accuracy and no matter how hard we try there
will always be infinitesimal variations between experiments. It is in the nature of turbulence to
amplify these variations so that, no matter how small the perturbations in the initial conditions,

5
Nevertheless, we provide a definition in Section 2.4 of Chapter 2!

12 | the ubiquitous nature of turbulence


Another random document with
no related content on Scribd:
upon the enemy was complete.

{612}

"At 2 p. m. on this date, the 11th, the surrender of the city


was again demanded. The firing ceased and was not again
renewed. By this date the sickness in the army was increasing
very rapidly as a result of exposure in the trenches to the
intense heat of the sun and the heavy rains. Moreover, the
dews in Cuba are almost equal to rains. The weakness of the
troops was becoming so apparent I was anxious to bring the
siege to an end, but in common with most of the officers of
the army I did not think an assault would be justifiable,
especially as the enemy seemed to be acting in good faith in
their preliminary propositions to surrender. On July 11 I
wrote General Toral as follows: 'With the largely increased
forces which have come to me, and the fact that I have your
line of retreat securely in my hands, the time seems fitting
that I should again demand of your excellency the surrender of
Santiago and of your excellency's army. I am authorized to
state that should your excellency so desire the Government of
the United States will transport the entire command of your
excellency to Spain.' General Toral replied that he had
communicated my proposition to his general-in-chief, General
Blanco.

"July 12 I informed the Spanish commander that Major-General


Miles, commander-in-chief of the American Army, had just
arrived in my camp, and requested him to grant us a personal
interview on the following day. He replied he would be pleased
to meet us. The interview took place on the 13th, and I
informed him his surrender only could be considered, and that
as he was without hope of escape he had no right to continue
the fight. On the 14th another interview took place, during
which General Toral agreed to surrender, upon the basis of his
army, the Fourth Army Corps, being returned to Spain, the
capitulation embracing all of eastern Cuba east of a line
passing from Acerraderos on the south to Sagua de Tanamo on
the north, via Palma Soriano. It was agreed commissioners
should meet during the afternoon to definitely arrange the
terms. … The terms of surrender finally agreed upon included
about 12,000 Spanish troops in the city and as many more in
the surrendered district. It was arranged the formal surrender
should take place between the lines on the morning of July 17,
each army being represented by 100 armed men. At the time
appointed, I appeared at the place agreed upon with my general
officers, staff, and 100 troopers of the Second Cavalry under
Captain Brett. General Toral also arrived with a number of his
officers and 100 infantry. We met midway between the
representatives of our two armies, and the Spanish commander
formally consummated the surrender of the city and the 24,000
troops in Santiago and the surrendered district. After this
ceremony I entered the city with my staff and escort, and at
12 o'clock noon the American flag was raised over the
governor's palace with appropriate ceremonies."

Annual Reports of the War Department, 1898,


volume 2, pages 157-159.

UNITED STATES OF AMERICA: A. D. 1898


(July-August: Army administration).
Red-tape and politics.
Their working in the campaign.

"The Cuban campaign had been foreseen by intelligent officers


for more than a year, but the department which clothes the
army had taken no steps toward providing a suitable uniform
for campaigning in the tropics until war was declared. The
Fifth Army Corps, a comparatively small body of 17,000 men,
was concentrated at Tampa on the railroad within reach of all
the appliances for expediting business. Between April 26, when
war was declared, and, June 6, when the corps embarked for
Cuba, sufficient time elapsed to have clothed 1,000,000 men if
the matter had been handled in the same manner a wholesale
clothing firm would handle similar business. Yet the corps
went to Cuba wearing the winter clothing it had brought on its
backs from Montana, Wyoming, and Michigan. It endured the heat
of the tropics clad in this, and was furnished with light
summer clothing by the department to wear for its return to
Montauk, where the breezes were so bracing that the teeth
chattered even when the men were clad in winter clothing. The
only reason for this absolute failure to properly clothe the
army was that the methods of the department are too slow and
antiquated for the proper performance of business. There was
no lack of money. It was a simple case of red-tape delays.
There can be no doubt that the intention was that the summer
clothing should be worn in Cuba and that there should be warm
clothing issued at Montauk. It was issued after the troops had
shivered for days in their light clothes. The delays
unavoidably connected with an obsolete method caused great
suffering that should not have been inflicted upon men
expected to do arduous duty. A sensible man would not put a
heavy blanket on a horse to do draught work on a hot day; but
the red tape of an antiquated way of doing business caused our
soldiers to wear heavy woolen clothes in torrid heat, when every
nerve was to be strained to the breaking point in athletic
exertion. This is not pointed out in a fault-finding spirit.
The men are proud to have been in the Fifth Corps and to have
endured these things for the country and the flag; but these
unnecessary sufferings impaired the fighting strength of the
army, caused much of the sickness that visited the Fifth
Corps, and might have caused the failure of the whole
expedition. …

