You are on page 1of 296

Rough

Surfaces
Second Edition
Rough
Surfaces
Second Edition

Tom R.Thomas
Production Engineering Department,
Chalmers University of Technology, Sweden

Imperial College Press


Published by
Imperial College Press
203 Electrical Engineering Building
Imperial College
London SW7 2BT

Distributed by
World Scientific Publishing Co. Re. Ltd.
P 0 Box 128, Farrer Road, Singapore 912805
USA office: Suite lB, 1060 Main Street, River Edge, NJ 07661
U K ofice: 57 Shelton Street, Covent Garden, London WC2H 9HE

British Library Cataloguing-in-Publication Data


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

First edition published in 1982 by Longman Group UK Limited

ROUGH SURFACES, Second Edition


Copyright 0 1999 by Imperial College Press
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.

ISBN 1-86094-100-1

Printed in Singapore by Eurasia Press Pte Ltd


For Ann
CONTENTS

PREFACE xi
...
ACKNOWLEDGEMENTS Xlll

1. INTRODUCTION
1.1. Surface Roughness
1.1.1. What Causes Roughness?
1.1.2. Why Is Roughness Important?
1.2. Principles of Roughness Measurement
1.2.1. Range and Resolution
1.3. References

2. STYLUS INSTRUMENTS 11
2.1. Mechanical Instruments 11
2.2. Electrical Instruments 13
2.2.1. Stylus and Skid 15
2.2.2. Transducers 16
2.2.3. Pickup 18
2.2.4. Output Recording 19
2.3. Sources of Error 20
2.3. I . Effect of Stylus Size 20
2.3.2. Effect of Stylus Load 23
2.3.3. Other Sources of Error 25
2.4. Calibration and Standards 28
2.5. References 29

3. OPTICAL INSTRUMENTS 35
3.1. Profiling Techniques 36
3.1.1. Optical Sections 36
3.1.2. Optical Probes 37
3.1.3. Interferometers 44
3.2. Parametric Techniques 46
3.2.1. Specular Reflectance 47
3.2.2. Total Integrated Scatter 49

vii
viii Rough Surfaces

3.2.3. Angular Distributions 50


3.2.4. Direct Fourier Transformation 52
3.2.5. Ellipsorvietry 52
3.2.6. Speckle 54
3.3. References 56

4. OTHER MEASUREMENT TECHNIQUES 63


4.1. Profiling Methods 63
4.1.1. Taper Sectioning 63
4.1.2. Electron Microscopy 64
4.1.3. Capacitance 66
4.1.4. Scanning Microscopies 68
4.2. Parametric Methods 71
4.2.1. Mechanical Methods 71
4.2.2. Electrical Methods 77
4.2.3. Fluid Methods 80
4.2.4. Acoustic Methods 83
4.3. References 84

5. OTHER MEASUREMENT TOPICS 91


5.1. 3D Measurement 91
5.2. Relocation 95
5.3. Replication 97
5.4. In-Process Measurement 100
5.4.1. Optical Techniques 102
5.5. References 106

6. DATA ACQUISITION AND FILTERING 113


6.1. Data Acquisition 113
6.2. Filtering 115
6.2.1. Envelope Filters 125
6.3. References 130

7. AMPLITUDE PARAMETERS 133


7.1. Extreme-Value Parameters 134
7.2. Average Parameters 138
7.3. The Height Distribution 139
Contents ix

7.4, Bearing Area 144


7.5, References 147

8. TEXTURE PARAMETERS 151


8.1, Random Processes 152
8.2, The Profile as a Random Process 157
8.3, Practical Computation 159
8.4. Fractal Roughness 162
8.5. References 168

9. SURFACES IN THREE DIMENSIONS 171


9.1, Filtering 172
9.2, Parameters 173
9.3, Random Processes in Three Dimensions 177
9.4, The Surface as a Random Process 180
9.5, Practical Computation 185
9.6. Anisotropy 188
9.7. References 195

10. APPLICATIONS: CONTACT MECHANICS 199


10.1. The Contact of Rough Surfaces 201
10.2. Rough Contact Mechanics 205
10.2.1. Contact of Curved Sulfaces 212
10.2.2. Joint Stiffness 214
10.3. The Plasticity Index 2 15
10.4. References 220

11. TRIBOLOGY 225


11.1. Friction 225
11.2. Lubrication 227
11.3. Wear 229
11.4. Seals 235
11.5. References 238

12. SOME OTHER APPLICATIONS 247


12.1. Contact Resistance 247
12.2. Noise and Vibration 250
X Rough Su$aces

12.3. Fluid Flow 25 1


12.4. Dimension and Tolerance 255
12.5. Abrasive Machining 257
12.6. Bioengineering 259
12.7. Geomorphometry 26 1
12.8. References 263

INDEX 269
PREFACE

This book is intended for graduate scientists and engineers who need to know
somethmg more about roughness, how to measure and describe it and what
practical problems it might cause them. It assumes a general familiarity with
scientific and engineering terms and concepts and a mathematical level nowhere
above that of a final-year engineering course. For a non-mathematical introduction
to the subject at an undergraduate level, the reader is commended to the book by
Mummery cited extensively in the text. A more comprehensive and rigorously
mathematical account, to which the reader will also often be referred in these
pages, is that of Professor Whitehouse.
The first edition of Rough Surfaces, the first comprehensive monograph on
the subject in English, was published in 1982 as a multi-author work. Several of
the original authors have since enjoyed professional careers of some distinction,
reflecting the increased importance of the subject since that time. Advances in the
intervening period have required the addition of much new material and the
updating of most of the original work, so that the present book is almost entirely a
new production. Some of the new material is based on lectures which I have given
at Chalmers University over the last few years.
The book treats roughness primarily as an engineering phenomenon,
reflecting its author’s interests and background in tribology and production
engineering. I am very conscious, however, of the scientific and technical
communities, of hydrodynamicists, geographers, optical engineers and many
others, for whom roughness is equally important, and I have tried to keep the
discussion as general as possible to reflect their needs and preoccupations. This is
considerably helped by the conceptual unity of the subject; techniques of
characterisation used successfully by the atomic-force microscopist can often be
applied virtually without change to the problems of the geomorphologist, and vice
versa. For those with more specialist interests, there are monographs (on
scattering, for instance, by Ogilvy, and Bennett and Mattsson) which will be
referred to where appropriate in the text. The subject naturally divides into three
parts: measurement, characterisation and applications, and this division will be
followed in the structure of the book.
A large number of people have contributed to this book in various ways and I
thank them all for their help and for their encouragement over what must have

Xi
xii Rough Surfaces

seemed a very long period. I am grateful to my colleagues at Chalmers University


for many valuable discussions and other lundnesses, and particularly to Bengt-
Goran Rosen and Robert Ohlsson for their fruitful collaboration, and to Professor
Ralph Crafoord for extending to me the hospitality and facilities of the Production
Engineering Department. I am also grateful to the University itself for the award
of a Jubilee Professorship during 1997/98 which considerably aided the writing of
this book.
ACKNOWLEDGEMENTS

I am grateful for permission to use copyright material from the following copyright
holders:

American Institute of Aeronautics and Astronautics for my Figs. 8.8 and 12.2
from Thomas & Sayles 1975, Prog. Astronaut. Aeronaut. 29, 3-20 Figs. 5 & 8;
American Physical Society for my Fig. 4.5 from Hansma & Tersoff 1987, J. Appl.
Phys. 61, Rl-R23 Fig. 1, my Fig. 4.6 from Alexander et al. 1989, J. Appl. Phys.
65, 164 Fig. 1, my Fig. 10.6 from Archard 1961, J. Appl. Phys., 32, 1420-1425
Fig. 2; American Society of Mechanical Engineers for my Fig. 6.14 from Olsen
1963, Proc. Int. Prod. Eng. Res .Con$, Pittsburgh, 8, 655-658 Fig. 3, my Figs.
7.11, 9.10 and 9.11 from Sayles & Thomas 1979, J Lubr Techno1 101, 409-417
Figs. 2, 11, 12, my Fig. 7.16 from Williamson et al. 1969 in: Surface mechanics.
Ling, F.F. (ed.), 24-35 Fig. 2, my Figs. 9.8 and 9.9 from Nayak 1971, J. Lubr.
Tech., 93, 398-407 Figs. 4-6, 10, my Fig. 10.3 from Majumdar & Bhushan 1990,
Journal of Tribology 112, 205-216 Fig. 11, my Fig. 10.10 from Greenwood &
Tripp 1967, J. Appl. Mech. 34, 153-159 Fig. 5, my Fig. 12.9 from Thomas et al.
1980, J. Biomech. Eng., 102, 50-57 Fig. 3; American Society for Testing
Materials for my Fig. 4.10 from Doty 1975, 42-61 Fig. 1, and my Fig. 4.13 from
Henry & Hegmon 1975, 3-17 Fig. 1, both in Surface Texture versus Skidding:
Measurements, Frictional Aspects and Safety Features of Tire-pavement Interactions,
STP 583; ARRB Transport Research Ltd. for my Fig. 12.10 from Potter et al.
1992, Road & Transport Research 1, 6-27 Fig. 2; British Hydromechanic Research
Association for my Fig. 11.9 from Thomas et al. 1975, Proc. 7th. Int. Con$ on
Fluid Sealing, Paper J32, my Fig. 12.3 from Thomas & Olszowski 1974, Proc.
6th. Int Gas Bearing Symp. D6, 73-92 Fig. 5; British Standards Institution for my
Fig. 6.6 from BS1134 Part 1 1988 Fig. 20, my Fig. 6.9 from BS1134 Part 2 1972
Fig. 4; Cassell plc, London, for my Fig. 2.1 from Galyer & Shotbolt 1990,
Metrologyfor engineers 5e Fig. 9.3; Chalmers University, Goteborg, for my Figs.
2.6 and 5.2 from Desages & Michel 1993 Figs. 2.9, 3.5, 3.6, 3.8; Elsevier Science,
Oxford, for my Fig. 1.5 from Stedman 1987, Prec. Engng., 9, 149-152 Fig. 2, my
Fig. 3.16 from Vorburger & Teague 1981, Precis. Eng. 3,61-83 Fig. 19, my Fig.
7.1 from Thomas & Charlton 1981, Precis. Eng. 3, 91-96 Fig. 3, my Fig. 10.2
from Sayles & Thomas 1976, Appl. Energy, 2 , 249-267 Fig. 1, my Figs. 2.7 and
2.8 from Radhakrishnan 1970, Wear, 16, 325-335 Figs. 1 & 9, my Figs. 5.4 &
xiii
xiv Rough Surfaces

11.6 from Thomas 1972, Wear, 22, 83-90 Figs. 2 & 4, my Fig. 5.6 from George
1979, Wear, 57, 51-61 Fig. 4, my Figs. 5.8 & 5.9 from Clarke & Thomas 1979,
Wear, 57, 107-116 Figs. 2, 4, 5, my Fig. 6.4(b) from Thomas 1975, Wear, 3 3 ,
205-233 Fig. 1, my Fig. 6.16 from Fahl 1982, Wear 83, 165-179 Fig. 3, my Fig.
6.17 from Shunmugam 1987, Wear 117, 335-345 Fig. 3c, my Fig. 8.6b from
Thomas & Sayles 1978, Tribology International, 11, 163-168 Fig. 2, my Fig. 8.7
from Thwaite 1978, Wear, 51, 253-267 Fig. 3, my Fig. 10.8 from So & Liu 1991,
Wear 146, 201-218 Fig. 8, my Fig. 10.9 from Woo & Thomas 1979, Wear, 5 8 ,
331-340 Figs. 1 & 2, my Fig. 10.11 from Wu & Zheng 1988, Wear, 121, 161-172
Fig.1, my Fig. 11.1 from Koura & Omar 1981, Wear, 73, 235-246 Fig. 11, my
Fig. 11.2 from Ogilvy 1993, Wear 160, 171-180 Fig. 6, my Fig. 11.8 from
Golden 1976, Wear, 42, 157-162 Fig. 3, my Fig. 3.6 from Brown 1995, Int. J.
Mach, Tool Manufact. 35, 135-139 Fig. 1, my Fig. 4.15 from Wager 1967, Int. J.
Mach. Tool Des. Res., 7, 1-14 Fig. 5, my Fig. 12.7 from Sayles & Thomas 1976,
Int. J. Prod. Res 14, 641-655 Figs. 3 & 6, my Figs. 1.3 & 1.6 from Thomas
1998, Int. J. Mach. Tool Manufact. 38, 405-41 1 Figs. 1 & 2, my Fig. 9.16 from
Zahouani 1998, Int. J. Mach. Tool Manufact. 38, Fig. 11, my Figs. 8.13 & 10.13
from Rostn et al. 1998, Int. J. Mach. Tool Manufact. 38, Figs. 2 & 3, reprinted
with permission; the European Commission for my Table 9.1 from Stout et al.
1993, The development of methods f o r the characterisation of roughness in 3
dimensions, EUR 15178 EN Fig. 12.22; Feinpriif Perthen GmbH, Gottingen, for
my Figs. 7.2, 7.3 and 8.2 from Sander 1991, A practical guide to the assessment of
surfice texture Figs. 12a, 13, 27; Hallwag Verlags GmbH, Ostfildern, for my Fig.
3.15 from Lonardo 1978, Ann. CIRP27, 531-533 Fig. 5, my Fig. 4.7 from Goch
& Volk 1994, CIRP Ann. 43, 487-490 Fig. 6; Hommelwerke GmbH,
Schwenningen, for my Figs. 6.11, 6.12, 6.15, 7.4, 7.7, 7.8, 7.12-7.15 from
Mummery 1990, Sugace texture analysis: the handbook, Figs. 2.10, 2.1 1, 3.2, 3.3,
3.4, 3.6, 3.13, 3.17, 3.18, 3.19, 3.25, 3.26, 3.29, 3.30; I. F. S. (Publications)
Ltd., Bedford, for my Fig. 5.7 from Dutschke & Eissler 1978, Proc. 3rd. Con$ on
Automated Inspection & Product Control 19-30 Fig. 1; Indian Institute of
Technology, Madras, for my Fig. 4.14(b) from Radhakrishnan & Sagar 1970, Proc.
4th. All-India Machine Tool Design & Research Con$ Fig. 1; Industrial Press Inc.,
New York, for my Figs. 2.4, 3.2 and 4.14 from Farago 1982, Handbook of
dimensional measurement 2e, Fig. 6.1, Tables 15.2 & 15.4, used with permission;
IOP Publishing Ltd., Bristol and the authors for my Figs. 3.3, 3.8 and 3.9 from
Whitehouse 1994, Handbook of surfme metrology, Figs. 4.112 & 4.1 17, my Fig.
4.3 from Bugg & King 1988, J. Phys. E: Sci. Instrum., 21, 147-151 Fig. 2, my
Fig. 4.9 from Powell 1957, J . S c i . Instrum., 34, 485-492 Fig. 1; Institution of
Acknowledgements xv

Electrical Engineers for my Figs. 2.9 and 2.10 from Reason 1944, J. Inst. P r d .
Engrs., 23, 347-372 Figs. 4 & 16; Institution of Mechanical Engineers for my Fig.
2.2 from Lackenby 1962, Proc. I. Mech. E., 176, 981-1014 Fig. 1, my Fig. 3.1
from Keller 1967/8, Proc. I. Mech. E., 182, Part 3K, 360-367 Fig. 3.5, my Fig.
3.1 1 from Westberg 1967/68, Proc .I .Mech .E., 182, Part 3K, 260-273 Fig. 25.1,
my Fig. 7.5 from Hydell 1967/8, Proc. I. Mech. E., 182, Part 3K, 127-134 Fig.
15.7, my Fig. 8.4 from Peklenik 1967/68, Proc. I. Mech. E., 182, Part 3K, 108-
126 Fig. 24.18, my Fig. 11.5 from Leaver et al. 1974, Proc. 1. Mech. E., 1 8 8 ,
461-469 Fig. 1, my Fig. 11.7 from Thomas 1978, Proc. 4th Leeds-Lyon S y m p . ,
99-108 Fig. 5, by permission of the Council of the Institution;
International Business Machines for my Fig. 4.4 from Binnig & Rohrer 1986, ZBM
Journal of Research and Development 30; 355-369 Fig.1; Japanese Society of
Precision Engineers for my Fig. 2.12 from Nara 1966, Bull. Jap. SOC. Precision
Engng., 1, 263-273 Figs. 5 & 8; Kluwer Academic Publishers, Dordrecht, with
kind permission for my Fig. 2.14 from Song 1988, Sulface Topography, 1, 29-40
Fig. 2, my Fig. 3.4 from Bristow 1988, Sulface Topography, 1, 281-285 Fig. 1,
my Fig. 3.5 from Sayles et al. 1988, Sulface Topography, 1, 219-227 Fig. 1, my
Figs. 4.1 and 4.2 from Garbini et al. 1988, Surface Topography, 1, 131-142 Figs.
1 & 6, my Figs. 4.1 1 and 4.12 from Lieberman et al. 1988, Sulface Topography,
1, 115-130 Figs. 3 & 6 , my Fig. 11.4 from Chandrasekaran 1993, J. Mat. Sci.
Lett. 12, 952-954 Fig. 5; Lasertec Corporation for my Fig. 3.7; Macmillan
Magazines Ltd., Basingstoke, and the authors, for my Fig. 8.9 from Sayles &
Thomas 1978, Nature 271, 431-434 Fig. 2, reprinted with permission; Macmillan
Press, Basingstoke, and the authors, for my Fig. 2.11 from Agullo & Pages-Fita
1974, Proc. 15th. Int. Machine Tool Des. & Res. Conf., 349-362 Fig. 9, my Fig.
12.6 from Thomas 1973, Proc. 13th Int. Machine Tool Des. & Res. Conf., 303-
308 Fig. 3b, reprinted with permission; the McGraw-Hill Companies, New York,
for my Fig. 5.1 from Terman 1937, Radio Engineering 2e, Fig. 430, my Fig. 12.4
from Hunsaker & Rightmire 1947, Engineering applications offluid mechanics Fig.
43; McGraw-Hill Publishing Co., Maidenhead, for my Fig. 6.5 from Golten 1997,
Understanding signals and systems Fig. 7.8; Optical Society of America for my Fig.
3.10 from Creath 1987, Applied Optics 26, 2810-2815 Fig. 6, my Fig. 3.12 from
Birkebak 1971, Appl. Opt. 10, 1970-1979 Fig. 1, my Fig. 3.17 from Fujii & Lit
1978, Appl. Opt. 17, 2690-2705 Fig. 1, my Fig. 3.13 from Bennett & Mattsson
1989, Introduction to sulface roughness and scattering Fig. 17; Plenum Publishing
Corporation, New York, and the authors, for my Fig. 10.4 from Russ 1994, Fractal
surfaces, p. 67 Fig. 9, my Fig. 12.8 from Longfield et al. 1969, Biomed. Engng. 4,
517-522 Fig. 1; Random House (UK) Ltd. for my Figs. 1.2 and 6.4a from Brooker
XVi Rough &$aces

1984 ed., Manual of British standari in engineering metrology (Hutchinson), pp.


184 Fig. 1 and 191 Fig. 10.10; the Royal Society and the authors for my Fig. 9.12
from Greenwood 1984, Proc. Roy. SOC. Lond. A393, 133-157 Fig. 7, my Fig.
10.5 from Pullen & Williamson 1972, Proc. R. SOC.Lond. A327, 159-173 Fig. 3;
Science History Publications Ltd. for my Fig. 1.1 from Thom & Thom 1972, J.
Hist. Astron. 3, 11-26 Fig. 5; SociCtC Belge des Mecaniciens for my Fig. 3.3 from
Whitehouse 1975, Rev. M. Mec. 21, 19-28 Fig. 11; Society of Manufacturing
Engineers for my Fig. 2.13 and Table 11.1 from Thomas et al. 1975, S.M.E. Paper
IQ75-128 Fig. 2; Society of Photo-optical Instrumentation Engineers and the
authors for my Fig. 3.14 from Church et al. 1977, Opt. Eng. 16, 360-374 Fig. 11,
my Fig. 9.15 from Thomas 1991, Proc. SPZE 1573, 188-200 Figs. 5 & 6 ;
Society of Tribologists and Lubrication Engineers for my Fig. 10.7 from Lee &
Ren 1996 Trib. Trans. 39, 67-74 Fig. 13, my Fig. 11.3 from Akamatsu et al.
1992, Trib. Trans. 35, 745-750 Fig. 8; Taylor Hobson Pneumo, Leicester, for my
Fig. 6.7 from Whitehouse & Reason 1965, The equation of the mean line of surface
texture found by an electric wavefilter, Figs. 11-14, my Figs. 7.6 and 9.14 from
Dagnall 1980, Exploring surjiie texture, Figs. 17 & 46; VCH Verlagsgesellschaft
mbH, Weinheim, and the author for my Table 4.1 from DiNardo 1994, Narwscale
characterisation of surfaces and inte$aces, p. 145 Table 2.

I have attempted to trace the copyright holder of all the material reproduced in
this book and apologize to copyright holders if permission to publish in this form
has not been obtained.
CHAPTER 1

INTRODUCTION

1.1. Surface Roughness

We are used to the idea that materials have intrinsic properties such as density,
conductivity and elastic modulus. Surfaces, representing material boundaries, have
perhaps rather more insubstantial properties, but we still think of some of these
properties as intrinsic, like colour. There are other properties, however, which are
easy to define but whose value seems to depend on the technique or scale of
measurement: hardness, for instance. Roughness seems to be such a property, with
the added difficulty that it is not always so easy to define as a concept.
"I can't define roughness, but I know it when I see it". When we speak of
"rough country", "a rough road", "a rough fabric", we imply very different scales of
feature in each case, but we understand well enough what sort of surface is meant.
The Concise OED has more than a column of meanings, starting with "of uneven
or irregular surface, not smooth or level or polished, diversified or broken by
prominences.. . .. . .. .coarse in texture.. . To develop this idea a little, it seems to
.'I.

have something to do with our scale of view. If the countryside, or the stretch of
road, or the patch of fabric which we observed, simply rose steadily in our field of
view, or contained a single prominence, we would not refer to it as "rough"; if it
contained, say, ten prominences, we probably would so categorise it; if it contained
a hundred such, we certainly would. So already, it seems, the ideas of sampling
interval and sample size which we will introduce later are emerging and are bound
up with the very concept of roughness at quite a fundamental level.

I . 1.1. What Causes Roughness?

A geographer might find this a strange question. To him or her, accustomed to the
scale of features of the natural world where roughness is the norm, a better
question would be, "What causes flatness (or straightness)?". Generations of
students have been accustomed by their education to the normality of ideal straight
lines and flat surfaces. This has been exacerbated by the wide use of computer-

1
2 Rough Surfaces

aided drafting software, where straightness and flatness are the default
assumptions. This is a conceptual problem not only for scientists and engineers,
but for everyone with a Eurocentric education. The idea of straightness is built in
to all Indo-European languages as a concept linked with good, power and approval,
where "right", "rule", "regal" all have the same root. This is so deeply embedded
in our mental software that we are disturbed by the alignements of Brittany, whose
megalithic architects spent thousands of person-years constructing adjoining lines
of huge boulders, kdometres long, equidistant but nowhere straight (Fig. 1.1).
Clearly they understood the concept of straightness (otherwise how could they have
made the lines equidistant?), but culturally it held no importance for them.

-.,.. ...... 0 50 lOOm

....................
............... ... .. ... ....
.........
............
..........
....
.. 1 .........
.-
....
:*: :

- ._. .....
*.*. .... -.----....-. - . ..._ * .
.-.
. . .......
."
C

.... ...........----- .
.I
U
.
. *. *
'
U
.

....
..L
,
........
15.-
.... ........
... .....
.. .*..
--. .. .....
,...............-L.(..*
* ' ....-.*
'4.

........
.L.. .
..)
a .nr

-
.. . . .
..
...- . ......
.............. , . .*-........
--rss.*
................
. _.,.--- .......
*.
- .
.*

...........
-.I..