"The difficulty here depicted was one which beset the


department at every turn in the whole campaign. It is a
typical case. Transports, tentage, transportation—it was the
same in everything. With the most heroic exertions the
department was able to meet emergencies only after they had
passed. This was caused partly by lack of ready material, but
mainly by an inelastic system of doing business which broke
down in emergencies. This, in turn, was caused mainly by the
illiberal treatment accorded to this, as well as to every
other department of the army by Congress. It uniformly cuts
mercilessly all estimates of this, as of every other
department, and leaves no margin of expenditure or chance of
improvement. It dabbles in matters which are purely technical
and require the handling of expert executive talent. …

"Plans for war should be prepared in advance. This was


especially true of the last war, which had been foreseen for
years and considered a probability for several months. All
details should have been previously worked out, all
contingencies foreseen before hostilities began. Such plans
would require some modifications, of course, but would form a
working basis.
{613}
Neither Santiago nor Manila Bay would have been foreseen; but
any plans for war would have involved the consideration and
solution of the following problems: How to raise, arm, equip,
organize, mobilize, clothe, feed, shelter, and transport large
bodies of soldiers. The point where the battle might occur
would be a mere tactical detail to be worked out at the proper
time. The above problems could all be solved in time of peace and
should have been solved. The general staff performs this
function in foreign armies, but we had no such body in our
service and nothing to imperfectly take its place. …

"The most urgently needed reform is the absolute divorcement


of the army in all of its departments from politics. … No
department of the army should be more exempt from political
influence than the staff. This points at once to the most
urgent reform, viz., make the commanding general the real
working head of the army, instead of the Secretary of War. No
good results have come to the service by the extension of the
Secretary's powers in Grant's first administration. Most of
the evils of the service can be traced to the fact that the
general commanding has since that time been practically
deprived of his proper functions, and the real head of the
army has been a politician."

Lieutenant J. H. Parker,
Our Army Supply Department and the need of a General Staff
(Review of Reviews, December, 1898).

UNITED STATES OF AMERICA: A. D. 1898 (July-August: Cuba).


The War with Spain.
Sickness in the American army at Santiago.
Its alarming state.
Hurried removal of troops to Montauk Point, Long Island.

"After the surrender of General Toral's army General Shafter


urged the War Department from time to time to hasten the
shipment of the Spanish prisoners to their homes, in order
that the American Army, whose condition was now deplorable,
might be transported to the United States. At this time about
half the command had been attacked by malarial fever, with a
few cases of yellow fever, dysentery, and typhoid fever. The
yellow-fever cases were mainly confined to the troops at
Siboney, and the few cases found among the troops at the front
were at once transferred to that place. … There was great
fear, and excellent grounds for it, that the yellow fever, now
sporadic throughout the command, would become epidemic. With
the command weakened by malarial fevers, and its general tone
and vitality much reduced by all the circumstances incident to
the campaign, the effects of such an epidemic would
practically mean its annihilation. The first step taken to
check the spread of disease was the removal of all the troops
to new camping grounds. … It was directed that the command be
moved in this way every few days, isolating the cases of
yellow fever as they arose, and it was expected that in a
short time the yellow fever would be stamped out. … But the
effect produced on the command by the work necessary to set up
the tents and in the removal of the camps increased the number
on the sick report to an alarming degree. Convalescents from
malarial fever were taken again with the fever, and yellow
fever, dysentery, and typhoid increased. It was useless now to
attempt to confine the yellow-fever cases to Siboney, and
isolation hospitals were established around Santiago. It was
apparent that to keep moving the command every few days simply
weakened the troops and increased the fever cases. Any exertion
in this heat caused a return of the fever, and it must be
remembered that the convalescents now included about 75 per
cent. of the command. The Commanding General was now directed
to move the entire command into the mountains to the end of
the San Luis railroad, where the troops would be above the
yellow fever limit; but this was a physical impossibility. …

"The situation was desperate; the yellow-fever cases were


increasing in number, and the month of August, the period in
which it is epidemic, was at hand. It was with these
conditions staring them in the face, that the officers
commanding divisions and brigades and the Chief Surgeon were
invited by General Shafter to discuss the situation. As a
result of this conference the General sent the following
telegram giving his views [and those of the General Officers
and Medical Officers]. … 'In reply to telegram of this date
[August 3], stating that it is deemed best that my command be
moved to end of railroad, where yellow fever is impossible, I
have to say that under the circumstances this move is
practically impossible. The railroad is not yet repaired,
although it will be in about a week. Its capacity is not to
exceed 1,000 men a day, at the best, and it will take until
the end of August to make this move, even if the sick-list
should not increase. An officer of my staff, Lieutenant Miley,
who has looked over the ground, says it is not a good camping
ground. … In my opinion there is but one course to take, and
that is to immediately transport the Fifth Corps and the
detached regiments that came with it, and were sent
immediately after it, with the least delay possible, to the
United States. If this is not done I believe the death-rate
will be appalling. I am sustained in this view by every
medical officer present. I called together to-day the General
Officers and the senior Medical Officers and telegraph you
their views.' …

"On August 4th instructions were received from the War


Department to begin the removal of the command to Montauk
Point, Long Island. Some of the immune regiments were on the
way to Santiago, and other regiments were at once ordered
there to garrison the district as General Shafter's command
was withdrawn. The first of the fleet of vessels to return the
Spanish troops arrived in time to be loaded and leave August
9th, and by the end of the month nearly all were transported.

"After the surrender the relations between the American and


Spanish troops were very cordial. There could be little or no
conversation between individuals, but in many ways the respect
each had for the other was shown, and there seemed to be no
hatred on either side. Most of the Spanish officers remained
in their quarters in town, and they shared in the feeling
displayed by their men. Salutations were generally exchanged
between the officers, and American ways and manners became
very popular among the Spaniards. …
{614}
"By the 25th of the month General Shafter's entire command,
with the exception of a few organizations just ready to
embark, had departed, and, turning over the command to General
Lawton, he sailed that day with his staff on the 'Mexico,' one
of the captured transports, and at noon September 1st went
ashore at Montauk Point, Long Island."

J. D. Miley, In Cuba with Shafter,


chapter 12 (New York: Charles Scribner's Sons).

UNITED STATES OF AMERICA: A. D. 1898 (July-August: Philippines).


Correspondence between the General commanding
United States forces at Cavite and Manila,
and Aguinaldo, the Filipino leader.
On the 4th of July, General Thomas M. Anderson, then
commanding the "United States Expeditionary Forces" at Cavite
Arsenal, addressed the following communication to "Señor Don
Emilio Aguinaldo y Famy, Commanding Philippine Forces":

"General: I have the honor to inform you that the United


States of America, whose land forces I have the honor to
command in this vicinity, being at war with the Kingdom of
Spain, has entire sympathy and most friendly sentiments for
the native people of the Philippine Islands. For these reasons
I desire to have the most amicable relations with you, and to
have you and your people co-operate with us in military
operations against the Spanish forces. In our operations it
has become necessary for us to occupy the town of Cavite as a
base of operations. In doing this I do not wish to interfere
with your residence here and the exercise by yourself and
other native citizens of all functions and privileges not
inconsistent with military rule. I would be pleased to be
informed at once of any misconduct of soldiers under my
command, as it is the intention of my Government to maintain
order and to treat all citizens with justice, courtesy, and
kindness. I have therefore the honor to ask your excellency to
instruct your officials not to interfere with my officers in
the performance of their duties and not to assume that they
can not visit Cavite without permission."