... .. *.......
*-

-.
., .,.. -..
...I,
*
c

......
.
.-.
...
...
U .
" .
* .
..*.
- .
I "-0
*-*-
. ...... ........ -..
''''**w*w ~ .u
I.. 5.. .'.."."....I-.
.4*.. .*#a
*-*
..... lyy
%..
~~~

Figure 1 . 1 . Part of the Camac alignments, adapted tiom Thom & Thorn (1 972). Each mark represents a
menhir; a typical menhir is 2m high and lm in diameter.

The fact is that roughness is the natural state of surfaces, and left to its own
devices Nature will make sure they are rough. The roughness of a surface is a
measure of its lack of order. Disorder is entropy under another name, and if we
consider a solid surface as a closed system then the Second Law of
Thermodynamics predicts that its entropy will tend to a maximum. To reduce its
roughness we must reduce its entropy, and the Second Law tells us that we can
only do this by doing work. Thus if we transpose the axes of the well-known figure
which relates machining time to roughness, we see that it is nothing but an entropy
diagram (Fig. 1.2). Many, perhaps most, natural surfaces are fractal (see Chapter
9), and it is characteristic of fractal surfaces that their roughness increases without
limit. So it is that in a universe of fractal surfaces man's attempts to reduce entropy
by imposing straightness and smoothness extend over only a very small range of
dimensions (Russ 1994).
Introduction 3

work

Figure 1.2. (a) Relationship of surface texture to production time (Brooker 1984); (b) The same figure
replotted as work reducing entropy

1.1.2. Why is Roughness Important?

When I was starting research in the field of surface roughness years ago I was
advised against it by a distinguished academic engineer on the grounds that
roughness is essentially a second-order effect in physical systems, and would
therefore never assume an important place in engineering science. Time has, I
think, vindicated my judgement rather than hs, for two rather interesting reasons
(Thomas 1988). The first is that while it is certainly true that roughness is a
second-order effect, it is a second-order effect across a very wide spectrum of
technical activity; not just tribologists and production engineers, but cartographers,
radar engineers, highway and aircraft engineers, hydrodynamicists and even
bioengineers find increasingly that from time to time it obtrudes into their
particular specialty (Fig. 1.3). The second is that all the easy first-order problems
have been solved: whatever happened, for instance, to heavy electrical engineering
as an academic discipline? We live increasingly in a world where second order
effects present the major remaining challenge; any fool can make an internal
4 Rough &$aces

combustion engine that works, the trick is to make one that will run for a million
miles at 100 miles to the gallon.

100 '
1970 1975 1980 1985 1990 1995

Figure 1.3. Cumulative number ofpublications on surface roughness (Thomas 1997)

1.2. Principles of Roughness Measurement

By "measurement" we mean something more than mere inspxtion. We will define


measurement in the present context as a process which gives, or is capable of
giving, quantitative information about individual or average surface heights. Thus
we exclude many forms of optical examination. These may give information about
the existence and direction of the lay on machined surfaces, or about the presence
and spacing of feed and chatter marks or other individual defects, but this does not
fall within our definition.
There are some general considerations in choosing any measuring
instrument: cost, ease of operation, size and robustness. There is also the issue of
whether a measurement is comparative or absolute. In addition, for roughness
measuring instruments, it is necessary to decide whether or not the instrument
should make physical contact with the surface, and whethei it needs to be able to
measure an area of a surface or only a section or profile through it. Most important
of all are the horizontal and vertical range and resolution.
Some of these criteria are self-explanatory, but the issue of comparative
versus absolute measurement is worth a few moments' digression. Many roughness
measuring instruments, for instance stylus instruments, give absolute
measurements of local heights. Thus they can be calibrated against secondary
length standards such as slip gauges and so in principle at least are traceable to
primary standards. Other instruments, for instance glossmeters, give average
values of some surface parameter, which may depend on material properties and
may vary from one finishing process to the next. Such instruments must be
Introduction 5

calibrated against an absolute instrument used under the same conditions. Under
these conditions they may still be traceable, but in a much more tightly restricted
way. This is likely to be of some practical importance in a manufacturing
environment where the roughness instrument is part of a quality system under I S 0
9000. Vorburger and Teague (1981) classify these two kinds of instruments as
"profiling" and "parametric" techtuques.
Sectional measurement is usually quicker, simpler and easier to interpret than
areal measurement, and all current roughness standards, as we shall see later, are
written in terms of sectional measurements. For many practical purposes sectional
measurements are adequate, and sectional techruques should be preferred unless
there is some good reason to the contrary. However, most engineering interactions
of surfaces, including all contact phenomena, are areal in nature, and the
information necessary to describe their function must similarly be areal. Often this
information can be inferred mathematically from sectional information, as
discussed in later chapters. The problem arises when the events about which
information is sought are comparatively rare. For instance, under most
engineering loads the area of real contact between two surfaces is likely to be less
than 1% of their nominal area. Thus a sectional measurement may underestimate
or miss altogether the features of practical interest (Fig. 1.4). The special problems
of areal, or three-dimensional (3D), measurement will be discussed separately
below.

Figure 1.4. A sectional measurement may underestimate the number and size .If important features or miss
them altogether
6 Rough Sur$aces

I.2.1. Range and Resolution

We can say crudely (Thomas 1978) that roughness exists in two principal planes:
at right angles to the surface, when it may be characterised by some kind of height,
and in the plane of the surface, identified as "texture" by Reason (195 1). There are
thus two sets of limitations which we need to discuss with reference to each
roughness measuring instrument or technique: the largest and smallest differences
of height which it will resolve, and the longest and shortest surface wavelengths
with which it can cope. It is important to remember that every instrument or
technique is subject to these limitations of resolution, and that the actual figures
involved will vary from instrument to instrument.
A very useful way of defining and comparing instrument performance at a
glance is due to Stedman (1987), who suggested plotting the horizontal range and
resolution of an instrument as an envelope in two-dimensional space (Fig. 1.5). If
z,, and z,,, are the maximum and minimum heights that can be measured by the
instrument, and similarly A,, and are the longest and shortest wavelengths,
these will define a rectangle in z-/2 space outside which the instrument will not
measure. But the practical operating envelope is subject to further restrictions, for
instance the steepest slope em, which the instrument will measure, and the
sharpest curvature C,, which it will follow. These may also be displayed in z-/2
space if some assumption is made about the form of the surface. For mathematical
convenience Stedman assumes a sinusoidal surface

z = Rpsin(2xx//2)

where Rp is the amplitude. Slopes and curvatures are then the first and second
differentials respectively:

B = ( 2 xRp/A) cos (2 n x / R )

C = - (4 ?Rp/A2) sin ( 2 x x / A )

The maxima of these functions occur when the trig functions are unity, so taking
logs,

logRp = 10g(8,,/2x) + log1

logRp = log(Cm,/4?) + 210gA


Introduction 7

On a logarithmic plot these are lines of slope 1 and 2 respectively which further
restrict the operating envelope (Fig. 1S ) .

Figure 1.5. Envelope of instrument pe.rfomance in amplitude-wavelengthspace (after Stedman 1987)

Other restrictions on the operating envelope could be devised; Stedman


includes, for instance, the effect of angular uncertainties in the instrument
reference plane. The assumption of sinusoidal surfaces is also somewhat of a
simplification. Nevertheless, Stedman diagrams are such a straightforward and
convenient way of defining and comparing instrument performance that they will
be used throughout to illustrate our discussion of measurement techmques,
omitting for the sake of simplicity the restrictions due to slopes and curvatures.
Readers should be aware, however, that these Stedman diagrams represent a
maximum working envelope, not all of which may be available to the same
instrument at the same time. Also, the examples given will usually refer to a
particular instrument, that is to an exemplification of the technique rather than as a
general statement about the technique. For instance, it is possible to build stylus
instruments on quite different scales of size, and it would be misleading to try to
present a generic Stedman diagram which would cover all possible realisations of
the stylus principle.
It is interesting to construct a Stedman dagram to examine summarily the
global coverage of topographical measurement techniques (Fig. 1.6). A single
envelope covers the current range of roughness measuring instruments, stylus,
optical and AFM. Electron microscopy extends this coverage to shorter
8 Rough Surfaces

wavelengths and larger amplitudes, though it is difficult (though not impossible;


for a discussion see Whitehouse 1994) to extract quantitative height information.
At the upper end of the roughness range, coordinate measuring machines take
over. A final envelope of surveying techniques covers everything from
autocollimators at the low-range end to satellite ranging techniques at the upper
end.

I Inm 1w 1mm lm

Figure 1.6. Global Stedman diagram (Thomas 1997).

Fig. 1.6 is not intended to be comprehensive or definitive, and for the sake of
simplicity many useful measurement techniques have been omitted. It does make
the point, however, that there are large areas in range-resolution space which are
not accessible to any current technique. Does this matter?
The lower right area of Fig. 1.6 represents small amplitudes at long
wavelengths. At the moment there does not seem to be any technological
requirement in this area, but it is worth noting that the error in the Hubble
telescope mirror (Parks 1991) would only just come withm a shaded area; a future
generation of Hubbles might well fall outside. The upper left area perhaps
represents a more pressing practical problem, that of the unavailability of
techruques for measuring large amplitudes at short wavelengths. It is certainly
possible to think of existing artefacts whose topography falls in this area, for
instance a hairbrush! More importantly, there are very many biological structures
with large vertical but small horizontal dimensions, starting on a cellular scale
Introduction 9

(Boyan et a1 1996), continuing up through growing crops (Gilley & Kottwitz 1994)
and ending with forest canopies (Gallagher et al. 1992). Many such structures are
of great economic importance, and at the moment we have no way of describing
their topography comprehensively.
This task would demand a much greater ratio of vertical range to resolution
than is available from current instruments. But when we reflect on the
improvement in this ratio in recent years, the implications for the future are
promising (Thomas 1997). At the start of the 1980s few roughness instruments
offered a ratio of better than 103:1 (Farago 1982). The current state of the art
provides examples both of stylus instruments (Garratt 1982) and optical
instruments (Caber et al. 1993) with ratios of better than 105:1. There does not
seem to be any fundamental law of instrument design preventing further
improvements, if not in these techniques then perhaps in newer ones; chemical
balances, for instance, have for many years been constructed with a range 108
times their resolution (Cook & Rabinowicz 1963).
Bearing the above principles in mind, we will begin in the next chapter by
discussing the stylus instrument, still the most popular and widely used method of
measuring surface roughness. We will go on to consider the increasing use of
optical instruments, both profiling and parametric. Many other techniques of
measuring roughness have been developed, and some of the more popular of these
are highlighted and their use in scanning mode for 3D measurement in various
microscopy systems is discussed. Finally in this part of the book we examine some
associated measurement questions, such as replication and in-process
measurement.

1.3. References

Boyan, B. D., Hummert, T. W., Dean, D. D., Schwarz, Z., "Role of material
surfaces in regulating bone and cartilage cell response", Biomaterials 17, 137-146
(1996)
Brooker, K. ed., Manual of British standards in engineering metrology
(Hutchmson, London, 1984)
Caber, P. J., Martinek, S. J., Niemann, R. J., "A new interferometric profiler
for smooth and rough surfaces", Proc. SPIE 2088 (1993)
Cook, N. H., and Rabinowicz, E., Physical measurement & analysis
(Addson-Wesley, Palo Alto, 1963)
10 Rough Surfaces

Farago, F. T., Handbook of dimensional measurement 2e (Industrial Press,


New York, 1982)
Gallagher, M. W.; Beswick, K. M.; Choularton, T. W., "Measurement and
modelling of cloudwater deposition to a snow-covered forest canopy", Atmospheric
Environment 26A, 2893-2903 (1992)
Garratt, J.D., "Applications for a wide range stylus instrument in surface
metrology", Wear 83, 13-23 (1982).
Gilley, J.E.; Kottwitz, E.R., "Darcy-Weisbach roughness coefficients for
selected crops", Trans. ASAE 37, 467-471 (1994)
Parks, R. E., "The Hubble space telescope investigation", Optics & Photonic
News 2,28 (1991)
Reason, R. E., "Surface finish", Australasian Engr., 44, 48-64 (1951).
Russ, J. C., Fractal surfaces (Plenum Press, New York, 1994).
Stedman, M., "Basis for comparing the performance of surface-measuring
machines", Prec. Engng., 9, 149-152 (1987)
Thom, A,, and Thom, A. S., "The Carnac alignments", J. Hist. Astron. 3, 11-
26 (1972)
Thomas, T. R., "Surface roughness: the next ten years", Surface Topography
1, 3-9 (1988)
Thomas, T. R., "Trends in surface roughness", Trans. 7th. Int. Con$ on
Metrology & Properties of Engng. Surfaces (Goteborg, 1997)
Thomas, T.R., "Surface roughness measurement: alternatives to the stylus",
Proc. 19th. Machine Tool Design & Research ConJ, 383-390 (UMIST,
Manchester, 1978)
Vorburger, T.V., Teague, E.C., "Optical techniques for on-line measurement
of surface topography", Precis .Engng. 3, 61-83 (1981).
Whitehouse, D. J., Handbook of surface metrology (Institute of Physics,
Bristol, 1994)
CHAPTER 2

STYLUS INSTRUMENTS

We shall commence in this chapter by dealing with the techniques most commonly
used for roughness measurement: those based on the use of the stylus instrument.
The two most natural ways to establish the roughness of a surface are to look at it
and to run a finger over it. Whitehouse (1994) has pointed out that all roughness
measurement techniques can be classified as analogues of one or the other of these
elementary methods. The stylus instrument is the embodiment of this second way;
there were estimated to be 25,000 stylus roughness measuring instruments in the
U.S.A.alone (Young & Bryan 1974).

2.1. Mechanical Instruments

The principle of the phonograph or gramophone, where a sharp probe


traverses a surface and transforms its minute irregularities into another form of
energy, seems ideally suited in retrospect to apply to the measurement of surfaces.
Strangely enough it was a generation after the invention of the phonograph before
surfaces were first measured with a stylus instrument. One early instrument used
an optical lever to magnify the stylus movement (Schmaltz 1936). Another
amplified the vertical movement of the stylus mechanically by a system of levers
until it sufficed to cause visible fluctuations in a continuous scratch on a smoked-
glass plate (Reason 1944) (Fig. 2.1). This had the advantage that the stylus need
not move at constant speed. In another version (Abbott et al. 1938) the vertical
movements of the probe were transmitted to a mirror forming part of an optical
lever, and the deflections of a light beam thus amplified were recorded on a
moving photographic film.

An interesting variant of the mechanical stylus instrument is the Flemming


integrator (Way 1969). As the stylus moves over the rough surface an ingenious
mechanical arrangement sums its displacements in a downward direction only. If
this sum is divided by the length of traverse it is easy to show that the result is half
the mean absolute slope. The actual figure measured will depend on the stylus

11
12 Rough Surfaces

dimensions; F l e m i n g proposed the use of two styli of Merent radii to distinguish


between the slopes associated with roughness and those associated with waviness,
probably the first recorded recognition that mean slope is not an intrinsic property
of a surface.

FI
RO

HORIZONTAL MOTION
O F INSTRUMENT BODY
6 L E A F SPRING
SMOKED GLASS

Figure 2.1. Tomlinson roughness meter (Galyer & Shotbolt 1990)

Before we leave mechanical stylus instruments mention must be made of the


wall gauge (Lackenby 1962). This instrument was developed by the British Ship
Research Association (BSRA) to measure the roughness of :hips' hulls. These are
very much rougher than most machined surfaces and a larger instrument is
consequently appropriate. The BSRA wall gauge consists of a railway about 76 cm
long carrying a trolley on which the stylus and recording gear is mounted (Fig.
2.2). The stylus is a steel ball, and through an arrangement of levers its vertical
motion is amplified and recorded on a moving smoked-glass microscope slide.
T h s arrangement does not require the trolley to be moved at constant speed, which
is just as well because it is advanced along the track manually by the operator. The
slide is subsequently mounted in a projector with a cylindncal lens to increase its
Stylus Instruments 13

relative vertical magnification and measurements are made of the projected image
by hand.

As gauge moves Holder carrying


16 cm holder moves 7.6 cm standard 7 . 6 m x 2 . S m
giving horizontal glass slide coated
reduction of 10 :I with colloidal

Probe with 1.5 mm dia.


ball point constrained
10move normal to track
- -
10 cm travel of gauge
\
Pair of feet

Direction of

Figure 2.2. BSRA wall gauge (Lackenby 1962)

The device is thus an ingenious compromise. The measurement and analysis


are split into two separate tasks. The former can be carried out by relatively
unskilled personnel and the instrument is robust enough to survive the rigorous
environmental conditions of a dry dock. The latter needs skilled technical
assistance but can be performed entirely in the laboratory at leisure. These are
important considerations when it is borne in mind that up to 80 measurements may
be needed on a single hull. An electrically recording version of this instrument has
since been developed (Chuah et al. 1990).

2.2. Electrical Instruments

Finally, however, the obvious step was taken and the stylus was given a transducer
to convert its vertical movement into electrical oscillations. The Abbott
profilometer (Abbott & Firestone 1933) ushered in a new era in surface
measurement. (We shall refrain from using the word 'profilometer' hereafter,
firstly because it is apparently a regstered trade-mark in the United States and thus
14 Rough Surfaces

may not be used in a generic sense, and secondly because it is bad practice to mix
two classical languages.) In its original form it had all the important components
which stylus instruments have embodied ever since: a pickup, driven by a gearbox,
which draws the stylus over the surface at a constant speed; an electronic amplifier
to boost the signal from the stylus transducer to a useful level; and a device, also
driven at constant speed, for recording the amplified signal (Fig. 2.3).

Pickup
Gear-box
Datum
Stylus

Transducer
Amplifier

11111

1nm 1Iy” 1mm l r n

Data logger Chart recorder


-
Figure 2.3. Schematic stylus instrument

The vertical range of a stylus instrument depends on the dynamic range of the
transducer and can be as much as 1 mm or more. The vertical resolution is
ultimately limited by background mechanical vibrations and thermal noise in the
electronics; a commercial instrument is available with a claimed resolution of 2 nm
(Moody 1968). At this level of sensitivity the instrument is quite an efficient
seismograph and the most stringent precautions must be taken to ensure a stable
thermal and mechanical environment. The story is told that during its
development a mysterious transient, thought at first to be an electronic
malfunction, was finally traced, after the entire electronics had been unsuccessfully
rebuilt, by observing that its appearance coincided with the departure of the milk
train for London from a station several miles away!
The horizontal range of stylus instruments is set by the length of pickup
traverse; horizontal resolution depends on stylus dimensions. The disadvantages of
the stylus instrument are manifest: its bulk, its complexity, its relative fragility, its
Stylus Instruments 15

high initial cost, its limitation to a section of a surface, the necessity of a skilled
operator for any measurement out of the ordinary. The single advantage which
outweighs all these is the availability of an electrical signal, which can be subjected
to all the conditioning processes of modem electronics to yield any desired
roughness parameter, or can be recorded for display or subsequent analysis. The
stylus instrument is thus by far the most popular method of surface measurement
and outnumbers all other instruments combined. It is also the instrument in terms
of which all national roughness standards are defined. It is therefore appropriate to
discuss its component parts in more detail.

2.2.1. Stylus and Skid

In early instruments the stylus was often a phonograph needle (Abbott &
Goldschmidt 1937; Barash 1963) and more recently a sewinq needle has been used
as a stylus (Gray & Johnson 1972). However, phonograph needles were found to
be too large and too heavily loaded, and caused unacceptable surface damage.
Diamond styli are now universally employed. In many instruments they are cones
of 90" included angle and tip radius 4-12 pm (Williamson 1947). However, in one
of the most popular stylus instruments the stylus is a truncated pyramid. The angle
between the faces is 90" and the dimensions of the rectangular flat at the tip vary; a
small flat is classed as a high-resolution stylus, while an average one is about 3 pm
x 8 pm (Jungles & Whitehouse 1970). The short edge is parallel to the direction of
motion. Thus the stylus cannot resolve a wavelength shorter than 6 pm (see later
chapters), and integrates over a narrow strip of surface 8 pm wide.
The slopes of most real surfaces are so gentle that penetration of valleys is not
usually a problem. However, this is not always the case. It is sometimes necessary
to measure surfaces which consist effectively of more or less smooth planes
containing relatively steep-sided and deep craters, such as those of wood
(Elmendorf & Vaughan 1958) or machining tools (Tsao et al. 1968). More
common still is the requirement to measure surfaces of abrasive composites, either
coated abrasives or grinding wheels P a u l et al. 1972; Deutsch et al. 1973;
Friedman et al. 1974, Fugelso & Wu 1977), where no ordinary stylus will penetrate
the gaps between the individual grits. All the above authors have described
instruments designed to overcome this problem by the employment of a stylus
vibrated by an electrical transducer in a vertical plane with an amplitude much
greater than the anticipated variation of height on the surface. It is clearly
necessary that the frequency of vibration should be different from the expected
16 Rough Su$aces

frequencies of variation of height so that it can be removed from the signal by


electronic filtering.
The quantity actually measured by the stylus transducer is the change in the
vertical separation of the stylus and the transducer. If the pickup is constrained to
move in a horizontal plane or in a curve of fixed radius then the transducer will
give the instantaneous height difference between the stylus and this independent
datum. Such a system involves a tedious levelling procedure, particularly so at
high vertical magnifications, and for many purposes it is more convenient to use a
skid. The transducer senses the difference in level between the stylus and the skid
and no skilled setting up is required.
The skid is attached to the pickup and rests on the surface either beside or in
line with the stylus (Figure 2.4). In-line skids, either in front of or behind the
stylus, are the usual arrangement, and are preferred for tracing inside small bores.
Straddling skids, either as buttons or as a V-shaped support member, provide
guidance parallel to the axial plane of parts with curved surfaces.. Farago (1982)
describes a number of detailed skid designs for various specialised inspection
applications.

Figure 2.4. (a) Aligned and (b) straddling skids (Farago 1982)

2.2.2. Transducers

As with the cartridge of a high-fidelity phonograph, the performance of a stylus


instrument is only as good as its transducer. Some of the less expensive stylus
instruments used a piezoelectric crystal as the transducer element. Changes in
pressure due to stylus movement cause a small change in charge, which can then
be amplified. In practice the charge leaks steadily away from the crystal at such a
rate that the transducer will not correctly transmit low frequencies, resulting in a
loss of accuracy at long surface wavelengths (Reason 1956).
Stylus Instruments 17

In another system, the stylus moves the anode of a triode through a flexible
diaphragm, thus causing a large change in the effective electrical resistance for a
comparatively small stylus displacement (Underwood & Bidwell 1953; Chinick
1968). Because of the fragility of the vacuum tube it is essential that excessive
anode movement be avoided. This is achieved by coupling the stylus to the anode
extension through a viscous liquid. For high-frequency small-amplitude
displacements the coupling behaves as if it were rigid, but permits increasing shear
as the frequency decreases, thus acting as a high-pass filter; unfortunately its
transmission characteristic, being viscosity-dependent, is a function of temperature.
Moving-coil transducers are sometimes used, but their output depends on the
velocity of the stylus rather than its position and must be integrated to give an
amplitude. As the integrating circuits are inefficient at low frequencies this also
has a built-in high-pass filter (Reason 1956; Chinick 1968). A capacitance
transducer has also been described (Miyazaki 1965) where the stylus vertically
displaces, through a lever arm, one plate of a capacitor whose other plate is a
conducting liquid. Difficulties are reported, not surprisingly, in levelling the
system.

\ & Knife edge

Figure 2 . 5 . Schematic LVDT transducer

Many modern stylus instruments use a linear variable differential transformer


(LVDT). In a typical modified LVDT two coils are wound on opposite arms of an
E-shaped core (Fig. 2.5). An armature attached to the stylus and pivoted about the
centre arm increases the air gap of one coil as it decreases that of the other, causing
a differential change in inductance. This alters the output of a bridge circuit
excited by an oscillator at a frequency much higher than the maximum anticipated
frequency of stylus displacement. Thls camer frequency must subsequently be
removed from the signal by demodulation. The advantage of this system is that it
18 Rough Surfaces

will measure low frequencies right down to a static stylus displacement (Reason
1956; Chinick 1968).
Garratt (1982) describes an interferometric transducer. The stylus arm is a
pivoted lever, at one end of whtch is the stylus and at the other end of which is a
reflector which acts as the measurement arm of a Michelson laser interferometer.
The vertical resolution is 5 nm and the range is 2 mm, giving a ratio of range to
resolution of 5 x lo5.

2.2.3. Pickup

One instrument manufacturer has used an endless rubber belt driving a trolley
carrying the stylus and transducer along an optically flat guideway. Another has
used a linkage which in fact drives the pickup in a pair of arcs forming a very
shallow letter W. This is adequate when the stylus is used with a skid, but causes
some difficulty if a smooth surface is measured relative to an absolute datum. It is
sometimes inconvenient to move the pickup over the surface and instead the pickup
is held motionless and the test piece is moved below the stylus. Stages are
commercially available for this purpose (see the discussion of 3D measurement.

Figure 2.6. Pickup speed measured as a hnction oftime for a commercial stylus instrument (Desages &
Michel 1993). AC: Traverse length; AB:run-up length; BC: evaluation length

It is essential that the pickup be driven at constant horizontal speed while


measuring, as the overall design of many instruments assumes that equal intervals
of time correspond to equal intervals of horizontal distance. The actual length over
which measurements are made, the evaluation length (BS1134, 1981), is shorter
Stylus Instruments 19

than the total length of pickup travel, the traverse length, to allow an initial run-up
for pickup acceleration and a final run-out for the pickup to slow down and stop
(Fig. 2.6). The commercial instrument tested in Fig. 2.6 in fact consistently slowed
down by about 3% after the first quarter of the evaluation length.
Very often it is necessary to measure workpieces whose surfaces are not flat.
If they have a section of constant curvature this can be achieved by constraining the
stylus to move in an arc of the same curvature, and a number of ingenious
mechanical devices are available commercially for this purpose. If the test piece is
circular or cylindrical, matters are much easier. Instruments are commercially
available which will either rotate the component by rollers against a fixed stylus or
will move the stylus in an arc of variable radius around the workpiece.
The slower the stylus moves, the finer the detail that can be resolved (subject
to the limitations of stylus size) and accessories for commercial instruments have
been available which will reduce the speed to less than 5 p d s . At the lowest
speeds it is apparently difficult to guarantee constant speed because of stick-slip in
the translation mechanism. At the other end of the scale, the stylus cannot be
traversed too fast or it will lose contact where the surface falls away steeply. The
effect of this will be to skew the slope distribution, that is to record negative slopes
as being more gentle than they really are. The exact proportion of slopes
misrecorded depends on the interaction of the pickup dynamics and stylus load
with the geometry of the particular surface being measured (discussed in detail by
Whitehouse 1994). It is dlficult to avoid thls effect completely if measurements
are to be made in a reasonable time, and the design of most commercial
instruments is therefore a compromise. The fastest speed of traverse normally
employed is about 1 mm/s; a stylus instrument which has been reported to traverse
at 5 mm/s will not follow slopes steeper than about 6 degrees (Morrison 1995).

2.2.4. Output Recording

After removal of the carrier, if any, and amplification, it is necessary to transform


the signal again into some form which the operator can understand. In some early
instruments the amplified signal drove a loudspeaker (Harrison 1931) or an
oscilloscope (Abbott & Firestone 1933). Neither of these could provide a
permanent record, and it quickly became and has remained the practice to use a
chart recorder. Clearly it is essential that the recorder should run at constant speed
to avoid distortion of the signal, but this constant speed may be increased to
increase the horizontal magnification of the recorded signal. (The same effect may
20 Rough Sudaces

of course be achieved by slowing the pickup traverse speed relative to that of the
recorder.) Chart recordings of surface profiles are clear, unambiguous and self-
explanatory; the only difficulty in interpretation arises from the distorted ratio of
vertical to horizontal magnification, a problem which we discuss below.
Surface parameters of various kinds can be measured from the chart
recordings, but this is rather slow, and would be unsuitable for a quality-control
application where a large number of workpieces must be measured in a short time.
Most stylus instruments, therefore, either have additional averaging circuitry of
some kind which displays a selected parameter directly on a meter, or are
connected to a microcomputer which performs the same function. The definitions
of these parameters and the details of the procedures for calculating them are
discussed in a later chapter.

2.3. Sources of Error

2.3.1. Effect of Stylus Size

The stylus is not a mathematical point but an artefact of finite dimensions. This
implies that the stylus must fail to follow peaks and valleys faithfully and hence
must produce a distorted record of the surface. How serious is this distortion?

/--\

Figure 2.7. Distortion of measured profile due to f ~ t dimensions


e of stylus tip (exaggerated)
(Radhakrishnan 1970)

The so-called "traced profile" (IS0 3274, 1996) recorded by the stylus
instrument is the locus of the centre of the stylus. If the contacting portion of the
Stylus Instruments 21

stylus is assumed to be spherical in section the effective profile will correspond to


the contacting envelope (Hullproflo of the E or envelope system (von Weingraber
1957) where the radius of the rolling circle is that of the spherical portion of the
stylus tip. The radius of curvature of a peak may be exaggerated, while a valley
may be represented as a cusp (Fig. 2.7).
To appreciate the likely magnitude of this effect it is necessary to consider
more closely the geometry of the stylus. According to I S 0 3274 a stylus may have
an included angle of 60' or 90' and a tip radius of curvature of 2, 5 or 10 pm. The
measurement of the actual tip dimensions is extremely difficult as they approach
the limit of resolution of optical techniques (Williamson 1947; Jungles &
Whltehouse 1970); Williamson succeeded in measuring the tips of four conical
styli and reported average radii of curvature of from 2.5 pm to 53 pm.
It would seem, therefore, that a profile containing many peaks and valleys of
ra&us of curvature 10 pm or less, or many slopes steeper than 45O, would be likely
to be more or less badly misrepresented by a stylus instrument. Much play has
been made with the problem of measuring the Caliblock and similar specimens,
roughness standards on which a very nearly rectangular one-dimensional profile is
etched (Reason 1951; Peres 1953), and clearly indeed such a profile can never be
followed very closely by a standard stylus.

0.51
2.5
I
5
1
10
1
20 50
1 I
la,
I
200
Tracing stylus radius &m

Figure 2.8. Effect of stylus tip radius on measured roughness for various machined surfaces (Radhakrishnan
1970): (1) planed (2) electro-eroded (3) milled (4) ground; (5) electrochemically sunk, ( 6 )honed
22 Rough Surfaces

In measurements of real surfaces, however, this difficulty is largely illusory,


as their slopes are for the most part very gentle at the scale on which the stylus
measures them. This is confirmed by the experimental findings of a number of
workers (Williamson 1947; Radhakrishnan 1970) that styli of standard dimensions
do not significantly misrepresent the average roughness of the surface (Fig. 2.8).
Whitehouse (1974) has reached the same conclusion by a different stochastic
argument. Arguments which rely on a representation of the surface as a single
pure sinusoid (Nakamura 1966), though they may reach similar conclusions, are to
be avoided.
The misconception has arisen from the way in which profiles are commonly
presented. A slope of 1" is almost imperceptible to the human eye, and the
instrument manufacturers have therefore found it convenient to exaggerate the
vertical magnification over the horizontal on their chart recorders. This has given
generations of metrologists a totally false impression of surface microgeometry.
The contact of two rough surfaces, far from resembling Bowden's famous analogy
of "Austria turned upside down on top of Switzerland" (Bowden & Tabor 1950)
more closely resembles Iowa on top of the Netherlands (Thomas 1973). To
emphasize this point Reason (1944) showed a conventional chart recording of a
ground surface with the vertical scale exaggerated 35:l over the horizontal; the
stylus, distorted correspondingly (Fig. 2.9) appears now as a mere sliver. On the
same figure he drew the same recording at a 1: 1 ratio. As he himself remarks, it is
difficult to reconcile the two representations even with the aid of fiducial marks.
There is no room here to reproduce the whole of Reasonk figure, which unfolded
from the original paper to a length of a metre or so!

25:1 X Y
3 -

Figure 2.9. Effect of horizontal compression of chart recording on profile presentation (adapted from Reason
1944):(a) true appearance of section XY'; (b) representation on chart recording
Stylus Instruments 23

2.3.2. Effect of Stylus Load

As the dimensions of the stylus are finite, so also is the load on it. Although the
load is small, 0.75 mN according to I S 0 3274, the area of contact is also so small
that the local pressure may be sufficiently high to cause significant local elastic
downward deformation of the surface being measured. In some cases the local
pressure may exceed the flow pressure of the material and plastic deformation, i.e.
irreversible damage, of the surface may result.
It is fairly easy to calculate the elastic behaviour of a surface under a chisel-
shaped stylus. The average vertical deflection 6 of a homogeneous isotropic elastic
half-space of Hertzian modulus E' and Poisson's ratio v by a rigid indenter of
rectangular cross-section ab under load W is (Timoshenko & Goodier 1951):

S = mW(l-9)/Erd(ab)

where m is a dimensionless constant whose numerical value is a function of ah.


Taking m = 0.9, W = 1 mN, v = 0.33, E' = 2 x lo5 N/mm2 for steel gives 6 = 0.83
nm. Clearly there is no danger here of appreciable error in measurements on any
metal. A calculation for the spherical stylus tip specified in I S 0 3274 (1996)
would gwe a similar result.
The problem with metals at least, then, if any, is not elastic but plastic
deformation. This has been investigated by Quiney et al. (1967), Tucker and
Meyerhoff (1969) and Guerrero and Black (1972). Quiney et al. were able to make
a scratch about 1.7 pm deep on an aluminium surface by using a stylus force about
10 times higher than the recommended standard. Tucker and Meyerhoff presented
scanning electron micrographs of a lead surface damaged by a stylus, but admitted
that no scratches could be found on harder materials. Guerrero and Black
presented a number of scanning electron micrographs of so-called stylus damage,
and calculated that stylus scratches on a steel surface were as much as 50 nm deep;
however, their assumptions concerning stylus tip geometry are rather questionable.
The trouble seems to be that the stylus generally does make a visible scratch,
and the existence of this scratch is held to be prima facie evidence of unacceptable
damage. Here the key word is "unacceptable". If we can show, for instance, that
the surface is everywhere deformed by the same amount, the output of the
instrument will then be a true profile displaced downward by a constant distance.
Reason et al. (1944) traversed a profile with a stylus load of 0.06 mN, repeated the
measurement with an increased load of 0.8 mN and then returned to 0.06 mN for a
third traverse; the profiles were nearly identical (Fig. 2.10). Schwartz and Brown
24 Rough Surfaces

(1966) found that the stylus measurements of a step in a silver film on a glass
substrate agreed with interferometric measurements to within 24 nm, while Estill
and Moody (1966) found no more than 43 nm deformation even on a soft gold
film. Williamson (1968), in a series of careful experiments, could find no evidence
of the information from stylus measurements being affected by plastic deformation.
As he remarked elsewhere, a bulldozer traversing a range of hills would leave a
scar visible from many miles up, but a recording barometer carried on the vehicle
would return a profile of the topography accurate enough for most practical
purposes.

Figure 2.10. Effect of stylus load (Reason 1944): (a) profile measured at 6 rng load; (b) relocated profile at
80 mg load (c) relocated profile at 6 mg load again

2.3.3. Other Sources of Error

A question was raised above concerning the dynamic response of the stylus:
whether there is any possibility of it losing contact with steep reverse slopes of the
surface as a consequence of the speed of traverse. Again this is to overestimate the
average steepness of real surfaces. Nakamura (1966) and Damir (1973) have
considered the effect of stylus geometry on its dynamic response, but the input
surface models in both cases are somewhat unrealistic. Nakamura's results,
however, indicate that with typical surface conditions, and with styli dimensions as
quoted in the international standards, the errors incurred are negligible. Funck et
al. (1992) found that speed of traverse significantly affected stylus measurements of
roughness on wood surfaces, but it is not clear whether these effects were due to the
timedependent mechanical properties of wood.
Another possibility of error lies in the lateral deflection of the stylus by
asperities. Verkerk et al. (1978) recommend that the ratio of axial to lateral
stiffness should be less than 0.00 I . AguIIo and Pages-Fita (I 974) have shown (Fig.
2.11) that lateral deflection can amount to as much as 1 pm between extremes
Stylus Instruments 25

when traversing a rough surface. However, the RMS excursion is more like 0.3
pm, and in any case, as we have already seen, a typical stylus is in effect
integrating over a strip 8 pm wide.

Figure 2.1 1. Effect of lateral deflection of stylus (Agullo & Pages-Fita 1974): (a) artificial two-dimensional
square-wave surface traversed at an angle to the lay; (bJ and (c) typical manufactured surfaces of different
roughness
26 Rough Su$aces

It has been claimed that the skid itself can cause damage to the surface.
Tucker and Meyerhoff (1969) observed deformation of soft surfaces such as lead
and niobium by a skid. Here the same argument applies as was used in the case of
stylus deformation: will it actually affect the measurements? According to Reason
et al. (1944), the difference in the deformation of a peak on a steel surface relative
to that of a hollow amounts to about 40 nm. It seems unlikely that this will be an
important source of error on hard surfaces.

“1.
b

a
a

Reciprocalwavelength

Figure 2.12. Effect of skid (Nara 1966):(a) profile as seen by (top to bottom) stylus with absolute datum;
stylus with skid; skid with absolute datum: (b) power spectra of (a): solid line. stylus with absolute datum;
broken line: stylus with skid
Stylus Instruments 27

More importantly, the skid acts as a mechanical high-pass filter. This has
two consequences. Firstly, information about longer wavelengths is lost; this is
irremediable and if these wavelengths are deemed to be relevant to the problem
under investigation then a skid must not be used. Secondly, the filter introduces a
phase lag meason et al. 1944) which might be supposed to distort the appearance
of the surface. Experiment suggests, however, that this distortion is not obtrusive
(Nara 1966) (Fig. 2.12a). It appears also that the mechanical filter embodied by
the skid has quite a sharp cut-off (Fig. 2.12b) and that the power spectrum, which
of course contains no phase information, is relatively unaffected at shorter
wavelengths. Ishigaki & Kawaguchi (198 1) conclude that varying the separation
&stance of skid and stylus has little effect.
The question of damage to soft surfaces has been discussed above, but what of
measurements on surfaces whch easily yield elastically? Elastomers are widely
used in engineering, particularly as elements in static or dynamic seals. The
surface finish of the metal elements is known to have an important effect on
sealing properties, and it seems reasonable to assume that the finish of the
elastomer will also play a part. To assume that all the asperities on the elastomer
are somehow squashed flat is to beg the question; no matter how compliant the
elastomer or heavy the load, a surface wavelength will exist below which the
elastomer will not conform. Calculations such as those described above suggest
that a stylus under its standard load will deform the surface of an elastomer by
more than 100 pm. Of course it does not follow that measurements are therefore
impossible; if every element of the surface is displaced vertically by exactly the
same amount then we will still record a true profile. However, as most commercial
elastomers are composite materials this may be rather a severe requirement.

- I mm

Figure 2.13. Stylus measurement of a compliant surface (Thomas et al. 1975): (a) profile of elastomer
cooled below its glass transitiontemperature; (b) a subsequent relocated measurement at room temperature
28 Rough &$aces

To investigate compliant yielding under the stylus, measurements were made


on an elastomer at two different temperatures using a relocation table (Thomas et
al. 1975). The elastomer was first frozen below its glass transition temperature by
a stream of evaporating liquid nitrogen. In t h s condition it is no more compliant
than steel. When it had warmed up to room temperature it was measured again
(Fig. 2.13), and no significant differences in the two profiles were apparent. The
explanation of this remarkable and useful result is not clear, but there is some
evidence that a thin surface layer of the elastomer may be much less compliant
than the bulk material.

2.4. Calibration and Standards

Because the stylus instrument was the earliest roughness mzasuring instrument to
achieve general acceptance, and because it is still so widely used, roughness
standards are still written largely with stylus instruments in mind. Production
instruments are usually calibrated from secondary calibration specimens, of which
IS0 5436 (1985) distinguishes four main types. Type A, used for checking vertical
magnification, has wide grooves of a known depth. Type B, for checking the
condition of the stylus tip, has a series of narrow grooves of various depths and
widths. Type C, for checking parameter meters, has repetitive grooves of