On the following day Aguinaldo replied:

"General: Interpreting the sentiments of the Philippine


people, I have the honor to express to your excellency my most
profound gratefulness for the sympathy and amicable sentiments
with which the natives of these islands inspire the great North
American nation and your excellency. I also thank most
profoundly your desire of having friendly relations with us,
and of treating us with justice, courtesy, and kindness, which
is also our constant wish to prove the same, and special
satisfaction whenever occasion represents. I have already
ordered my people not to interfere in the least with your
officers and men, orders which I shall reiterate to prevent
their being unfulfilled; hoping that you will inform me of
whatever misconduct that may be done by those in my command,
so as to reprimand them and correspond with your wishes." …

To this communication General Anderson returned the following


on the 6th: "General: I am encouraged by the friendly
sentiment expressed by your excellency in your welcome letter
received on the 5th instant to endeavor to come to a definite
understanding, which I hope will be advantageous to both. Very
soon we expect a large addition to our forces, and it must be
apparent to you as a military officer that we will require
much more room to camp our soldiers, and also storeroom for
our supplies. For this I would like to have your excellency's
advice and co-operation, as you are best acquainted with the
resources of this country. It must be apparent to you that we
do not intend to remain here inactive, but to move promptly
against our common enemy. But for a short time we must
organize and land supplies, and also retain a place for
storing them near our fleet and transports. I am solicitous to
avoid any conflict of authority which may result from having
two sets of military officers exercising command in the same
place. I am also anxious to avoid sickness by taking sanitary
precaution. Your own medical officers have been making
voluntary inspections with mine, and fear epidemic diseases if
the vicinity is not made clean. Would it not be well to have
prisoners work to this end under the advice of the surgeons?"

On the 9th of July General Anderson reported to the War


Department at Washington: "General Aguinaldo tells me he has
about 15,000 fighting men, but only 11,000 armed with guns,
which mostly were taken from the Spaniards. He claims to have
in all 4,000 prisoners. When we first landed he seemed very
suspicious, and not at all friendly, but I have now come to a
better understanding with him and he is much more friendly and
seems willing to co-operate. But he has declared himself
dictator and president, and is trying to take Manila without
our assistance. This is not probable, but if he can effect his
purpose he will, I apprehend, antagonize any attempt on our
part to establish a provisional government."

On the 17th the American commander caused another


communication to be addressed to "General Emilio Aguinaldo" as
follows: "Sir: General Anderson wishes me to say that, the
second expedition having arrived, he expects to encamp in the
vicinity of Paranaque from 5,000 to 7,000 men. To do this,
supply this army and shelter, will require certain assistance
from the Filipinos in this neighborhood. We will want horses,
buffaloes, carts, etc., for transportation, bamboo for
shelter, wood to cook with, etc. For all this we are willing
to pay a fair price, but no more. We find so far that the
native population are not willing to give us this assistance
as promptly as required. But we must have it, and if it
becomes necessary we will be compelled to send out parties to
seize what we may need. We would regret very much to do this,
as we are here to befriend the Filipinos. Our nation has spent
millions of money to send forces here to expel the Spaniards
and to give good government to the whole people, and the
return we are asking is comparatively slight. General Anderson
wishes you to inform your people that we are here for their
good, and that they must supply us with labor and material at
the current market prices. We are prepared to purchase 500
horses at a fair price, but can not undertake to bargain for
horses with each individual owner. I regret very much that I
am unable to see you personally, as it is of the utmost
importance that these arrangements should be made as soon as
possible."