sinusoidal or triangular section. Type D has a pseudo-random profile which
extends the whole width of the standard to provide a more realistic, but less
accurate, overall system check.
Type C specimens of triangular section were prodused in the U.S.A. by
General Motors under the trade name Cali-Block (Young & Scire 1972) with an
included angle of 150 degrees. The production of sinusoidal Type C specimens has
been described by Sharman (1967/8) and by Teague et al. (1982), who found that
the rigorous NIST specification could best be satisfied by diamond turning.
Square-wave sections have been proposed for Type C specimens (Berger 1988) but
interaction with stylus geometry makes these prone to error (Peres 1953).
The best-known Type D specimens are those produced at the Physikalische-
Technische Bundesanstalt by Hasing (1965). These have a measuring area of
random profiles, obtained by grinding, in the direction of traverse, which repeat
every 4 mm, and cover a range of roughnesses from 1.5 pm down to 0.15 pm..
Song (1988) has improved on the PTB design by adding a smooth reference surface
at each end of the traverse to provide a datum for the skid, and has extended the
range of roughness down to 12 nm (Fig. 2.14).
Stylus Instruments 29

Figure 2.14. Type D modified PTB roughness calibration specimen with 8 consecutive identical profiles
(Song1988)

But how are these secondaq calibration specimens to be calibrated


themselves? Small height increments can be measured directly by interferometry
(Spragg 1967/8, IS0 5436). In most instances, however, the most suitable
instrument for the purpose is another stylus instrument, which itself must be
traceably calibrated, that is there must be an unbroken and documented chain of
calibration from a primary standard of length, as required by IS0 9000 (1987).
For the static deflection of the stylus and transducer, traceability may be provided
by a series of gauge blocks arranged in steps ( I S 0 5436), which serves to check the
linearity of the transducer as well as its sensitivity. If the steps are too coarse, they
can be scaled down by a reducing lever (Spragg 1967/8). Dynamic calibration of
the stylus instrument is often effected by a vibrator which can be driven at variable
frequency (Van Hasselt & de Bruin 1962/3, Parkes 1969, Bendeli et al. 1974); the
traceability of such vibrators is discussed by Barash & Reznikov (1983).

2.5. References

Abbott, E. J. and Firestone, F. A,, "Specifying surface quality", Mech.


Engng., 55, 569-572 ( 1 933).
30 Rough Sudaces

Abbott, E. J., Bousky, S. and Williamson, D. E., "The profilometer", Mech.


Engng. 60, 205-216 (1938).
Abbott, E. J. and Goldschmidt, E.," Surface quality", Mech. Engng., 59,
pp. 8 13-25. ( 1937).
Agullo, J. B. and Pages-Fita, J., "Performance analysis of the stylus technique
of surface roughness assessment: a random field approach", Proc. 15th Int.
Machine Tool Des. & Res. Con$, 349-362 (Birmingham University, 1974).
Barash, M. M., "Measuring the finish of rough surfaces", Int. J. Mach. Tool
Des. Res., 3,97-I00 (1963).
Barash, V. Y. and A. L. Reznikov, "Standard vibrator in metrological
certification of contact methods of roughness measurement", Measurement
Technology, 26, 658-661 (1983)
Baul, R. M., Graham, D. and Scott, W., "Characterization of the working
surface of abrasive wheels", Tribology, 2, 169-176 (1972).
Bendeli, A,, Duruz, J. and Thwaite, E.G., "A surface simulator for the precise
calibration of surface roughness measuring equipment", Metrologia, 10, 137-143
(1974).
Berger, J., "A new surface roughness standard fabricated using silicon
technology", Surface Topography, 1, 4 1-47 (1988)
Bowden, F. P. and Tabor, D., The friction and lubrication of solids Part 1,
(Oxford University Press, 1950).
BS 1134, "Assessment of surface texture Part 1. Methods and
instrumentation" (British Standards Institution, London, 1988)
Chinick H. P., "LVDT puts precision in surface texture measurement",
Cutting Tool Engng., 20, 13 (1968).
Chuah, K. B., Dey, S. K., Thomas, T. R., Townsin, R. L., "A digital hull
roughness analyser", Int. Workshop on Marine Roughness & Drag (RINA, London,
1990)
Damir, M. N. H., "Error in measurement due to stylus kinematics", Wear, 26,
219-227 (1973).
Desages, F. and Michel, O., "Calibration of a 3-D surface roughness
measuring device", Production Engng. Dept Report (Chalmers University,
Goteborg, 1993)
Deutsch, S. J., Wu, S. M. and Straklowski, C. M., "A new irregular surface
measuring system", Int. .J. Mach. Tool Des. & Res., 13,29-42 (1973).
Elmendorf, A. and Vaughan, T. W., "A survey of methods of measuring
smoothness of wood", Forest Products J., 8, 275-82 (1958).
Stylus Instruments 31

Estill, W. B. and Moody, J. C . , "Deformation caused by stylus tracking on


thin gold film", Z.S.A. Trans., 5, 373-378 (1966).
Farago, F. T., Handbook of dimensional measurement 2e, (Industrial Press,
New York, 1982)
Friedman, M. Y., Wu, S. M., and Suratkar, P. T., "Determination of
geometric properties of coated abrasive cutting edges", Trans. A.S.hLE.: J .Engng
.Ind., 96B, 1239-1244 (1974).
Fugelso, M., Wu, S. M., "Digital oscillating stylus profile measuring device",
Irzt. J. Mach. Tool Des Res 17,191-195 (1977)
Funck, J. W.; Forrer, J. B.; Butler, D. A,; Brunner, C. C.; Maristany, A. G.,
"Measuring surface roughness on wood: a comparison of laser-scatter and stylus-
tracing approaches", Proc. SPIE 1821, 173-184 (1993)
Galyer, J. F. W., and Shotbolt, C . R., Metrologyfor engineers 5e (Cassell,
London, 1990)
Garratt, J. D., "A new stylus instruent with a wide dynamic range for use in
surface metrology", Prec. Engng. 4, 145-151 (1982)
Gray, G. G. and Johnson, K. L., "The dynamic response of elastic bodies in
rolling contact to random roughness of their surfaces", J. Sound & Vibration, 22,
323-342 (1972).
Guerrero, J. L. and Black, J. T., "Stylus tracer resolution and surface damage
as determined by scanning electron microscopy", Trans. A.S.M.E. J . Eng. Ind.,
94B, 1087-1093 (1972).
Hamson, R. E. W., "A survey of surface quality standards and tolerance costs
based on 1929-1930 Precision-Grinding practice", Trans. ASME 53, 11-25 (193 1).
Hasing, J., "Herstellung und Eigenschaften von Referenznormalen fur das
Einstellen von Oberflachenmefigeraten",Werkstattstechnik 55, 380-382 (1965)
Ishigaki, H. and I. Kawaguchi, "Effect of a skid on the accuracy of measuring
surface roughness", Wear, 68, 203-21 1 (1981).
I S 0 3274, "Geometric product specifications - surface texture: profile method
nominal characteristics of contact (stylus) instruments" (International
Organisation for Standardization, Geneva, 1996)
I S 0 5436, "Calibration of stylus instruments" (International Organisation for
Standardization, Geneva, 1985)
IS0 9000, "Quality management and quality assurance standards"
(International Organisation for Standardization, Geneva, 1987)
Jungles, J. and Whitehouse, D. J., "An investigation of the shape and
dimensions of some diamond styli", J. Phys: Sci. Instrum., 3E,437-440 (1970).
32 Rough Suvfaces

Lackenby, H., "The resistance of ships, with special reference to skin friction
and surface condition", Proc. I. Mech. E., 176, 981-1014 (1962).
Miyazaki, K., "Electronic method, based on the surface of a liquid, for
measuring flatness", Microtechnic, 19,74-76 (1965).
Moody, J. C., "Measurement of ultrafine surface finishes", I.S.A. Trans. 7,
67-7 1 (1968).
Morrison, E., "A prototype scanning stylus profilometer for rapid
measurement of small surface areas", Int. J. Mach. Tools Manufact. 35, 325-331
(1995)
Nakamura, T., "On deformation of surface roughness curves caused by finite
radius of stylus and tilting of stylus holder arm", Bull. Jap. Soc. Precision Engng.,
1, 240-248 (1966).
Nara, J., "On CLA value obtained with direct reading surface roughness
testers - effects of skid and high pass filter", Bull. Jap. Soc. Precision Engng., 1,
263-273 (1966).
Parkes, D. H., "Calibration, certification and traceability of surface roughness
measuring equipment", A.S. T.M.E. Tech. Paper IQ69-505 (1969).
Peres, N. J. C., "Geometrical considerations arising from the use of square
wave calibration standards of surface finish", Aust. J. Appl. Phys., 4, 380-388
(1953).
Quiney, R. G., Austin, F. R. and Sargent, L. B., "The neasurement of surface
rougbness and profiles on metals", A.S.L.E. Trans. 10, 193-202 (1967).
Radhakrishnan, V., "Effect of stylus radius on the roughness values measured
with tracing stylus instruments", Wear, 16, 325-335 (1970).
Reason, R. E., "Surface finish and its measurement", J. Inst. Prod. Engrs .,
23, 347-372 (1944).
Reason, R. E., "Surface finish", Australasian Engr., 44, 48-64 (195 1).
Reason, R. E., Hopkins, M. R. and Garrod, R. I., Report on the measurement
of surface finish by stylus methods", (Taylor Hobson, Leicester, 1944)
Reason, R. E., "Significance and measurement of surface finish part 2: how
transducers affect instrument performance; how to select proper cutoff values",
Grinding &Finishing, 2, 32-36 and 41 (1956).
Sayles, R. S., Thomas, T. R., Anderson, J., Haslock, I. and Unsworth, A,,
"Measurement of the surface microgeometry of articular cartilage", J.
Biomechanics, 12, 257-267 (1979)
Schmaltz, G., Technische Oberfliichenkunde (Springer-Verlag, Berlin, 1936)
Stylus Instruments 33

Schwartz, N. and Brown, R., "A stylus method for evaluating the thickness of
thin films and substrate surface roughness", Trans. of the 8th National Vacuum
Symp., 836-845 (1966).
Sharman, H. B., "Calibration of surface texture measuring instruments",
Proc. I. Mech. E., 182, Part 3K, 319-326 (1967/68).
Song, J. F., "Random profile precision roughness calibration specimens",
Surface Topography, 1, 29-40 (1988)
Spragg, R. C., "Accurate calibration of surface texture and roundness
measuring instruments", Proc. I. Mech. E., 182, Part 3K, 397-405 (1967/68).
Teague, E. C., Scire, F. E., Vorburger, T. V., "Sinusoidal profile precision
roughness specimens", Wear 83, 61-73 (1982)
Thomas, T. R., "Influence of roughness on the deformation of metal surfaces
in static contact", Proc. 6th. Int .Conf on Fluid Sealing, B3, 33-48 (BHRA Fluid
Engineering, Cranfield, 1973).
Thomas, T. R., Holmes, C. F., McAdams, H. T. and Bernard, J. C., "Surface
features influencing the effectiveness of lip seals: a pattern - recognition approach",
S.M.E. Paper IQ75-128, (1975).
Timoshenko, S., and Goodier, J. N., Theory of elasticity (McGraw-Hill, New
York, 1951)
Tsao, K. C., Husein, A. B. and Wu, S. M., "Cutting tool crater wear
measurement by the lapping-comparator technique", Znt .J. Mach. Tool Des .Res.,
8, 15-26 (1968).
Tucker, R. C. and Meyerhoff, R. W., "An SEM study of surface roughness
measurement", Proc. 2nd Annual Scanning Electron Microscopy Symp., 389-396
(Illinois Inst. of Technol., Chicago, 1969).
Underwood, A. F. and Bidwell, J.B., "New instrument for roughness
measurement", Mach. & Tool Blue Book, 49, 202-215 (1953).
Van Hasselt, R., and de Bruin, W., "Comparative investigation of industrial
surface-roughness measuring instruments", Ann. CIRP 11, 193 (1962/3)
Verkerk, J.; Orelio, J. M. B.; Willemse, H. R., "Ratio of axial to lateral
stiffness, a quality parameter for stylus surface profile traciqg instruments", Int J
Mach Tool DesRes 18, 107-116 (1978)
Von Weingraber, H., "Suitability of the envelope line as a reference standard
for measuring roughness", Microtecnic, 11,6-17 (1957) .
Way, S., "Description and observation of metal surfaces", Proc. Con$ on
Friction & Surface Finish, 2e, 44-75 (MIT, Cambridge, 1969).
Whitehouse, D. J., Handbook of surface metrology (Institute of Physics,
Bristol, 1994)
34 Rough Surfaces

Whitehouse, D. J., "Theoretical analysis of stylus integration", Ann. C.I.R.P.


23, 81-82 (1974).
Williamson, D. E., "Tracer-point sharpness as affecting roughness
measurements", Trans. A.S.M.E., 69, 3 19-323 (1947).
Williamson, J. B. P., "Topography of solid surfaces", in Ku P. M. ed.,
Interdisciplinary approach to friction and wear, SP-181, 85-142 (NASA,
Washington, 1968).
Young, R. D. and Bryan, J. B., "The role of NE3P in the US National
Measurement System for surface finish", Ann. C.I.R.P., 23, 183-184, (1974)
Young, R. D. and Scire, F. E., "Precision reference specimens of surface
roughness: Some characteristics of the Cali-Block", J .Res .Nut .Bur .Stand.,
C76C, 2 1-23 (1 972).
CHAPTER 3

OPTICAL INSTRUMENTS

When electromagnetic radiation is incident on a rough sudace a proportion of its


energy, depending on the local physical properties of the surface, will be reflected.
The reflected beam will carry information about the roughness on which the design
of an instrument may be based. This information may appear in several different
ways.
The radiation may be reflected either specularly or diffusely or both (Fig.
3.1). Reflection is totally specular when the whole energy in the incident beam
obeys Snell's law, that is, the angle of reflection is equal to the angle of incidence,
and a surface which reflects radiation in this manner is said to be smooth.
Reflection is totally hffuse when the energy in the incident beam is distributed as
the cosine of the angle of reflection (Lambert's law). In practice, matters are not as
simple as this. Reflections from most real surfaces are neither completely specular
nor completely diffuse. Clearly the relationship between the wavelength of
radiation and the texture of the surface will affect the physics of reflection; thus a
surface which is smooth to radiation of one wavelength may behave as if it were
rough to radiation of a different wavelength (Ogilvy 1991).

Figure 3.1. Modes of reflection of electromagnetic radiation from a solid surface (Keller 1967/8). (a)
Combined specular and diffuse; @) specular only; (c) diffuse only.

35
36 Rough Surfaces

The angular arc through which reflected energy is scattered, and the
proportion of specular to diffuse reflection, both depend on the surface roughness.
Instruments which measure these angles and ratios directly are glossmeters or
scatterometers. Other instruments may extract more detailed roughness
information by further optical processing.
A general account of optical roughness measuring techniques is given by
Bennett & Mattsson (1989). Reviews and comparisons of optical roughness
measurement techniques have been made by Vorburger and his co-workers at NIST
(Young et al. 1980, Teague et al. 1981, Vorburger 1992). Vorburger & Teague
(1981) give more than 200 references to optical work. There is also a lengthy
discussion of optical techniques in Whitehouse (1994). The following summary
relies heavily on the above accounts. We will follow Vorburger & Teague (1981)
in dividing optical techniques into profiling and parametric. Profiling techniques
are associated with specular reflection, parametric techniques mainly with diffuse
reflection.

3.1. Profiling Techniques

3.1.1. Optical Sections

In the light-section microscope the image of a slit is thrown on to the surface at an


incident angle of 45" and viewed by a microscope objective at a reflected angle of
45" (Fig. 3.2). The reflected image will appear as a straight line if the surface is
smooth, and as an undulating line if the surface is rough. The relative vertical
magnification of the profile is the cosecant of the angle of incidence, in this case
1.4. Resolution is about 0.5 pm and it is quite easy to measure peak-to-valley
roughness. Light-section microscopes have been commercially available.
The image need not be of a slit; a straight-edge, such as a razor blade, will
suffice as object (Kayser 1943; Way 1969). Shaw and Peklenik (1963) used a
variation of this technique to measure the roughness of a razor-blade edge by
projecting its image on to an inclined screen. They reported discrepancies in the
apparent magnification which they attributed to the finite thickness of the razor
blades; this agrees with the suggestions made above concerning the integrating
effect of profile width. In a later development, Howes (1974) modified the light-
section technique to observe the virtual image of the slit, from which local slopes
Optical Instruments 37

‘f.
I lnm lprn Imm l m

Figure 3.2. Principle of light-section microscope (Farago 1982).

and curvatures can be measured. The technique can be applied to surfaces too
rough for the standard light-section microscope.
Johnson et al. (1993) measured terrain roughness by projecting a bar of light
from a flash gun onto a rough surface and photographing its image at an oblique
angle. The photographic negative was digitized back in the laboratory. Horizontal
and vertical resolutions were said to be 1 mm and 2 mm respectively; horizontal
and vertical ranges were 0.5 m and 1 m respectively. A similar system, but using a
video camera for data acquisition, has been described by Davies et al. (1994).

3.1.2. Optical Probes

One possible method of optical measurement is simply to use the light beam as a
non-contacting stylus for profile measurement. The most straightforward method
is to detect the change in the angle of specular reflection as the surface is translated
under an incident beam (Ramgulam et al. 1993). While adequate for its intended
purpose, the measurement of textile roughness, the lateral and vertical resolutions
of 25 pm and 10 pm respectively make such a system unsuitable for finer surfaces.
A number of more sophisticated variations exist. In constructing their Stedman
dagrams below, it will be assumed unless otherwise stated that the range in the
plane of the surface is 100 mm for a generic translation stage. Whitehouse (1994)
summarises the designs of several other optical profilers in addition to the ones
described below.
The first method employs the so-called Foucault knife-edge test (Dupuy
1967/68). An image of a spot is formed on the surface (Fig. 3.3). By using a half-
silvered mirror the image of the spot can be imaged itself to the knife-edge. A field
lens is placed here to image the objective lens on to a screen, I. If the conjugates of
the lens, 0, are at the knife-edge and surface respectively, a uniform disc appears
38 Rough Sur&aces

on the screen. If the surface is moved away the knife-edge intercepts the rays of
light in a different way resulting in a non-uniformity of light on the screen. This
non-uniformity is a measure of the distance moved by the surface. Electrically
maintaining the focus by means of cells A and B (which produce signals to move
the objective or knife-edge) and monitoring the movement gives the required
transducer effect. Vertical and horizontal resolutions of 0.01 pm and 0.5 pm
respectively, and a vertical range of 60 pm, are claimed for this instrument, a
specification which compares favourably with that of many stylus instruments.
Whitehouse (1975) has pointed out, however, that the system has a poor frequency
response and is unduly sensitive to tilt. A development of this system described by
Thwaite (1977) is claimed to have vertical and horizontal resolutions of 1 nm and
1 pm respectively

1-1 mechanism

paition I Polirioa 2 P d o n3

Figure 3.3. Dupuy optical probe (Whitehouse 1975): (a)Schematic layout; (b, simplified view; (c) view of
objective.
Optical Instruments 39

Another optical probe utilizing a somewhat similar principle was described


by Keller (1967/68). A spot object is imaged on to the surface and the diffuse
reflection at some arbitrary angle, which can be varied to increase sensitivity,
impinges on a pair of photoresistors differentially connected to give a null signal
when the spot is in focus. The out-of-balance signal drives a servo system which
alters the height of the probe until focus is again achieved. The vertical sensitivity
and range are respectively 2.5 pm and several millimemes. The horizontal
resolution, however, is only about 5 mm, as the probe is designed primarily for the
measurement of errors of form. Also intended for form measurement is the system
described by Ennos & Virdee (1986). Autocollimation of a reflected laser beam
measures the local slopes, which must then be integrated to obtain a height profile.
The lateral range is 15 mm but the lateral resolution is only 0.1 mm, and a single
slope measurement takes 3 s. The vertical range is 0.8 pm and the vertical
resolution is 0.1 nm.
A system based on Nomarski microscopy is described by Bristow (1988). A
laser beam passes to a translation stage which holds two mirrors (Fig.3.4). The
mirrors are in a pentaprism arrangement so that small mechanical motions of the
stage will not affect the 90' turning angle of the beam. The beam then passes

Figure 3.4. Long-pathlength optical profiler (Bristow 1988)

through a Nomarski (modified Wollaston) prism which shears it into two


orthogonally polarized components. The objective focusses both beams onto the
40 Rough Surfaces

surface, the two beams separated by about a quarter of the focal spot diameter.
After reflection at the surface the beams spatially recombine at the Nomarski prism
retaining their polarization identities as they pass to the turning mirrors and the
non-polarizing beamsplitter. The beams are finally split info their respective
components by the polarizing beamsplitter and directed to either of two detectors.
The surface height difference is related to the phase difference between the
two beams focussed on the surface and is proportional to the voltage difference
between the two detectors. The spot diameter on the surface ranges between 1 and
1.8 pm in diameter depending on the choice of the objective. Translation of the
turning mirrors causes the focal spots to scan across the surface, with a maximum
scan length of 100 mm. The surface slope is calculated at each point, sampled 1
pm apart, and the profile is calculated by integrating the slope data. A vertical
resolution of 0.025 nm and range of 2 pm is claimed.

W t R DIODE lrn, , , , , , , , ,

T
L

C O U l MATOR

FOCUS1NC
ENS

CLASS C U T E

SRCIMEN

Figure 3.5. CD player as an optical profiler (Sayles et al. 1988).

The sensor of a compact disc player is a kind of optical stylus whose


operating envelope overlaps with the characteristics of machined surfaces, and it is
not surprising that a number of systems have been based on its focus-detection
principle. Wehbi & Roques-Carmes (1986) reported a system using both white
light and laser sources which split the reflected beam so that it fell on two
detectors, the ratio of their outputs being proportional to the vertical distance by
Optical Instruments 41

which the local surface was out of focus. The principle of operation was
demonstrated successfully but the performance of the actual system was too poor
for a practical instrument. All focus-detection systems seem to suffer from a
problem in coping with sudden sharp surface discontinuities.
Sayles et al. (1988) adapted an actual CD reader, in which the out-of-balance
signal from the two detectors servos the moveable objective back into focus (Fig.
3.5). They extended the vertical range to 30 mm by axially displacing the light
source with a stepping motor. Vertical and horizontal resolutions of 0.1 pm and 1
pm were reported. The system worked well enough at long surface wavelengths,
but spurious short wavelengths were generated by inappropriate damping of the
servo.
In a commercial realisation of the focus-detection principle (Brown 1995),
the vertical displacement of the objective is measured independently (Fig. 3.6).
The light source is an infrared laser diode, and a separate optical system permits
simultaneous viewing of the workpiece for setting-up purposes. Vertical range and
resolution are said to be 0.1 mm and 6 nm respectively.

7, 1 analogue output 1: Laser diode


2: Prismwithbeam
splitter

UBC14 Controller 1
3: IRmirror
4: Window
5: Photodiodes
6: Leafspring
7: Coil
8: Magnet
'4 -I
1nm
u
inm Im
lpm Imm

(1
F------
9: Collimator lens
10: Objective
11: Tube
12: Light barrier
measurement
system
13: Test surface
14: PC control card
-' 15: Microscopewith
illumination
16: CCD camera
2
!
!3 13

Figure 3.6. UBM optical profiler (Brown 1995).


42 Rough S u ~ a c e s

The confocal microscope (Hamilton & Wilson 1982) is a focus detector of


sorts. A pinhole is interposed in the detector path so that the axial position of
sharpest focus is also the position of maximum intensity (Fig. 3.7). The distance
between the workpiece and objective is varied incrementally and the workpiece is
scanned at each axial position. By recordmg the axial position of greatest intensity
for each pixel, a composite picture of great depth of field can be built up. In one
realisation, the vertical range and resolution are 600 pm and 10 nm respectively,
and the horizontal range and resolution are 0.4 mm and 0.25 pm respectively.
Confocal microscopy was not developed primarily for roughness measurement but
has been adapted for this purpose (Lange et al. 1993, Sandoz et al. 1996).

REAL TIME IMAGE [b)

FOCUS PLANE
\ PIN HOLE
SAMPLE DETECTOR

Figure 3.7. Confocal microscope: (a) schematic @) construction of image by stacking sections

In the polarizing interferometer of Downs et a1 (1985, 1989), which needs no


independent reference, a birefringent lens preceding the microscope objective splits
the light into two polarisation components (Fig. 3.8). One component, the probe
beam, is focussed on the surface, and the other polarisation component, the
reference beam, is unfocussed. The phase difference between these two
components, the probe beam and the area-averaging reference beam, yields the
surface profile. The bandwidth is limited between the focussed spot size of 1 pm
Optical Instruments 43

Beam exoonder

Polorizer L-c

Non-polar~zmg
beam solltter

Error :eh:s,ve
Objective
I m c
I.zl.1
1 ,. 2 slgna' detector
'\ / I
Lens vibrator

Figure 3.8. Downs polarizing interferometer (Whitehouse 1994)

and the reference spot size of 10 pm, but the vertical resolution is an impressive
0.05 nm.
Another polarising instrument has been described by Sommargren (198 la, b).
A heterodyne laser beam is divided by a Wollaston prism into its two polarisation
components which are focussed at different places on the workpiece. (Fig. 3.9).
The reflected beams are then recombined at the prism. The reference beam is
coincident with the axis of rotation of the workpiece, and the measuring beam
traces a circular path on the surface as the workpiece is rotated. The resulting
surface profiles of about 1 mm circumference are defined by the phase difference
between the probe beam at the traversing point and the stationary reference beam.
Vertical range and resolution are 0.5 pm and 0.1 nm respectively.
Olsen and Adams (1970) described an instrument based on a rather different
principle for measuring the profiles of ocean waves from a low-flying aircraft. An
amplitude-modulated laser beam is reflected from the surface at normal incidence.
Changes in height will produce a path difference causing a phase shift between the
transmitted and reflected signals. The vertical resolution and range are 15 mm and
3 m respectively, and the horizontal resolution is 18 mm, but the effective signal-
44 Rough Surfaces

i - -
Ywoble :..-.-J
- 7
I . _] Rotoiable A 1 2 Dlate
neutral fblter
I Fmed M i plote
+
-
-
Reterenre poth

/9
.c-
// Laser

to-noise ratio is only about 20: 1. The reference datum is the time-averaged height
of the aircraft, in effect a high-pass filter.
An elegant application of the phase-measuring principle has been reported by
Pettigrew and Hancock (1978). A laser resonating simultaneously in several
different modes is employed as the light source for a Michelson interferometer. A
phase detector is used which is sensitive to the beat frequency between any two of
these modes. The effective wavelength is then very much longer than either of the
two beating wavelengths, giving a vertical range of up to 24 +m for a phase change
of 2n. The spot dameter is only 2.5 pm. The vertical resolution is not quoted but
appears to be of the order of 10 nm.

3.1.3. Interferometers

The distinction between profilers and interferometers is to some extent arbitrary.


A number of the instruments in the preceding section are based on an interference
Optical Instruments 45

principle, while several of the interferometer systems described below are referred
to by their designers as profilers. Many beautiful interferograms of rough surfaces
were produced by Tolansky (1960, 1970a, b), but the usefulness of interferometry
for roughness measurement was originally limited by several factors. In the
absence of coherent light sources it was difficult to obtain fringes of sufficient
sharpness and contrast. If the amplitudes of surface roughness were greater than
the wavelength of the incident light, there was no easy way of distinguishing
between fringes of different order. Finally, there was no convenient way of
converting the surface map of the interferogram into spot heights suitable for
quantitative analysis.
These difficulties are confronted in a phase-shifting interferometer originally
conceived by Wyant and his co-workers (1986). Fringes from a Michelson, Linnik
or Mireau interferometer, depending on the required magnification, are imaged
onto a charge-coupled diode and the resulting intensity variations are stored in a
computer (Fig. 3.10). A slightly different design (Biegen & Smythe 1988) uses a
Fizeau interferometer. The entire interferometer optics are shifted axially relative
to the workpiece by a piezoelectric transducer to change the optical path, giving a
different set of fringes. Three such sets of measurements give enough recorded
information to solve the intensity equations point by point for local heights.

Q
Reid +piece
SbD

PZT transducer
Mhu mi (microprocessorcontrolled
chrance wrtace

Minu hterfemmete
amspliier plate
...._...:.
:: .'..I ;

fbrtsurtsce

Figure 3.10. Phase-shft interferometer (Creath 1987).


46 Rough SurJaces

The vertical resolution is about 0.1 nm, but early realisations of this principle
were limited in vertical range to half the wavelength of the illuminating light, say
about 0.3 pm (Bennett & Mattsson 1989). Combining measurements made at
different wavelengths extended the vertical range to 15 pm (Creath 1987). In a
more recent development, the vertical range is extended still further by combining
phase-shift interferometry with what is described iis vertical scanning
interferometry (Caber et al. 1993). The visibility of fringes from a white-light
source drops off rapidly from its maximum value at minimum optical path
difference. If the interferometer is translated axially, these maxima, and hence the
local height, can be extracted on a point-by-point basis by signal processing. A
typical commercial realisation of this technique claims horizontal range and
resolution of 0.2 mm and 0.4 pm respectively, vertical range and resolution of 150
pm and 1 nm. The maximum measurable slope at this magnification is stated to be
14". A somewhat similar system, rather misleadingly described as "coherence
radar", has been described by Hausler & Neumann (1992).

3.2. Parametric Techniques

Scattering is a subject of interest to optical engineers, radar engineers and many


others, and has an extensive literature of its own. The present account will only
attempt to cover it in sufficient depth to treat its application to roughness
measurement. The mathematical foundations of scattering theory are set out in
Beckmann & Spizzichino (1963) and more recently by Ogilvy (1991).
Experimental techniques are discussed at length by Bennett & Mattsson (1989).
The following account follows very closely the review of Vorburger & Teague
(1981).
When a beam of light is reflected by a rough surface, the intensity and pattern
of the scattered radiation depend on the roughness heights, the spatial wavelengths
and the wavelength of the light. In general, short and long surface wavelengths
diffract the light into large and small angles respectively relative to the specular
direction. For most surfaces, there is a broad spectrum of surface wavelengths, and
the light is therefore diffracted into a range of angles.
Five main mechanisms of interaction may be distinguished between rough
surfaces and electromagnetic radiation (Vorburger & Teague 1981). For surfaces
whose roughness is much less than the wavelength of the incident radiation, most
of the reflected light propagates in the specular direction. As the roughness
increases, the intensity of the specular beam decreases while the diffracted
Optical Instruments 47

radiation increases in intensity and becomes more diffuse. I a addition, the angular
distribution of diffuse radiation consists of a fine grainy structure called speckle,
whxh shows up as intensity contrast between neighbouring points in the scattered
field. Finally, the light wave may undergo a change in its polarization state upon
reflection from the surface. All of these phenomena, the relative intensity or
reflectance in the specular direction, the total intensity of the scattered light, the
diffuseness of the angular scattering pattern, the speckle contrast, and the
polarization, depend on the surface roughness, and all five have served as the bases
for potential surface measuring instruments.

3.2.1. Specular Reflectance

One way of assessing specular reflectance is to measure what is called image


clarity: the ability of the surface to reflect a clear image of a row of posts
(Elmendorf & Vaughan 1958) or of a grid (Westberg 1967168). To quantify this,
Halling (1954) designed an instrument in which the reflection of a series of vertical
bars in the surface is observed at the specular angle. This angle is gradually
decreased until the bars can no longer be distinguished. He found that the
roughness was inversely proportional to the cosine of the angle of extinction.

Figure 3.11. Dflerent designs of glossmeter and definitions of their associated measurements (Westberg
1967168).
48 Rough Sufaces

A simpler approach is to measure the intensity of the specular beam.


Commercial instruments following this general approach are sometimes called
glossmeters, though gloss is a tenn not easily defined; Westberg (1967/68) quotes a
review of 44 different definitions of gloss. He himself has reviewed the design,
construction and operation of a number of designs of glossmeter at some length,
and distinguishes six basic combinations of specular and diffuse measurement (Fig.
3.11).
The measurement of roughness relies on the inverse correlation which exists
between the specular reflectance p and the roughness Rq. For rougher surfaces (Rq
> ;iy / 10 ) the true specular beam effectively disappears so p is no longer
measurable. However, even in these circumstances there are a number of empirical
studies that show an inverse correlation between the light intensity scattered in the
specular direction and the surface roughness (Tanner & Fahoum 1976, Murray
1973, Spurgeon & Slater 1974). The main advantage of this method is speed. If
the instrument is properly calibrated, a single measurement immediately yields a
value for Rq. Therefore, the specular reflectance method is ideal for routine
comparisons of similar surfaces. The technique is capable of studying isotropic or
anisotropic surfaces and the ultimate vertical resolution is about 1 nm
(Cunningham et al. 1976).

Figure 3.12. Roughness measured by specular reflectance and stylus methoas by six different groups of
workers (adapted kom Birkebak 1971).
Optical Instruments 49

There are a number of disadvantages to techniques based on specular


reflection. Agreement with stylus measurements is not very good (Fig. 3.12). The
measured intensities must be normalized for each material under inspection.
Measurements are affected by the direction of the lay (Vashisht & Radhakrishnan
1974), though this dependence dlsappears at normal incidence (Ollard 1949).
Also, even for fairly small apertures the diffuse component will be significant
unless p/pois fairly close to unity. This implies that Rq must be much less than the
optical wavelength 4, which means that visible light is suitable only for measuring
the smoothest surfaces; for most engineering surfaces infrared radiation must be
used. A final drawback of this technique is that it is primarily a function of surface
amplitude and is not suitable for measuring surface wavelengths.

3.2.2. Total integrated scatter

The total integrated scatter (TIS) method is complementary to specular reflectance.


Instead of measuring the intensity of the specularly reflected light, one measures
the total intensity of the diffusely scattered light (Bennett & Mattsson 1989). TIS
has much the same strengths and weaknesses as specular reflectance, i. e. the
technique is fast, is based on an approximate theory, is practical only for surfaces
smoother than the wavelength of the incident radiation, and has similar bandwidth
limitations. The fundamental difference between the two is that the TIS is more
dlrectly related to Rq. TIS is a fundamental quantity for the functioning and
testing of optical components (Bennett et al. 1979). In many optical applications,
if the TIS can be measured accurately and routinely, the roughness need not be.
Using the technique for practical roughness measurements presents several
problems, however. To begin with, the experimental set-up is more elaborate than
that required for specular reflectance. Moreover, the bandwidth limitation arises
from the fact that the collecting optics normally contain some sort of aperture or
stop to prevent the specular beam from being detected along with the scattered
radiation. Such an aperture or stop inevitably also blocks a portion of the scattered
light and hence limits the surface wavelengths that can be detected (Bennett et al.
1979).
The net result is that the technique does not seem to have been successfully
applied to surfaces with Rq greater than about 10 nm. Even in the regime of very
smooth surfaces, the roughness results obtained from TIS, although self-consistent,
do not seem to agree well with measurements made by non-scattering techniques
(Fig. 3.13). It appears then that TIS is an important technique for rapid
50 Rough SurJaces

measurements of the optical scattering characteristics of very smooth surfaces with


short wavelengths. Like specular reflectance, TIS is not suitable for measuring
surface wavelength parameters.

=m
200

8 0 NWC TIS
0 BALZERS TIS
A LIVERMORE OHP

2t
0 N W C WYKO TOPO-20
A 0 NWC TALYSTEP

I I I I I I I I I
20 50 100 200
AVERAGE IVIS ROUGHNESS. TIS d l

Figure 3.13. Summary of measurements on five roughness standards made by: 0 0 , TIS instruments; AO,
optical profilers; 0,stylus instrument (Bennett & Mattsson 1989).

3.2.3. Angular Distributions

In principle, the entire angular distribution of the scattered radiation contains a


great deal of information about the surface topography. In addition to rms
roughness, measurements of the angular distributions (Tanner & Fahoum 1976,
Stover 1976, Thwaite 1979) can yield other surface parameters such as the average
wavelength or the average slope. The angle of incidence is normally held constant
and the angular distribution is measured by an array of detectors or by a movable
detector and is stored as a function I(@ where is the angle of scattering. A good
example of the resulting angular distributions for a diamond-turned specimen
illuminated by coherent light is shown in Fig. 3.14 (Church et al. 1977). The
upper curve shows the angular distribution in the plzne of incidence and
perpendicular to the predominant lay of the surface. It contains an intense specular
beam at 8, =0, a broad scattering distribution due to the random component of the
roughness, and a series of dwrete lines due to a periodic component of the
roughness caused by the feed rate of the diamond tool. The lower curve is an
Optical Instruments 51

angular distribution measured parallel to the lay direction and it shows another
broad distribution characteristic of the random roughness pattern in this direction.
In principle then, one can distinguish between effects due to periodic and random
roughness components and can detect the directional properties of surfaces.

6'
lo-'
.-
e lo-'
sB ,o-Q
lo-'
lo-*
lo-'
10.6
- 4 o ~ w - z I p o0 w 200 30" SOD
Scattering angle

Figure 3.14. Angular distribution of light scattered &om a diamond-turnedsurface (adapted from Church et
al. 1977). Upper curve, across lay; lower curve, along lay.

The kind of surface information that may be obtained from angular


Qstributions depends on the roughness regime. For Rq >>% and surface spatial
wavelengths il >> A,. one is working in the geometrical optics regime where the
scattering may be described as purely specular scattering from a series of facets
(Church 1979). The angular distribution is therefore related to the surface slope
distribution, and its width is a measure of the characteristic slope of the surface.
As Rq and il decrease, the distribution becomes a much more complicated
function of both surface slopes and heights and is difficult to interpret. Some work
has been done by Leader (1979), Smith & Hering (1970) and Chandley (1976) for
surfaces with Rq A., -Leader's comparisons between calculated and measured
angular distributions yield interesting, qualitative idormation about the
topography of painted and dielectric surfaces.
The most interesting regime is where Rq << because then it can be shown
theoretically that the angular distribution should directly map the power spectral
density G (2x14 of the surface topography. The bandwidth is limited at the low
end by the maximum scattering angle of 90" and at the high end by the minimum
scattering angle which can be measured with respect to the specular beam, about
0.5" in practice (Church 1979), thus the maximum surface wavelength is only
52 Rough Surfaces

about 100 times the minimum wavelength. The angular distribution technique
may be better suited to measuring wavelength parameters than roughness
amplitude parameters.
In summary, the scattering methods described so far are generally limited by
available theories to studies of surfaces whose roughness is much less than the
wavelength of the incident radiation. There are a few studies, mostly empirical,
which have pushed beyond this limit. With a HeNe laser as the light source, the
above constraint means that these techniques have been used mainly on optical
quality surfaces where Rq < 0.1 pm. Within that limited regime, they can provide
high speed quantitative measurements of the RMS roughqess of both isotropic
surfaces and those with a pronounced lay. With rougher surfaces, angular
distribution may be useful as a comparator for monitoring both amplitude and
wavelength surface properties.

3.2.4. Direct Fourier Transformation

Ribbens and Lazik (1968) have described a technique in which a transparent


replica of a rough surface is made. A photographic film is illuminated through the
replica and developed; the local transmittance of the developed film is a function of
the thickness at the corresponding position on the replica The power spectral
density of the film transmittance can be measured directly and hence the power
spectrum of the original surface can be determined.
Anderson (1969) proposed a modification of this technique in which the
replica itself is used as the object of the Fourier transforming lens. He pointed out
that the relationships assumed by Ribbens and Lazik are correct only for fairly
smooth surfaces. However, he suggested that rougher surfaces could be
accommodated by reducing the effective variations of optical path length by
interposing a liquid of appropriate refractive index, a device also proposed by
Nagata et al. (1973). Thwaite (1979) used a direct optical transform in reflection
on a number of periodic surfaces and reported good agreement between the power
spectra measured in this way and by stylus instruments.

3.2.5. Ellipsometry

Ellipsometry measures the change in the polarization state of a beam of light when
it is reflected from a surface. For an extensive review, see Azzam & Bashara
Optical Instruments 53

(1977). In traditional null ellipsometry the important quantities are the angle of
incidence of the beam of light and the rotational positions of the polarizing and
analysing elements in the light path that produce a null in the detector. These
measurements allow one to determine the ratio of the complex reflection
coefficients for the p- and s-components of the electromagnetic field.
Ellipsometry measurements of the index of refraction are sensitive to a
number of surface properties including composition; surface structurt such as
damage, defects or surface crystal faces; temperature; strain state ; and surface
roughness. Since the roughening of surfaces can significantly change the results
from the ellipsometer, thereby obscuring the other surface changes to be observed,
the question arises whether ellipsometry can be used to measure the surface
roughness of engineering surfaces directly. Investigations so far have been largely
empirical and they are in disagreement.

Figure 3.15. Variation of Lonardo's(1978) parameter a with roughness for various angles of incidence

Lonardo (1978) studied ellipsometry as a potential tool for in-process


detection of surface roughness during manufacture. He worked with ground and
polished steel surfaces of roughness, measured by stylus techniques, ranging
approximately from 0.01 pm to 1.1 pm, and a derived ellipsometry parameter was
found to vary almost linearly with Ra (Fig. 3.15). Overall results show that
ellipsometry parameters cannot be related simply to roughness and slope.
Vorburger and Ludema (1980) and Williams et al. (1988) reached a similar
conclusion.
If the key problem of quantlfication could be solved, ellipsometry would have
several strengths as a surface measurement technique. The sensitive roughness
range of the technique seems to be approximately 0-1 pm, which is in the range of
54 Rough Sudaces

interest for machined surfaces. Also, ellipsometry measurements should not be


sensitive to fluctuations in the scattered-light intensity due to surface vibration
since the technique measures the polarization state rather than the intensity of the
scattered light.

3.2.6. Speckle

When a rough surface is illuminated with partially coherent light, the reflected
beam consists in part of random patterns of bright and dark regions known as
speckle. These patterns can be interpreted in terms of Huygen's principle whereby
the intensity at a field point is caused by the interference of wavelets, scattered
from different points within the illuminated area, with their phases randomized by
height variations of the surface. The spatial pattern and contrast of the speckle
depend on the optical system used for observation, the coherence condition of the
illumination, and the surface roughness of the scatterer. A review (Briers 1993)
cites more than 100 references on roughness-related aspects of speckle.
Two broad classes of measurement methods for determining the roughness
properties of a surface from speckle patterns can be discussed: speckle contrast and
speckle pattern decorrelation. In both methods the roughness properties are
obtained from the speckle patterns by empirically relating either the contrast or the
degree of pattern correlation to the roughness of the surface under study. The
empirical relationship is then interpreted with first order theories of speckle pattern
formation, which generally assume that only single scattering of the
electromagnetic wave takes place and that the scattering surfaces can be