To this communication there seems to have been no written


reply until the 24th; and, on the 20th, the Chief
Quartermaster reported to General Anderson "that it is
impossible to procure transportation except upon Señor
Aguinaldo's order, in this section, who has an inventory of
everything. The natives have removed their wheels and hid
them." On the 23d General Anderson repeated his request, as
follows:

{615}

"General: When I came here three weeks ago I requested your


excellency to give what assistance you could to procure means
of transportation for the American Army, as it was to fight
the cause of your people. So far we have received no response.
As you represent your people, I now have the honor to make
requisition on you for 500 horses and 50 oxen and ox carts. If
you can not secure these, I will have to pass you and make
requisition directly on the people. I beg leave to request an
answer at your earliest convenience."

The next day Aguinaldo replied: "I have the honor to manifest
to your excellency that I am surprised beyond measure at that
which you say to me in it, lamenting the nonreceipt of any
response relative to the needs (or aids) that you have asked
of me in the way of horses, buffaloes, and carts, because I
replied in a precise manner, through the bearer, that I was
disposed to give convenient orders whenever you advised me of
the number of these with due anticipation (notice). I have
circulated orders in the provinces in the proximity that in
the shortest time possible horses be brought for sale, but I
cannot assure your excellency that we have the number of 500
that is needed, because horses are not abundant in these
vicinities, owing to deaths caused by epizootic diseases in
January and March last. Whenever we have them united (or
collected), I shall have the pleasure to advise your
excellency. I have also ordered to be placed at my disposal 50
carts that I shall place at your disposition whenever
necessary, always (premising) that you afford me a previous
advice of four days in anticipation."
Meantime, General Anderson had written to the War Department,
on the 18th: "Since reading the President's instructions to
General Merritt, I think I should state to you that the
establishment of a provisional government on our part will
probably bring us in conflict with insurgents, now in active
hostility to Spain. The insurgent chief, Aguinaldo, has
declared himself dictator and self-appointed president. He has
declared martial law and promulgated a minute method of rule and
administration under it. We have observed all official
military courtesies, and he and his followers express great
admiration and gratitude to the great American Republic of the
north, yet in many ways they obstruct our purposes and are
using every effort to take Manila without us. I suspect also
that Aguinaldo is secretly negotiating with the Spanish
authorities, as his confidential aid is in Manila. The city is
strongly fortified and hard to approach in the rainy season.
If a bombardment fails we should have the best engineering
ability here." And, again on the 21st, he had written: "Since
I wrote last, Aguinaldo has put in operation an elaborate
system of military government, under his assumed authority as
dictator, and has prohibited any supplies being given us,
except by his order. As to this last I have written to him
that our requisitions on the country for horses, ox carts,
fuel and bamboo (to make scaling ladders) must be filled, and
that he must aid in having them filled. His assumption of
civil authority I have ignored, and let him know verbally that
I could, and would, not recognize it, while I did not
recognize him as a military leader. It may seem strange that I
have made no formal protest against his proclamation as
dictator, his declaration of martial law, and publication and
execution of a despotic form of government. I wrote such a
protest, but did not publish it, at Admiral Dewey's request,
and also for fear of wounding the susceptibilities of
Major-General Merritt, but I have let it be known in every
other way that we do not recognize the dictatorship. These
people only respect force and firmness. I submit, with all
deference, that we have heretofore underrated the natives.
They are not ignorant, savage tribes, but have a civilization
of their own; and although insignificant in appearance, are
fierce fighters, and for a tropical people they are
industrious. A small detail of natives will do more work in a
given time than a regiment of volunteers."

On the 24th General Anderson received from the Philippine


leader a very clear and definite statement of his attitude
towards the "Expeditionary Forces of the United States," and
the intentions with which he and the people whom he
represented were acting. "I came," he wrote, "from Hongkong to
prevent my countrymen from making common cause with the
Spanish against the North Americans, pledging before my word
to Admiral Dewey to not give place [to allow] to any internal
discord, because, [being] a judge of their desires, I had the
strong convictions that I would succeed in both objects,
establishing a government according to their desires. Thus it
is that in the beginning I proclaimed the dictatorship, and
afterwards, when some of the provinces had already liberated
themselves from Spanish domination, I established a
revolutionary government that to-day exists, giving it a
democratic and popular character as far as the abnormal
circumstances of war permitted, in order that they [the
provinces] might be justly represented, and administered to
their satisfaction. It is true that my government has not been
acknowledged by any of the foreign powers, but we expected
that the great North American nation, which struggled first
for its independence, and afterwards for the abolition of
slavery, and is now actually struggling for the independence
of Cuba, would look upon it with greater benevolence than any
other nation. Because of this we have always acknowledged the
right of preference to our gratitude.