characterized by a Gaussian distribution of heights with a correlation length much
less than the dimensions of the scattering region. It is doubtful that these
assumptions are fully satisfied in most practical scattering problems. The results
given below, however, demonstrate that withm these two classes of measurement
techniques are methods to determine the roughness of a surface over the large
range of 10 nm to 30 pm with a reasonable degree of confidence.
In speckle contrast measurements the intensity variations are quantified in
terns of an average contrast defined as the normalized standard deviation of
intensity variations at the observation plane. The intensity variations are
determined by either moving the specimen and thereby the speckle pattern past a
fixed detector with an aperture smaller than the speckle size, or by moving the
detector through the speckle field. High-contrast speckle patterns are produced
when all the interfering wavelets have sufficient phase difference ( > 271 ) to give
Optical Instruments 55

complete destructive interference at some points in the pattern and the illumination
has a h g h degree of spatial and temporal coherence. Spatial coherence means that
the phases of the electromagnetic field at two points spaced across the propagating
wavefront are highly correlated, and temporal coherence means that the phases of
the field at two points spaced along the direction of light propagation are highly
correlated.
Speckle contrast is unity for fully coherent monochromatic light illuminating
a surface whose roughness is much larger than 4 so that the wavelet phases are
uniformly distributed over the interval from 0 to 27c. Correspondingly, for coherent
monochromatic illumination, as the reflecting surface becomes smoother and less
complete destructive interference occurs, the contrast V decreases toward zero.
Experiments to relate surface roughness to the contrast of speckle patterns
produced by coherent monochromatic illumination (Fig 3.16) showed that a strong
linear correlation exists between V and Ra determined by stylus profilometry for Ra
values up to 0.13 pm.

0 m o D a J 3 a a Q D Q y I

2Ra (microns)

Figure 3.16. Maximum average contrast of speckle intensity variations as a hnction of roughness for
surfaces of various metals and finishes (Vorburger & Teague 1981).

The second broad class of techniques for relating surface roughness and
speckle is speckle pattern decorrelation measurement. Here two speckle patterns
are obtained from the test surface by illuminating it with different angles of
incidence or different wavelengths of light. Correlation properties of the speckle
patterns are then studied by recordmg the patterns on the same photographic plate
by double exposure or by photoelectric detection of the two patterns. The primary
attribute of this type of speckle measurement is that Rq values as large as 30 to 50
pm can be measured. Fujii & Lit( 1978) applied a speckle decorrelation technique
56 Rough Sur$aces

to the measurement of a range of ground glass and ground metal surfaces. They
found good correlation between roughness deduced from correlation measurements
and roughness measured by stylus instruments over a range of roughness from 0.13
pm to 6 pm (Fig. 3.17).

Figure 3.17. Roughness deduced &om speckle decorrelation measurementscompared with stylus roughness
measurements for glass (circles) and metal (triangles) surfaces (Fujii & Lit 1978).

As the examples of this subsection have demonstrated, speckle patterns are


rich in information about the microtopography of a test surface, though the field
has not yet yielded techniques for obtaining characterizations of roughness other
than the Rq value. The range of Rq values measurable with speckle techniques and
their apparent insensitivity to the type of material and type of surface forming
process indicate that these techniques have a high potential for roughness
measurements. However, Briers (1993) has noted that speckle techniques have not
in general been converted into practical instruments; he describes them as a
"solution in search of a problem".

3.3. References

Anderson, W. L., "Surface roughness studies by optical processing methods",


Proc. I.E.E.E., (Letters), 57, 95 (1969).
Optical Instruments 57

Azzam, R. M. A,, and Bashara, N. M., Ellipsometry and polarised light


(North Holland, Amsterdam, 1977)
Bailey, W., "Optical inspection of cylinder bores", Trib. Int. 10, 319-322
( 1977)
Beckmann, P. and Spizzichino, A,, The scattering of dectromagnetic waves
porn rough surfaces (Pergamon Press, Oxford, 1963).
Bennett, J. M., Burge, D. K., Rahn, J. P., Bennett, H. E., "Standards for
optical surface quality using total integrated scattering", Proc. SPIE 181, 124-128
(1979)
Bennett, J. M., and Mattsson, L., Introduction to surface roughness and
scattering (Opt. SOC.Am., Washington, 1989)
Biegen, J. F. and Smythe, R. A,, "High-resolution phase-measuring laser
interferometric microscope for engineering surface metrology", Surface
Topography, 1, 287-299 (1988)
Birkebak, R. C., "Optical and mechanical RMS surface roughness
comparison", Appl. Opt. 10, 1970-1979 (1971)
Briers, J. D., "Surface roughness evaluation", in Speckle metrology, R. J.
Sirohi ed., (Marcel Dekker, New York, 1993).
Bristow, T. C., "Surface roughness measurements over long scan lengths",
Surface Topography, 1, 281-285 (1988)
Brown, A. J. C., "Rapid optical measurement of surfaces", Int. J. Mach. Tool
Munufact. 35, 135-139 (1995)
Caber, P. J., Martinek, S. J., Niemann, R. J., "A new interferometric profiler
for smooth and rough surfaces", Proc. SPIE 2088 (1993)
Chandley, P. J., "Determination of the autocorrelation function of height on a
rough surface from coherent light scattering", Opt. Quantum Electron. 8, 329-333
(1976).
Church, E. L., "The measurement of surface texture and topography by
differential light scattering", Wear, 57, 93-105 (1979).
Church, E. L.; Jenkinson, H. A.; Zavada, J. M., "Measurement of the finish
of diamond-turned metal surfaces by differential light scattering", Opt. Eng. 16,
360-374 (1977).
Church, E. L.; Jenkinson, H. A.; Zavada, J. M., "Relationship between
surface scattering and micro-topographic features", Opt. Eng. 18, 125-131(1979).
Creath, K., "Step height measurement using two-wavelength phase-shifting
interferometry", Applied Optics 26, 2810-2815 (1987)
58 Rough &$aces

Cunningham, L. J., and Braundmeier, A. J., "Measurement of the correlation


between the specular reflectance and surface roughness of Ag films", Phys. Rev.
14B, 479-488 (1976)
Dainty, J. C . , "The statistics of speckle patterns", in E. Wolfed., Progress in
Optics 14 (North Holland, Amsterdam, 1976)
Davies, T.; Kun, X; Luxmoore, A. R., "Digital measurement of surface
profiles by automated optical sectioning", Measurement Science & Technology 5,
710-715 (1994)
Downs, M. J., Mason, N. M., Nelson, J. C. C., "Measurement of the profiles
of super smooth surfaces using optical interferometry", Proc. SPIE 1009, 14-17
(1989)
Downs, M. J., McGivern, W. H., Ferguson, H. J., "Optical system for
measuring the profiles of super-smooth surfaces", Prec. Engng. 7, 2 11-2 15 (1985)
Dupuy, O., "High-precision optical profilometer for the study of micro-
geometrical surface defects",. Proc .I .Mech .E., 182, Part 3K, 255-259 (1967/68).
Edwin, R. P., "Light scattering as a technique for measuring the roughness of
optical surfaces", J. Phys. E6, 55-59 (1973)
Elmendorf, A. and Vaughan, T. W., "A survey of methods of measuring
smoothness of wood", Forest Products J.,8, 275-282 (1958).
Elson, J. M., and Bennett, J. M., "Relation between the angular dependence
of scattering and the statistical properties of optical surfaces", J. Opt. SOC.Am. 69,
31-39 (1979)
Elson, J. M., Rahn, J. P., Bennett, J. M., "Light scattering from multilayer
optics: comparison of theory and experiment",Appl. Opt. 19, 669-675 (1980)
Ennos, A. E. and M. S. Virdee, "High accuracy profile measurement of quasi-
conical mirror surfaces by laser autocollimation",Precis. Engng. 1, 5-8 ( I 982)
Farago, F. T., Handbook of dimensional measurement 2e (Industrial Press,
New York, 1982)
Fujii, H., and Lit, J. W. Y., "Surface roughness measurement using
dichromatic speckle pattern: an experimental study", Appl. Opt. 17, 2690-2705
(1978)
Halling, J., "A reflectometer for the assessment of surface texture", J. Sci.
Instrum., 31, 3 18-320 (1954).
Hamilton, D. K.; Wilson, T., "Three-dimensional surface measurement using
the confocal scanning microscope", Appl Phys 27,2 1 1-2 13 (1982).
Hard, S., and Nilsson, O., "Laser heterodyne apparatus for roughness
measurements of polished surfaces", Appl. Opt. 17, 3827-383 1 (1978)
Optical Instruments 59

Hausler, G., and Neumann, J., "Coherence radar - an accurate 3D sensor for
rough surfaces", Proc. SPIE 1822, 200-205 (1992)
Howes, V. R., "An angle profile technique for surface studies",
Metallography 7, 43 1-440 (1974).
Johnson, F.; Brisco, B.; Brown, R. J., "Evaluation of limits to the
performance of the surface roughness meter", Canadian Journal of Remote Sensing
19,140-145 (1993)
Kayser, J. F., "Optical cut method for the determination of surface
roughness", Foundv Trade J., 70, 137-138 (1943)
Keller, B. E., "Non-contact surface contour analyser", Proc. 1. Mech. E., 182,
Part 3K, 360-367 (1967/68).
Lange, D. A.; Jennings, H. M.; Shah, S. P., "Analysis of surface roughness
using confocal microscopy", Journal of Materials Science 28,3879-3884 (1993)
Leader, J. C., "Analysis and prediction of laser scattering from rough-surface
materials", J. Opt. SOC.Am. 69, 610-619 (1979)
Lonardo, P. M., "Testing a new optical sensor for in-process detection of
surface roughness", Ann. CZRP 27, 53 1-533 (1978)
Murray, H., "Exploratoly investigation of laser methods for grinding
research", Ann. CIRP 22, 137-139 (1973)
Nagata, K., Umehara, T. and Nishiwaki, J., "The determination of RMS
roughness and correlation length of rough surface by measuring spatial coherence
function", Japan. J. Appl. Phys., 12, 1693-1698 (1973).
Ogilvy, J. A,, Theory of wave scatteringfrom random rough surfaces (Adam
Hilger, Bristol, 1991)
Ollard, E. A., "Surface reflectometer for evaluating polished surfaces", J.
Electrodepos. Tech. SOC.,24, 1-8 (1949)
Olsen, W. S. and Adams, R. M., "A laser profilometer", J. Geophysical Res.,
75,2185-2187 (1970).
Parry, G., "Some effects of surface roughness on the appearance of speckle in
polychromatic light", Opt. Comm. 12, 75-81 (1974)
Pettigrew, R. M., and Hancock, F. J., "An optical profilometer", Proc.
NELEX Conf: (Nat. Engng. Lab., Glasgow, 1978)
Ramgulam, R. B.; Amirbayat, J.; Porat, I., "Measurement of fabric roughness
by a noncontact method", Journal of the Textile Institute 84,99-106 (1993)
Ribbens, W. B. and Lazik, G. L., "Use of optical data processing techniques
for surface roughness stuhes", Proc. I.E.E.E. ('Letters) .56, 1637-1638 (1968).
60 Rough Surfaces

Sandoz, P., Tribillon, G., Gharbi, T., Devillers, R., "Ruughness measurement
by confocal microscopy for brightness characterisation and surface waviness
visibility evaluation", Wear 201, 186-192 (1996)
Sayles, R. S., Wayte, R. C., Tweedale, P. J. and Briscoe, B. J., "The design,
construction and commissioning of an inexpensive prototype laser optical
profilometer", Surface Topography, 1 , 219-227 (1988)
Shaw, M. C. and Peklenik, J., "A light projection technique for studying
surface topology", Ann. C.Z.R.P.,12, 93-7 (1963).
Smith, T., "Effect of surface roughness on ellipsometry of aluminium", Surf:
Sci. 56, 252-259 (1976)
Smith, T. F. and Hering, R. G., Tomparison of bidirectional reflectance
measurements and model for rough metallic surfaces", Proc. 5th Symp.
Thermophys. Properties, 429-435 (ASME, New York, 1970).
Sommargren, G . E., "Optical measurement of surface profile", Precis. Eng. 3,
131-136 (1981a)
Sommargren, G . E., "Optical heterodyne profilometry", Appl. Opt. 20, 610-
618 (1981b)
Spurgeon, D. and Slater, R. A. C., "In-process indication of surface
roughness using a fibre-optics transducer", Proc. of the 15th Int. Machine Tool
Des. & Res. Conf.',Birmingham, 339-347 (1974).
Stover, J. C . , "Spectral-density function gives surface roughness", Laser
FOCUS12,83-85 (1976).
Tanner, L. H. and Fahoum, M., "A study of the surfacc parameters of ground
and lapped metal surfaces, using specular and diffuse reflection of laser light",
Wear, 36,299-3 16 (1976)
Teague, E. C., Vorburger, T. V., Maystre, D., Young, R. D., "Light scattering
from manufactured surfaces", Ann. C I W 30,563-569 (1981).
Thwaite, E. G., "The direct measurement of the power spectrum of rough
surfaces by optical Fourier transformation", Wear, 57, 71-80 (1979).
Thwaite, E. G., "The roughness of surfaces",Australian Physicist (November
1977)
Tolansky, S . , Multiple-beam interference microscopy of metals (Academic
Press, London, 1970a).
Tolansky, S . , Multiple-beam interferometry of surfazes and films (Dover
Publications, Inc., New York, 1970b).
Tolansky, S., Surface microtopography (Longmans, London, 1960).
Vashisht, S. K. and Radhaknshnan, V., "Surface studies with a gloss meter",
Tribology Znt., 7, 70-76 (1974).
Optical Instruments 61

Vorburger, T. V., and Ludema, K. C., "Ellipsometry of rough surfaces", Appl.


Opt. 19, 561-569 (1980)
Vorburger, T. V., "Methods for characterising surface topography", in
Tutorials in Optics, 137-151 (Opt. Soc. Am., Washington, 1992)
Vorburger, T. V.; Teague, E. C., "Optical techniques for on-line
measurement of surface topography", Precis. Eng. 3,61-83 (1981).
Way, S., "Description and observation of metal surfaces", Proc. Con$ on
Friction & Surface Finish, 2e, 44-75 (MIT, Cambridge, 1969).
Wehbi, D. and C. Roques-Carmes, "Physical limitations of optical
defocussing technique," Wear 109, 287-295 (1986)
West, R. N., and Stocker, W. J., "Automatic inspection of cylinder bores",
Metrology &Inspection 9, 9-10 (1977)
Westberg, J., "Development of objective methods for judging the quality of
ground and polished surfaces in production", Proc .I .Mech .E., 182, Part 3K, 260-
273. (1967/68).
Whitehouse, D. J., "Modern trends in the measurement of surfaces", Rev. M.
Mec., 21, 19-28 (1975).
Whitehouse, D. J., Handbook of surface metrology (Institute of Physics,
Bristol, 1994)
Williams, M. W., Ludema, K. C. and Hildreth, D. M., "Mueller matrix
ellipsometry of practical surfaces", Surface Topography, 1, 357-372 (1988)
Wyant, J. C., Koliopoulos, C. L., Bhushan, B., Basila, D., "Development of
a three-dimensional noncontact dlgital optical profiler", Trans. ASME: J. Trib.
108, 1-8 (1986)
Young, R. D.; Vorburger, T. V.; Teague, E. C., "In-process and on-line
measurement of surface finish", Ann. CIRP 29,435-440 (1980).
CHAPTER 4

OTHER MEASUREMENT TECHNIQUES

We have considered stylus and optical methods in some detail because equipment
using these methods comprises the major part of the installed base of roughness
measuring instruments. However, a large number of other techniques have been
used for the measurement of surface roughness. Some are mainly of historic
interest, but may still be worth studying because the principles involved might
suggest an application to some measurement problem insoluble by other means.
Others are novel and still under development, and cover ranges or offer other
special features which complement stylus and optical methods. We will continue
to make the convenient distinction between profiling techniques, which can yield
point-by-point information about the surface topography, and parametric
techniques, which give directly some average measure of the surface roughness.

4.1. Profiling Methods

4.1.1. Taper Sectioning

In taper sectioning, as its name implies, a section is cut through the surface to be
examined at a shallow angle, thus effectively magnifying height variations by the
cotangent of the angle, and subsequently examined by optical microscopy. The
technique was first described by Nelson (1969) and has since been employed by a
number of other workers (Broadston 1944; Tarasov 1945; Darmody 1946; Shaw
and Peklenik 1963; Dorinson 1965). Practical details are dlscussed at length by
Nelson and by Rabinowicz (1950).
It is necessaly to support the surface to be sectioned with an adherent coating
which will prevent smearing of the contour during the sectioning operation. This
coating must adhere firmly to the surface; must have a similar hardness; should not
dlfise into the surface; and should not be affected by any subsequent etching. For
steel these requirements are met by electroplating with nickel to a thickness of 0.5
mm. The specimen is then ground on a surface grinder at an angle of between 1

63
64 Rough Surfaces

and 6 degrees, depending on the required magnification, till the interface between
coating and substrate has advanced halfway along the specimen. The taper section
so produced is lapped, polished and possibly finally lightly etched or heat tinted to
provide good contrast for the optical examination.
The main advantage claimed for taper sectioning is its accuracy; indeed Shaw
and Peklenik (1963) have gone so far as to describe it as 'probably the most
accurate method that has ever been devised for studying the profile of a surface'.
This seems rather an excessive claim for a technique whose vertical resolution is
admitted by the Same authors to be only 0.25 pm. Great play is made, however, by
Tarasov (1945) among others, of the ability of the method to show deep scratches
which a stylus will not penetrate. The only measurement which can conveniently
be made from a micrograph of a taper section is the peak-to-valley height, and
Tarasov compares this measurement with the RMS roughness found by a stylus
instrument for a number of surfaces. According to other results quoted by Shaw
and Peklenik, a comparison of taper-section profiles with those of a stylus
instrument revealed larger peak-to-valley roughness in every case, up to a
maximum of 100 per cent discrepancy.
This is not surprising when the integrating effect of taper sectioning is taken
into account. Tarasov quotes peak-to-valley roughness of between 1 pm and 5 pm
at a relative vertical magnification of 25. His sections must therefore have
represented profiles of effective width from 25 pm to 125 pm. This compares with
8 pm for the width of a typical stylus (Jungles and Whitehouse 1970), which
according to Guerrero and Black (1972) normally 'sees' an even narrower strip as it
tends to ride on one edge only. Taper sectioning is thus equivalent, at a
conservative estimate, to measuring 3-5 profile lengths from a stylus instrument.
As the greater the profile length the higher the probability of encountering a high
peak or deep valley, it is small wonder that the taper section gives a larger
measurement. The other disadvantages of the technique are too obvious to need
comment; we have discussed it at this length mainly because of the inflated claims
made for its accuracy.

4.1.2. Electron Microscopy

Electron microscopy is thought of as primarily a technique for visualisation, but the


very short wavelengths of electron beams offer the possibility of very high
resolutions for quantitative work. Transmission electron microscopy (TEM),
dealing as it does with specimens thin enough to pass an electron beam, is of little
Other Measurement Techniques 65

interest from the point of view of roughness measurement. Scanning electron


microscopy (SEM) is at first sight more promising. An electron beam incident on
the surface excites the emission of secondary electrons which are then detected. As
with the STM, specimens must be electrically conducting and measured under
vacuum. However, SEM has some vely attractive potential advantages. The depth
offield is unusually large: features a few Angstroms high stand out plainly, and the
vertical range is of the order of mm. The horizontal resolution is limited only by
the beam diameter to a few nm. The beam is deflected virtually instantaneously by
field coils, obviating the need for mechanical translation. The steepest slope which
can be measured is about 15 degrees (Whitehouse 1994). For a general review of
electron microscopy see Grundy & Jones (1976).
The resulting intensity signal bears a striking resemblance to a profile
measured by a stylus instrument. Unfortunately this resemblance is more apparent
than real. Secondary electrons originate from regions 100 nm or so below the
specimen surface, and local convexities increase the local electron flux,
exaggerating topographic features. Point-by point comparisons of secondary
electron images with stylus measurements (Samuels et al. 1974) show rather poor
agreement. It is possible to extract true height information by analysing stereo
pairs (Howell & Bayda 1972, Lee & Russ 1989), but this is laborious and may lose
much of the potential high resolution.
A number of workers have tried to get around this difficulty in other ways.
One technique of SEM analysis employed a modified detector which followed the
line-of-sight properties of back-scattered electrons (McAdams 1974). These
electrons, which travel in straight-line trajectories, are thresholded according to
direction and produce an electron optical sectioning of the surface along the critical
trajectory direction. Surface elevation is therefore recorded as variations in
detector position, as opposed to signal intensity in the standard SEM. A multiple-
detector technique (Lebiedzik & White 1975) produced reasonable agreement
between measurements of average roughness. Holburn & Smith (1982) employed
an autofocussing technique with a claimed vertical resolution of 1 pm. Myshkin
and his co-workers (Kholodilov et al. 1985, Grigor'ev et al. 1988) have obtained
roughness parameters from SEM measurements of secondary electron emission,
but it is not clear whether they were able to reconstruct a true profile.
Rasigni et al. (1981) have studied transmission electron micrographs of
carbon replicas of calcium fluoride films and other surfaces, using a
microdensitometer. They show that the micrograph transmittance is approximately
proportional to the slope of the surface elements, which enables determination of
the surface profile by integration of the microdensitometer data. This technique is
66 Rough Sugaces

able to measure roughnesses of 2 nm with a lateral resolution of better than 1 nm


over an area of about 1 pm x 1 pm.

4. I.3. Capacitance

Capacitance techniques are discussed in more detail in the section on parametric


techniques, but there are two designs of capacitance-based profilers which will be
discussed here. Garbini et al. (1988) use a so-called fringe-field technique. A thin
electrode held normal to the specimen is translated in a direction parallel to the
plane of the electrode (Fig. 4.1). Lateral range is 6 mm; vertical range is not stated
but appears to be about 10 pm.

Figure 4.1. Fringe-field capacitance probe (Garbini et al. 1988)

Although the electrode is only 0.3 pm wide, the capacitance between it and
the specimen is influenced by regions of the specimen adjacent to the electrode in a
manner analogous to the weighting function of a low-pass filter, so it is rather
difficult to determine the lateral resolution. The equivalent stylus width is the
length w of the electrode (about a millimetre in the practical realisation), so
"profile" measurements are only meaningful on a surface produced by a process
such as shaping or turning which is basically two-dimensional. In spite of these
limitations, the instrument agrees with stylus measurements over a restricted range
of roughness (Fig. 4.2) and appears to work well at relatively high (25 m d s )
Other Measurement Techniques 67

translation speeds. The fringe-field capacitance technique has also been used for
parametric measurement (Nowicki & Jarkiewicz 1997).

- 3

'I
0 I 2 3 4 5

Surface roughness. RB Cm)

Figure 4.2. Maximum, average and minimum roughnesses for various test samples (Garbini et al. 1988).
(0)Stylus instrument; (x) fringe-field profilometer.

An instrument of much higher resolution is described by Bugg & King


(1988). In the scanning capacitance microscope the electrode is a fine vertical
wire. Most of the capacitance between the electrode and the specimen is due, as in
the fringe-field device, to the surrounding field, but the effect of this field is
ingeniously removed by vibrating the wire in a vertical plme and measuring the
differential capacitance. In practice a servo arrangement is used to keep the
separation of transducer and specimen constant during translation. The height
variation is given by the servo signal, thus obviating the inherent non-linearity of
the transducer (Fig. 4.3). In a commercial realisation of this instrument (Bugg
1991), lateral range and resolution are given as 26 mm and 10 pm respectively.
The vertical resolution is O.lpm, and the vertical range is said to be 5 mm,
presumably with the aid of some vertical translation device.
68 Rough Surfaces

Figure 4.3. Scanning capacitance microscope ( B u g & King 1988)

4.1.4. Scanning Microscopies

Although the techniques described below are all basically profiling techniques,
they are generally used in a raster scan mode to make area measurements. The
lateral displacements required are too small for conventional mechanisms
requiring relative motion between their components, so piezo drives are used for
translations in all three axes. Piezo drives are well suited for this, but are limited
in lateral range to some fraction of a millimetre. They are also susceptible to
hysteresis, which can then appear as an apparent form error.
T h s wide variety of techniques shares many common elements (Teague
1988): servo control of tip-specimen spacing to maintain constant reaction; precise
mechanical scanning of the tip with respect to the specimen; high sensitivity of the
output to tip-specimen spacing, requiring stiff microscope structures and isolation
from mechanical noise; lateral resolution determined by tip dimensions, with a
resultant emphasis on the problem of probe formation.
If a conducting probe is placed very close to a conducting surface a small
potential difference across the gap will encourage electrons to cross the gap by
quantum tunnelling. The resulting current is highly sensitive to the width of the
gap. As the probe is translated across the rough surface the width of the gap, and
thus the tunnelling current, changes. This is the principal of the scanning
tunnelling microscope (STM) (Binnig & Rohrer 1986) (Fig. 4.4).
Other Measurement Techniques 69

Figure 4.4. Principle of STM (Binnig & Rohrer 1986)

There are two possible modes of operation (Fig. 4.5). The probe can be
servoed to maintain a constant gap as it is translated, in which case the restoring
servo voltage is a measure of the local height. This is relatively slow but can more
easily follow the rougher surfaces. Alternatively, the probe can be maintained at a
constant height and the change in tunnelling current can be measured. This is
quicker but works best on smooth surfaces.

CONSTANT CURRENT Y M E
I CONSTANT NIGHT YXIE

Figure 4.5. Alternative modes of operation of STM (adapted from Hansma & Tersoff 1987)

The vertical resolution is a few Angstroms, and if a sufficiently fine probe is


used the horizontal resolution is in principle atomic. The vertical range is limited
in practice by the exponential fall-off in the tunnelling current as the gap increases.
Also, the specimen must be in a vacuum chamber and must be of a conducting
material. Within these limitations the S T M is a very powerful and flexible
70 Rough Surfaces

instrument and many commercial versions are available. Fu et al. (1992) have
described an STM with a lateral range of 0.5 mm and lateral resolution of 1 nm. A
recent review (DiNardo 1994) lists 400 or so references to STM and related
techniques.
In the atomic force microscope (AFM) (Binnig & @ate 1986), a probe
mounted on a cantilever is repelled by the van der Waal's forces as it travels over
the rough surface. The force deflects the cantilever and the deflection is sensed,
either by an STM (Binnig & Quate 1986) or by the angular displacement of a
reflected laser beam (Fig. 4.6), giving the added amplification of an optical lever
(Alexander et al. 1989). Either repulsive forces or attractive electrostatic forces
can be used, sometimes in the same instrument. Again the measured force may
either be recorded directly or used as the control parameter for a feedback circuit
which maintains the force at a constant value (McClelland et al. 1987).

Figure 4.6. AFM with optical lever mounted on the cantilever (Alexander et al. 1989)

The AFM avoids two disadvantages of the STM, its restriction to conducting
specimens and the requirement for a hard vacuum. Note that although the AFM
appears at first sight to be a kind of small-scale stylus instrument, the probe does
not actually make contact with the surface. Translation arrangements, ranges and
resolutions are similar to those of the STM, and the AFM is also commercially
available in a number of models.
In the scanning near-field acoustic microscope (SNAM), the friction of the
air and other damping effects in the small gap between a vibrating tip and the
measured surface change the frequency of vibration (Goch & Volk 1994). A
standard diamond tip fixed to the constantly excited tuning fork from a wristwatch
Other Measurement Techniques 71

is guided along the surface at a separation of about 100 nm. A servo moves the
fork up and down to maintain constant frequency, hence constant separation (Fig.
4.7), so the servo signal gives the varying surface height. Vertical resolution is
about 1 nm, but lateral resolution is limited to about 0.5 l m by the radius of
curvature of the diamond tip.

,nm I F 4- tm inspected surface

Figure 4.7. Schematic of an acoustic microscope (Goch & \ olk 1994)

DiNardo (1994) describes a number of other techniques operating on a


similar scale to STM and AFM and using similar translation systems (Table 4.1).
These all yield topographic information of some kind so could in principle be used
to measure roughness.

4.2. Parametric Methods

4.2.1. Mechanical Methods

The basic idea of a tactile test is that a probe of some kind is run across the surface
to be measured and the friction between the surface and the probe is compared with
that from a similarly machined surface of known roughness. The simplest and
cheapest probe is the human fingernail, and it is surprisingly effective. Indeed one
American engineer (Broadston 1947) waxed lyrical at the thought that every
machinist carried $100,000 worth of surface-measuring equipment about his
person, i.e. 10 fingers at $10,000, the cost of a good stylus instrument, each.
The human fingernail is more sensitive to some frequencies than to others
(Abbott & Goldschmidt 1937), so there is presumably an optimum speed with
which it should be drawn along the surface. Schlesinger (19 12) performed some
72 Rough Surfaces

Table 4.1. Near-field surface characterisationprobes (adapted from DiNardo 1994).

Instrument Measurement Lateral Reference


principle resolution
......
Scanning thermal surface < 30 nm Williams &
profiler (STP) temperature Wickramasinghe
(1986)

Scanning chemical thermoelectric < 1 nm Williams &


potential microscope voltage Wickramasinghe
(SCPM) (1991)

Optical absorption effects of light < 1 nm Weaver et al. (1989)


microscope (OAM) absorption

Scanning ion ion current < 0.1 nm Hansma et al. (1989)


conductance microscope
(SICM)

Laser force microscope probe vibration N.A. Whitehouse (1994)


(LFM) amplitude
Attractive force van der Waals < 100 nm DiNardo (1994)
microscope force

Charge force microscope electrostatic < 200 nm Whitehouse (1994)


( C M , EFM) force

Magnetic force magnetic force < 100 nm Whitehouse (1994)


microscope @EM)

Scanning near-field near-field < 25 nm Pohl et al. (1984)


optical microscope optical
(SNOM or NSOM) reflection

Photon scanning near-field 50 nm Reddick et al. (1990)


tunnelling microscope optical
transmission
Other Measurement Techniques 73

careful tests in which subjects were asked to differentiate between pairs of test
pieces of increasingly different roughness. He found that for some finishes
differences in roughness of as little as 20 per cent could be detected by the majority
of his test panel (Fig. 4.8). A similar experiment by Haesing (1961) found a
correlation between the subjects' assessments and stylus readings which, not
unexpectedly, was stronger for peak height than for average roughness.

30

20

0-025 0.05 0.1 0.2 0.4 0.8 1.6 3.2 6.4 I3 25

Ra (microns)

Figure 4.8. Tactile comparison (Schlesinger 1942). Percentage difference in roughness which could be
assessed by 9 out of 10 testers: A, lapped, honed and ground B, milled C, turned and shaped.

This technique is of some scientific interest as it appem to be sensitive only


to a narrow band of wavelengths, namely to those corresponding to the thickness of
the probe. The maximum height difference detectable is presumably set by the
protrusion of the nail beyond the fingertips while the minimum is set by the
sensitivity of the human nervous system. It is quicker, cheaper and simpler than
any other method providing a suitable range of reference specimens is available,
but to give reliable results a fair amount of experience is probably necessary on the
part of the tester. Watanabe & Fukui (1995, 1996) have shown that the subjective
sensation of roughness is affected by vibrating the measured surface at ultrasonic
frequencies.
A development is the Mecrin tester, where a thin flexible steel blade is
pushed along the surface at a gradually increasing angle till it buckles (Rubert
1967/68). As the device presumably relies on just failing to overcome static
friction it must be sensitive to the mean slope. It is not clear what is gained in
performance over the fingernail for the extra complexity.
74 Rough Su8aces

Another friction method is based on the retardaticn of the swing of a


pendulum due to friction between the tested surface and a smooth shoe attached to
the pendulum. The pendulum is released from a position 30 degrees from the
vertical and the apparatus simultaneously begins to eject a paper tape at constant
speed. The length of tape ejected before the pendulum comes to rest is taken as a
measure of the finish of the test piece (Jost 1944). Dynamic friction is influenced
by at least two surface parameters, RMS roughness and mean slope (Hirst &
Hollander 1974), so the instrument must be used with reservations even as a
comparator. The long-wavelength cut-off will depend on the nominal contact area,
which in turn will depend on the load and the geometry of the contacting surfaces.
The vertical resolution will depend ultimately on the finish of the shoe. The
apparatus is easy and quick to operate but rather expensive to build; its
measurements will also be affected by the cleanliness of the test surface.
In a method employing spherical contact a ball of radius r rolls down an
inclined plane as soon as the angle a of tilt exceeds a value which increases with
the roughness of the planes (Bikerman 1970). If Rp is the peak-to-valley
roughness,

Rp = r(1 - cos a)

This relation is stated to hold approximately for surfaces rougher than 1 pm.
Although this method makes use of the phenomenon of static friction the equation
shows that it is independent of material parameters. Assuming contact is elastic,
the area measured is probably not larger than 10 pm in diameter. The method
gives an absolute measurement. It is very cheap, robust and simple to use, and
might be suitable for production-line gauging, though it might be easier to set up in
the form of a rolling cylinder.
The thetameter is a device which presses a smooth steel sphere into the test
surface under a known load. The increase in load required to increase its
penetration by a fixed amount is measured (Tornebohm 1936). The 'theta' of its
title is the effective change in the Hertzian elastic modulus of the test piece due to
its roughness. A rigorous theory of the elastic contact of a sphere with a rough
plane was not developed until many years later by Greenwood and Tripp (1 967),
apparently in ignorance of the existence of an instrument bssed on this principle.
The effect of roughness prevails only at light loads, and then not in a simple
relationship with load. Asperity density and curvature are also involved. As the
instrument reading is the result of a combination of at least two surface parameters
Other Measurement Techniques 75

it is probably not suitable even for a comparison without exhaustive calibration


against a less ambiguous instrument using the range of surfaces to be tested.
On the other hand the thetameter is reasonably cheap, robust, quick and easy
to use and reproducible. Its sensitivity will decrease with decreasing roughness
and the limit of vertical resolution will be set by the load-measuring device. The
long-wavelength cutoff will depend on the Hertzian contact area. No dimensions
are given in the reference, but an estimate of 10 pm diameter for the contact area
seems plausible.
In another technique involving contact with a sphere, two 16 mm diameter
metal balls are mounted in a block of thermally insulating material (Fig. 4.9)
(Powell 1957). One is completely recessed while the other protrudes slightly.
They are connected by a differential thermocouple. The apparatus is placed in an
oven till it attains constant temperature and then removed and placed, under a load
of about 1 N, on the surface whose roughness is to be measured. The protruding
ball will cool faster because it is in contact with the test surface, and this is
quantified as a voltage reading from the differential thermocouple after a certain
time has elapsed from contact.

Phosphor-bronze balls
Balsa wood
m'
I I ,--

Figure 4.9. Thermal comparator (Powell 1957).

Clearly, this technique relies on far too many parameters, most of them
difficult to quantifj, to be suitable for an absolute determination of roughness. The
rate of cooling of the ball not in contact will depend on its initial temperature and
on various material and atmospheric properties. The increased rate of cooling of
the ball in contact will depend on the thermal conductance of the contact, which in
turn will depend on the thermal conductivities of the contacting materials and their
elastic moduli or relative hardness, depending on whether the contact is elastic or
76 Rough Surfaces

plastic, in addition to the surface properties and the load (Thomas & Probert 1972).
The relevant surface properties are probably the RMS roughness and the mean
slope (Thomas & Probert 1970).
The sensitivity increases as the test surfaces become smoother. The useful
upper roughness limit is probably about 4 pm RMS. The lower limit would
probably be set by the roughness of the balls themselves. The range of wavelengths
measured depends on whether contact is elastic or plastic, but the long-wavelength
cut-off for this particular instrument is probably about 10 pm. An instrument
based on this principle might be useful on a production line as a golno-go gauge,
possibly with a built-in heater; it would be relatively cheap and robust and very
simple to operate. Thermal comparators for general-purpose use have been
commercially available.
A scraping technique uses molten asphalt poured on to the surface whose
roughness is to be measured (Blkerman 1970). After cooling and solidlfication, the
excess is scraped off with a razor blade, leaving only the asphalt below the highest
peaks over the area A measured. Its volume V is determined and the ratio v/!
taken as RpI2. The ratio of VIA to the RMS roughness measured with a stylus
instrument was reported as between 2.4 and 1.9. The height resolution is
presumably set by the straightness of the edge of the razor blade. The method is
simple and cheap but rather tedious.

I 1 II
g 1.15

0.10 I
-

I I I
10
I
I
20
I
30
I
40

Test number r
Figure 4.10. Scatter in sand-patch measurements of roughness (Doty 1975)

On somewhat similar lines, an open-bottomed vessel is placed on the surface


to be measured and filled with fine sand to a predetermined level. The volume of
sand required to fill the interstices of the surface is deduced by comparing the
amount of sand needed with that needed on a perfectly flat smooth surface, and
Other Measurement Techniques 77

hence the average depth of the surface roughness is calculated @oty 1975). This
test is intended for road surfaces, but could be used for finer surfaces if a finer
particle size were employed.
The longest and shortest wavelengths are clearly set by the diameter of the
vessel and the diameter of a sand grain respectively. The latter dimension also
limits the height resolution in theory, though in practice the mass or volume
measurement of the differential quantity of sand would probably be the limiting
factor. The upper height limit is set only by the height of the vessel. The method
is simple, cheap and fairly quick, but is unduly sensitive to long wavelengths, as
confirmed by the very large scatter in reported results (Fig. 4.10).

4.2.2. Electrical Methods

Capacitance profilers have been dealt with above, but the capacitance principle,
being an areal rather than a sectional phenomenon, is better suited to parametric
applications. The capacitance between two conducting elements is directly
proportional to their area and the dielectric constant of the medium between them
and inversely proportional to their separation. If a rough surface is regarded as the
sum of a number of small elemental areas at different heights it is fairly easy to
work out the effective capacitance between it and a smooth plate for various
deterministic surface models (Sherwood & Crookall 1967/68; Ten Nape1 & Bosma
1970/71).
Unfortunately, real surfaces are rarely deterministic. The capacitance of a
condenser, one of whose plates is rough with a probability distribution of surface
heights p(z), is proportional to

-m

where h is the separation of the mean planes.


The difficulty is at once apparent: unless we have some grounds to truncate
the height distribution at a height less than the mean plane separation we will
always end up with infinite capacitance. Some numerical solutions for a Gaussian
height distribution truncated at various arbitrary heights indicate that the
capacitance is very sensitive to the mean plane separation and to the height of the
highest point on the surface (Thomas 1978). Instruments for measuring surface
78 Rough Su$aces

roughness based on a capacitance principle have been commercially available


(Fromson et al. 1976, Brecker et al. 1977, Risko 1981). The readings of Fromson's
instrument correlated well with stylus measurements over a rather restricted range
of roughness, but as the roughness increased the relationship became non-linear.
A more versatile capacitance-based instrument was described by Lieberman
et al. (1988). The electrode is a flexible diaphragm which when pressed against
the rough surface conforms to more and more peaks as the load P increases (Fig.
4.11). Modelling the diaphragm as a two-dimensional elastic beam, the degree of
conformity is calculated iteratively from stylus profile traces by assuming the
maximum load is that which gives stable capacitance for an electrode in contact
with a sinusoidal surface of wavelength 0.8 mm. The effective plate separation can
then be calculated and agrees reasonably well with measured values. Vertical
resolution is stated to be 25 nm.

P =0 (Rigid Beam)

Figure 4.1 1. Progressive stages of the sagging-beam calculation for modelling the compliance of the
capacitance probe (Liebeman et al. 1988).

When these values are compared with stylus measurements of a wide range of
roughnesses and finishing processes (Fig. 4.12), although the surface model is two-
dimensional, it seems to work just as well for finishes without a lay. However,
while the overall trend is clear, the scatter is such that many individual
measurements disagree with stylus values by 50% or more. This is a pity, as the
method, being fast and non-destructive, is potentially well-suited to production
applications.
The inductance between two magnetic surfaces will also be a hnction of their
roughness, again because inductance falls off with increasing separation.
Radhakrishnan ( 1977a) has measured the inductance between a magnetic
recording head and a number of rough surfaces, and compared his results with
Other Measurement Techniques 79

../ '

Figure 4.12. Capacitance versus stylus roughness far 41 surfaces, with best fit straight line (Lieberman et al.
1988)

stylus measurements. Useful correlation with average roughness was obtained only
when the comparison was restricted to a particular machining process. A stronger
correlation was reported with peak density, again indicating a sensitivity to local
maxima. As the measurement is quick and cheap this technique also has
possibilities for quality control, though of course it is restricted to magnetic
materials.
Skin resistance is a also a phenomenon affected by roughness. Alternating
electric currents of high frequencies are shifted from the central to the peripheral
annuli of a wire; thus, the major part of the current flows in a surface layer, which,
for copper, would be about 0.4 pm thick at 25 GHz (Bikerman 1970). Thus, at
high frequencies, the thickness of the actively conducting region is of the order of
magnitude of the height of surface hills. Consequently, the experimental resistivity
of a wire deviates from that calculated under the assumption of no rugosity, and the
degree of roughness can be deduced from this deviation.
The largest height difference which can be detected decreases as the
fiequency increases. The smallest detectable height difference will depend on the
resolution of the resistance change. The long-wavelength cut-of€ will be set by the
wavelength of the a.c. current and the velocity of its propagation in the metal; for
the above frequency it would be about 5 mm for a steel wire. The method is
suitable only for measurement of wire specimens, but for these it might well be the
only practicable technique.
80 Rough Su#aces

4.2.3. Fluid Methods

In the outflow meter, an open-bottomed vessel with a compliant annulus at its


lower end is placed in contact with the surface to be measured and filled with water
to a predetermined level. The time taken for a given volume of water to escape
through the gap between the compliant seal and the rough surface is measured.
Moore (1965) has defined the ratio of the total effective cross-sectional area of the
gap to the perimeter as the mean hydraulic radius (MHR). According to his
analysis, which assumes laminar flow, the time of escape is inversely proportional
to the fourth power of the MHR and should therefore be very sensitive to changes
in surface texture.
This device was intended to measure road surfaces, though there seems no
reason why the principle should not be applied to much smoother surfaces. To
work out the vertical and horizontal ranges is rather difficult, as the roughness
measured is actually that of an annular section. The compliant seal acts as a high-
pass filter whose cut-off depends in a complicated way on its elastic properties and
on the roughness itself. A low-pass cut-off of sorts is set by the width of the
annulus. The maximum height range is governed only by the rate of discharge of
water through an open pipe, and the minimum height resolution by the operator's
patience!

Timer contacts
Ground contact

Figure 4.13. Sectional view of an outflow meter (Henry & Hegmon 1975).
Other Measurement Techniques 81

A more sophisticated version has been described (Henry & Hegmon 1975)
(Fig. 4.13) in which a marked temperature dependence has been found for smooth
but not for coarse surfaces. This leads the authors to suggest that flow is turbulent
for coarse surfaces and that Moore's analysis is therefore invalid. The device is
fairly simple and fairly cheap and should give reproducible results, though no one
seems to have compared its measurements with those of more orthodox equipment
on the same surfaces.
If the outflow meter employs a compressible rather than an incompressible
fluid its theory becomes a little more complicated, but the basic strategy remains
the same: to measure the effective cross-sectional area of outflow. However, as the
viscosity of compressible fluids is so much lower an instrument employing one is
more suited to engineering surfaces. Pneumatic gauges are used extensively in
manufacturing industry; the general principles of their design, and various
practical realisations, are discussed at length by Farago (1982). They were first
employed to measured surface roughness by Nicolau (1937), but the theory was not
worked out till rather later (Graneek & Wunsch 1952).

Figure 4.14. (a) Principal elements of a pneumatic gauging system (Farago 1982): (1) continuous supply of
pressurised air; (2) pressure reducing valve; (3) metering device; (4)pressure indicator; (5) gauge head (6)
specimen surface: (b) various nozzle cross-sectionsused for roughness measurement, outer diameter 25-30
mm (Radhakrishnan & Saga 1970)

The essential elements are a gauging nozzle in proximity to the test surface
connected through a metering device to a source of air at constant pressure (Fig.
82 Rough Su$aces

4.14a). Air escapes between the gauging nozzle and the rough surface, thus
lowering the pressure downstream of the metering device. A circular nozzle was
orignally used, but various other shapes have been tried including slits and
cruciform sections (Fig. 4.14b). A sufficiently fine slit should provide something
nearly like a profile measurement, with the high-pass cutoff set by its length. The
horizontal and vertical resolutions presumably depend in principle on the gas laws,
but in practice the useful horizontal range of measurement is set by the range of
linearity of the relationship between pressure and roughness. Measurements have
been reported (Graneek & Wunsch 1952; Wager 1967) which correlate very well
with roughness readings from a stylus instrument over a range from 0.1 pm to 5
pm (Fig. 4.15). Tanner (1979, 1980, 1981, 1982) has described a pneumatic
Wheatstone bridge for roughness measurement.

Figure 4.15. Variation ofpneumatic gauge reading with roughness (Wager 1967)

This pneumatic technique seems to have possibilities which would repay


development. With a circular nozzle it is unduly sensitive to long wavelengths, but
offers the compensating advantage of giving a reading independent of orientation
of a surface with a lay. Its reading depends on height parameters only. Up till now
it has only been used as a comparator, but a more rigorous theory could probably be
devised to give an absolute interpretation. It is cheap, robust, simple, quick and
non-contacting, in fact well suited to a production line. When the sensitivity is
improved by electrical pressure sensing and the nozzle is miniaturised, it can even
be used as a profiler, as described in the section on in-process measurement.
Other Measurement Techniques 83

In the oil-droplet method, a droplet of oil of volume V is placed on to a rough


solid and squeezed with an optical flat (Bikerman 1970). If the greatest area of the
oil patch which can be achieved is A , then V/A is the average thickness of the
patch. This is claimed to be approximately equal to half Rp, though this
relationship almost certainly depends both on the finish of the surface and on the
size of the oil droplet. The vertical resolution will depend on the waviness of the
optical flat. The method is simple and cheap, but slow and not very reliable.
In the stagnant-layer method, a plate is covered with a non-volatile liquid and
suspended vertically (Bikerman 1970). The mass and thus also the volume Vof the
liquid remaining on the plate after time t is determined from time to time. The
experimental function V = f (t) deviates from that predicted by hydrodynamics in
such a manner as if a stagnant layer were present. The deduced thickness of the
layer is from one to two times the RMS roughness. The height resolution will
depend on the resolution of the weighing arrangements; a difference in roughness
of 0.1 pm on an area of 100 cm2 would cause a change of mass of only about 1 mg
in a total mass of perhaps 100 g. The method is not very simple and requires quite
expensive equipment; it is slow and not very reliable; and it is only suitable for
workpieces of the appropriate geometry.
Using a somewhat similar principle, an oil drop is timed as it flows down a
fixed length of an inclined test piece (Kamnev 1966). The theory is not discussed,
but one might expect both the mean slope and the RMS roughness to affect the
performance. This is confirmed by a difference in behaviour between surfaces
finished in various ways. Roughness was found to be proportional to the 3.7 power
of the time for filed and milled surfaces but the 1.74 power for etched surfaces.
Sensitivity falls off rapidly as the surface gets rougher. The long-wavelength cut-
off must be set by the diameter of the drop. The method is simple, cheap and
surprisingly reproducible, but rather slow.

4.2.4.Acoustic Methods

Acoustic radiation interacts with rough surfaces in ways analogous to the various
interactions of electromagnetic rahation, for instance by backscattering (Ogilvy
1988, Blakemore 1993). In addition acoustic waves can be transmitted through a
rough interface, and the transmission may yield information about the roughness
itself (Nagy & Adler 1987, Pecorari et al. 1992, 1995b) and the real area of contact
(Krolikowski & Szczepek 1992, Polijaniuk & Kaczmarek 1993, Pecorari et al.
1995a).
84 Rough Su$aces

If the acoustic wavelengths are short compared to the surface wavelengths


then specular reflection may occur, and the topography of underwater surfaces is
investigated in this way by sonar, giving vertical resolutions of better than a meter
and lateral resolutions of a few meters (Russ 1994). On a smaller scale, Blessing &
Eitzen (1988) have obtained profiles of machined surfaces acoustically, but with
rather poor resolution. Acoustics show more promise for parametric measurement
of roughness. De Billy et al. (1976), using backscattering techniques, have
measured roughnesses in the range from 3 pm to 100 pm, and Stor-Pellinen &
Luukkala (1995) have investigated the possibility of measuring the roughness of
paper using ultrasound. Chiang et al. (1994) have suggested that ultrasound could
measure the roughness of articular cartilage in vivo.
Gezanhes et al. (1982) measured the roughness of cmcrete surfaces using
acoustical speckle correlation. The surface to be measured was illuminated by two
coherent ultrasonic waves at different angles of incidence. The correlation between
the two scattered waves depends on the roughness, the angles of incidence and the
acoustic wavelength. With a so-called "acoustic interferometer" constructed on
this principle they were able to measure roughnesses of between 25 pm and 2.4
mm.

4.3. References

Abbott, E. J. and Goldschmidt, E., "Surface quality", Afech. Engng., 59, 813-
825 (1937).
Alexander, S., Hellemans, L., Marti, O., Schneir, J., Elings, V., Hansma, P.
K., "An atomic-resolution atomic-force microscope implemented using an optical
lever", J. Appl. Phys. 65, 164 (1989).
Bikerman, J.J., Physical surfaces, (Academic Press, New York, 1970).
Binnig, G., and Quate, C. F., "Atomic force microscope", Phys. Rev. Lett. 56,
930-933 (1986).
Binnig, G.; Rohrer, H., "Scanning tunnel microscopy", IBM Journal of
Research and Development 30, 355-369 (1986)
Blakemore, M., "Scattering of acoustic waves by the rough surface of an
elastic solid", Ultrasonics 31, 161-174 (1993)
Blessing, G. V. and Eitzen, D. G., "Surface roughness sensed by ultrasound",
Surface Topography, 1, 143-158 (1988)
Brecker, J. N., Fromson, R. E.; Shum, L. Y., "Capacitance-based surface
texture measuring system", Ann. C I W 26,375-377 (1977).
Other Measurement Techniques 85

Broadston, J. A., "Control and measurement of surface finishes"' Steel, 120,


No. 2, pp.82-3, 116, 118 and 121 (1947).
Broadston, J. A., "Measuring methods described for surface roughness
specification", Prod. Engng. 15, 806-810 (1944).
Bugg, C. D., "Noncontact surface profiling using a novel capacitive
technique: scanning capacitance microscopy", Proc. SPIE 1573, 2 16-224 (1 99 1)
Bugg, C.D. and King, P.J., "Scanning capacitance microscopy", J. Phys. E:
Sci. Instrum., 21, 147-151 (1988)
Chiang, E. H.; Adler, R. S.; Meyer, C. R.; Rubin, J. M.; Debrick, D. K.;
Laing, T. J., "Quantitative assessment of surface roughness using backscattered
ultrasound: the effects of finite surface curvature", Ultrasound in Medicine and
Biology 20, 123-135 (1994)
Darmody, W, J., "Tapering for surface inspection", Am. Mach., 90, 134-135
(1946).
de Billy, M., Cohen-Tenoudji, F., Jungman, A,, Quentin, G. J., "Possibility of
assigning a signature to rough surfaces using ultrasonic backscattering diagrams",
IEEE Trans Sonics Ultrason SU-23,356-363 (1976).
DiNardo, N. J., Nanoscale characterisation of surfaces and interfaces,
(VCH, Weinheim, 1994)
Dorinson, A., "Microtopography of finely ground steel surfaces in relation to
contact and wear", A.S.L.E. Trans., 8, 100-108 (1965).
Doty, R. N., "Study of the sand patch and outflow meter methods of pavement
surface texture measurement", in Rose, J. G. ed., Surface Texture versus Skidding:
Measurements, Frictional Aspects and Safety Features of Tire-pavement
Interactzons, STP 583, 42-61 (ASTM, 1975)
Farago, F. T., Handbook of dimensional measurement 2e (Industrial Press,
New York, 1982)
Fromson, R. E.; Shum, L. Y.; Brecker, J. N., "Universal surface texture
measuring system", SME IQ76-597, (1976)
Fu, J., Young, R. D., Vorburger, T. V., "Long-range scanning for scanning
tunnelling microscopy", Rev. Sci. Instrum. 63, 2200-2205 (1992)
Garbini, J. L., Jorgensen, J. E., Downs, R. A. and Kow, S. P., "Fringe-field
capacitive profilometry", Surface Topography, 1, 13 1-142 ( 1 388)
Gezanhes, C ; Calaora, A.; Condat, R., "Acoustical measurement of surface
roughness by speckle correlation", Signal Process n 2-3 (Apr 1982)
Goch, G., Volk, R., Tontactless surface measurement with a new acoustic
sensor", CIRP Ann. 43,487-490 (1994)
86 Rough Surfaces

Graneek, M. and Wunsch, H. L., "Application of pneumatic gauging to the


measurement of surface finish", Machinery, 81, (1952).
Greenwood, J. A. and Tripp, J. H., "The elastic contact of rough spheres",
Trans. A.S.M.E: J.Appl. Mech., 34E, 153-159 (1967).
Grigor'ev, A. Ya., Myshkin, N. K., Semenyuk, N. F. and Kholodilov, O.V.,
"Evaluating specific surface area by the secondary electron emission method",
Trenie i Iznos, 9, 793-798 (1988)
Grundy, P. J., and Jones, G. A,, Electron microscopy in the study of
materials, (Edward Arnold, London, 1976)
Guerrero, J. L. and Black, J. T., "Stylus tracer resolution and surface damage
as determined by scanning electron microscopy", Trans. A.S.M.E: J. Eng. Ind.,
94B, 1087-1093 (1972).
Haesing, J., "Determining surface finish of workpieces by means of surface
standards", Microtecnic, 15, 24-28 (1961).
Hansma, P. K., and Tershoff, J., 5canning tunnelling microscopy", J. Appl.
Phys. 61, Rl-R23 (1987).
Hansma, P. K., Drake, B., Marti, O., Goud, S. A. C., Prater, C. B., Science
243, 641-643 (1989)
Henry, J. J. and Hegmon, R. R., "Pavement texture measurement and
evaluation", in Rose, J. G. ed., Surface Texture versus Skidding: Measurements,
Frictional Aspects and Safety Features of Tire-pavement Interactions, STP 583, 3-
17 (ASTM, 1975)
Hirst, W. and Hollander, A. E., '*Surfacefinish and damage in sliding", Proc.
R. SOC.Lond. A337,379-394 (1974).
Holburn, D. M. and Smith, K. C. A., "Topographical analysis in the SEM
using an automatic focusing technique", J. Microscopy, 127, 93-103 (1982)
Howell, P. G. T. and Bayda, A., Proc. 5th. SEMSymp., Chicago, 1972
Jost, H . P., "A case for the qualitative inspection of surface finish", Machy.
65,483-486 (1944)
Jungles, J. and Whitehouse, D. J., "An investigation of the shape and
dimensions of some diamond styli", J. Phys. E: Sci. Instrum., 3,437-440 (1970).
Kamnev, V. V., "Integral evaluation of surface roughness", Meas. Tech., 2,
261-263 (1966).
Kholodilov, 0. V., N. K. Myshkin and A. Y. Grigor'ev, "Microtopography
evaluation with scanning electron microscope", Soviet Journal of Friction Wear, 6,
133-136, (1985)
Krolikowski, J.; Szczepek, J., "Phase shift of the reflection coefficient of
ultrasonic waves in the study of the contact interface", Wear 157, 51-64 (1992)
Other Measurement Techniques 87

Lebiedzik, J. and White, E. W., "Multiple detector method for quantitative


determination of microtopography in the SEM", Proc. 8th Annual Symp. Scanning
Electron Microscopy, 101-188 (Illinois Inst. of Technol., Chicago, 1975).
Lee, J. H., and Russ, J. C., "Metrology of microelectronic devices by stereo
SEM", J. Computer-assisted Microscopy 1,79-90 (1989)
Lieberman, A. G., Vorburger, T. V., Giaque, C. H. W., Risko, D. G. and
Rathbun, K. R., "Comparison of capacitance and stylus measurements of surface
roughness", Surface Topography, 1,115-130 ( 1988)
McAdams, H. T., "Scanning electron microscope and the computer: new
tools for surface metrology", Modem Machine Shop, 82-9 1 (1974).
McClelland, G. M., Erlandsson, R., Chiang, S., "Atomic force microscopy:
general principles and a new implementation" in Review of progress in
quantitative nondestructive evaluation, 6B, D. 0. Thompson and D. E. Chimenti
eds., 1307-1314, (Plenum, New York, 1987).
Moore, D. F., "Drainage criteria for runway surface roughness", J. Roy.
Aeronaut. SOC.,69, 337-342 (1965).
Nagy, P. B. and L. Adler, "Surface roughness induced attenuation of reflected
and transmitted ultrasonic waves", Journal of the Acoustical Society of America,
82, 193-197 (1987)
Nelson, H. R.,"Taper sectioning as a means of describing the surface contour
of metals", Proc. ConJ on Friction & Surface Finish, 2e, 217-237 (MIT,
Cambridge, 1969).
Nicolau, M. P., "Application du micrometre solex a la mesure de l'etat des
surfaces", Mecanique, 80-83 (Mar-Apr 1937)
Nowicki, B., and Jarkiewicz, A,, "The in-process surface roughness
measurement using fringe field capacitive method", Trans. 7th Int. Con$ on
Metrology & Properties of Engng Surfaces, 325-332 (Gothenburg, 1997)
Ogilvy, J. A., "Computer simulation of acoustic wave scattering from rough
surfaces", J. Phys. D: Appl. Phys., 21,260-277 (1988)
Pecorari, C., Mendelsohn, D. A,; Adler, L., "Ultrasonis wave scattering from
rough, imperfect interfaces. Part I. Stochastic interface models", Journal of
Nondestructive Evaluation 14, 109-116 (1995a)
Pecorari, C., Mendelsohn, D. A,; Adler, L., "Ultrasonic wave scattering from
rough, imperfect interfaces. Part 11. Incoherent and coherent scattered fields",
Journal of Nondestructive Evaluation 14, 117-126 (1995b)
Pecorari, C.; Mendelsohn, D. A.; Blaho, G.; Adler, L., "Investigation of
ultrasonic wave scattering by a randomly rough solid-solid interface", Journal of
Nondestructive Evaluation 11,211-220 (1992)
88 Rough Surfaces

Pohl, D. W., Denk, W., Lam, M., Appl. Phys. Lett. 44,651-653 (1984)
Polijaniuk, A,, Kaczmarek, J., "Novel stage for ultrasonic measurement of
real contact area between rough and flat parts under quasi-static load", Journal of
Testing &Evaluation 21, 174-177 (1993)
Powell, R. W., "Experiments using simple thermal comparator for
measurement of thermal conductivity, surface roughness and thickness of foils or
of surface deposits", J.Sci. Instrum., 34, 485-492 (1957).
Rabinowicz, E., "Taper sectioning, A method for the examination of metal
surfaces",Metal Industry, 76, 83-86 (1950).
Radhakrishnan, V. and Sagar, V., "Surface roughness assessment by means
of pneumatic measurement", Proc. 4th. All-India Machine Tool Design &
Research Con$, (Indian Inst. Tech., Madras, 1970)
Radhakrishnan, V., "Application of inductive heads for non-contact
measurement of surface finish", Proc. Int. ConJ Prod. Eng. 2 (Inst. of Eng.,
Calcutta, 1977).
Rasigni, M., Rasigni, G . , Palmari, J.-P., Llebaria, A,, "Study of surface
roughness using a microdensitometer analysis of electron micrographs of surface
replicas:- 1. surface profiles", J. Opt. SOC.Am. 71, 1124-1133 (1981).
Reddick, R. C. R., Warmack, R. J., Chilcott, D. W., Sharp, S. L., Ferrell, T.
L., Rev. Sci. Instrum. 61, 3669-3677 (1990)
Risko, D. G., "Quick, non-destructive method for measuring surface finish
using capacitance", Carbide Tool J 13,26-29 (1981)
Rubert, M. P., "Functional assessment of surface roughness", Proc. I. Mech.
E., 182, Part 3K, 350-359 (1 967/68).
Russ, J. C., Fractal surfaces, (Plenum Press, New York, 1994).
Samuels, J. M., Hoover, M. R., Tarhay, L., Johnson, G . C . , White, E. W.,
"Quantitative SEM and raster profilometer analysis of fracture surfaces", in Bradt
R. C. et al. Eds., Fracture mechanics of ceramics, (Plenum Press, New York,
1974)
Schlesinger, G., Surface finish, (Inst. of Prod. Engrs., London, 1942).
Shaw, M. C. and Pekleruk, J., "A light projection technique for studying
surface topology", Ann. C.I.R.P., 12, 93-97 (1963).
Shenvood, K. F. and Crookall, J. R., "Surface finish assessment by an
electrical capacitance technique", Proc. I. Mech. E., 182, Part 3K, 344-349
(1967/68).
Stor-Pellinen, J., Luukkala, M., "Paper roughness measurement using
airborne ultrasound", Sensors and Actuators, A: Physical 49, 37-40 (1995)
Other Measurement Techniques 89

Tanner, L. H., "A self-balancing pneumatic Wheatstone bridge for surface


roughness measurement", Wear 83, 37-47 (1982)
Tanner, L. H., "An improved pneumatic Wheatstone bridge for roughness
measurement", J. Phys: Sci. Instrum. 13E., 593-594 (1980)
Tanner, L. H., "Pneumatic Wheatstone bridge for surface roughness
measurement", J. Phys: Sci. Instrum. 12E, 957-960 (1979)
Tanner, L. H.,"A self-balancing pneumatic potentiometer and Wheatstone
bridge with electrical readout. Applications to surface roughness measurement,
pneumatic gauging and to measurement of pressure difference ratios", Precis. Eng.
3,201-207 (1981).
Tarasov, L. P., "Relation of surface roughness readings to actual surface
profile", Trans. A.S.M.E, 67, 189-194, (1945).
Teague, E. C . , "Scanning tip microscopies: an overview and some history", in
G. W. Bailey ed., Proc. 46th. Annual Meeting of the Electron Microscopy Society
ofAmerica, 1004-1005 (San Francisco Press, San Francisco, 1988)
Ten Napel, W. E. and Bosma, R., "The influence of surface roughness on the
capacitive measurement of film thickness in elastohydrodynamic contacts", Proc. I.
Mech. E., 185,635-639 (1970/71).
Thomas, T. R. and Probert, S. D., "Correlations for thermal contact
conductance in vacuo", Trans. Am. SOC.Mech. Engrs., 94C, 176-180 (1972)
Thomas, T. R. and Probert, S. D., "Thermal contact resistance: The
directional effect and other problems", Int. J. Heat Mass Transfer, 13, 789-807
(1970).
Thomas, T. R., "Surface roughness measurement: alternatives to the stylus",
Proc. 19th. MTDR ConJ, 383-390 (UMIST, Manchester, 1978)
Tornebohm, H., "Modern tolerance requirements and their scientific
determination", Mech. Engng., 58, 41 1-417 (1936).
Wager, J. G., "Surface effects in pneumatic gauging", Int. J. Mach. Tool Des.
Res., 7, 1-14 (1967).
Watanabe, T.; and Fukui, S., "Control of tactile surface-roughness sensation
using ultrasonic vibration", Trans. JSME C62, 1329-1334 (1996)
Watanabe, T.; and Fukui, S., "Method for controlling tactile sensation of
surface roughness using ultrasonic vibration", Proc. IEEE Int. Con$ on Robotics
andAutomation 1, 1134-1139 (IEEE, Piscataway, NJ, 1995)
Weaver, J. M. R., Walpita, L. M., Wickramasinghe, H. K., Nature 342, 783-
785 (1989)
mt e house , D. J., Handbook of surface metrology, (Institute of Physics,
Bristol, 1994)
90 Rough Surfaces

Williams, C. C. and Wickramasinghe, H. J., J. Vac. Sci. Technol. B9, 537-


540 (1991)
Williams, C. C. and Wickramasinghe, H. J., Proc. Ultrasonics Symp. 393-
397 (IEEE, Piscatawy, NJ, 1986)
CHAPTER 5

OTHER MEASUREMENT TOPICS

There are a number of important topics connected with the measurement of


roughness which cannot readily be considered under the various headings of
instrument design. Before we leave the subject of measurement and move on to
characterisation, it is convenient to collect these topics here for discussion. We
begin with the special problems of 3D measurement (3D characterisation will be
discussd at length later). We go on to consider the Miculties of repeatedly finding
a small area on a surface in a sequence of measurements, and the techniques
necessarily employed if for some reason it is not possible or convenient to bring the
instrument to the surface under investigation. Finally we discuss the measurement
or inspection of roughness during manufacture.

5.1. 3D Measurement

By 3D measurement we mean more than a measurement of average area properties


such as that given by glossmeters. We mean an actual determination of surface
relief over an area, so that at least in principle a topographic map can be
constructed. On a surveying scale this was traditionally done by triangulation; that
is, starting from an original baseline, the position of individual points on a
landscape was determined by measuring their angular displacement from two fixed
points and solving the resulting triangle. Interestingly, this is basically a digital
method, though not of course based on a rectangular grid. More recently, height
information has been extracted from pairs of stereo photographs of the same area
taken at slightly different angles by stereophotogrammetry, a technique which has
been applied to surface roughness (Unsworth & Hepworth 1971). Contour maps
on a smaller scale may be obtained directly by interferometry (Tolansky 1960,
1970a,b) but do not by themselves give quantitative height information. The
volume of data from 3D measurements is so large and its processing so tedious that
in practice, whatever the original measuring technique, relief information is at
some stage extracted in the form of individual heights scanned on a rectangular
grid.

91
92 Rough Sudaces

There is a paradox inherent in 3D measurement in the sense in which we


have defined it above. To approximate as closely as possible to spot height
measurement the footprint of the probe or sensor on the surface must be as small as
possible. How can this small window on the surface map an area? The traditional
answer is a raster scan, used for many years by television engineers to build up
pictures on cathode ray tubes. The raster moves the transducer over the surface in
a number of closely spaced parallel lines (Fig. 5.1). If the output of the transducer
is processed and displayed according to an appropriate protocol, a picture of the
surface being scanned will be built up line by line.

1Aperture

%nes indicafing path and


direction foJlowedby aperiwe

Figure 5.1. Raster scanning of an area (Terman 1937). The "aperture"correspondsto the footprint of a
roughness measuring probe.

In a television the scanning device is an electron beam, easy to move, and the
display is analogue. For roughness measurement, movement is not usually so easy
and acquisition and processing of data are digital. The aim of the measurement
procedure is to build up a matrix of individual height readings in the computer
whch can then be processed numerically, either to create graphical representations
of the surface or to extract quantitative information about the relief. In principle
such a scan need not be Cartesian; polar scans (Edmonds et al. 1977, Newman et
al. 1989) and even spiral scans (Mollenhauer 1973) have been employed, but the
non-Cartesian spacing is inconvenient for subsequent analysis.
Thus there are no 3D measuring instruments as such: all so-called 3D
instruments are basically profilers, and any of the profiling instruments described
in previous chapters can in principle be used for 3D measurement using raster
Other Measurement Topics 93

scanning. Individual scanning devices have individual strengths and weaknesses


which have already been discussed. But there are some general problems, of which
the first is translation in the plane of the measured surface. In electron microscopy
this is achieved by deflecting the electron beam. In scanning interferometers the
fringes from an area are imaged onto an array of detectors which is simply read out
sequentially by the computer. In STM's and similar instruments a deflecting
voltage is applied to a piezo element.
In general, however, it is more convenient to keep the measuring probe
stationary and move the workpiece. This is partly due to the ready availability of
precision x-y translation tables already developed for optical engineering. The
requirements for such tables are formidable, as they must not only offer precisely
controlled movement in very small increments but also give accurate positional
information and provide an absolute reference datum of height. Desages & Michel
(1993) measured two kinds of error on a proprietary translation table used for
roughness measurements. A laser interferometer measured the errors in x and y
separately at various table positions (Fig. 5.2a, b). Positional errors of more than
60 pm were detected in a 280 x 280 mm area.

Figure 5.2. Positional errors (a) in x, (b) standard deviation of (a), (c) in z,of a translation table (Desages &
Michel 1993).
94 Rough Surfaces

They went on to measure the error in z by mounting an optical flat on the


table and translating it in the x-direction (Fig. 5 . 2 ~ ) .Height errors of more than a
micron were detected, with a marked periodicity. Both positional and height errors
occur at wavelengths of the order of millimetres, and their effect would probably be
filtered out in many roughness calculations. Note also that there are considerable
differences in the time taken by different instruments to map the same area; while
data acquisition by a scanning interferometer is virtually instantaneous, a scanning
stylus profiler may take several hours. This is a period of time long enough for
changes in the environment, for instance temperature changes, to affect the
measuring system. Such changes usually appear as spurious errors of form and
again will probably be filtered out in roughness calculations.
In Fig. 5.1 the probe is moved in a true raster, that is with each measuring
scan always in the same direction and followed by a return passage when no
measurements are taken. This presents no difficulties when the probe is an
electron beam whose movement is effectively instantaneous, but with a
mechanically translated probe the return passage may represent a significant time
overhead. Instrument designers often attempt to reduce this overhead by
employing boustrophedon scanning (pouozpocpqFov = "as the ox turns", i.e. in
ploughing a field), where measurements are taken on the return stroke also
(Teague 1988). With many probes this causes no particular problem, but users
should be aware that styli are usually designed to be dragged, and the dynamic
effects of pushing them instead can introduce serious measuring errors (Sayles &
Thomas 1976).
Analysis of 3D data will be dealt with in a later chapter. Before analysis can
begin, however, there is a problem of visualising the data. The amount of data
collected in a single area measurement is so large that it is difficult to make a
preliminary appreciation unless the data is processed by software to appear in some
visual form. 3D visualisation software is now widely available in mathematical
packages such as MathematicaTMand MatlabTMand even in general-purpose
spreadsheets like ExcelTM,and most 3D roughness measuring systems are bundled
with their own proprietary software. At the very least a customer should expect to
be able to see his data as a contour map or as an isometric wire-frame view, often
with perspective and rendering available as well (Fig. 5.3). There should be
facilities to zoom in on areas of arbitrary size, to change the number and heights of
contours, to reject obviously defective data points, and to export the data in a
portable format.
The facility to reject data points, though desirable to clean up measurements
which may be impossible to repeat, should be treated with care. Problem data
Other Measurement Topics 95

points usually represent regions where the local surface is too high, or too low, or
too steep, for the sensor to follow. If the user is only interested in average or
typical features, this may not matter, but extreme-value roughness parameters may
be sigrufcantly affected by missing data. Any area measurement which contains
more than say 10% of rejected heights should be discarded unless there are
pressing reasons to the contrary. If changing the instrument settings does not
improve the rejection rate, the wrong instrument is being used. Particular care
should be taken to read the small print of the program manual; the default settings
of some software packages simply interpolate over the top of bad data without
flagging it, a recipe for drawing wrong conclusions.

a
b

I Y

Figure 5.3. (a) 0.12 mm x 0.12 mm of a plateau-honed surface mapped with an AFM, height contours at 1
p intervals; @) isometric view of (a); (c) 40 pn x 20 p &om (a), contours at 0.2 pn intervals; (d)
isometric view of @)

5.2. Relocation

There are many situations, particularly in research, where it would be very useful
to look at a single profile through a surface before and after some experiment such
96 Rough Sui$aces

as running-in to see what changes have occurred to the surface geometry. It is


clearly essential that exactly the same section is traversed each time, otherwise the
changes observed could be attributed to the displacement of the profile and small
but significant changes might not be observed at all. This requirement is very
difficult to achieve in practice; the action of returning the profiling transducer to
the start of its traverse is often enough to displace it laterally by a few microns, and
as a typical stylus is only 8 pm wide, and other sensors may well be smaller, this is
sufficient to invalidate the result.
The problem was overcome by Williamson and Hunt (1968) who designed
what they called a relocation table. The table is bolted to the bed of the stylus
instrument, and the specimen stage is kinematically located against it at three
points and held in position pneumatically. The stage can be lowered and removed,
an experiment of some kind performed on the specimen, and the stage replaced on
the table. Relocation of the stylus then occurs to within the width of the original
profile. It is necessary to raise the stylus during the return stroke of the pickup,
and of course the specimen may not be removed from the stage during the course of
the experiment.. Fig. 5.4 shows a succession of relocated profiles of a surface
which was initially rough turned (Thomas 1972). After each measurement the
stage was transferred to the table of a surface grinder and a slightly deeper cut
taken. The progressive dtsappearance of the peaks and the persistence of the
valleys can be followed in great detail.

Figure 5.4. Relocated profiles of an initially turned surface at progressive stages of grinding (Thomas 1972)
Other Measurement Topics 97

The original device has been widely copied and used to study wear in rolling
and sliding contact (Grieve et al. 1970; Efeoglu et al. 1993a, b), contact mechanics
(OCallaghan & Probert 1973), rolling of sheet metal (Atala & Rowe 1975),
running-in (Stout et al. 1977), effect of successive coats of paint (King & Thomas
1978), and impact wear (Engel & Millis 1982). Detailed designs of relocation
table have been described by Edmonds et al. (1977) and Sherrington & Smith
(1993).
When we move on from profile measurement to area measurement there is a
tendency to assume that relocation is no longer so important because of the
increased statistical reliability of the much larger data set. For measurement of
average roughness parameters this is probably true in many cases, but there are
other situations where a more detailed examination of the relocated surface is
necessary. Several workers have found it necessary to use relocation for 3D work
(Bengtsson & Ronnberg 1984, 1986, Jeng & Lalonde 1992) The very homogeneity
of a well-machined surface can make it difficult to return to exactly the same area
on an apparently featureless plain. A 99% match in x and y implies a mismatch of
2% in area, quite enough to interfere with extreme-value calculations; in fact
Newman et al. (1989), scanning wear scars with a stylus instrument, found linear
relocation within 0.1% necessary to avoid sigdicant error in scar volumes. Rather
than rely on mechanical relocation, Newman et al. marked the workpiece with a
pattern of indentations which could be realigned visually, a practice also followed
by Blunt et al.( 1994).

5.3. Replication

Replication of the original surface to be measured is needed with some optical


methods to provide a transparent specimen (Anderson 1969, Dyson 1955,
Herschman 1945, Lech et al. 1984). It is also needed in electron microscopy
(Andersson 1974, Butler 1973, Chan et al. 1976) to provide a conducting specimen
from a non-conducting workpiece. Its use with stylus instruments is generally to
obtain measurements on parts which are not easily accessible, such as internal
surfaces (Timms & Scoles 1948) or underwater surfaces (Sawyer 1953), or which
cannot conveniently be brought to the instrument, such as gear teeth (Timms &
Scoles 1948, Young & Clegg 1959), crankshafts (Davis 1979), the rollers from
steel mills (Pearson & Hopkins 1948), ships' hulls (Karlsson 1978, King et al.
1981, King 1982), large optical components (Gourley et al. 1985) and human teeth
98 Rough Sudaces

(Mathia et al. 1989a,b). It has also been used with compliant surfaces in the belief
that direct measurement would damage or misrepresent the surface (Dawson et al.
1967/68).
The principle is usually to place the surface to be measured in contact with a
liquid which will subsequently set to a solid, hopefully faithfully reproducing the
detail of the original as a mirror image, what might be termed a negative.
Materials such as plaster of Paris and dental cement have been employed, but it is
now customary to use a polymerizing liquid. The vital question is how closely the
replica reproduces the features of the original. Lack of fidelity may arise from
various causes. The liquid may not wet the surface completely; usually it will first
be necessary to degrease the surface carefully. If the surface is itself already wet, as
may be the case for biological specimens, there may be problems of diffusion or
even of chemical reaction during setting. Portions of the replica may adhere to the
surface as they are parted unless a release agent is used. In any case the replica is a
negative and a stylus instrument does not respond to a valley bottom in the same
way as to a peak, so a further positive replica may need to be made (Fig. 5.5). In
the case of transparent replicas, optical techniques often rely on detecting an
optical path difference which is a function of refractive index. Misinterpretation
can occur here due to inhomogeneity of the replica or to changes in refractive
index due to temperature. A rigid replica may not reproduce short wavelengths
faithfully, while a flexible replica may not be faithful to long wavelengths.

Figure 5 . 5 . Plateau-honed cylinder liner and positive replica (Ohlsson & Rosen 1993)

One series of careful comparisons made (Sayles et al. 1979) has found
replicas of an optical flat of negligible measured roughness to show roughnesses of
between 0.03 pm and 0.13 pm, and replicas of machined surfaces to disagree in
Other Measurement Topics 99

roughness with the originals by up to 17 per cent. Shunmugam & Fbdhakrishnan


(1976), Narayanaswamy et al. (1979) and George (1979) have attempted to
compare the power spectra of replica and original. This approach can show
dramatically the range of wavelengths over which the replication material is
effective (Fig. 5.6).

Wavelength in p m Wavelength in p m

Strand dais resin B Warwick chemicals plymaster

-I
10 I@l in I00
Spatid frequency in cycledmm -1 Spatial frequency in cycled mm-'

Wavelength In p m Wavcleneth in am

Parent wrface hefore and after


I Strand glass rcsin C acrulile replica

I I ,
-I I in In1 - I - IIY)
Spatial frequency in cycler/mm-' Spatial frcqucncy in cyclcrimm-1

Figure 5.6. Comparison of power spectra of original and replica for three diEerent replicating materials
(George 1979). Effect of the act of replication on the original surface is also shown. Bands are 50%
confidence limits

Replicas of calibration specimens have been made by electroforming (Song et


al. 1988). Again this is a two-stage process ending up with a positive replica. The
originals were essentially two-dimensional, that is to say the same "random" profile
was reproduced across the whole width of the specimen, so the issue of relocation
100 Rough Su$aces

did not arise in comparing replica with original. Flatness did not reproduce well,
and the nickel replicas were not as hard as the originals, but comparison of a
number of different roughness parameters showed agreement usually within 2%,
which is of the order of the traceable uncertainty of roughness measurements. This
technique is probably too specialized and laborious for routine use in quality
control.

5.4. In-process Measurement

A method of measuring the surface roughness of a component during its


machining would clearly be valuable, either to terminate the machining process as
soon as the required finish was obtained and thus increase throughput per machine,
or to take part in some adaptive-control loop. Slow attainment of the required
finish, for instance, might signal the need for tool replacement. The requirements
for in-process measurement are fairly stringent. Measurement must be continuous
and rapid and should preferably provide an electrical signal of some kind. The
sensor must be robust and relatively insensitive to environmental changes such as
temperature and the intermittent presence of films or sprays of lubricating or
cutting fluids. It should be small and adaptable to as wide a range as possible of
workpiece shapes and sizes. For obvious reasons the sensor should preferably not
contact the workpiece. Young et al. (1980) conclude that only area (i.e.
parametric) measurements made by optical instruments will satisfy a similar list of
requirements, but in fact, as we will see below, a variety of methods have been used
with varying degrees of success.
A similar problem applies to inspection (Thomas 1997). Of course the
majority of roughness instruments sold are used for inspection purposes, usually
off-line and on a statistical sampling basis. But modem trends of quality control
increasingly require 100% inspection (Kennedy et al. 1987), and this is now the
norm for many other product parameters. Most existing roughness instruments are
too slow for 100% inspection of mass-produced components. For a component
produced at a rate of 3000 an hour, not excessive for many production lines, the
total time available for setting up, data acquisition and processing is not much
more than a second. Although stylus instruments have made impressive advances
in speed (Morrison 1995), the only current techniques which can approach the
speeds required for inspection are capacitance (Garbini et al. 1988, Table 5.1) or
optical. For many comparative purposes these work well enough; difficulties arise
when it is necessary to validate their performance in terms of legal standards,
Other Measurement Topics 101

which are still exclusively written in terms of stylus instruments and their
limitations.
Pneumatic gauging looks promising for in-process work, and it has been
established (Wager 1967) that the dynamic effects of the moving workpiece are not
serious. This is rather surprising, as the theory of the pneumatic gauge (Graneek
& Wunsch 1952) assumes isothermal conditions whereas fluctuations at, say,
turning speeds are more likely to be adiabatic. The pneumatic gauge is robust and
the air jet will help to clear unwanted surface fluid. It also measures a parametric
roughness integrated over the entire path of movement of the surface. Its
dlsadvantages are that a nozzle fixed relative to the workpiece will be unduly
affected by waviness, and that it is too insensitive to measure fine finishes. By
miniaturising the nozzle and pressure transducers, Woolley (199 1) has succeeded
in making a pneumatic profiler whch will resolve height differences of 12 nm at
very high rates of translation (Table 5 . l), though the lateral resolution is no better
than 75 pm.

I.Casing
2. Wheel flanges
3. Workpiccc 13
4. Bearing
5. Leaf-springjoint 4
6 . Leaf springs
7. Guide
8. Tracer pin 14

9. Leaf springs
10. Ferrite
I I . Coil
12. Shock absorber
13. Inductive signal
14. Mount

Figure 5.7. Sectional schematic ofrotating stylus device for in-process roughness measurement (Dutschke &
Eissler 1978).

The stylus instrument, which might be thought a priori unsuitable, has


proved itself a serious contender in the measurement of turned surfaces (Dutschke
& Eissler 1978). The measuring device is a steel cylinder which rotates in contact
with the workpiece being machined. At every revolution of the cylinder a stylus,
piercing a hole in its circumference, is deflected, producing an electrical signal
102 Rough Su$aces

which is read out through telemetry (Fig. 5.7). Surprisingly good correlations with
orthodox roughness measurements were reported. A development of this design by
Zhao & Webster ( 1 989) achieved translation speeds of 1 . 1 m/s (Table 5.1).
Acoustic techniques have also been applied to in-process measurement of
roughness. Blessing & Eitzen (1988) measured the amplitude of ultrasonic back-
scattering from stationary and moving surfaces. Roughnesses of between 1 pm and
40 pm were successfully determined at speeds of up to 5 m/s (Table 5.1). This
work has since been extended by Coker and Shin (Shin et al. 1995, Coker & Shin
1996).
The issue of speed would certainly seem to favour parametric methods, most
of which are effectively instantaneous by comparison with machining speeds.
However, for in-process measurement the critical speed is not the cutting speed but
the speed of translation of the workpiece, which is generally much lower. Profiling
techniques are steadily increasing in speed (Table 5.1) and are now well within the
range of translation speeds for many machining techniques.

Table 5.1. Profilers for in-process measuremenf in order oftranslation speed (adapted &om Thomas 1997).

Technique Speed Reference


( d s >

Capacitance 0.025 Garbini et al. 1988

Stylus 1.1 Zhao & Webster 1989

Optical 1.7 Mitsui 1986

Ultrasound 5 Blessing & Eitzen 1988

Pneumatic 52 Woolley 199 1

5.4.I . Optical Techniques

Optical techmques for in-process measurement have been reviewed extensively by


workers at NIST (Young et al. 1980, Vorburger & Teague 1981). More recently,
Mitsui (1986) has described a number of optical instruments designed for in-
Other Measurement Topics 103

process use. Instrumentation can be removed to a safe and convenient place by


using fibre-optic techniques to interpose a flexible conduit (Adkins 1969; Spurgeon
& Slater 1974, Takeyama et al. 1976, Mitsui 1986). Then there are the problems
of variation in reflectivity and optical path length due to the intervention of fluid
spray or films or particles of swarf, or an actual alteration in colour due to thermal
changes during machining. Some of these variations might be removed by signal
processing, for instance by taking the first or second differential of the reflected
intensity of a continuously chopped signal (Tipton & Roberts 1967/68). A more
promising technique might be polarization of coherent light; here the only
measured quantity is the ratio of polarization of incident and reflected light, which
is insensitive to any amplitude changes in the signal (Gee et al. 1975).
The advent of coherent light sources has led to their application to in-process
measurement of roughness (Fad1 & Parsons 1978; West & West 1978, Shiraishi
1980, Yanagi et al. 1986, Persson 1992). Takaya and Miyoshi (Takaya et al. 1995,
Miyoshi et al. 1995) have used Fraunhofer diffraction to measure the roughness of
polished silicon wafers in-process to better than 1 nm, calibrating their optical
system against a stylus instrument. Optical stylus methods have also been used.
Mitsui et al. (1985) built a focus-detection system based on astigmatic focussing
with a vertical range and resolution of 1 pm and 1 nm respectively. Bodschwinna
& Bohlmann (1991) measured the finish of milled crankshaft housings with a
proprietary optical stylus; although this system was part of the production line it
does not actually seem to have been used in-process. Kiyono et al. (1994) describe
a system using two optical styli simultaneously, one of which has a greater vertical
range but poorer lateral resolution. The coarse stylus thus acts as an optical skid
for the fine stylus, filtering out waviness and also cancelling out extraneous sources
of noise.
In the laser scanning analyser (Clarke & Thomas 1979) a laser beam is
reflected from a polygonal mirror rotating at high speed, down on the workpiece
surface, where it is reflected into a fixed photodetector receiver with wide aperture
(Fig. 5.8). The detector output is amplified and applied to the vertical deflection
coils of an oscilloscope whose time base is provided by the rotation of the mirror.
Localized variations in the reflectance of the surface thus appear as changes in
signal strength whose position on the workpiece can be established from the time
elapsed since start of the current scan. The spot diameter can be set from 200 prn
upwards.
When used to measure roughness the receiver aperture is masked to a narrow
slit and the angular reflectance function is produced as the spot scans a strip whose
width is dependent on the range from the surface and the angular width required.
104 Rough Sur$aces

At a given moment in any scan the fixed detector is receiving light scattered from
the single point on the strip which happens at that instant to be illuminated by the
deflected beam. The picture &splayed on the oscilloscope screen is therefore a
symmetrical curve whose height at any point is proportional to the intensity of light
scattered into the corresponding angle, and whose maximum corresponds to the
reflection received at the specular angle. If the curve is characterized by its width
at half the maximum amplitude the reflectance is effectively normalized. When
this half-width is plotted against measurements by a stylus instrument for a range
of surfaces correlation is better with slope than with roughness (Fig. 5.9).

Figure 5.8. Schematic of a laser scanning analyser (Clarke & Thomas 1979).

Another optical technique intended for in-process roughness measurement,


this time by ellipsometry, has been described by Lonardo (1978). The quantity