"Debtor to the generosity of the North Americans, and to the


favors we have received through Admiral Dewey, and [being]
more desirous than any other person of preventing any conflict
which would have as a result foreign intervention, which must be
extremely prejudicial, not alone to my nation but also to that
of your excellency, I consider it my duty to advise you of the
undesirability of disembarking North American troops in the
places conquered by the Filipinos from the Spanish, without
previous notice to this government, because as no formal
agreement yet exists between the two nations the Philippine
people might consider the occupation of its territories by
North American troops as a violation of its rights.

"I comprehend that without the destruction of the Spanish


squadron the Philippine revolution would not have advanced so
rapidly. Because of this I take the liberty of indicating to
your excellency the necessity that, before disembarking, you
should communicate in writing to this government the places
that are to be occupied and also the object of the occupation,
that the people may be advised in due form and [thus] prevent
the commission of any transgression against friendship.
{616}
I can answer for my people, because they have given me evident
proofs of their absolute confidence in my government, but I
can not answer for that which another nation whose friendship
is not well guaranteed might inspire in it [the people]; and
it is certain that I do this not as a menace, but as a further
proof of the true and sincere friendship which I have always
professed for the North American people, in the complete
security that it will find itself completely identified with
our cause of liberty."

In the same strain, on the 1st of August, Aguinaldo wrote to


United States Consul Williams, as to a "distinguished friend:"

"I have said always, and I now repeat, that we recognize the
right of the North Americans to our gratitude, for we do not
forget for a moment the favors which we have received and are
now receiving; but however great those favors may be, it is
not possible for me to remove the distrust of my compatriots.
These say that if the object of the United States is to annex
these islands, why not recognize the government established in
them, in order in that manner to join with it the same as by
annexation? Why do not the American generals operate in
conjunction with the Filipino generals and, uniting the
forces, render the end more decisive? Is it intended, indeed,
to carry out annexation against the wish of these people,
distorting the legal sense of that word? If the revolutionary
government is the genuine representative by right and deed of
the Filipino people, as we have proved when necessary, why is
it wished to oppress instead of gaining their confidence and
friendship?

"It is useless for me to represent to my compatriots the


favors received through Admiral Dewey, for they assert that up
to the present the American forces have shown not an active,
only a passive, co-operation, from which they suppose that the
intentions of these forces are not for the best. They assert,
besides, that it is possible to suppose that I was brought
from Hongkong to assure those forces by my presence that the
Filipinos would not make common cause with the Spaniards, and
that they have delivered to the Filipinos the arms abandoned
by the former in the Cavite Arsenal, in order to save
themselves much labor, fatigue, blood, and treasure that a war
with Spain would cost. But I do not believe these unworthy
suspicions. I have full confidence in the generosity and
philanthropy which shine in characters of gold in the history
of the privileged people of the United States, and for that
reason, invoking the friendship which you profess for me and
the love which you have for my people, I pray you earnestly,
as also the distinguished generals who represent your country
in these islands, that you entreat the Government at
Washington to recognize the revolutionary government of the
Filipinos, and I, for my part, will labor with all my power
with my people that the United States shall not repent their
sentiments of humanity in coming to the aid of an oppressed
people.

"Say to the Government at Washington that the Filipino people


abominate savagery; that in the midst of their past
misfortunes they have learned to love liberty, order, justice,
and civil life, and that they are not able to lay aside their
own wishes when their future lot and history are under
discussion. Say also that I and my leaders know what we owe to
our unfortunate country; that we know how to admire and are
ready to imitate the disinterestedness, the abnegation, and
the patriotism of the grand men of America, among whom stands
pre-eminent the immortal General Washington."