used to characterize the roughness is the deviation between the value of one of the
ellipsometric angles for a smooth reference surface and its value for the rough
surface measured. He found good agreement, approximately linear in some cases,
between this deviation and the average roughness as measured by stylus. He also
investigated the effect of contaminants on in-process ellipsometric measurements
Other Measurement Topics 105

of roughness during grinding, and found the error to be generally less than 10 per
cent.

Figure 5.9. Variation of half-width with (a) roughness (b) mean absolute slope (Clarke & Thomas 1979). A:
milled B: turned; C: spark eroded; D: shaped E: ground;; F: criss-cross lapped G: parallel lapped.

Several instruments measure the ratio of the specular intensity to the intensity
at an off-specular angle. Since this ratio generally dec:eases with increasing
surface roughness, it could provide a measure of the roughness itself. Peters ( 1 965)
used this technique with the detector held 40 degrees off specular to determine the
roughness of cylindrical parts while they were being ground. His results show
good correlation between the diffuseness and roughness over a range of roughness
up to 0.3 pm. Even under different lubrication conditions (oil, water, dry) the
results are well fitted bv a single curve.
A similar instrument was developed by Corey (1978) to measure roughness in
the range 0.2 pm to 2 pm for high-speed quality control of the surface finish of
machined hemispherical parts. Essential features of the instrument are its non-
destructive capability and its ability to scan the entire surface of the part. The
instrument uses the ratio of the intensity measured 15 degr2es off-specular to the
specular intensity to yield a value for roughness. In order to make meaningful
roughness measurements for a particular type of surface, a set of roughness
specimens that have been manufactured in a way similar to the test specimens with
known roughness values are required.
Another system, developed by Takeyama et al. (1976), measures the ratio of
the specular intensity at the surface-normal to the back-scattered intensity 30
degrees off-normal. This system is designed to be quite insensitive to surface
vibration with its use of fibre optics bundles to transmit the incident and reflected
106 Rough Surfaces

light. Indeed, the measured signals from spinning parts are fairly stable with time.
Takeyama et al's ratio measurements were performed on machined surfaces with
high peak-to-valley roughness ranging from 5 to 80 pm. They found that the
experimental curves of intensity ratio against roughness were a function of the tool
radius used to machine the surfaces. This dependence on the manufacturing
process again implies a need for a set of calibration specimens. Takeyama et a1
claimed that the curves of intensity ratio versus roughness were independent of the
surface material, but this claim does not seem to be supported by all of their data.

5.5. References

Adkins, H., "A look at surface finish", Am. Mach., 113, 111-116 (1969).
Anderson, W. L., "Surface roughness studies by optical processing methods",
Proc. I.E.E.E., (Letters), 57, 95 (1969).
Anderson, S., "Plastic replicas for optical and scanning electron
microscopy", Wear, 29, 271-274 (1974)
Atala, H. F. and Rowe, G. W., "Surface roughness changes during rolling",
Wear, 32, 249-268 (1975).
Bengtsson, A. and A. Ronnberg, "Absolute measurement of running-in.",
Wear, 109, 329-342, (1986)
Bengtsson, A. and A. Ronnberg, "Wide range three-dimensional roughness
measuring system", Precision Engineering, 6 , 141-147, (1984)
Blessing, G. V. and Eitzen, D. G., "Surface roughness sensed by ultrasound",
Surface Topography, 1, 143-158 (1988)
Blunt, L., Ohlsson, R., Rosen, B.-G., "A comprehensive comparative study of
3D surface topography measuring instruments", in P. Hedenqvist, S. Hogmark and
S. Jacobson eds., Proc. 6th. Nordic Symp. On Tribology, Uppsala (1994).
Bodschwinna, H., and Bohlmann, H., "Online surface roughness
measurement in production lines for process control", 12th. IMEKO World
Congress (Beijing, 1991)
Butler, D. W., "A stereo electron microscope technique for microtopographic
measurements", Micron, 4, 410-424 (1973)
Chan, E. C.; Marton, J. P.; Brown, J. D., "Evaluation of surface roughness of
metal films by transmission electron microscopy and ellipsometry", J. Vac. Sci.
Technol. 13,981-984 (1976).
Clarke, G. M., T. R. Thomas, "Roughness measurement with a laser
scanning analyser", Wear, 57, 107-116 (1979).
Other Measurement Topics 107

Coker, S. A., and Shin, Y. C., "In-process control of surface roughness due to
tool wear using a new ultrasonic system", Int. J. Machine Tools & Manufacture 36,
411-422 (1996)
Corey, H. S., "Surface finish from reflected laser light", Proc. SPIE 153, 27
(1978)
Davis, F. A,, "Replica techniques in the study of crankshaft journal
topography", Automotive Engr. 46-47 (ApriWMay 1979)
Desages, F. and Michel, O., Calibration of a 3 - 0 surface roughness
measuring device, (Prodn. Engng. Dept., Chalmers University, Gothenburg, 1993)
Dutschke, W. and Eissler, W., "A new sensor for measuring the surface
roughness in-process on a grinding machine", Proc. 3rd. Con. on Automated
Inspection & Product Control, 19-30 (Nottingham University, 1978)
Dyson, J., "Examining machined surfaces by interferometry", Engineering,
179, 274-276 (1955).
Edmonds, M. J., A. M. Jones, P. W. O'Callaghan and S. D. Probert, "A three-
dimensional relocation profilometer stage", Wear, 43, 329-340 (1977)
Efeoglu, I.; Amell, R.D.; Tinston, S.F.; Teer, D.G., "Mechanical and
tribological properties of titanium nitride coatings formed in a four magnetron
closed-field sputtering system", Surface & Coatings Technology 57, 61-69 (1993a)
Efeoglu, I.; Arnell, R. D.; Tinston, S. F.; Teer, D. G., "Mechanical and
tribological properties of titanium aluminum nitride coatings formed in a four
magnetron closed-field sputtering system", Surface & Coatings Technology 57,
117-121 (1993b)
Engel, P.A. and H. B. Millis, "Study of surface topography in impact wear",
Wear, 75,423-442 (1982).
Fadl, M. F. A,, and Parsons, F. G., "Electro-optical flaw detection", Proc.
3rd. C o n . on Automated Inspection & Product Control, 111- 118 (Nottingham
University, 1978)
Garbini, J. L., Jorgensen, J. E., Downs, R. A. and Kow, S. P., "Fringe-field
capacitive profilometry", Surface Topography, 1, 131-142 (1988)
Gee, S., King, W. L., and Hegmon, R. R., "Pavement texture measurement by
laser: a feasible study", in Surface texture versus skidding: measurements,
frictional aspects and safety features of tire-pavement interactions, STP 583, 29-
41 (ASTM, 1975).
George, A. F., "A comparative study of surface replicas", Wear, 57, 51-61
(1 979).
108 Rough Su6ace.s

Gourley; D., H. E. Gourley and J. M. Bennett, "Evaluation of the


microroughness of large flat or curved optics by replication." Thin Solids Films,
124, 277-282, (1985)
Graneek, M. and Wunsch, H. L., "Application of pneumatic gauging to the
measurement of surface finish", Machinery, 81, 701-707 (1952).
Grieve, D. J., Kaliszer, H., and Rowe, G. W., "A normal wear process
examined by measurements of surface topography", Ann. C.I.R.P., 18, 585-592
(1970).
Herschman, H. K., "Replica method for evaluating finish of a metal surface",
Mech. Eng. 67, 119-122 (1945).
Jeng, Y.-R., Lalonde, G. A., "3-D surface topography measurement system
and its applications", in Special Publications 936, 175-182 (SAE,Warrendale, PA,
1992)
Karlsson, R.I., "Effect of irregular surface roughness on the frictional
resistance of ships", Proc. Int. Symp. on Ship Viscous Resist. 9, 1-20 (Gothenburg,
1978)
Kennedy, C. W., Hoffman, E. G., Bond, S. D., Inspection and gaging 6e
(Industrial Press, New York, 1987).
King, M. J. and Thomas, T. R., "Stylus measurement of the microgeometry of
a coated surface", J. Coatings Tech., 50, 56-61 (1978)
King, M. J., "The measurement of ship hull roughness", Wear 83, 385-397
(1982).
King, M. J., Chuah, K. B., Olszowski, S. T. and Thomas, T. R., "Roughness
characteristics of plane surfaces based on velocity similarity laws", ASME Paper
81-FE-34 (1981)
Kiyono, S., Yamatani, M., Ohe, A., Huang, P., Suzuki, H., "Critical angle
type optical stylus with optical skid (2nd report) - construction by two optical
sources and one receiving optical system", Seimitsu Kogaku Kaishi/Journal of the
Japan Society for Precision Engineering 60, 114-118 (1994)
Lech, M., I. Mruk and J. Stupnicki, "Comparison of tribological parameters
of surfaces determined by the stylus method and by the immersion method of
holographic interferometry." Wear, 93, 167-179 (1984)
Lonardo, P. M., "Testing a new optical sensor for in-process detection of
surface roughness", Ann CIRP 27, 531-534 (1978)
Mathia, T. G., Brugirard, J. L., Duarte, J. and Maurin-Perrier, B., "Enamel
and hydrocolloide dental replica surfaces: Part 1. Statistical characterisation of
enamel topography" Surface Topography, 2, 157-172 (1989a)
Other Measurement Topics 109

Mathia, T. G., Brugirard, J. L., Balleydier, M., Duarte, J. and Maurin-


Perrier, B., "Enamel and hydrocolloide dental replica surfaces: Part 2. New
statistical criteria for evaluation of replica fidelity", Surface Topography, 2, 173-
191 (1989b)
Mitsui, K., "In-process sensors for surface roughness and their applications.",
Precision Engineering, 8, 212-220, (1986)
Mitsui; K., N. Ozawa and T. Kohno, "Development of a high resolution in-
process sensor for surface roughness by laser beam.", Bulletin of the Japan Society
of Precision Engineering, 19, 142-143, (1985)
Miyoshi, T., Takaya, Y., Saito, K., "Nanometer measurement of silicon wafer
surface texture based on Fraunhofer diffraction pattern", CIRP Ann. 44, 489-492
(1995)
Mollenhauer, C., "Surface topography measurement techniques", Proc. Int.
Con$ on Surface Technol., Pittsburgh, 173-186 (SME, 1973).
Morrison, E., "A prototype scanning stylus profilometer for rapid
measurement of small surface area", Int. J. Mach. Tools Manufact. 35, 325-33 1
(1995)
Narayanasamy, K., V. Radhakrishnan and R. G. Narayanamurthi, "Analysis
of surface reproduction characteristics of different replica materials", Wear, 57, 63-
69 (1979)
Newman, P. T., Radcliffe, S . J. and Skinner, J., "The accuracy of
profilometric wear volume measurement on the rough LClB coated surfaces of an
articulating joint", Surface Topography, 2, 59-77 (1989)
O'Callaghan, P. W. and Probert, S. D., "Effects of static loading on surface
parameters", Wear, 24, 133-145 (1973).
Ohlsson, R., and Rosen, B.-G., "On replication and 3D stylus profilometry
techniques for measurement of plateau-honed cylinder liner surfaces", in R. J.
Hocken ed., Proc. ASPEAnnual Meeting, Seattle, 146-149 (1993)
Pearson, J. and Hopkins, M. R., "Plastic replicas for surface-finish
measurement",J. Iron & Steel Inst. 67-70 (May, 1948).
Person, U., "Real time measurement of surface roughness on ground surfaces
using speckle-contrast technique", Optics and Lasers in Engineering 17, 6 1-67
( 1992)
Peters, J., "Messung des Mitterauswertes Zylindrischer Teile Wahrend des
Schleifens", VDI-Berichte 90, 27 (1965)
Sawyer, J. W.,"Method for recording roughness of submerged surfaces", Am.
SOC.Nav. Eng. J., 65, 816-821 (1953).
110 Rough &$aces

Sayles, R. S., Thomas, T. R., "Mapping a small area of a surface", J. Phys. E:


Sci. Instrum. 9,855-861 (1976).
Sherrington, I.; Smith, E. H., "Design and performance assessment of a
Kelvin clamp for use in relocation analysis of surface topography", Precision
Engineering 15, 77-85 (1993)
Shin, Y. C.; Oh, S. J.; Coker, S. A., "Surface roughness measurement by
ultrasonic sensing for in-process monitoring", Trans. ASME: J. Eng. Ind. 117,
439-447 (1995)
Shiraishi, M., "In-process measurement of surface roughness in turning by
laser beams", ASME Paper 80-WAPROD- 17 (1980)
Shunmugam, M. S. and Radhaknshnan, V., "An analysis of the reference
lines of the surface profile and its true replica", Wear, 40, 155-163 (1976)
Spurgeon, D. and Slater, R. A. C., "In-process indication of surface
roughness using a fibre-optics transducer", Proc. 15th Int. Machine Tool Des. &
Rex Con$, Birmingham, 339-347 (1974).
Song, J. F., Vorburger, T. V. and Rubert, P., "Comparison between precision
roughness master specimens and their electroformed replicas. ", Precision
Engineering 14, 84-90 (1992)
Stout, K. J., King, T. G. and Whitehouse, D. J., "Analytical techniques in
surface topography and their application to a running-in experiment", Wear, 13,
99-1 15 (1977)
Takaya, Y., Miyoshi, T., Arai, M., Hayashi, K., Setaka, M., "Development of
random micro-roughness measuring apparatus based on Fraunhofer diffraction -
subnanometer measurements by the new error calibration method", Seimifsu
Kogaku Kaishi/Journal of the Japan Society for Precision Engineering 61, 377-
381 (1995)
Takeyama, H., Sekiguchi, H., Murata, R., Matsuzaki, H., "In-process
detection of surface roughness in machining", Ann CIRP 25,467-471 (1976)
Teague, E. C., "Scanning tip microscopies: an overview and some history", in
G. W. Bailey ed., Proc. 46th. Annual Meeting of the Electron Microscopy Society
ofAmerica, 1004-1005 (San Francisco Press, San Francisco, 1988)
Terman, F. E., Radio Engineering 2e (McGraw-Hill, New York, 1937)
Thomas, T. R., "Trends in surface roughness", Trans. 7th Int. Con$ on
Metrology & Properties of Engineering Surfaces (Chalmers University,
Gothenburg, 1997)
Thomas, T. R., "Computer simulation of wear", Wear, 22, 83-90 (1972).
Timms, C. and Scoles, C. A., "Some applications of the plastic replica
process to surface finish measurement", Machinery, 73, 871-875 (1948)
Other Measurement Topics 111

Tipton, H. and Roberts, J. I., "New optical method of assessing surface


quality", Proc. I. Mech. E., 182, Part 3K, 274-278 (1967/68).
Tolansky, S., Multiple-beam interferometry of surfaces and f h s , (Dover
Publications, Inc., New York, 1970a).
Tolansky, S., Multiple-beam interference microscopy of metals, (Academic
Press, London, 1970b).
Tolansky, S., Surface microtopography, (Longmans, London, 1960).
Unsworth, A., and Hepworth, A,, "A new stereo-adapter for use with the
scanning electron microscope", J. Microscopy 94, 252 (1971)
Vorburger, T. V.; Teague, E. C., "Optical techniques for on-line
measurement of surface topography", Precis. Eng. 3,61-83 (1981).
Wager, J.G., "Surface effects in pneumatic gauging", Znt. J. Mach. Tool Des.
Rex, 7, 1-14 (1967).
West, R. N., and West, P., "New applications of laser scanners for on-line
product inspection", Proc. 3rd. Conf on Automated Inspection & Product Control,
133-138 (Nottingham University, 1978)
Williamson, J. B. P. and Hunt, R. T., "Relocation profilometry", J. Phys .E:
Sci .Instrum., 1, 749-752 (1968).
Woolley, R. W., "Pneumatic method for making fast, high-resolution.
noncontacting measurement of surface topography", Proc. SPIE 1573, 205-2 15
(1991)
Yanagi; K., T. Maeda and T. Tsukada, "Practical method of optical
measurement for the minute surface roughness of cylindrical machined parts.",
Wear, 109, 57-67, (1986)
Young, A. P. and Clegg, B. H., "Replica method for examining surface
profiles" Rev. Sci. Instrum., 30, 444-446 (1959).
Young, R. D.; Vorburger, T. V.; Teague, E. C., "In-process and on-line
measurement of surface finish", Ann. CIRP 29,435-440 (1980).
Zhao, Y. W. and Webster, J., "An in-process roughness measuring system for
adaptive control of plunge grinding", Surface Topography, 2, 247-26 1 ( 1989)
CHAPTER 6

DATA ACQUISITION AND FILTERING

6.1. Data Acquisition

The use of digital techniques is now so widespread that one is unlikely to find any
new roughness measurement instrument which relies solely on analogue methods.
Even if the transducer itself is non-electrical, processing and presentation are likely
to involve digital electronics. The process of analogue-to-digital conversion (ADC)
amounts to the representation of the continuous analogue signal by a series of
discrete numbers. This discretisation occurs in two ways (Fig. 6.1). In the
amplitude domain, the signal is split into a number of levels parallel to the plane of
the surface. This process is called quantisation. In the frequency or wavelength
domain, the instantaneous value of the signal is recorded at equal intervals in the
plane of the surface. This process is called sampling.

Amplitude domain: quantisation

Frequency domain: sampling

Figure 6.1. Analogue-to-digital conversion: quantisation and sampling

113
114 Rough Suflaces

The number of quantisation levels is determined by the resolution of the ADC


hardware, usually expressed as a number of bits (powers of 2). Thus an 8-bit
converter will quantize to 256 levels and so on, that is it will resolve amplitude
variations to 1 part in 256. There are several points to watch here. One is that
quantisation is likely to represent the limiting resolution of the overall
measurement system; there is no point spending money on a transducer capable of
resolving to one part in a thousand if its output is to be processed by an 8-bit ADC.
The second is that the computing arrangements may themselves embody several
components, and the limiting quantisation of the system will be the lowest of any
of the components. A 16-bit ADC followed by an 8-bit processor, or a processor
running software which will only represent numbers to 8 bits, is still an 8-bit
system.

u ” ul
L
Figure 6.2. Too few quantisation levels cause loss of detail

In deciding on an appropriate quantisation level it must be remembered that


the quoted figure represents the entire range of the transducer, what for analogue
instruments would be called the full-scale deflection (FSD). In practice, an
instrument is usually set up so that the hghest peak in the entire signal is safely
below FSD and the lowest valley is similarly above zero. Thus it can easily happen
that for a nominal quantisation of 8 bits, for two-thirds of the length of a signal
with a Gaussian amplitude distribution the variation in height is actually
represented by only about 50 levels. This can lead to a sih:ation (Fig. 6.2) where
small peaks which may be functionally significant are missed altogether. The
remedy is obviously to increase the quantisation, but remembering that there will
be a corresponding price to pay in longer conversion times and larger data storage
requirements. Most commercial systems use at least 16 bits, and quantisation of
less than 10 or 12 bits is not recommended.
Similarly if the signal is sampled too often it will lead to data storage and
processing difficulties. But if it is sampled too seldom, then any wavelengths
Data Acquisition & Filtering 115

present in the signal which are shorter than the sampling interval may be
misinterpreted as longer wavelengths of the same amplitude (Fig. 6.3a). This
effect is known as aliasing. According to the Nyquist sampling theorem (Wade
1994), the shortest measurable wavelength AN is twice the sampling interval. The
effect of aliasing is to mirror the power spectrum of the aliased frequencies about
the Nyquist frequency oN, (Fig. 6.3b), so that a real frequency oN + o appears as
an aliased frequency of wN - w.

'Apparent waveform
Sample points

Figure 6 .3 . Aliasing: (a) short wavelengths misinterpreted as longer wavelengths by sampling too seldom; (b)
aliased frequencies reflected about Nyquist frequency falsify power spectrum.

The sampling interval is usually matched to the dimensions of the probe.


The simplest way to sample is "on-the-fly", that is to keep the translation
mechanism in constant motion and sample the resultant signal at equal intervals of
time, assuming that these represent equal intervals of horizontal distance. This
will only be true if the translation is at constant speed, which is not always safe to
assume (see Section 2.2.3). It must also be remembered that the actual ADC
process takes a finite time, and that this time represents an integration of fine
surface detail over a finite length of the signal, though this length is usually small
compared with the sampling interval. The alternative to on-the-fly sampling is the
added complication of an independent measurement of horizontal displacement.

6.2. Filtering

It very rarely happens that the output of any instrument is perfectly matched to the
application for whch it is intended. Sometimes the instrument produces too little
116 Rough Surfaces

information for the intended purpose. More often it produces too much
information, and the useful information has to be extracted or the extraneous
information suppressed. In electrical engineering terms, theuseful information is
the signal and the extraneous information is noise, and separating the noise from
the signal is an essential preliminary to characterisation. This process of
separation is calledjltering.
The concept of filtering is borrowed from electronic engineeering, where the
signal and the noise are both treated as essentially sinusoidal, continuous (i.e.
indefinitely long signals are available for measurement and analysis) and stationary
(i.e. increasing the length of the signal does not change the information present in
it), and the problem therefore resolves into one of sorting groups of sinusoids. This
approach is ill-suited to describing surfaces for several reasons: common
experience tells us that surfaces are not sinusoids; the instruments which measure
them do not produce continuous signals; and many real surfaces are not stationary.
Nevertheless, because the properties of signals have been widely worked out in
sinusoidal terms, it will be convenient for the time being to use this terminology.
Restricting ourselves to two dimensions for the time being, we may think of a
generalised measured profile consisting of a continuous spectrum of surface
wavelengths. The width of the spectrum, i.e. the range of wavelengths, is fixed by
the measuring instrument itself. The instrument will be unable to ”see” any
wavelengths longer than its traverse length, and in fact the longest wavelength
reliably represented in the spectrum will be only a fraction of the traverse length
(see the later discussion on measurement of power spectra). At the other end of the
spectrum, the instrument will be unable to see any profile wavelengths smaller than
the dimensions of its own sensor.
But there is no a priori reason why the spectrum of wavelengths produced by
the instrument, designed as a general purpose device to suit a range of applications,
should coincide with the spectrum of wavelengths associated with any particular
application. In engineering metrology, for instance, the spectrum was once divided
into six (DIN 4760, 1982), of which only the first three classifications need detain
us here. The longest wavelengths are associated with errors of form; shorter
wavelengths constitute waviness; and the shortest wavelengths are called
roughness (Fig. 6.4a). DIN 4760 prescribes that the length of an error of form
should be at least 1000 times its amplitude, and the ratio of wavelength to
amplitude of the waviness should be between 1OO:l and 1OOO:l. This division is
quite arbitrary; wavelengths associated with errors of form on a machined surface
would be described as roughness on a ship’s hull, for instance. In manufacturing
engineering, the problem is usually to remove the waviness and errors of form from
Data Acquisition & Filtering 117

the signal so that the remaining property of the surface wtiich is assessed is the
roughness. Note again that there is no generally agreed wavelength which divides
roughness from waviness, it is a matter for subjective assessment.

of danimni
Catternl
Error of

1 I Wavelength

Figure 6.4. (a) Surface characteristics(Brooker 1984) and (b) their associated power spectrum (Thomas
1975)

One important reason for this removal is that for almost all real surfaces, the
longer the wavelength, the larger the amplitude (Fig. 6.4b). This is a typical
property of self-af€ine fractals, as we shall see in a later chapter, and it has the
consequence that the numerical value of any parameter which depends on the
amplitude properties of a profile will be dominated by the longest wavelengths
present. For many practical purposes, then, filtering means hzgh-puss filtering. In
the language of communication engineering, a high-pass filter stops long
wavelengths (low frequencies) but passes short wavelengths (high frequencies).
Using this terminology permits us access to the large body of existing work
on digital filter design (see for example Wade 1994, Rorebaugh 1997, Golten
1997). Although this body of work is mainly concerned with low-pass filters,
which stop short wavelengths but pass long wavelengths, the conceptual arguments
are similar, as a high-pass filter is equivalent to subtracting the output of a low-
pass filter from the original signal. Consider the internal computer representation
of a profile, after analogue-todigital conversion, as a series of discrete heights.
Conceptually a filter may be thought of as a sequence of weighting terms (the
118 Rough Sur$aces

impulse response) which is moved along the profile z(i), multiplying it term by
term and thus smoothing it as it goes (Fig. 6.5).

function
WelgM1ng m

Figure 6.5. Convolution as a low-pass filter (Golten 1997)

The sequence itself may be chosen arbitrarily, say of length 2m + 1 terms, and the
output of the filter z’(i) is the convolution of the impulse response with the original
signal :

The transmission coeflcient at a given wavelength is the ratio of the


amplitude of the filtered signal at that wavelength to that of the unfiltered signal.
The cutoffof the filter is the wavelength at which the transmission coefficient is
some specified fraction of the input amplitude (Fig. 6 . 6 ) . The roll-off
characteristic of the filter describes how sharply the transmission coefficient varies
on either side of the cutoff. Clearly for roughness work as sharp a roll-off as
possible is desirable. Eqn. 6.1 implies that an infinitely sharp cutoff, where
transmission is 100% right up to the cutoff and zero immediately after it, is not
realisable. Thus all real high-pass filters let some long-wmelength power bleed
through and simultaneously stop some short-wavelength power which should have
been admitted. All filter design is therefore a compromise. The large number of
hfferent designs of filter which have been proposed all have particular advantages
and disadvantages.
Data Acquisition & Filtering 119

025 06 25 80
Wnvelerqth in mm

Figure 6 .6 . Transmission curves showing roll-off charactenstics at four standard cutoff lengths for a 2CR
high-pass analogue filter (BS 1134, 1988)

Causal filters can only operate with historic data, that is for values of z (x + i)
for which i < 0. All physically realisable filters are causal, because a filter
operating in real time cannot "see" future data. But this is not a limitation for
many roughness measurements, where pre-recorded data are being filtered, and
non-causal filters may operate on values of z(x + i) wit]? positive values of i .
Recursive filters are modified by a feedback from the output. Recursive filters are
also called inJnite impulse response (IIR) filters because their weighting sequence
cannot be represented by a fixed number of terms, whereas the weighting sequence
of non-recursive filters is finite (FIR). IIR filters may potentially be unstable,
whereas FIR filters are always stable. Because of the feedback mechanism implicit
in IIR filters, fewer terms are needed to achieve a similar roll-off performance and
so IIR filters are more efficient. On the other hand, linear phase response
characteristics are impossible to achieve with IIR filters, whereas non-causal FIR
filters do not introduce any phase distortion. The relative advantages of IIR and
FIR filters for roughness assessment are discussed at length by Medhurst (1989).
To determine the frequency response of the filtx, that is how the
transmission coefficient varies with wavelength, it is necessary to take the Fourier
transforms of the weighting sequence and the input signal and multiply them
together. Their product is the Fourier transform of the frequency response, which
may be recovered by inversion. This multiplication in the frequency domain is
formally equivalent to convolution in the distance domain. So it is possible to
proceed in the reverse direction, to decide on a required frequency response and
hence to deduce the appropriate weighting sequence.
120 Rough Su$aces

To begin with filters were analogue, operating on the actual instrument


output in real time with physical components. The first internationally agreed
filter (IS0 3274, 1975) was a cascade of two resistor-capacitor circuits, the so-
called 2CR filter (Fig. 6.6), with a cutoff of 75%. The mean line with respect to
which roughness parameters were calculated was taken as the zero voltage level of
the output (the so-called M-system). This filter had an asymmetrical weighting
function, the practical effect of which was to cause phase distortion of the output
signal (Fig. 6.7).

Figure 6.7. Phase distortion caused by 2CR filter (Whitehouse & Reason 1965)

To avoid this distortion, Whitehouse (1967/68) proposed the use of a phase-


corrected (PC) filter. He noted that a suitable weighting function would need to be
concentrated in a central lobe and die away quickly on either side, though not so
quickly as to produce undesirable harmonics. The effects on various machined
surfaces of phase-corrected filters were computed by Shunmugam &
Radhakrishnan (1976a). Whitehouse considered various weighting functions,
including a Gaussian weighting function which satisfied the above criteria, but
practical realisation was limited by the existing state of technology. With the
advent of fast cheap digital processing, Gaussian PC filters are now practicable and
have become the standard (DIN 4777, 1990). I S 0 11562 (1996) prescribes the
weighting function as

where a = 0.4697 is a numerical constant. The cutoff is now defined as 50%


transmission so that the mean line of the profile may be extracted by subtracting
Data Acquisition & Filtering 121

the high-pass filtered profile from the original, and the transmission coefficient at
wavelength A, that is the ratio of the unfiltered amplitude a. to the filtered
amplitude, is

an/ao= 1 - exp f -71 (&/A) }

As well as desirable phase characteristics, the PC filter also has a sharper roll-off
than the 2CR filter (Fig. 6.8).

Figure 6.8. Frequency responses of 2CR and PC filters compared for a cutoff of 0.8 mm

There are many other possible filtering techniques, some of which are in
common use. Most are not capable of representation in terms of Eqn. 6.1, so
although it is usually possible to define their cutoff, it is rarely possible to specie
their frequency response. Earlier standards described the decomposition of the
profile into a number of consecutive equal sample lengths (BS1134, 1972). To
each sample length a separate straight mean line is fitted, by least squares or an
equivalent method, then all the mean lines are joined in a single straight line (Fig.
6.9). This is an effective high-pass filter, but its operation cannot be specified in
terms of Fourier transforms. Also, unless a separate algorithm is used to match the
ends of the profile segments, the resulting discontinuities will give rise to spurious
short wavelengths in the output spectrum.
122 Rough Su$aces

Figure 6.9. High-pass filtering by fitting straight lines to consecutive sample lengths (BSll34, 1972)

A polynomial filter (Fig. 6.10) can use any one of a number of well-known
numerical techniques to fit a least-squares polynomial to the input profile. This
and the previous filter are interesting as examples of techniques which will work
adequately for roughness measurements with their short signal lengths, but would
be unsuitable for the continuous signals of communications engineering.

Figure 6.10. Profile (a) fitted with a 12" order polynomial (b) after subtractingthe polynomial

Another filtering technique whose behaviour is difficult to quantify is the


valley suppression filter (Schneider et al. 1988), prescribed by DIN 4776 (1990)
and I S 0 13565 (1996) for use with surfaces containing deep valleys. Filtering
takes place in three stages (Fig. 6.11). First, a high-pass PC filter is applied to
determine the mean line. Next, all valleys below the mean line are removed and
the profile is filtered again. Finally, the original profile is referred to the mean line
from the second filter.
The practical implementation of a filter algorithm is fraught with problems.
How wide should the impulse response be? If it is too narrow, it will be fast but the
filter will ring. If it is too wide, it will be slow, it may not have enough profile to
process in order to give a stable result, or alternatively if it overlaps the ends of the
profile it may create end effects. Unfortunately the width of the weighting function
Data Acquisition & Filtering 123

Step 1

step 3

Figure 6.1 1 . The three stages of the valley suppression filter (Mummery 1990)

is not specified in the standards, and manufacturers tend to make their own
decisions on commercial grounds, decisions whch they are not likely to share with
their customers. This is called "method divergence" and is just one reason why
software from two suppliers, both conforming to the letter of the standards, may
give quite different roughness values for the same profile (Fig. 6.12). However,
from Eqn. 6.2, the standard deviation of a Gaussian filter 0Y 0.19Ac. If the width
is taken as ?C 3 s equivalent to 99.5% of the possible area of the weighting
function, then 6 0 = 1.I&; that is, conveniently, the width of the window is about
the length of the cutoff.
Low-pass filters, which let through long wavelengths and stop short
wavelengths, are not so much used in roughness work &cause the instrument
sensor, as already pointed out, is its own low-pass filter, and because in any case
short wavelengths, with their small amplitudes, have little effect on amplitude-
based parameters. If for whatever reason a low-pass filter is required, a simple
running average is often enough. If something more elaborate is required, many
mathematical software packages offer spline filters. There is a good deal to be said
for a first difference check on new profile data in any case. If this throws up
occasional slopes of say 30 degrees or greater, when the steepest slope which the
sensor will measure is 10 degrees, then there is a problem which needs dealing
with before any further processing.
International standards prescribe a range of preferred cutoff wavelengths,
roughly in the ratio of 3 : l . The cutoffs most used on measuring instruments for
124 Rough Sur$aces

Unfiltered profile

2 Rc fitter

Phase Corrected
filter (DIN 4777)

ValleySup ression
mer (DIN8776)

Figure 6.12. The effect of three different standard filters on the same profile (Mummery 1990)

. .:
. .
. .
. .
a .
.. - ...
.. ..
. ..
. ... . . *
.. . . _.
.I

1.

2 4 6 U 10 12 14 16 18 28
Rq (microns) at 2.5 mm cutoff

Figure 6.13. Roughness of 200 profiles measured on the same surface at high-pass cutoffs of 2.5 mm and 50
mm (Medhurst 1989)
Data Acquisition & Filtering 125

production engineering are 2.5 mm, 0.8 mm, and 0.25 mm. As has already been
mentioned, amplitude-dependent roughness parameters increase with high-pass
cutoff wavelength (often approximately as its square root, see later), so when a
roughness parameter is quoted, on a drawing or elsewhere, it is necessary also to
spec@ the cutoff. If no cutoff is specified, a default value of 0.8 mm is assumed.
This curious distance apparently arises because it was the width of the field of view
in the earliest light-section microscopes, though I cannot trace a reliable source for
this story.
It is sometimes suggested that choice of cutoff is not particularly important if
measurements are being used for comparative purposes only. It is argued that if
roughness values are ranked in a particular order when measured at one cutoff,
then they will probably be ranked in the same order when measured at any other
cutoff. This amounts to claiming that there is a strong correlation between
measurements made at different cutoffs. Fig. 6.13 should be enough to disabuse
most readers of this misapprehension.

6.2.1. Envelope Filters

The M- or mean-line system, now undisputed, was once in contention with the so-
called E- or envelope system. In the E-system (von Weingraber 1957), the
reference lines are defined by the loci of centres of circles of different radii rolled
along the profile. The locus of the centre of the larger circle gives the curve of
form (Formprojl) while that of the smaller circle gives the contacting profile
(Hullprojil) (Fig. 6.14). The geometrical projle is now drawn; this is defined as a
profile of the surface determined by the design, neglecting errors of form and
surface roughness. The area between the geometrical profile and the curve of form
represents the errors of form; the area between the curve of form and the contacting
profile represents the secondary texture or waviness; and the area between the
contacting envelope and the effective profile (defined as the nearest instrumental
approximation to the true profile) represents the primary texture or roughness. The
E-system mean line is defined as the contacting envelope displaced downwards by
a distance such that the areas enclosed by the effective profile above and below it
are equal.
The advantages of the E-system over the M-system are claimed to be that the
E-system is physically more significant in that many engineering properties of a
surface are determined by its peaks, and that the mean line is easier and quicker to
construct graphically. The contacting profile and curve of form can be obtained
126 Rough Sur$aces

Geometrical profile
I

Locus of centre for


rolling circle
Curve of form
w w - - (Formprofil)

Contacting envelope
Errors of form (Hiillprofil)
Waves Effective profile
Peak to valley

Figure 6.14. Terminology ofthe E- (envelope) system of reference lines in which the filters are two circles of
radius r and R rolling along the profile (Olsen 1963)

mechanically by using skids of appropriate shape, and instruments so designed


were apparently commercially available at one time. Algorithms for computing the
various reference lines have also been published (Shunmugam & Radhakrishnan
1976b). The skids were in fact spheres rather than circles, causing an integrating
effect which can alter the form of the various reference lines (Shunmugam &
Radhakrishnan 1974). Standard radii were 25 mm for roughness and 250 mm for
waviness, though other radii have been proposed (Radhakrishnan 1972).
The system does, however, exhibit certain disadvantages. On a visual
display, where the vertical magnification is exaggerated, the rolling circle becomes
a "rolling ellipse". The mean line, composed as it is of a succession of intersecting
arcs, is mathematically discontinuous. Attempts to reconcile the two systems
(Ishlgaki & Kawaguchi 1981) have not been successful as it is difficult to represent
the envelope action as a weighting function.
A successor to the E-system has gained wider acceptance. This is the French
system of so-called motif analysis, which is now the subject of an international
standard (IS012085, 1996). Motif analysis is more than just a filtering technique,
it is a whole new method for classifying surfaces, but it will be convenient to deal
with it here. It is basically a set of algorithms which purports to embody the
collective expertise, based on subjective visual judgement of profiles, of inspectors
Data Acquisition & Filtering 127

Figure 6.15. Motifanalysis (Mummery 1990). For explanation see text.


128 Rough Surfaces

in the French automobile industry. The conceptual foundation of the method is


described by Fahl (1982). An experienced inspector, scanning a profile chart
recording by eye, picks out a number of characteristic featuies or motifs which he
knows are typical of the surface and on which the fitness of the surface for its
purpose may be assessed. This is essentially a pattern-recognition process at which
the human brain is known to excel. Motif analysis is an attempt to reduce this
complex and almost instinctive process to a list of instructions.

T?<0,6 TR

Figure 6.16. The four conditions for motif combination (Fahl 1982). For explanation see text

The process consists of a sequence of steps (Fig. 6.15). (1) The profile is
divided into a number of windows, each normally 250 pm wide, The peak-to
valley height of the profile within each window is calculated and averaged over all
the windows. (2) Peaks less than 5% of the average peak-to-valley height are
eliminated. (3) The profile is divided into stylized patterns, each of which must
Data Acquisition & Filtering 129

contain a valley bounded by two peaks. (4) The patterns are combined according to
a complex and arbitrary set of rules, given below, so as tr, minimise their total
number. (5) The profile is now represented by a series of linked patterns. (6) The
peak-to valley height of the profile within each window is again calculated and
averaged over all the windows. Any peaks or valleys which are more than 1.65
standard deviations from the mean are reduced to this value, and the average peak-
to-valley height is once more calculated. (7) The upper envelope line is determined
by joining the tops of the roughness patterns. (8) Waviness patterns are
determined from the upper envelope line, using the same criteria for motif
combination, but this time starting with windows 2.5 mm wide.
In arriving at numerical values the rules for combination of motifs are clearly
crucial. There are four conditions (Fig. 6.16): two adjacent motifs cannot be
combined if their common middle peak is larger than thc two outer peaks; (b)
combined motifs cannot be wider than 500 pm; (c) adjacent motifs may not be
combined if the characteristic depth (the smaller of the two individual peak-to-
valley heights within each motif) of the result is less than the characteristic depth
of each of the original motifs; (d) adjacent motifs may not be combined if at least
one has a characteristic depth less than 60% of the combined characteristic depth.
These conditions are to be applied iteratively until no more motifs can be
combined.
I1101

Wavelength, l l f k jm

Figure 6.17. Comparison of the power spectra of various attempts to filter a profile of a ground surface
(Shunmugam 1987). Continuous line, unfiltered profile; chained line, 2CR filter mean line; dashed line,
rolling circle envelope; dotted line, motif upper envelope
130 Rough Su$aces

Considerable claims have been made for motif analysis. Boulanger (1992)
describes the method as a filter for separating waviness from roughness with "an
absolutely sharp cutoff'. Dietzsch et al. (1997) compare an unfiltered profile
assessed by motif analysis with a profile high-pass filtered according to the M-
system, and not surprisingly find that the M-system has removed the waviness.
Clearly the process described above is not readily represented in algebraic terms,
though Scott (1992) has developed a formally rigorous definition of a motif.
However, Shunmugam (1987) has tried to compare the motif method's
performance with that of other filters for specific surfaces. In Fig. 6.17, where the
power spectrum of the upper envelope line from a m o analysis ~ is compared, for
the same profile, with the envelope of a rolling circle of 3.2 mm radius and the
mean line of a 2CR high-pass filter with 250 pm cutoff, the roll-off of the motif
envelope does not seem to be particularly sharp.

6.3. References

Boulanger, J., "'Motifs' method. Interesting complement to I S 0 parameters


for some functional problems", International Journal of Machine Tools &
Manufacture 32,203-209 (1992)
Brooker, K. ed., Manual of British standards in engineering metrology
(British Standards Institution, London, 1984)
BS 1134 Part 1, Assessment of surface texture: methods and instrumentation
(British Standards Institution, London, 1988)
BS 1134 Part 2, Method for the assessment of surface texture: general
information and guidance (British Standards Institution, London, 1972)
Dietzsch, M., Papenfuss, K., Hartmann, T., "The motif-method - a suitable
description for functional, manufactural and metrological requirements", Trans.
71h. Int Con$ On Metrology & Properties of Engng Surfaces, 231-238
(Gothenburg, 1997)
DIN 4760, Surface irregularities - terms and definitions - classification
system (Deutsches Institut f i r Normung, Berlin, 1982)
DIN 4776, Measurement of surface roughness: parameters Rk, Rpk, Rvk,
Mrl, Mr2 for describing the material portion (projile bearivg length ratio) in the
roughness profile: measuring conditions and evaluation procedures (Deutsches
Institut f i r Normung, Berlin, 1990)
DIN 4777, Profile filters for electric stylus instruments - phase-corrected
filters (Deutsches Institut f i r Normung, Berlin, 1990)
Data Acquisition & Filtering 131

Fahl, D., "Motifcombination", Wear 83, 165-179 (1982)


Golten, J., Understandingsignals and systems (McGraw-Hill, London, 1997)
Ishigaki,H. and I. Kawaguchi, "Analysis of the initial contact height when a
curved body approaches a rough flat surface", Wear, 54,273-289 (1979)
I S 0 3214, Instruments for the measurement of surface roughness by the
profile method - contact (stylus) instruments of consecutive projle transformation
- contact profile meters, system M (ISO, Geneva, 1975)

IS0 11562, Geometricproduct specijications (GPS) surface texture: projle


~

method - metrological characteristics of phase correctfilters (ISO, Geneva, 1996)


I S 0 12085, Geometric product specification (GPS) surface texture: profile
~

method motifparameters (ISO, Geneva, 1996)


~

I S 0 13565-1, Geometric product specifcations - surface texture: profile


method surfaces having strati$ed functional properties: Part 1. Filtering and
~

general measuring conditions (ISO, Geneva, 1996)


Medhurst, J. S., The systematic measurement and correlation of the frictional
resistance and topography of ship hull coatings, with particular reference to
ablative antifoulings, PhD thesis, Newcastle University (1989)
Mummery, L., Surface texture analysis: the handbook, (Hommelwerke,
Muhlhausen, 1990).
Olsen, K.V., "The standardization of surface roughness", Proc. Znt. Prod.
Eng. Res .Con$, Pittsburgh, 8,655-658 (1963)
Radhakrishnan, V., "Selection of an enveloping circle radius for E-system
roughness measurement", Int. J. Mach. Tool Des. & Res. 12, 151-159 (1972).
Rorabaugh, C. B., Digital filter designer's handbook 2e (McGraw-Hill, New
York, 1997)
Scott, P. J., "Mathematics of motif combination and their use for functional
simulation", International Journal of Machine Tools & Manufacture 32, 69-73
( 1992)
Shunmugam, M. S., "Comparison of motif combination with mean line and
envelope systems used for surface profile analysis", Wear 117,335-345 (1987)
Shunmugam, M. S., Radhaknshnan, V., Tomparison of the mean lines of
surface profiles obtained from different electrical filters", Zsr. J. Technol. 14, 253-
260 (1976)
Shunmugam, M. S. and Radhakrishnan, V., "Two- and three-dimensional
analyses of surfaces according to the E-system", Proc. I. Mech. E., 188, 691-699
(1974).
132 Rough Surfaces

Shunmugam, M. S.; Radhakrishnan, V., "Comparison of different methods


for computing the two-dimensional envelope for surface finish measurements",
Comput Aided Des. 8,89-93 (1976).
Thomas, T. R., "Recent advances in the measurement and analysis of surface
microgeometry", Wear, 33, 205-233 (1975).
Von Weingraber, H., "Suitability of the envelope line as a reference standard
for measuring roughness", Microtecnic, 11, 6-17 (1957) .
Wade, G., Signal coding and processing 2e (Cambridge University Press,
1994)
Whitehouse, D. J. and Reason, R. E. The equation of the mean line of surface
texture found by an electric wave filter", (Rank Precision Industries, Leicester,
1965).
Whitehouse, D. J., "An improved type of wavefilter for use in surface finish
measurement", Proc. I. Mech. E., 182, Part 3K, 306-3 18 (1967/68).
CHAPTER 7

AMPLITUDE PARAMETERS

In presenting roughness parameters we will proceed in more or less historical


order. We will begin with amplitude parameters, starting with extreme-value
parameters and going on to average parameters and properties of the height
distribution. We will then introduce the height distribution and its moments,
which will lead us to a discussion of multiprocess surfaces and their description.
Turning to texture parameters, we will discuss autocorrelation and the correlation
length, leading to the power spectrum and its properties. This will enable us to
deal with surfaces in three dimensions and to touch on the problems of
nonstationarity. Fractals will be suggested as the logical treatment of
nonstationary surfaces, and finally we will consider the difficulties presented by
anisotropy.
First a note of caution. Because roughness is implicated in so many
technological and scientlfic problems, many workers without a background in
metrology have invented new roughness parameters to describe their own
application, not realising that conventions for the description of roughness already
exist. The consequence of this is what Whitehouse (1982) has termed the
"parameter rash, a proliferation of parameters, possibly running into hundreds,
many of which are redundant in the sense that they are only slightly different ways
of describing the same property, or that they may be derived from each other. No
completely comprehensive account of these parameters exists, and the following
account tries to concentrate on the main parameters of theoretical or historical
importance or which are defined in standards. Hopefully this will help the reader,
on coming across a new parameter, at least to place it into context.
One problem in specifying numerical values of roughness parameters is their
large uncertainty compared to other common engineering length measurements,
because of their inherent statistical nature. In the United Kingdom calibration
system the smallest uncertainty permitted to be claimed for a measurement of Ra
was f 4% at one time (my own calibration laboratory was permitted *7% by the
accrediting body). This may be compared with a typical calibration uncertainty of
better than *0.001% for the length of a gauge block. A common measure of
uncertainty is the coefficient of variation, that is the ratio of the standard deviation

133
134 Rough Sudaces

to the mean value of a set of measurements. Fig. 7.1 compares the coefficients of
variation of a number of commonly used roughness parameters from 10 parallel
profiles measured on a ground surface. Few of these are within sight of 4%: this is
the reality of roughness measurement. Readers who study this figure will perhaps
come to share a certain impatience with standards committees which argue for
years about minuscule points of parameter definition.

Skewness
0 5 0 4 0 3 0 2 01 0 01 0 2 0 3 0 4 0 5

50 40 30 20 10 0 10 20 30 40 50
Coefficient of variation. %

Figure 7.1. Coeficients of variation of a number of roughness parameters measured on a ground surface
across (left) and along (right) the lay, showing also the effect of high-pass filtering (Thomas & Charlton
1981)

7.1. Extreme-value Parameters

The development of parameters to describe roughness followed closely on the


development of instruments to measure roughness. The first reliable measuring
instrument was the light-section microscope,and the only parameter which this
could conveniently measure was the vertical separation of the highest peak and
lowest valley of the unfiltered profile, Pt (DIN 4771, 1977) (Fig. 7.2). (Roughness
parameters are normally denoted by a capital and some following lower-case
letters. Sometimes the following letters are written as subscripts and sometimes on
the line, by analogy with the abbreviations for dimensionless numbers in fluid
Amplitude Parameters 135

dynamics. We will follow the I S 0 convention of writing all amplitude-dependent


parameters with following letters on the line.)

Referenca line

I Evaluation length ,I m

Figure 7.2. Pt is the vertical distance between two parallel straight lines enveloping the unfiltered profile
within the evaluation length Zm (Sander 1991)

When electrical filters became available it was possible to define the


maximum roughness depth Rt (DIN 4762, 1960) in similar terms for the filtered
profile (Fig. 7.3). When Rt is evaluated over only a single sample length it is
known as Ry. Rt does not provide much useful informaticn by itself, so is often
split into Rp, the height of the highest peak above the mean line, and Rm, the depth
of the lowest valley below the mean line.

Upper enveloping line

Evaluation length ,I

Figure 7.3. Rt is the analogue ofPt for a filtered profile (Sander 1991)

Extreme-value parameters are what the statisticians call inefficient


parameters, that is they are unrepresentative of the surface because their numerical
values vary so much from sample to sample, as we can see from Fig. 7.1.
Averaging over several consecutive sampling lengths reduces this variation but it is
still too large to be practical for most purposes. Rt has largely been replaced by the
136 Rough Surfaces

ten-point height Rz, defined as the vertical separation of the average of the 5
highest peaks, and the average of the 5 lowest valleys. Confusingly, there are two
versions of Rz in common use (Fig. 7.4). Rz(IS0) (IS0 4287, 1984) is the average
distance between the 5 highest peaks and the 5 deepest valleys within the
assessment length. Rz(DZN) (DIN 4762, 1989) is the mean of 5 individual values
of Ry for 5 consecutive sample lengths. Needless to say, these two definitions in
general give two different numerical values for the same profile, though Sander
(1991) claims that "for economic reasons" some instruments which pretend to
measure the IS0 parameter actually measure the DIN parameter. A variation of
&(DIN) is R3z, based on the third highest peaks and third lowest valleys. This is
unfortunately far from an exhaustive account of extreme-value parameters; for a
more extensive list see the useful monographs by Mummery (1990) and Sander
(1991).

(a 1

Figure 7.4. Definitions of& according to (a) I S 0 (b) DIN (Mummery 1990)

Even more elaborate peak-to-valley definitions have been devised, notably the
Allison system (Hydell 1967/8) developed by General Motors, and the Swedish H-
system (SMS 47, 1994). Each system positions the peak and valley reference lines
by truncating the higher peaks and lower valleys. The Allison or "general surface
texture" system (Fig. 7.5) truncates the 10% highest peaks and lowest valleys,
similarly the Swedish technique involves a 5% truncation for the upper reference
line and a 10% truncation for the lower line. In the Allison system the reference
lines are termed the upper and lower GST lines and enclose the general surface
texture which is considered as the workable surface between initial wear-in and a
severe wear condition. The difference between the reference lines is used as the
roughness parameter (GST). Outside the GST reference lines the system is
Amplitude Parameters 137

reported to offer a control on the highest peaks and lowest valleys by specifying a
permissible peak height and a a permissible valley depth.

Figure 7.5. The Allison system ofprofile characterisation (Hydell 1967/8)

In discussing extreme-value parameters there is a general problem of the


definition of a peak. Consider Fig. 7.6. Most readers would agree that the large
features A and B are peaks, and probably the smaller feature f, but what about
points a-e and g? Clearly the numerical values of peak-rzlated parameters will
depend on the answer to this question, which does not seem to be generally agreed.
The simplest mathematical definition of a peak is a local maximum in z(x), but this
has two undesirable consequences. Firstly, a peak so defined can occur below the
mean line ( e g point g in Fig. 7 . 6 ) . Secondly and more seriously, the profile as we
deal with it is not a continuous function, and the discrete form of our peak
definition is a point higher than its nearest neighbours. This implies that the
number of peaks depends on the sampling interval( point e is a peak at sampling
interval A, but not at A*), and also on the quantisation level (Whitehouse 1978).
Nevertheless we shall use this definition of a peak from here on unless noted to the
contrary.

Figure 7.6. Peak definition: point e is a peak at sampling interval A1 but not at A1 (adapted from Dagnall
1980)
138 Rough Surfaces

7.2. Average Parameters

A parameter which represented the average properties of a profile would clearly be


an improvement, and the extraction of such parameters became practical as soon as
electrical instruments became available. The Americans passed the electrical
signal through an AC voltmeter which responded to the root mean square average
of the signal, so they defined the RMS roughness:

where L is the evaluation length (in practice usually several cutoff lengths for a
filtered profile). The Europeans preferred to pass the AC signal through a rectifier
so that it could be used to charge up a capacitor, so they defined the centre-line
average (CLA) roughness:

The total area of the material-filled profile above the mean line should be
equal to the total area of voids below the mean line (Fig. 7.7). The CLA roughness
was known for a whle in the USA as the arithmetic average (AA) roughness, but
finally, as Rq ceased to be used, both terms dropped out of use and Ra is now
known simply as the average roughness.

R .-

Figure 7.7. Derivation ofthe average roughness Ra (Mummery 1990)

The victory of Ra is rather a pity from a scientific point of view, as it has no


theoretical application, whereas Rq, as we shall see below, has quite a fundamental
Amplitude Parameters 139

mathematical significance. It is easy to show that R a B q = 23f2n = 0.9 for the


special case of a sinusoid (and RdRq = (2/7r)'/* = 0.8 for a Gaussian height
distribution, see below). In practice, for many surfaces it is possible to ignore the
difference and use Ra and Rq interchangeably, as the inherent statistical variation
in all roughness measurements, to which we referred earlier, is of the same order.
This is not necessarily true for surfaces with more valleys than peaks (or vice versa,
whch is much rarer), as the squared term in Rq makes it more sensitive to outliers,
and the ratio Ra'Rq can then differ significantly from unity (King & Spedding
1982). Unfortunately Ra does not discriminate very well between profiles of quite
different characters (Fig. 7.8).

Figure 7.8. Three profiles with similar Ra values (Mummery 1990)

7.3. The Height Distribution

Unlike the extreme-value parameters, the average parameters contain no


information about the spatial or textural variation of the profile, that is the
variation in height from point to point of the surface. Logically, therefore, one
could go a step further and discuss the variation of surface heights in terms of a
statistical distribution, completely free of textural considerations. Theoretical
representation of surface relief in statistical terms was well-established by the end
of the Fifties (e.g. Longuet-Higgins 1957, Iwaki & Mori 1958, Bennett & Porteus
1961). The drudgery of mechanical calculation required by a statistical approach
postponed its more general application until the advent of digital techniques, when
140 Rough Surfaces

a number of workers took the method up independently at about the same time (e.g.
Kubo 1965, Whitehouse & Reason 1965, Greenwood & Williamson 1966).
Instead of a height z(x) at a point x, we now consider the probability density
p(z) of a distribution of heights (Fig. 7.9). The probability of a height lying in the
between z and z+dz is p(z) dz,and the cumulative probability that a height will be
below some level h is just

P(h) = J -m
h p(z)dz

p(zj 0 50% loo%

Figure 7.9. Profile height distribution p(z) and cumulative height distribution P(z)

The distribution p(z), whose mean we assume is at the height of the mean line
of the profile, may be characterised by its central moments

The second moment ,uz is the variance and describes the excursions of the
distribution from its mean, in our terms the roughness. Because the variance is in
units of (height) squared it is customary to use dp2 instead; this is the standard
deviation of the distribution, usually denoted by o, and is formally identical to the
RMS roughness Rq.
Many surfaces have more or less symmetrical height distributions, and many
workers have found it convenient to assume that such surfaces can be represented
by the well-known Gaussian distribution (Fig. 7.10):
Amplitude Parameters 141

p (z)dz=P(z<h<z+dz)

Hatched area =P(hS

z
I

Figure 7.10. Gaussian probability density and cumulative probability functions, with abscissa normalised by
Rq

If this is so, then it gives us access to the very large existing body of statistical
results using the Gaussian distribution, whose properties are extensively tabulated.
Certainly many surface height distributions look very closely Gaussian (Fig. 7.1l),
and Williamson et a1.(1969) have argued that a process of surface formation
consisting of a large number of random independent events, such as shot-blasting,
will produce a Gaussian height distribution by the action of the central limit
theorem. But are matters quite so straightfonvard?
142 Rough Surfaces

Surface height ($)

Figure 7.1 1. Distribution of 403,200 heights c om 5 nun x 7.5 mm of a ground surface, Rq = 2.4 pm,
compared to a Gaussian distribution (Sayles & Thomas 1979)

The usual way of establishing whether data fit a particular distribution is to


2
apply a goodness-of-fit test, such as the test, which calculates the probability that
a particular sample comes from a given parent population. The larger the sample,
the smaller the differences between sample and population must be to satisfy the
2
criterion. When a test is applied to the very large sample of Fig. 7.11, it fails;
the sample is not from a Gaussian population. Indeed it has been shown that the
technically important process of grinding should not produce a Gaussian height
distribution (Sayles & Thomas 1976). There is a further complication: some
authors confuse distributions of peaks on a profile, or summits on a surface, with
distributions of all surface heights, and assume that all these distributions are
Gaussian. In fact they are quite separate distributions, and it has been shown
(Wlutehouse & Archard 1970, Nayak 1971) that if the distribution of all surface
heights is Gaussian, then the distributions of peaks and summits must be non-
Gaussian.
Should we then abandon the convenience of the Gaussian assumption?
Probably not. Goodness-of-fit tests should be made on statistically independent
observations, and surface heights, as we shall see in a later section, are internally
correlated and hence not independent in the sense required. The effective sample
size must therefore be much less, and the criterion for agreement correspondingly
less rigorous, though it is difficult to quantlfy this. In any case the Gaussian
assumption is rather robust, and even for undeniably non-Gaussian surfaces it
yields statistical predictions which are in good agreement with experiment (Sayles
Amplitude Parameters 143

& Thomas 1979). We will continue therefore with due caution to assume Gaussian
properties where it is mathematically convenient. We should bear in mind that this
assumption is most questionable in the tails of a dlstribution, that is for the highest
peaks and lowest valleys. Unfortunately the highest peaks are the regions most
critical in such applications as contact mechanics. Attempts have been made to fit
other distributions to surface height data, including beta ( e g Tzeng & Saibel
1967, Dakshina Murthy & Raghavan 1972, Whitehouse 1978, Spedding et al.
1980) and log-normal, gamma and Rayleigh distributions (Izmailov & Kuriya
1983), but they do not appear to have attracted much subsequent imitation.

Figure 7.12. (a) Skewness and (b) kurtosis (Mummery 1990)

The third central moment p3 is the skewness, usually normalised as Sk = p3 /


Rq3. Because it is an odd moment, it is a sensitive measure of the degree of
asymmetry of the distribution, as its name suggests (Fig. 7.12a). Most machined
surfaces tend to be at least slightly negatively skewed, because peaks are more
easily removed than valleys. The fourth moment is the kurtosis, normalised as
K = / Rq4, and describes the overall shape of the distribution (Fig. 7.12b). A
Gaussian distribution has a kurtosis of 3. Distributions which have a sharper
central peak and longer tails are called leptokurtic. Distributions which are flatter
are platykurtic. Although many software packages for roughness analysis offer
kurtosis as well as skewness, not much practical use has been found for it. If
144 Rough Surfaces

numerical values are available for skewness and kurtosis, the ratio Ra/Rq can be
calculated (King & Spedding 1982).

7.4. Bearing Area

In 1933 Abbott & Firestone defined the bearing area fraction at a given height
above the mean line as the proportional length of all the plateaux which would
result if the surface were abraded away down to a level plane at that height (Fig.
7.13). The sum of the lengths of individual plateaux at a particular height,
normalised by the total assessment length, is the bearing ratio tp (DIN 4762 Part 1,
1960). If the bearing area fraction defined in this way is measured over a range of
heights and plotted against height, the result is the bearing area or material ratio
cuwe. Values of tp are sometimes specified on drawings, but this can lead to large
uncertainties if the bearing area curve is referred to the highest and lowest points
on the profile.

Figure 7.13. Derivation ofthe material ratio curve and the parameter tp (Mummery 1990)

We can now see that the bearing area at height h is simply the complement of
the cumulative probability bstribution function:

t* = [;p(z)dz = 1-
Amplitude Parameters 145

For a Gaussian height distribution the function P(h) is tabulated in terms of Rq,
hence if Rq is known the numerical value of tp at any height may be found
immediately by inspection.
The bearing area is a useful tool in characterising a large group of surfaces of
some practical importance. Many technical surfaces, particularly those employed
in machine components involving tribological interactions, are not produced in a
single operation but in a sequence of machining operations. Typically the initial
operation establishes the general shape of the surface with a more or less coarse
finish, and subsequent operations refine this finish to give the final properties
required by the design. Such a sequence of operations may remove the peaks of the
original process and superimpose a finer texture on the resulting plateaux, but
leaving the deep valleys of the initial process untouched. Such processes were
termed interrupted $finishes (Martz 1949, Williamson et al. 1969) and more
recently multiprocess (Whitehouse 1985) or stratijed (IS0 13565, 1996) surfaces.
Characteristically their height distributions are negatively skewed, making it
difficult for a single average parameter such as Ra to represent them effectively for
speclfication or quality control purposes.

Figure 7.14. Derivation of (a) the core roughness Rk and (b) the peak height Rpk and valley dzpth Rvk
(Mummery 1990)
146 Rough &$aces

The bearing area curve should contain all the information needed to
categorise these surfaces. One way of extracting this information has been
proposed by Schneider et al. (1988) (Fig. 7.14). A straight template covering 40%
of the total is offered to the central and flattest portion of the bearing area curve
and moved until its slope is a minimum. This straight line is projected through the
axes, and the height which separates the two intercepts is defined as the core
roughness Rk (Fig. 7.14a). The area above the intercept is now considered, and a
right-angled triangle of the same area is constructed. The height of this triangle is
the peak height Rpk. A similar construction at the other intercept finds the valley
depth Rvk (Fig. 7.14b). This technique has been incorporated into a German
standard (DIN 4776, 1990) and is now an international standard (IS0 13565,
1996).
This technique certainly yields numerical values of the various parameters
which discriminate successfully between two different surfaces with the same Fb
(Fig. 7.15). It has been criticized by Zipin (1990) on the grounds that the upper
and lower curved portions of the bearing area have nothing to do with peaks and
valleys or stratified textures but are simply mathematical properties of the Gaussian
distribution. A further and more serious criticism might be that the construction is
quite arbitrary and lacks any theoretical basis.

Figure 7.15. Comparison ofRk, Rpk, Rvk for different profiles with the same Ra value (Mummery 1990)

Zipin himself proposes a more soundly based procedure originally due to


Williamson (1967/68), who drew attention to &he possible application for
roughness characterisation of the cumulative probability scale, the vertical axis of
which is deliberately distorted so that a Gaussian distribution plots as a straight
line (Fig. 7.16b). The slope of the line is the standard deviation of the distribution.
Amplitude Parameters 147

or in our terms Rq. If the height distribution of a multi-process surface is plotted


on such a scale, it appears as two straight lines of different slopes, intersecting at a
height which corresponds to the depth to which the final finishing process has
superimposed itself (Fig. 7.16a, c). The two slopes are proportional to the two Rq
values which may thus be determined immediately (Malburg & Raja 1993). If the
straight lines are not very well-conditioned, their slopes and point of intersection
may be found by fitting a hyperbola to the probability plot (ISODIS 13565-3,
1995). In the fitting process it should be borne in mind that not all points are of
equal weight; the points in the central region of the distribution represent many
times more data values than the points in the tails.

99 i v

"i
v
v
v

-K" w

A
A
w

.
A
0 ?
A 0 w

I'

Figure 7.16. Cumulative distributions of heights from three surfaces produced by two-stage processes,
plotted on a scale distorted so that a Gaussian distribution appears as a straight line (Williamson et al. 1969):
the second process makes the upper region of the surface (a)rougher (b) ihe same (c) smoother

7.5. References

Abbott, E. J. and Firestone, F. A,, "Specifying surface quality", Mech.


Engng., 55, 569-572 (1933).
Bennett, H. E. and Porteus, J. O., "Relation between surface roughness and
specular reflectance at normal incidence", J. Opt. Soc. Am., 51, 123-129 (1961).
148 Rough Sudaces

Dagnall, H., Exploring surface texture (Rank Taylor Hobson, Leicester,


1980).
Dakshina Murthy, H. B. and Raghavan, M. R., "Compliance of rough
cylinders in compression", Wear, 20, 353-369 (1972).
DIN 4762 Part 1, Assessment of 2nd to -fth order irregularities of surface
conjguration by means of sections of surfaces: definitions relating to reference
systems and dimensions (Deutsches Institut fur Normung, Berlin, 1960)
DIN 4762 Part 2, Assessment of 2nd to Jib order irregularities of surface
configuration by means of sections of surfaces: evaluation of profile sections
based on geometrically ideal reference projile (Deutsches Institut f i r Normung,
Berlin, 1960)
DIN 4762, Surface roughness - terms and dejnitions (Deutsches Institut fiir
Normung, Berlin, 1989)
DIN 4771, Messung von der Profiltiefe Pt von Oberfachen (Deutsches
Institut f i r Normung, Berlin, 1977)
DIN 4776, Measurement of surface roughness: parameters Rk, Rpk, Rvk,
Mrl, Mr2 for describing the material portion (Projle bearing length ratio) in the
roughness profile: measuring conditions and evaluation procedures (Deutsches
Institut fir Normung, Berlin, 1990)
Greenwood, J. A. and Williamson, J. B. P., "Contact of nominally flat
surfaces", Proc. Royal SOC.A295, 300-319 (1966).
Hydell, R. R., "Today's need - functional surface roughness control", Proc. I.
Mech. E., 182, Part 3K, 127-134 (1967/68).
I S 0 13565, Geometric product specijkations - surface texture: profile
method - surfaces having stratij2ed functional properties (ISO, Geneva, 1996)
IS0 4287, Surface roughness - terminologv - part 1: surface and its
parameters (ISO, Geneva, 1984)
ISO/DIS 13565-3, Surface texture (projle method) - characterisation of
surfaces having stratified functional properties - part 3: height characterisation
using the material probability curve of surfaces consisting of two vertical random
components (ISO, Geneva, 1995)
Iwaki, A. and Mori, M., "On the distrihution of surface roughness when two
surfaces are pressed together", Bull. J.S.M.E., 1, 329-337 (1958).
Izmailov, V. V. and M. S. Kuriva, "Use of the Beta-distribution for
calculating the contact characteristic of rough-bodies", Trenie i Iznos, 4, 983-990,
(1983)
King, T. C. and T. A. Spedding, "On the relationships between surface
profile height parameters", Wear 83, 91-108 (1982)
Amplitude Parameters 149

Kubo, M., "Instrument for the measurement of slope and height distribution
of surface roughness", Rev. Sci. Instrum., 36, 236-237 (1965).
Longuet-Higgins, M. S., "The statistical analysis of a random, moving
surface", Phil. Trans. Royal SOC.,A249,321-387 (1957).
Malburg, M. C., and Raja, J., "Characterisation of surface texture generated
by plateau honing process", Ann. CIRP 42, 637-639 (1993).
Martz, L. S., "Preliminary report of developments in interrupted surface
finishes", Proc. I . Mech. E ., 161, 1-9 (1949).
Mummery, L., Surface texture analysis: the handbook (Hommelwerke,
Muhlhausen, 1990).
Nayak, P. R., Tandom process model of rough surfaces", Trans. A.S.M.E.
Ser. F. J. Lubr. Tech., 93, 398-407 (1971).
Sander, M., A practical guide to the assessment of surface texture (Feinpruf
Perthen, Gottingen, 1991).
Sayles, R. S., Thomas, T. R., "Stochastic explanation of some structural
properties of a ground surface", Int J Prod Res 14,641-655 (1976).
Sayles, R. S.; Thomas, T. R., "Measurements of the statistical microgeometry
of engineering surfaces", J Lubr Technol Trans ASME 101,409-417 (1979).
Schneider, U., Steckroth, A,, Fbu, N. and Hiibner, G., "An approach to the
evaluation of surface profiles by separating them into functionally different parts",
Surface Topography, 1, 343-355 (1988)
SMS Rapport 47, Ytjamnhet: nya parametrar, angivning pa ritnning och
projildjupet H (Sveriges Mekanstandardiering, Stockholm, 1994)
Spedding, T. A.; King, T. G.; Watson, W.; Stout, K. J., "Pearson system of
distributions: Its application to non-gaussian surface metrology and a simple wear
model", J Lubr Technol Trans ASME 102,495-500 (1980).
Thomas, T. R.; Charlton, G., "Variation of roughness parameters on some
typical manufactured surfaces", Precis. Eng. 3 , 9 1-96 (198 1).
Tzeng, S. T. and Saibel, E., "Surface roughness effect on slider bearing
lubrication", A.S.L.E. Trans., 10, 334-338 (1967)
Whitehouse, D. J. and Archard, J. F., "The properties of random surfaces of
significance in their contact", Proc. R. SOC.Lond., A316, 97-121 (1970).
Whitehouse, D. J. and Reason, R. E., The equation of the mean line of
surface texture found by an electric wave j l t e r (Rank Precision Industries,
Leicester, 1965).
Whitehouse, D. J., "Beta Functions for surface typologie?", Ann. CIRP, 27,
491-497 (1978)
150 Rough Su$aces

Whitehouse, D. J., "The digital measurement of peak parameters on surface


profiles", J. Mech. Eng. Sci.,20, 221-227 (1978)
Whitehouse, D. J., "The parameter rash - is there a cure?", Wear 83, 75-78
(1982)
Whitehouse, D. J., "Assessment of surface finish profiles produced by multi-
process manufacture", Proc. Instn. Mech. Engrs., 199, B4, 203-270 (1985)
Williamson, J. B. P., "The microtopography of surfaces", Proc. I. Mech. E.,
182, Part 3K, 21-30 (1967/68).
Williamson, J. B. P., Pullen, J. and Hunt, R. T., "The shape of solid
surfaces", in: Surface mechanics. Ling, F.F. (ed.). 24-35 (A.S.M.E., New York,
1969).
Zipin, R. B., "Analysis of the Rk surface roughness parameter proposals",
Prec. Eng., 12, 106-108 (1990)
CHAPTER 8

TEXTURE PARAMETERS

From the preceding chapter it is plain that a description of a surface restricted to its
variations in amplitude will be incomplete for practical purposes. If stranded on a
mountain top at nightfall, it is useful to know that the average height of the
surrounding terrain is 1000 metres, but it is more importan: to know how quickly
the height changes with position. In terms of roughness, it is necessary to find
some convenient means of describing the variation of relief in the plane of the
surface. How, for instance, are we to distinguish between the two profiles of Fig.
8.1, which are visibly different but for which all the amplitude parameters are the
same?

T T

Figure 8.1. Two surfaces with the same amplitude parameters but different ''textures", and their respective
autocorrelationfunctions

Historically various measures were proposed which could be measured readily


from the output of a stylus instrument. The high-spot count HSC is the number of
peaks per unit length of a profile, and its reciprocal, the high-spot spacing or mean
spacing of profile irregularities Sm, is the mean distance between peaks (Sander
1991). This immediately returns us to the problem of peak definition; too large a
peak will result in a parameter too sensitive to waviness, too small a peak will

151
152 Rough Sur$aces

result in a parameter swamped by the fine detail. In ISODIS 4287 (1984) a peak
is defined, for this purpose only, as the distance between two mean line crossings
(Fig. 8.2).

Mean line

s, = + % + Snu + S, mm
I

Figure 8.2. Mean spacing ofprofile irregularities (Sander 1991)

An alternative is the profile length ratio, fr (DIN 4762, 1989), also known as
the profile roughness parameter RL (Gokhale & Drury 1993), defined as the ratio of
the developed length of the profile to its nominal length. This is rather hard to
measure and if profile slopes are small it is rather insensitive to changes in slope.
It is also instrument-dependent; the finer the detail revealed by the sensor, the
longer the real length.
Sensor dependence is also a problem in measuring profile slopes, and for the
same reason. Using the mean slope is no use, as its value will tend to zero unless
the surface has teeth pointing one way; the mean of the moduli of the slopes da, or
alternatively the FWS slope, must be used instead (IS0 DIS 4287, 1984), but the
sampling interval, on which the slope depends, is not defined in standards. Spragg
& Whitehouse (1974) proposed a parameter 2&a/Aa which they called the
average wavelength h.By analogy there is a corresponding RMS wavelength Aq
(IS0 DIS 4287, 1984), but neither of these parameters seems to have won wide
acceptance.

8.1. Random Processes

In 1945 Womersley & Hopkins suggested that differences in surface texture might
be more systematically described by a correlogram, that is by investigating the
correlation between pairs of points on a profile as the separation of the pairs is
vaned. The technology to implement their suggestion did not then exist, but by the
late sixties Peklenik (1967, 1967/8) was computing various random-process
Texture Parameters 153

functions for machined surfaces. These are functions developed to describe time
series and used in communications engineering for signal processing, principally
the autocovariance function (ACVF) and its Fourier transform, the power spectral
density function (PSDF). For a profile of length L the ACVF is

L-S
1
R(Z) = ~

L-Z
Iz(x)z(x + 7)dx
0

where z(x) and z(x+Q are pairs of heights separated by a delay 7 (Fig. 8.3). The
ACVF has dimensions of height squared.

Figure 8.3. Construction ofthe autocovariance function

This is often normalised as the autocorrelation function (ACF)

where R(0) is the variance Rqz of the height distribution. The ACF is
dimensionless with an initial value of unity. The ACVF and ACF are very
sensitive to periodic components of the surface and will detect these even when
they are obscured by random components (Fig. 8.4). Lf the profile is entirely
random, the autocorrelation function will decay asymptotically to zero at a rate
which depends on the open-ness of the texture, and will thus distinguish between
the two profiles of Fig. 8.1.
Like the bearing area in the last chapter, the ACF is a function, not a number,
and needs to be characterised numerically in some way if it is to be of practical use.
The correlation length is defined as the length over which it decays to some
fraction of its initial value, sometimes taken as a tenth, sometimes as lle. Pairs of
points separated by distances greater than the correlation length are statistically
independent. Note that random and Gaussian are not synonyms, though often used
154 Rough Sur$aces

as if they were. Random simply means uncorrelated; a table of random numbers


would be expected to have a rectangular distribution (otherwise some lottery
numbers would be luckier than others!), and it is perfectly possible to imagine a
periodic profile with a Gaussian height distribution.

0.05

'0 2 4 6 8 10 pm

Figure 8.4. (a) Profile, @) height distribution, (c) ACF, (d) PSDF for a profile of a fine turned surface
(Peklenk 1967/68). Note the periodicity due to the turning feed detected by the ACF.

The corresponding function in the frequency domain is the power spectrum


or power spectral densify function (PSDF). The Wiener-Khinchine relation
defines the PSDF (Fig. 8.4) as the Fourier transform of the ACVF (Bendat &
Piersol 1966):

(8.3)
Texture Parameters 155

where w = 2x / /z is an angular frequency with dimensions of reciprocal length.


The PSDF is a probability density function like the height distribution, and its
dimensions are frequently a source of confusion. A single value of the PSDF has
dimensions of height squared per unit frequency, and like the height distribution it
has moments m,:

m, = ia,
a2
w"G(w)do

The zeroth moment mo is the variance of the height Istribution. Thus if the
power spectrum is integrated over a pass-band of surface frequencies between q
and m,the area under the function is the square of the RMS roughness, Rg2. The
second and fourth moments are the variances of the distributions of slopes and
curvatures respectively.

Sayles & Thomas (1977) proposed the use of the structure fimction

L-7
1
L-2
1
S( 2) = - {z(x) - z(x + 2)12rn
0
(8.5)

as an alternative to correlation for surface roughness investigations. The structure


function is related to the autocorrelation function by

It contains the same information as the ACF and the PSD, biit offers some practical
advantages: it is stable and easy to compute, it does not impose a periodogram
model on the surface, and it does not require prior hgh-pass filtering. Another
useful property is that for a random stationary profile it tends asymptotically to a
value of 2Rq2 as 2 + 00. The structure function can show functional changes more
clearly than the ACF (Fig. 8.5).
Smith & Walmsley (1979, Yolles et al. 1982) have proposed the use of Walsh
functions as an alternative to Fourier analysis for surface profiles. Based on square
waves, Walsh functions are binary hence lend themselves to fast computation.
Whitehouse (1994) suggests that they are particularly suitable for representing
surfaces with sharp discontinuities,but points out that they require many terms to
represent periodic surfaces. Another alternative is the use of autoregressive
156 Rough Sur$aces

moving-average ( M A ) techniques, which treat the surface as the output of a


system (the manufacturing process) which modifies white noise (DeVor & Wu
1971). ARMA models have been used to describe surfaces prduced by abrasive
machining (Pandit et a1 1976), but Whitehouse again points out that they do not
represent periodic processes well, and accurate models make heavy computational
demands (Kovacevic & Zhang 1992).

--
c 1.0
9

0.0

Figure 8.5. (a) Autocorrelation functions and (b) structure functions ofthe same pair ofworn and unworn
profiles
Texture Parameters 157

8.2. The Profile as a Random Process

Whitehouse and Archard (1970) investigated a model of a random profile with a


Gaussian distribution of heights and an exponentially decaying autocorrelation
function,

where p* is the correlation length. Defining a peak as a point higher than its two
nearest neighbours, they obtained results for a number of statistical properties of
the profile’s geometly. These are worth quoting at length, as they are among the
more important results in the literature:

Ratio of the number of peaks to the number of all heights:

Ratio of mean peak height to Rq:

Ratio of variance of peak heights to Rq2 (adapted from Whitehouse 1978: the
equation in Whitehouse & Archard 1970 appears to be misprinted)

Mean peak curvature x 2 I Rq:

Variance of profile curvature (i.e. second differential of profile) x 2 / Rg:


6 - 8p + 2p2
158 Rough Sut$aces

Mean absolute slope:

It is instructive to present these results in terms of the delay t-, which for this
purpose we may treat as the sampling interval. As the sampling interval decreases,
the number of peaks decreases from a third to a quarter of the total number of
heights, the mean peak height asymptotically approaches the profile mean line, and
the peak standard deviation tends to Rq (Fig. 8.6a). The variation of slopes and
curvatures is rather more dramatic (Fig. 8.6b); as the sampling interval gets
smaller, the slope increases as its square root, while the pe& radius of curvature
decreases as the 3/2 power. From these figures it is clear that the numerical result
obtained for any texture parameter depends on the sampling interval at which the
data is measured. The shorter the interval, the steeper the slopes and the more and
sharper the peaks. It is hard to overemphasize the importance of this result, which
implies that none of these texture parameters is an intrinsic property of the profile.
It follows that any measurement of texture parameters must quote the sampling
interval in order to be meaningful, and that it is impossible to compare results from
Merent sets of measurements unless the respective sampling intervals are known.

0.8

0.4 I /
0.2

0.01 0.10 1.oo 10.00


Sampling interval I correlation length

Figure 8.6a. The Whitehouse-Archardrelations: Ratio of peaks to all heights, and mean peak height and
standard deviation normalised by Rq, as a function of normalised sampling interval
Texture Parameters 159

I . pm

Figure 8.6b. The Whitehouse-Archard relations: variation of mean absolute slope and peak radius of
curvature with sampling interval, for a profile of roughness 1 pm and correlation length 50 pm (Thomas &
Sayles 1978)

8.3. Practical Computation

The practical problems of measuring distributions of slopes and curvatures are now
seen to be considerable. The steepest slope and sharpest curvature which can be
measured depend on the construction of the sensor; if a contact probe has an
included angle of 90 degrees then it will record any slope steeper than 45 degrees
as a 45 degree slope, unless an error trap is built into the acquisition software. The
numerical values obtained will of course depend on the sampling interval as
discussed above, and also on the numerical formula used for their calculation.
Seven-point, five-point and three-point central difference formulae will all give
different answers. The more points, the lower the uncertainty, but the higher the
overhead of computation time; increasing the number of points used to calculate a
slope or curvature amounts to low-pass filtering (Thwaite 1978). The effecl of
different numerical techniques is discussed by Whitehouse (1978). Direct
measurement of the mean peak radius of curvature (MPRC) presents a special
problem, in that a plateau will return an infinitely large radius of curvature, usually
160 Rough &$aces

resulting in the computer refusing to divide by zero. The solution is to measure its
reciprocal, the curvature, remembering of course that the numerical value of the
mean reciprocal peak curvature is not in general the same as that of the MPRC.
Direct measurement of power spectra from Eqn. 8.3 is time-consuming, and
instead the fast Fourier transform (FFT) is universally employed (e.g. Wade 1994).
This makes best use of the computer’s facility for modulo 2 arithmetic to speed up
the computation of a spectrum of n data points by a factor n / log2n, = 100 for n =
1024. It has the consequent disadvantage that n must be a power of 2, so if, say,
2000 data points have been laboriously collected the FFT can only use 1024 of
them. This is one reason why so many modern instruments are designed to acquire
an integral power of 2 data points.

Figure 8.7. Ensemble averaging of power spectral estimates from profiles on a ground glass surface improves
their stability (Thwaite 1978)

Computing power spectra has other problems. As the spectral wavelength


approaches the length of the record, fewer and fewer data points are available for
the estimate of power. Bendat (1958) has shown that this results in a systematic
underestimate of power at long wavelengths, resulting in a spurious peak in the
spectrum. When such long-wavelength peaks are seen in the literature they should
always be suspected unless some special physical argument can be advanced to
explain them. Bendat also showed that random errors in the spectral estimate
increase as the record length decreases; for record lengths typical of roughness
Texture Parameters 161

measurement, these errors may be of the order of the estimates themselves. Raw
estimates of spectral power may be "smoothed" by convolution with various
windows (see for example Bendat & Piersol 1966), but this requires a prior
assumption that no spectral peaks exist which the smoothing will obscure. Thwaite
(1978) recommends the technique of ensemble averaging over a number of profiles
(Fig. 8.7).
The autocorrelation function can be computed directly, but in practice is
usually computed by an inverse Fourier transform from the PSDF. Again delays of
the order of the record length should be avoided, as the estimates of correlation will
be based on very few data points and the function may go unstable. The correlation
length turns out, disappointingly, to be rather sensitive to high-pass filtering (Fig.
8.8), and so in spite of its fundamental theoretical significance has not been used
much for practical characterisation.

Figure 8.8. Autocorrelation functions of a profile of a shotblasted surface after high-pass filtering at cutoff
wavelengths of (a) 40 mm, (b) 20 mm, (c) 2 nun, (d) 0.6 mm (Thomas & Sayles 1975)

So far we have drawn attention to the difficulties of obtaining unambiguous


numerical values for texture parameters, and have tacitly assumed that amplitude
parameters are well-defined by comparison. Unfortunately amplitude parameters
too have their problems of definition. An assumption usually made in dealing with
time series is that they are stationary, that is that there exists some finite length of
record beyond which further measurement will reveal no new information. This
appears not to be the case for a wide range of natural and man-made surfaces
(Sayles & Thomas 1978), for the spectra of which power appears to increase
162 Rough Surfaces

indefinitely as approximately the square of the wavelength (Fig. 8.9). But, as we


have already seen, the variance of the height distribution, and hence Rq, is a
moment of the power spectrum. If the PSDF increases indefinitely, so must Rq and
other amplitude parameters. Thus we have arrived at a situation where many, if
not most, real surfaces, including the majority of technical surfaces, appear to have
no intrinsic geometrical properties whatever, and any numerical values we obtain
will depend entirely on the scale of measurement. This is a uniquely unsatisfactory
state of affairs in metrology; indeed elsewhere in engineering only fluid dynamics
suffers a similar indignity. Is there no way out of this impasse?

8.4. Fractal Roughness

The spectra of the surfaces of Fig. 8.9 can all be represented by a relation of the
type

where p i s a dimensionless constant, and B is a constant with dimensions of length


which Sayles & Thomas (1978) called the topothesy. These constants appear to be
intrinsic properties of the surface; if they define the power spectrum, it should be
possible to express most of the parameters which we have discussed in the last two
chapters in terms of them. Mandelbrot (1978, private communication) pointed out
that such surfaces are examples of fractals. Berry (1979) redefined the topothesy
as the horizontal separation of pairs of points on a surface corresponding to an
average slope of one radian.
Fractals are functions which are continuous but not differentiable
(Mandelbrot 1983). They possess the property of self-szmzfarity,that is they appear
the same at any scale of magnlfication (Fig. 8.10). Self-similar fractals can be
completely characterised by a single parameter, the fractal dimension D. Examples
of self-similar fractals in nature include fracture surfaces (Baran et al. 1992) and
natural terrain (Snow & Mayer 1992). However, man-made surfaces in general are
subject to the further restriction that they appear to measuring instruments to be
single-valued. T h s implies that smaller features must always have steeper slopes
than larger features, as Whltehouse and Archard (1970) concluded on quite
different grounds. Thus when the scale of observation is changed, a scaling factor
with dimensions of length must be introduced to restore the appearance of self-
Texture Parameters 163

Figure 8.9. Variation of power spectral density with wavelength for 23 natural and man-made surfaces
(Sayles & Thomas 1978). Solid line is best fit of slope 2.
164 Rough Su$aces

similarity. This scaling factor turns out to be the topothesy, and single-valued
fractals of this kind are described as self-afine.

Figure 8.10. Self-similarityof surface profiles (Thomas & Thomas 1986)

Most methods of calculating the fractal dimension were developed for self-
similar fractals, like coastlines, cracks in rocks and particles of powder (Russ
1994). They do not work very well for self-sine surfaces with gentle slopes. An
effective way to calculate the fractal parameters of a profile is to compute the
structure function. It can be shown that for a fractal profile (Russ 1994):

In other words, the structure function of a fractal profile obeys a power law, so it
plots as a straight line on a log-log scale (Fig. 8.11). This is an easy way of
establishing fractal behaviour, and from the slope and intercept of this straight line
both the fractal dimension D and the topothesy A can easily be calculated. This is