United States, 56th Congress, 1st Session,


Senate Document Number 208.

In an article published in the "North American Review,"


February, 1900, General Anderson discussed his relations with
Aguinaldo very frankly, in part as follows: "On the 1st of
July, 1898, I called on Aguinaldo with Admiral Dewey. He asked
me at once whether 'the United States of the North' either had
recognized or would recognize his government—I am not quite
sure as to the form of his question, whether it was 'had' or
'would.' In either form it was embarrassing. My orders were,
in substance, to effect a landing, establish a base, not to go
beyond the zone of naval co-operation, to consult Admiral
Dewey and to wait for Merritt. Aguinaldo had proclaimed his
government only a few days before (June 28), and Admiral Dewey
had no instructions as to that assumption. The facts as to the
situation at that time I believe to be these: Consul Williams
states in one of his letters to the State Department that
several thousand Tagals were in open insurrection before our
declaration of war with Spain. I do not know as to the number,
yet I believe the statement has foundation in fact. Whether
Admiral Dewey and Consuls Pratt, Wildman and Williams did or
did not give Aguinaldo assurances that a Filipino government
would be recognized, the Filipinos certainly thought so,
probably inferring this from their acts rather than from their
statements. If an incipient rebellion was already in progress,
what could be inferred from the fact that Aguinaldo and
thirteen other banished Tagals were brought down on a naval
vessel and landed in Cavite? Admiral Dewey gave them arms and
ammunition, as I did subsequently, at his request. They were
permitted to gather up a lot of arms which the Spaniards had
thrown into the bay; and, with the four thousand rifles taken
from Spanish prisoners and two thousand purchased in Hong
Kong, they proceeded to organize three brigades and also to
arm a small steamer they had captured. I was the first to tell
Admiral Dewey that there was any disposition on the part of
the American people to hold the Philippines, if they were
captured. The current of opinion was setting that way when the
first expeditionary force left San Francisco, but this the
Admiral had had no reason to surmise.

"But to return to our interview with Aguinaldo. I told him I


was acting only in a military capacity; that I had no
authority to recognize his government; that we had come to
whip the Spaniards, and that, if we were successful, the
indirect effect would be to free them from Spanish tyranny. I
added that, as we were fighting a common enemy, I hoped we
would get along amicably together. He did not seem pleased
with this answer. The fact is, he hoped and expected to take
Manila with Admiral Dewey's assistance, and he was bitterly
disappointed when our soldiers landed at Cavite. … A few days
thereafter, he made an official call, coming with cabinet and
staff and a band of music. On that occasion he handed me an
elaborate schedule for an autonomous government which he had
received from some Filipinos in Manila, with a statement that
they had reason to believe that Spain would grant them such a
form of government.
{617}
With this was an open letter addressed to the Filipino people
from Pedro Alexandre Paterno, advising them to put their trust
in Spain rather than America. The day before, two German
officers had called on Aguinaldo and I believed they had
brought him these papers. I asked him if the scheme was
agreeable to him. He did not answer, but asked if we, the
North Americans, as he called us, intended to hold the
Philippines as dependencies. I said I could not answer that,
but that in one hundred and twenty years we had established no
colonies. He then made this remarkable statement: 'I have studied
attentively the Constitution of the United States, and I find
in it no authority for colonies and I have no fear.' It may
seem that my answer was somewhat evasive, but I was at the
time trying to contract with the Filipinos for horses, carts,
fuel and forage. …

"The origin of our controversies and conflicts with the


Filipinos can … be traced back to our refusal to recognize the
political authority of Aguinaldo. Our first serious break with
them arose from our refusal to let them co-operate with us.
About nine o'clock on the evening of August 12, I received
from General Merritt an order to notify Aguinaldo to forbid
the Filipino insurgents under his command from entering
Manila. This notification was delivered to him at twenty
minutes past ten that night. The Filipinos had made every
preparation to assail the Spanish lines in their front.
Certainly, they would not have given up part of their line to
us unless they thought they were to fight with us. They,
therefore, received General Merritt's interdict with anger and
indignation. They considered the war as their war, and Manila
as their capital, and Luzon as their country. … At seven
o'clock I received an order from General Merritt to remove the
Filipinos from the city. … I therefore took the responsibility
of telegraphing Aguinaldo, who was at Bacoor, ten miles below,
requesting him to withdraw his troops and intimating that
serious consequences would follow if he did not do so. I
received his answer at eleven, saying that a Commission would
come to me the next morning with full powers. Accordingly the
next day Señors Buencomeno, Lagarde, Araneto and Sandeco came
to Division Headquarters in Manila and stated that they were
authorized to order the withdrawal of their troops, if we
would promise to reinstate them in their present positions on
our making peace with Spain. Thereupon I took them over to
General Merritt. Upon their repeating their demands, he told
them he could not give such a pledge, but that they could rely
on the honor of the American people. The General then read to
them the proclamation he intended to issue to the Filipino
people. …

"There is a great diversity of opinion as to whether a


conflict with the Filipinos could not have been avoided if a
more conciliatory course had been followed in dealing with
them. I believe we came to a parting of the ways when we
refused their request to leave their military force in a good
strategic position on the contingency of our making peace with
Spain without a guarantee of their independence."

T. M. Anderson,
Our Rule in the Philippines
(North American Review, volume 170, page 275).

UNITED STATES OF AMERICA: A. D. 1898 (July-August: Porto Rico).


Occupation of Porto Rico.

"With the fall of Santiago the occupation of Porto Rico became


the next strategic necessity. General Miles had previously been
assigned to organize an expedition for that purpose.
Fortunately, he was already at Santiago, where he had arrived
on the 11th of July with reinforcements for General Shafter's
army. With these troops, consisting of 3,415 infantry and
artillery, 2 companies of engineers and 1 company of the
signal corps, General Miles left Guantanamo on July 21st,
having 9 transports, convoyed by the fleet, under Captain
Higginson, with the 'Massachusetts' (flagship), 'Dixie,',
Gloucester,' 'Columbia' and 'Yale,' the two latter carrying
troops. The expedition landed at on July 25th, which port
was entered with little opposition. Here the fleet was joined
by the 'Annapolis' and the 'Wasp,' while the 'Puritan' and
'Amphitrite' went to San Juan and joined the 'New Orleans,'
which was engaged in blockading that port. The major general
commanding was subsequently reinforced by General Schwan's
brigade of the Third Army Corps, by General Wilson with a part
of his division and also by General Brooke with a part of his
corps, numbering in all 16,973 officers and men. On July 27th
he entered Ponce, one of the most important ports in the
island, from which he thereafter directed operations for the
capture of the island. With the exception of encounters with
the enemy at Guayama, Hormigueros [the Rio Prieto], Coamo, and
Yauco and an attack on a force landed at Cape San Juan, there
was no serious resistance. The campaign was prosecuted with
great vigor and by the 12th of August much of the island was
in our possession and the acquisition of the remainder was
only a matter of a short time. At most of the points in the
island our troops were enthusiastically welcomed.
Protestations of loyalty to the flag and gratitude for
delivery from Spanish rule met our commanders at every stage."

Message of the President of the United States


to Congress, December 5, 1898.

"During the nineteen days of active campaign on the island of


Puerto Rico a large portion of the island was captured by the
United States forces and brought under our control. Our forces
were in such a position as to make the positions of the
Spanish forces, outside of the garrison at San Juan, utterly
untenable. The Spaniards had been defeated or captured in the
six different engagements which took place, and in every
position they had occupied up to that time. The volunteers had
deserted their colors, and many of them had surrendered to our
forces and taken the oath of allegiance. This had a
demoralizing effect upon the regular Spanish troops. … The
loss of the enemy in killed, wounded and captured was nearly
ten times our own, which was only 3 killed and 40 wounded."

You might also like