the fractal equivalent of Eqn. 8.8, and the respective slopes and intercepts are
related by (Russ 1994):

p= 2 0 - 4 (8.10)

B = ( 2 ~ 2) r (~- p ) cos ( - p ~ / 2A3ffl


) (8.11)

where T i s a Gamma function.


Texture Parameters 165

1
log T

Figure 8.11.Fractal parameters and the structure function

Fractal descriptions of engineering surfaces have been given by a number of


workers (Wehbi et al. 1992, Stupak et al. 1991, Vandenberg & Osborne 1992,
Yordanov & Ivanova 1995, Brown et al. 1997). However, some of this work
should be treated with caution, as not all authors have distinguished between self-
similar and self-fine fractals, and different methods of calculation can yield
significantly M e r e n t numerical values of the fractal parameters (Russ 1994).
Many machined surfaces seem to be fractal at least at shorter wavelengths (Fig.
8.12). Fractal dimensions may vary between the theoretical limits of 1 (a straight
line) and 2 (a space-filling curve), whle topothesies may be represented by very
short lengths (Table S.l), the physical significance of which is not immediately
obvious (Russ 1994).

Figure 8.12. Structure functions of (a) spark-eroded (b) ground surfaces showing fiactal behaviour at short
wavelengths, measured with standard (open circles) and fine (filled circles) stylus instruments. Circles are
ensemble averages, error bars are standard deviations (Thomas& Thomas 1986).
166 Rough Surfaces

1-

bl-

-t ++' 0 0 1-

Figure 8.12 (continued). Structure functions of (c) bead-blasted (d) turned surfaces showing fractal
behaviour at short wavelengths, measuredwith standard (open circles) and fme (filled circles) stylus
instruments. Circles are ensemble averages, error bars are standard deviations (Thomas & Thomas 1986).

Table 8.1. Fractal parameters for some machined surfaces (Thomas & Thomas 1986)

Machining process D A (nm)

Grinding 1.17 0.338


Turning 1.18 0.274
Bead-blasting 1.14 0.427
;
In practice no real surface can be fractal over an infinite range of
wavelengths, because no natural or man-made process can operate over an infinite
range of wavelengths. A real surface will be formed by several different processes
each with its characteristic features. For instance, a mountain landscape may be
formed by erosion on a scale from kilometres down to meters. Below this scale the
landscape may be covered with vegetation, which may also be fractal down to a
scale of millimetres, but with completely different values of fractal properties.
Such surfaces are called multij?uctul (for an extensive discussion, see Russ 1994)
and typically will present a structure function as two or more straight lines of
different slope meeting at a more or less sharp discontinuity (Fig. 8.13). Such a
structure function is evidence of multifractal behaviour, and the wavelength
corresponding to the discontinuity will mark the transition between two different
Texture Parameters I67

mechanisms of surface formation. This transition point has been termed a corner
frequency (Majumdar & Tien 1990).

0) MID, unworn (7.4)

0,001 0,Ol I 10 100 lo00 10000

Figure 8.13. Structure functions ofregions ofthe same cylinder liner in worn and unworn conditions
showing multifi-actal behaviour with comer frequencies, measured with AFM and stylus instrument ( R o s h
et al. 1997).

Machined surfaces are likely to be multifractal because they are usually


produced by several processes. The approximate shape of the surface is produced
by casting or rough preliminary machining. This will produce the form and
waviness of the surface, possibly depending on the dynamics of the machine tool
(chatter, spindle runout and so on). One or more finishing processes will produce
the final roughness of the surface. One would expect corner frequencies to depend
on the dlmensions of the actual cutting element, which may produce surface
features far smaller than itself but cannot produce features any larger.
Fractal parameters may conveniently be extracted from bifractals by adapting
a method originally recommended for a similar problem, the extraction of Rq
values from bearing area curves of stratified profiles plotted as two intersecting
straight lines on a probability scale (IS0 DIS 13565-3, 1995). A pair of best
hyperbolae are fitted to the structure function by least squares (Fig. 8.14). The
intersection of the asymptotes gives the corner frequency, and the slopes, and hence
the fractal parameters, are found by differentiating the lower hyperbola at each end
of the structure function.
168 Rough Sur$aces

5.41 ' 1
5.61.

Figure 8.14. Extraction of fractal parameters from a bifractal structure function by hyperbola fitling ( h i n i
et al. 1998)

8.5. References

Baran, G.R.; Roques-Carmes, C.; Wehbi, D.; Degrange, M., "Fractal


characteristics of fracture surfaces",Journal of the American Ceramic Society 75,
2687-2691 (1992)
Bendat, J. S., Principles and applications of random noise theory (Wiley,
New York, 1958)
Bendat, J. S., Piersol, A. G., Measurement and analysis of random data
(Wiley, New York, 1966)
Berry, M. V., "Diffractals",J. Phys. A12, 781-797 (1979)
Brown, C., Johnsen, W. and Hult, K., "Scale-sensitivity, fractal analysis and
simulations",Trans. 7th. Int. Con$ On Metrology & Properties of Engng Surfaces,
pp. 239-243 (Goteborg, 1997)
DPJ 4762, Surface roughness - terms and dejnitions (Deutsches Institut fir
Normung, Berlin, 1989)
Texture Parameters 169

DeVor, R. E. and Wu, S. M., "Surface profile characterisation by


autoregressive-moving average models", Paper 7 1-WA/Prod-26 (ASME, New
York, 1971).
Gokhale, A,; Drury, W. J., "Surface roughness of anisotropic fracture
surfaces", Materials Characterization 30,279-286 (1993)
ISODIS 428711.2, Surface roughness terminology Part I : surface and its
~ ~

parameters (ISO, Geneva, 1984)


ISODIS 13565-3, Surface texture (projle method) ~characterisation of
surfaces having stratrfied functional properties - Part 3: Height characterisation
using the material probability curve of surfaces consisting of two vertical random
components (ISO, Geneva, 1995)
Kovacevic, R.; Zhang, Y. M., "Identification of surface characteristics from
large samples", Proc. I. Mech. E: Mech. Eng. Sci. 206C, 275-284 (1992)
Majumdar, A., and C. L. Tien, "Fractal characterisation and simulation of
rough surfaces", Wear 136, 313-327 (1990)
Mandelbrot, B. B., The fractal geometry ofnature 3e (Freeman, New York,
1983)
Panht, S. M., Suratkar, P. T. and Wu, S. M., "Mathematical model of a
ground surface profile with the grinding process as a feedback system", Wear, 39, .
205-217 (1976)
Peklenik, J., "Investigation of surface typology", Ann. C.I.R.P. 15, 381-385
(1967).
Peklenik, J., "New developments in surface characterization and
measurements by means of random process analysis", Proc. I. Mech. E., 182, Part
3K, 108-126 (1967/68).
Rosen, B.-G., R. Ohlsson and T. R. Thomas, "Nano metrology of cylinder
bore wear", Trans. 7th. Int. Conf: On Metrology & Properties of Engng Surfaces,
102-1 10 (Goteborg, 1997)
Russ, J. C., Fractal surfaces (Plenum Press, New York, 1994).
Sander, M., 4 practical guide to the assessment of surface texture (Feinpruf
Perthen, Gottingen, 1991).
Sayles, R. S. and T. R. Thomas, "The spatial representation of surface
roughness by means of the structure function: a practical alternative to correlation",
Wear 42,263-276 (1977).
Sayles, R. S. and T. R. Thomas, "Surface topography as a non-stationary
random process", Nature 271,43 1-434 (1978)
Smith, E. H. and W. M. Walmsley, "Walsh functions and their use in the
assessment of surface texture", Wear, 57, 157-166 (1979)
170 Rough Surfaces

Snow, R. S., and Mayer, L., "Introduction to special issue-Fractals in


geomorphology", Geomorphology 5, 1-4 (1992).
Spragg, R. C. and Whitehouse, D. J., "An average wavelength parameter for
surface metrology", Rev. M. Mec. 20, 293-300 (1974).
Stupak, P. R., C. Y. Syu and J. A. Donovan, "The effect of filtering
profilometer data on fractal parameters", Wear 154, 109-114 (1992)
Thomas, T. R. and Sayles, R. S., "Random-process analysis of the effect of
waviness on thermal contact resistance", Prog. Astronaut. Aeronaut. 29, 3-20
(1975)
Thomas, A. and Thomas, T. R., "Experimental measurement of surface
roughness on a sub-micron scale", Proc. Int. Con$ Modern Prod. & Prod.
Metrology, Tech. Univ., Vienna, 97.3-1 11.3 (1986)
Thomas, T. R. and Sayles, R. S., "Some problems in the tribology of rough
surfaces", Tribology International, 11, 163-168 (1978)
Thwaite, E. G., "The numerical interpretation of topography", Wear, 51, 253-
267 (1978).
Vandenberg, S., and C. F. Osborne, "Digital image processing techniques,
fractal dimensionality and scale-space applied to surface roughness", Wear 159,
17-30 (1992).
Wade, G., Signal coding and processing 2e (Cambridge University Press,
1994)
Wehbi, D.; Roques-Carmes, C.; Tricot, C., "Perturbation dimension for
describing rough surfaces", International Journal of Machine Tools &
Manufacture 32, 211-216 (1992).
Whitehouse, D. J., "The digital measurement of peak parameters on surface
profiles", J. Mech. Eng. Sci. 20,221-227 (1978)
Whitehouse, D. J., Handbook of surface metrology (Institute of Physics,
Bristol, 1994)
Whitehouse, D. J. and J. F. Archard, "The properties of random surfaces of
significance in their contact", Proc. R. Sac., A316, 97-121 (1970).
Womersley, J. R., & M. R. Hopkins, "Suggestions concerning the use of the
correlogram for the interpretation of measurements of surface finish", Journees des
etats de surface, 135-139 (1945)
Yolles, M. I., Smith, E. H., Walmsley, W. M., "Walsh theory and spectral
analysis of engineering surfaces", Wear 83, 151-164 (1982)
Yordanov, 0. I., and K. Ivanova, "Description of surface roughness as an
approximate self-sine random structure", Surface Science 331-333, 1043-1049
(1995).
CHAPTER 9

SURFACES IN THREE DIMENSIONS

The characterisation of surfaces in three dimensions of course increases the time


and effort of computation, and introduces some new topics such as anisotropy
which we have been able to disregard in the case of profiles. But much of the
analysis is a reasonably straightforward extension of the previous two-dimensional
discussion, and it will be convenient to deal with it in a similar order. First we
should consider any special problems of three-dimensional filtering, then we can
go on to discuss height and texture parameters as before, and introduce random
process theory and fractals in three dimensions. Practical computation will be
treated at some length as it poses some special problems in three dimensions.
Finally we need to look briefly at the problem of anisotropy which is unique to
three dimensions.
We begin by making an important distinction, following Williamson and
Hunt (1967/8). A local maximum on a profile will be called a peak, and a local
maximum on a surface will be called a summit. A profile will more often than not
pass over the shoulder of a surface feature rather than its summit. The shoulder
will, nevertheless, appear as a peak on the profile (Fig. 9.1). Thus, as we shall see
below, the dm-ibutions of peak and summit heights and curvatures may be quite
Merent.

Figure 9.1. Peaks on profilesare not in general summits on surfaces

171
172 Rough Surfaces

9.1. Filtering

The special problems of 3D filtering are discussed at length by Stout et al. (1993).
The first stage in processing raw data is to remove the DC level (often called the
piston term in the optical literature), the trend and any errors of form. In two
dimensions this is generally done by fitting a polynomial of appropriate order. In
3D it is more convenient to do this in stages. First the DC level and trend in x and
y are removed by fitting a least-squares mean plane

z * (x, y) = a + bx + cy (9.1)

and subtracting it from the data. From here on values of z will be assumed to be
referred to the mean plane unless stated to the contrary.
The coefficients a, b, c are found in the usual way by minimising the sum of
the squares of the differences between the mean plane and the data. The discrete
solutions are given in the section on computation. Form errors, if any, may then be
removed by fitting polynomials. Stout et al. give a general expression for a least
squares polynomial surface, but point out that in practice higher order polynomials
are rarely required. Often on inspection it suffkes to fit separate polynomials
serially in x andy (Fig. 9.2).

Figure 9.2. 3D filtering of a ground surface (x-direction is into the paper): (top left) raw data; (top right) 3'd
order polynomial in x fitted to remove form error, then (bottom left) aftex high-pass and (bottom right) low-
pass Gaussian filtering at 0.08 mm cutoff.
Surfaces in 3 0 173

Stout et al. identify general requirements for a 3D filter: it should be zero


phase so as to preserve the shape of surface features, and it should be separable, so
profiles in the x- and y-directions can be filtered separately, thus saving
computation time. They recommend the use of a Gaussian filter for general
purposes, and a modified zonal filter if sharp cutoffs are particularly needed, both
of which filters satisfy these requirements, and they recommend implementation by
FFT for h g h computational efficiency (Fig. 9.2).
Envelope filters also exist in three dimensions. The 3D analogue of a rolling
circle is a rolling ball. Algorithms for a rolling sphere were developed by
Shunmugam (Shunmugam & Radhakrishnan 1974a, b). Little difference was
found between the power spectra of rolling circle and rolling sphere on the same
profile (Shunmugam 1977). An extra difficulty presents itself in three dimensions.
If the rigorous geometric condition is imposed that the sphere is always supported
by the three highest points as it rolls in, say, the x-direction, then the locus of its
centre can no longer be constrained within the x-z plane but must wobble in the y-
direction. Also, the locus of the centre of the ball, which ir the reference surface,
will be unduly sensitive to high summits. For this reason ball filters have been
developed (Jonasson et al. 1997; Schmoeckel & Staeves 1997) where the virtual
ball is allowed to penetrate the surface to varying degrees. A 3D motif filter has
also been developed (Zahouani 1997), where the individual motif is a pit
surrounded by summits.

9.2. Parameters

Stout et al. (1993) have proposed a set of 14 parameters for characterising a surface
in 3D. As the so-called "Birmingham 14" are at present the only serious proposals
they are worth examining here in some detail. The set comprises 4 amplitude
parameters, 4 texture parameters, 3 hybrid parameters, and 3 so-called "functional"
parameters (Table 9.1).
They are presented below in continuous form, adapting the original discrete
forms where necessary; some of the discrete forms will be discussed later. The
amplitude parameters are RMS deviation Sq, ten-point height Sz, skewness Ssk and
kurtosis Sku, all defined by analogy with the two-dimensional forms, for instance

d:s
sq= - zZ(x,y)dxdy
A
(9.1)
174 Rough Sursaces

where the domain of integration is the area of measurement A . The ten-point


height is the difference in height between the mean of the 5 highest summits and
the 5 lowest valleys.

Table 9.1. The Birmingham 14 (Stout et al. 1993)

Amplitude Spatial Hybrid Functional


Parameters Parameters Parameters Parameters

RMS Density of RMS Slope Surface


Deviation Summits s4 Bearing Index
sq Sds Sb i

Ten Point Texture Mean Core Fluid


Height Aspect Summit Retention
sz Ratio Curvature Index
Str ssc Sci

Skewness Texture Developed Valley Fluid


Ssk Direction Area Ratio Retention
Std Sdr Index Svi

Kurtosis Fastest Decay


Sku Autocorrelation
Length
Sal

The texture parameters are the summit density Sds, the texture aspect ratio
Str, the texture direction Std and the shortest correlation length Sul. The summit
density is the number of local maxima of z (x, y) per unit area of the surface; thus a
local maximum at the bottom of a large valley will be included by the definition.
The texture aspect ratio is the ratio of the longest correlation length to the shortest
correlation length (Tsukada 62 Sasajima 1983) (Fig.9.3). Correlation length is
defined as the radial distance required for the area ACF to decay to 0.2. The
texture direction is the angle relative to the polar axis normal to the lay, for which
the angular variation of the area PSDF at some unspecified power is a maximum.
Su$aces in 3 0 175

b 1
0 3

0 2

0 1

001 ' '


0.4
' '
0 2 0 3 0'4 0 5 C m
0 6 0 mn.210 p / m

Figure 9.3. Contour of p (",y) = 0.2 for the surface of Fig. 9.2. The texture aspect ratio is the ratio of aa' to
bb', about 2.5. Note that the scales of this figure and Fig. 9.4 are Cartesian, not polar, so radial distances
must be scaled accordingly

The hybrid parameters are the RMS slope Sdq, the mean summit curvature
Ssc and the developed area ratio Sdr. The developed area ratio is the analogue of
the profile length ratio. The slope at any point x, y is

2
z'(x,y) = /(E)2+(g)
a &
so

The curvature at any summit is

Remembering that the sum of the curvatures of a surface at a point is equal to the
sum of the principal curvatures,
176 Rough SurJaces

where there are N summits.


The functional parameters are the surface bearing index Sbi, the core fluid
retention index Sci and the valley fluid retention index Svi. The surface bearing
index is the ratio of Sq to the height above the mean plane at 5% bearing area; for
a Gaussian height distribution Sbi = 0.61. The core and valley fluid retention
indices are respectively

and

(9.7)

where z* = z / Sq and the suffices refer to bearing area fractions. The statistical
approach to quantifying roughness makes it clear that the bearing length fraction
of a profile through a random surface must be identical to the bearing area fraction
of the whole surface, because the height distribution of any profile must be the
same as the surface height distribution. This may seem obvious but at one time
was disputed by a school of thought which claimed that the profile bearing length
fraction should be squared to obtain the bearing area fraction. It is of course
possible to construct a surface for which this is the case, for instance a rectangular
array of rectangular towers, but such a surface does not satisfy the condition that all
profiles through it have the same statistical properties.
The computation of some of these parameters will be discussed later. Some
general comments may be made here. As in two dimensions, the amplitude
parameters, and those depending on the autocorrelation function ,are sensitive to
long wavelengths, and the other texture and hybrid parameters are sensitive to
sampling interval. Stout et al. make no specific proposals for filtering on the
grounds that this should be carried out on a functional basis. In addition, the
summit density and mean summit curvature are sensitive to the definition of a
summit. The definition of texture direction is ambiguous; if a fine textured finish
were superimposed on a coarser finish, the texture direction as defined could vary
with the level of power (Fig. 9.4). It seems likely that the so-called "functional"
parameters are redundant; as they depend only on the amplitude distribution,
enough information to reconstruct them is probably combined in the RMS
Surfaces in 3 0 177

deviation, skewness and kurtosis. So far few instrument manufacturers seem to


offer the Birmingham 14; ironically, the 3D parameter most widely available on
proprietary systems is Sa, the 3D analogue of Ra, which is not even on the list.

i . t

I 3 9 rnrn.64 P/mrn 1 99 rnrn - 6 4 P-’rnm

Figure 9.4. Contours of equal power for a plateau honed surface at two diferent levels of power. The texture
direction is easier to discern ifthe level of power is defined.

9.3. Random Processes in Three Dimensions

Analogues of the autocovariance and autocorrelation functions, the power spectrum


and the structure function all exist in three dimensions. The 3D ACVF is

and the 3D ACF

where the normalising factor R(0, 0) is the variance Sq2 of the surface height
distribution. For an anisotropic surface the ACVF has the serious practical
drawback of being multi-valued at the origin. The ACF, being normalised, does
not suffer from thus drawback, but must be interpreted cautiously as it scales
differently in different angular directions (Fig. 9.5). Presentations like Fig. 9.5
make use of the symmetry properties of the ACF, i.e.
178 Rough Surfaces

A
m
v

Fig. 9.5. (a) Gritblasted, (b) ground and (c) plateau honed surfaces with their 3D autocorrelationfunctions
and power spectra (Amini et al. 1998)
Sufaces in 3 0 179

P (-L-rJ = p (rx,5) (9. 10)

to plot the origin at the centre of the figure, but the axes are Cartesian, not polar.
The 3D PSDF (Fig. 9.5) is

The 3D structure function is (Fig. 9.6)

Figure 9.6. 3D structure functions ofthe surfaces of Fig. 9.5, in the same order. Vertical scales are in pn2,
horizontal scales are in mm (Amini et al. 1998)

S(z, zJ is the expected value of the squares of the differences in height between all
the pairs of points on the surface whch are separated by d(r? + q2)(Fig. 9.7). If
the ratio between ry and r, is maintained at some constant value a, then a straight
line through (rx= 0, ry = 0) at an angle a to the r, axis is the ensemble average
structure function for all profiles which could be drawn at an angle tan-' a to the x-
axis. S(r,, 0 ) and S(0, 5) are the ensemble average SFs for all profiles parallel to
the x and y axes respectively. S( r- c) where c is a constant is the ensemble average
180 Rough Sufaces

of all pairs of heights separated by a distance r., on pairs of profiles themselves


separated by a distance c; a kind of cross-correlation.

Figure 9.7. 3D structure knction (Amini et al. 1998). Pairs of points z (1. j ) in the x,y plane separated by
d(h2 + zi2)

9.4. The Surface as a Random Process

The description of surfaces in three dimensions requires the study of functions of


several random variables. Nayak (1971), followed by Whitehouse & Phillips
(1978, 1982) and Greenwood (1984), pointed out the engineering significance of
the work of Longuet-Higgins (1957a, b, 1962). The description of ocean surFaces
led Longuet-Higgins to develop in a classical series of papers the theory of
statistical geometry. The full power of statistical geometry applies to situations in
which the surface height gradient and curvatures are random and furthermore the
surface is in motion. Herein we consider only cases in which the surface is static.
Nayak assumed the surface to be stationary and random, with a Gaussian
distribution of heights. Starting with profile statistics, he showed that a number of
properties of the profile statistical geometry could be expressed in terms of the first
three even moments of the profile power spectrum. The density of zero crossings,
that is the number of times per unit length that the profile crosses its own mean
line, which is equivalent to some definitions of the high-spot density, is given by
Surfaces in 3D 181

and the density of extrema, that is the number of local maxima and minima per
unit length, is similarly given by

The mean absolute profile slope is

(9.15)

He goes on to define a bandwidth parameter

a = rngn4/rn; = (9.16)

The bandwidth parameter is related to the breadth of the surface PSD. As a +


1.5, its limiting value, the spectrum gets narrower, and as a + 00, it gets broader.
The probability density of peak heights is a function of a; as a + co, the
distribution approximates to the distribution of all surface heights, while at the
other extreme as a + 1.5 the distribution is clearly non-Gaussian and the peak
height lies more than one standard deviation above the profile mean line (Fig. 9.8).
The mean curvature of peaks is also a function of a (Fig. 9.9). As a + co, the
curvature becomes independent of peak height, while as a -+ 1.5, the higher peaks
become sharper than the lower peaks.

-2.0 -1.0 0 1.0 2~ in -29 -19 0 1.0 2.0 3.0


ZIRs ZlRq

Figure 9.8. Probability densities of heights of (left) peaks (right) summits for various values of cz (Nayak
1971)
182 Rough Surfaces

-
$ 3
e'
z
:
$ 0

-3' ij'
,z 9
a
c
g E l
6
E l

0
0
1.0 I .o 0 1.0 2.0 3.0
z/Rq z/Rq

Figure 9.9. Dimensionlessmean curvatures of (left) peaks (right) summils for various values of a (Nayak
1971)

Extending the analysis to an isotropic surface, he showed that the first three
even moments of the surface power spectrum are identical to the corresponding
profile moments, and hence that the density of summits is given by

Sds = (1 I 67143) (m4/ m2) (9.17)

and the absolute mean surface slope is

A, = 47cm2i2) (9.18)

so
A, = (n/ 2) Ap (9.19)

The probability density of summit heights and the mean summit curvatures
follow trends similar to those of the peaks (Figs. 9.8, 9.9), and as a + 00, the peak
and summit distributions both tend to the distribution c;f all surface heights.
Othenvise, the distributions show distinct differences; the mean summit height is
much higher than the mean peak height as CL + 1.5. As a --+ co, peak and summit
curvatures tend to a constant value, the summit curvature a little larger than the
peak. But as a -+ 1.5, curvatures become sharper with increasing height, and the
peaks become a little sharper than the summits.
These are clearly important results. They imply that for a Gaussian isotropic
surface, much of the information needed to predict the practical behaviour of the
surface in, for instance, contact mechanics, may be obtained simply by measuring
the power spectrum of any profile. The theory is sufficiently robust that its
Sur$aces in 3 0 183

predictions are substantially in agreement with experiment even for visibly non-
Gaussian surfaces (Fig. 9.10)

z/Rq

Figure 9. D. Ground surface 5 x 7 mm, sampled on a 7.8 x 12 pxn grid (Sayles & Thomas 979). (a)
389803 heights; smooth line is Gaussian distribution with the same standard deviation: @) 34997 summit
heights; smooth line is Nayak's prediction

It is of some interest to see how the moments and their associated geometrical
parameters behave for the regions of real profile spectra which we can measure.
For power-law spectra of the form of Eqn. 8.8 the moment equations 8.4 become

(9.20)
OH

integrating over a pass-band of profile frequencies between a high-pass cutoff


and a low-pass cutoff m. If m is set by the sampling interval so that a = 251 / Ax
and a N m, and if -3 < p < -1, as is usually the case, then the first three even
moments are approximately
184 Rough Sur$aces

(9.21)

(9.22)

(9.23)

In other words, the roughness is independent of the sampling interval but


depends on the high-pass cutoff, while the slopes and curvatures are independent of
the high-pass cutoff but depend on the sampling interval. Furthermore, the summit
density and the bandwidth parameter become

(9.24)

and

(9.25)

So the bandwidth parameter depends on the ratio of cutoffs, as one might


expect, and the summit density decreases as the square of the sampling interval.
Note that the summit density is a function only of the sampling interval and the
rate of decay of the power spectrum, and so at any particular scale of measurement
it is more or less independent of the process by which the surface is produced
(Sayles & Thomas 1979). Eqns. 9.21-9.23 imply that the numerical values of the
moments, and hence the geometrical properties of the surface which depend on
them, are not an intrinsic property of the surface but are dependent on the scale of
measurement (Fig. 9.1l),as a number of workers have pointed out (Nayak 197 1,
Greenwood 1984, Majumdar & Bhushan 1990, Kant 1996). None of these
expressions converges, so in the absence of any natural short-wavelength limit to
the scale of surface features, it seems that as the scale of measurement gets smaller,
so the number of summits increases and they become steeper and sharper
indefinitely.
Surjaces in 3 0 185

Figure 9.1 1. Variation of spectralmoments with sampling &al on a grttblasted surface (Sayles &
Thomas 1979). Open Circles: 6om Da 0.; closed circles: from correlation; continuous lines: from
distributions

Power-law spectra are characteristic of fractal surfaces, and the moments can
be related to the fractal dimension and the topothesy through Eqns. 8.10 and 8.11,
hence the statistical geometry of the surface can be quantified from fractal
parameters. Note that from Eqns. 8.10, 9.24 and 9.25, the summit density and
bandwidth parameter are independent of the scaling factor and depend only on the
cutoffs and fractal dimension.

9.5. Practical Computation

The coefficients of the least-squares mean plane sampled at intervals Ax, Ay are

(7mn +m+n - 5)w - 6(n + 1)n - 6(m + l)v


a= (9.26)
mn(m + l)(n + 1)

b = -6 2u-(m-l)~
(9.27)
Ax mn(m - l)(m + 1)

c=-
6 2v-(n-l)~
Ay mn(n - l)(n + 1) (9.28)
186 Rough Surfaces

where

(9.29)
i=l J = I

n m

v= XT(Z- l)Zi, (9.30)


? = I j=1

n r n
w = zi,j (9.3 1)
i=l j = 1

For practical implementation of filtering algorithms the reader is referred to


specialist texts, e.g. Wade 1994, Golten 1997, Rorabaugh 1937.
Turning to the Birmingham 14 parameters, the discrete form of Eqn. 9.1 is

(9.32)

Other amplitude parameters such as Su may be evaluated similarly, e.g. the discrete
skewness is

(9.33)

Measurement of the ten-point height presents the same problem as


measurement of the summit density: the definition of a summit. The first difficulty
is that the number of summits counted per unit area depends strongly on the
sampling interval, as discussed above. The second, and associated, difficulty is
that a local maximum of a discrete process can only be defined in terms of its
neighbours, and a summit higher than, say, its 4 nearest neighbours may not be
higher than its 8 nearest neighbours (Fig. 9.12). Greenwood (1984) has calculated
that for Nayak's surface model, 22% of 5-point summits are not 9-point summits.
The dependence of summit properties on summit definition is discussed
exhaustively by Whitehouse (Whitehouse & Phillips 1978, 1982, Whitehouse
1994).
Surfaces in 3 0 187

Ridge

SUMMIT NO SUMMIT

Saddle point

SUMMIT N O SUMMI'I

Figure 9.12. Discrepancies in summit definition: A and B are higher than their 4 nearest neighbours, but not
lugher than their 8 nearest neighbours (Thomas 1997, after Greenwood 1984).

The numerical values of slopes and curvatures also depend on sampling


interval, as discussed in the previous section. Stout et al. recommend a two-point
formula to compute the slopes and a three-point central difference formula to
compute summit curvatures:

i z(i + 1, j ) + z(i - 1, j ) - 2x(i, j ) + z(i,j


(W2 (AYI2
+ 1)
- 1) + ~ ( ji , - 2z(i, j )

(9.35)

The discrete forms of the 3D ACVF and structure function (Eqns. 9.8 and 9.12) are
188 Rough Surfaces

and

The nearer the area ~5 approaches to the area of measurement A, the fewer data
points are available for the computation of R(rmrJ and the less reliable its
estimates will be, the limiting case being 42;5 < A. In practice the ACVF is often
obtained by Fourier transforming the power spectrum. The PSDF is best computed
by FFT, and again the reader is referred to texts on signal processing.
Nayak's moments can of course in principle be computed directly from the
power spectrum, but in practice it is rarely convenient to do this. There are three
other approaches which may be used instead (Sayles & Thomas 1979). After first
measuring Sq = dmo,the first and simplest method to find the higher moments is
from Fqns. 9.13 and 9.14 by Counting the number of peaks, valleys and mean line
crossings in some length of the profile. Unfortunately this turns out to be the least
accurate. The second method is to measure the variance of profile slope and
curvature distributions. The third method is to merentiatr: the profile ACVF at
the origin, as m2 and m4 are respectively its second and fourth differentials. In
discrete form, the first three points of the ACVF are needed:

(9.38)

2@R(0)- 4R(Ax) +R(2&))


m, = (9.39)
(W4

9.6. Anisotropy

So far in this chapter the topics which we have treated have all had their 2D
counterparts. We now come to a topic which has no counterpart in 2D because it
deals with the directional properties of surfaces. Up to this point in the book we
Surfaces in 3 0 I89

Figure 9.13. Results of some common machining processes


190 Rough &$aces

have usually assumed, explicitly or implicitly, that roughness is an isotropic


phenomenon, that is to say that surfaces will have the same microgeometric
properties no matter what direction they are measured in. In fact this is not at all
the case; most common machining processes produce surfaces with highly
dmctional properties (Fig. 9.13).
These anisotropic surfaces are said in traditional engineering parlance to
possess a lay. Lay is rather Micult to define mathematically but rather easy to
recognize visually; it is the principal direction or directions in which machining
marks seem to the eye to line up with each other. Aniscltropy is not a feature
unique to man-made surfaces; anyone who has tried to cross the South Wales
mining valleys from east to west instead of north to south will agree that the local
topography possesses a distinct lay.
Processes like surface grinding and shaping can produce a surface which is
almost two-dimensional, where a profile parallel to the lay will appear to be nearly
smooth. Face turning produces a surface with a circular lay, while some milling
and honing processes produce surfaces with a complex pattern which although
clearly visible is difficult to describe (Fig. 9.14). Furthermore, the initial and final
machining operations on a stratified surface may create lays in two distinct
directions; the upper load-bearing region of the surface may have ridges running in
one direction, while the deep valleys may tend to line up in quite a different
direction.
Clearly these directional properties are likely to affect the functional
behaviour of the surface. Summits which are long and narrow are likely to have
different load-bearing properties from more symmetrical summits, for instance,
and similarly the passage of fluid between contacting surfaces is likely to be
influenced by the lay. It becomes important, therefore, to find if possible some
quantitative way of characterising anisotropy. There are really two different
problems: to find the direction of the lay, and to quantify the degree of anisotropy.
Dealing with the first problem, Boudreau & Raja (1992) observed that in 3D
data from a raster scan, the lay will show up as a feature repeating between pairs of
parallel profiles. The delay in the repetition will depend in a simple geometric way
on the angle between the lay and the profiles; if the lay is circular, of course, this
angle will change from profile to profile. The delay, which they call the relative
shift, may be found by cross-correlating pairs of parallel profiles z(x,), z(x3 and
measuring the displacement of the peak of the cross-covariance function (Peklenik
& Kubo 1968, Kubo & Peklenik 1968)

(9.40)
Sur$aces in 3 0 191

TYPE LAY SYMBOL

Parallel
/c_I
Perpendicular

Crossed d.
Multi -directional

Particulate

Circular

Radial

Figure 9.14. Different kinds of lay and associated drawing symbols (Dagnall 1980; see also BS 1134:1988)
192 Rough Sudaces

from the origin. They succeeded in determining the angle of the lay for several
surfaces with parallel lays, and the characteristic radius of machining for several
surfaces with circular lays. For surfaces whose lay is not curved, the polar or
quasi-polar presentation of the power spectrum recommended by Stout et al. (1993)
can highlight the principal directions of the lay effectively (Fig. 9.4), and the
texture aspect ratio provides a measure of the degree of anisctropy (Fig. 9.3).
Longuet-Higgins' (1962) analysis of a surface as a random process included a
discussion of anisotropy. He found that in general 9 moments of the surface PSDF,
including 3 odd moments, were necessary to characterise an anisotropic surface,
but that these could be combined in only 7 independent combinations. A ratio of
these combinations defined a measure of anisotropy which he called the long-
crestedness lly. For an isotropic surface y = 1 and for a "two-dimensional" surface
on whch all the waves have infinitely long crests y = 0. McCool (1984) pointed
out that if profile slopes are measured at various angles to some arbitrary reference
on the surface, yis just the ratio of the maximum and minimum RMS slopes.

Figure 9.15. Isotropy and anisotropy (Thomas199 1). (a) isotropy: 1 mm x 1 mm of a shotblasted surface;
(b) weak anisotropy: outlined area of (a) stretched 7.5 times in the y-direction, (c) strong anisotropy: 1.2 x
1.3 mm of a ground surface
Surfaces in 3 0 193

Longuet-Higgins' anisotropic analysis was subsequently extended by Nayak


(1973) and Semenyuk (1986a, b). Nayak showed that the 7 invariants of Longuet-
Higgins could be derived from only 5 nonparallel profiles, which is still rather
discouraging for practical purposes. Bush et al. (1979) made further progress by
drawing an important distinction between strong and weak anisotropy. In the
general case of weak anisotropy, indwidual surface features are exaggerated in one
direction; a weakly anisotropic surface may be thought of as an isotropic surface
which has been stretched out in one direction (Fig. 9.15). In the special case of
strong anisotropy, on the other hand, the major axes of the surface features are
aligned. Bush et al. were able to show that a strongly anisotropic surface can be
characterised by only 5 independent parameters which may be obtained from two
profiles only, one parallel and one at right angles to the lay (though Rudzit 1984
recommends a 45 degree profile as easier to measure). Fortunately, many
machining processes are likely to produce strongly anisotropic surfaces.

8% 1'z 25 J% r4 1% w x
-16.6 -> 42.h -24.8 -) -28.-
98

8% 1% g% z
-28.9 -> -24.8~1~

Figure 9.16. Anisotropy of a milled surface at various height levels (Zahouani 1997)

Approaches based on, or conceptually similar to, moM analysis have been
used to characterise anisotropy. Grigoriev et al. (1997) used an approach based on
pattern-recognition algorithms adapted from image processing and succeeded in
classlfj.ing more than 100 AFM measurements of surfaces into 4 texture groups.
Zahouani (1997) developed 3D motifs and by selecting the characteristic height of
the motif was able to detect different directions of anisotropy at different levels of
1 94 Rough Su$aces

the surface (Fig. 9.16). Barre et al. (1997) adapted techniques from
geomorphology to detect analogues of catchment areas and watersheds on
machined surfaces. To give meaninml results it was necessary to increase the size
of motif so as to lose the finer detail. The process looks very much like low-pass
filtering, and appears to give results rather similar to the PSDF method of Stout et
al. discussed earlier.

18 0 ow1

5
17
-E 1 E-05
'I 1.6 X'
1E-06
L P
6 15
a 5 1E-07
5 14
Y
0
c
1 E-08
13
1 E-09
1 .2

':I o
,
10 M
, ,
30
,
40
,
50
,
60
Angle to x-axis (degrees)
70
, ,
a0
,
90
1E-10

1E-11
0 10 20 30 40 50 60
Angle to x-axis (degrees)
70 80 90

0.01 I, I I , ,
0 10 20 30 40 50 60 70 80 90
A"@s to x-axis (degrees)

Figure 9.17. Variation of fi-actaldimension and topothesy with angle to the x-axis for the surfaces of Fig. 9.5
(Amini et al. 1998). Circles: gritblasted; lozenges: ground; triangles and crosses: plateau honed, short and
long wavelengths respectively

The fractal dimension of an isotropic surface is well established to be 1 + the


fractal dimension of any profile through the surface ( e g Russ 1994), but what is
the fractal dimension of an anisotropic surface? Russ discusses this question at
length, and argues that the answers are different for weak and strong anisotropy.
For a weakly anisotropic surface, the fractal dimension of a profile will be
independent of the angle of measurement, but the topothesy will change. For a
Su$aces in 3 0 195

strongly anisotropic surface, both fractal dimension and topothesy will change; the
fractal dimension along the lay will be less than that across the lay (Thomas &
Thomas 1988). Davies & Hall (1998) predict that for a strongly anisotropic
surface the fractal dimension will be the same in all angular directions except
along the lay, where it will decrease. Rather confusingly, in the fractal literature
"isotropic" and ''anisotropic'' are sometimes used as synonyms for self-similar and
self-afhe respectively (e.g. Arvia & Salvarezza 1994).
In an attempt to investigate the effect of anisotropy on fractal parameters,
Amini et al. (1998) measured profile fractal dimension and topothesy as a function
of angle for the three surfaces of Fig. 9.5 by taking sections through the 3D
structure function. The gritblasted surface was isotropic, the ground surface was
strongly anisotropic and the plateau-honed surface was an example of a stratified
surface with a more complex anisotropy. For the gritblasted surface, both the
fractal dimension and the topothesy were found to be independent of the angle of
measurement (Fig. 9.17). For the ground surface, the fractal dimension rose from
a minimum across the lay to a more or less constant value, then fell to a sharp
minimum along the lay. The topothesy also showed a sharp minimum along the
lay, where it fell by a dramatic 6 orders of magnitude. Fractal parameters for the
plateau-honed surface showed minima parallel to the direction of the honing
scratches.

9.7. References

Amini, N., B.-G. Rosen and T. R. Thomas, "Fractal surfaces characterised by


a 3D structure function", Proc. Fractal 98 (in the press)
Arvia, A. J., and Salvarezza, R. C., "Basic aspects regarding irregular metal
surfaces and their application in electrochemistry", J. de Physique 4, C1, 39-53
(1994)
Barre, F., Lopez, J., Mathia, T. G., "New method for characterising the
anisotropy of engineering surfaces", Trans. 7'h. Int. Con$ On Metrology 8
Properties of Engng Surfaces, 479-486 (Goteborg, 1997)
Boudreau, B. D.; Raja, J., "Analysis of lay characteristics of three-
dimensional surface maps", International Journal of Machine Tools &
Manufacture 32, 171-177 (1992)
BS 1134 Part 1, Assessment of surface texture: methods and instrumentation
(British Standards Institution, London, 1988)
196 Rough Sugaces

Bush, A. W., Gibson, R. D. and Keogh, G. P., "Stroqgly anisotropic rough


surfaces", Trans. ASME: J. Lub. Tech., 101F, 15-20 (1979)
Dagnall, H., Exploring surface texture, (Rank Taylor Hobson, Leicester,
1980).
Davies, S., and Hall, P., "Fractal analysis of surface roughness using spatial
data", J. Roy. Statist. SOC.Ser. B 61 (in the press)
Golten, J., Understanding signals and systems (McGraw-Hill, London, 1997)
Greenwood, J. A., "Unified theory of surface roughness", Proc. Roy. SOC.
Lond. A393, 133-157 (1984)
Grigoriev, A. Y., Chizhik, S. A,, Myshkin, N. K., "Texture classification of
engineering surfaces with nanoscale roughness", Trans. 7th. Int. Conf: On
Metrology & Properties of Engng Surfaces, 3 19-324 (Gotebcrg, 1997)
Jonasson, M., Wihlborg, A., Gunnarsson, L., "Analysis of surface topography
changes in steel sheet strips during bending under tension friction test", Trans. 71h.
Int. Con$ On Metrology & Properties of Engng Surfaces, 38-46 (Goteborg, 1997)
Kant, R.,"Statistics of approximately self-afline fractals: random corrugated
surfaces and time series", Phys. Rev. E53, 5749-5763 (1996)
Kubo, M., and Peklenik, J., "An analysis of micro-geometrical isotropy for
random surface structures", Ann. CIRP 16,235-242 (1968)
Longuet-Higgins, M. S., "The statistical analysis of a random, moving
surface", Phil, Trans. Royal SOC.,A249,32 1-387 (1957a).
Longuet-Higgins, M. S., "Statistical properties of an isotropic random
surface", Phil. Trans. Royal Soc., A250, 157-174 (1957b).
Longuet-Higgins, M. S., "The statistical geometry of random surfaces", Proc.
13th Symp. on Appl. Maths, 105-143 (Amer. Maths. SOC.,1962).
Majumdar, A,; Bhushan, B., "Role of fractal geometry in roughness
characterization and contact mechanics of surfaces", Trans. ASME. Journal of
Tribology 112, 205-216 (1990)
McCool, J. I., "Characterisation of surface anisotropy", Wear 49, 19-31
(1978)
Nayak, P. R., 'IRandom process model of rough surfaces", Trans. A.S.M.E: J.
Lubr. Tech., 93F, 398-407 (1971).
Nayak, P. R., "Some aspects of surface roughness measurement", Wear, 26,
165-174 (1973).
Peklenik, J. and Kubo, M., "A basic study of a three-dimensional assessment
of the surface generated in a manufacturing process", Ann. C.I.R.P., 16, 257-265
(1968).
Sur$aces in 3 0 197

Rorabaugh, C. B., Digital filter designer's handbook 2e (McGraw-Hill, New


York, 1997)
Rudzit, Y. A. ,"Methodology of analyticexperimental- determination of
friction surface microtopographic parameters", Soviet Journal of Friction Wear, 5 ,
57-63, (1984)
Russ, J. C., Fractal surfaces, (Plenum Press, New York, 1994).
Sayles, R. S.; Thomas, T. R., "Measurements of the statistical microgeometry
of engineering surfaces", JLubr Techno1 Trans ASME 101,409-417 (1979).
Schmoeckel, D., and Staeves, J., "Function-oriented 3D filtering for
tribological assessment of sheet metal surfaces in deep-drawing", Trans. 7fh.Int.
Con$ On Metrology & Properties of Engng Surfaces, 438-4A4 (Goteborg, 1997)
Semenyuk, N. F., "Average values of total and mean summit curvatures and
heights of asperities of an anisotropic rough surface." Soviet Journal of Friction
and Wear, 7,47-56, (1986)
Semenyuk, N. F., "Summit height probability density and anisotropic rough
surface summit characteristics", Trenie i Iznos, 7, 1017-1024 (1986)
Shunmugam, M. S. and Radhakrishnan, V., "Computation of the three-
dimensional envelope for roughness measurement", Int. J. Mach. Tool Des. Res.,
14,211-216 (1974).
Shunmugam, M. S. and Radhakrishnan, V., "Two- and three-dimensional
analyses of surfaces according to the E-system", Proc. I. Mech. E., 188, 691-699
( 1974).
Shunmugam, M. S., "Effectiveness of the E-system in three-dimensional
roughness measurement", Proc. Int. Con?Prodn. Engng. 2, (Inst. Engrs. (India),
Calcutta, 1977)
Stout, K. J., Sullivan, P. J., Dong, W. P., Mainsah, E., Luo, N., Mathia, T.
and Zahouani, H., The development of methods for the characterisation of
roughness in 3 dimensions, EC Contract No.3374/1/01170/90/2, Phase I1 Report,
Vol.1 (March 1993)
Thomas, T. R., "Trends in surface roughness", Trans. 7". Int. Con$ On
Metrology & Properties of Engng Surfaces, (Goteborg, 1997)
Thomas, T. R. and Thomas, A. P., "Fractals and engineering surface
roughness", Surface Topography, 1, 1-10 (1988)
Thomas, T. R., "Some problems in the characterisation of surface
microtopography", Proc. SPIE 1573, 188-200 (1992)
Tsukada, T. and Sasajima, K., "An assessment of anisotropic properties in
three-dimensional asperities", Bull. Japan SOC.of Prec. Eng., 17, 26 1-262 (1983)
198 Rough Sudaces

Wade, G., Signal coding and processing 2e (Cambridge University Press,


1994)
Whitehouse, D. J., Handbook of surface metrology, (Institute of Physics,
Bristol, 1994)
Whitehouse, D. J. and Phillips, M. J., "Discrete properties of random
surfaces", Phil. Trans. R. SOC.Lond., A290, 267-298 (1978)
Whitehouse, D. J. and Phllips, M. J., "Two-dimensional discrete properties
of random processes", Phil. Trans. R. Soc. Lond., A305, 441-468 (1982)
Williamson, J. B. P. and Hunt, R. T., "Relocation profilometly", J. Phys. E;
Sci. Instrum., 1, 749-752 (1968).
Zahouani, H., "Spectral and 3D motifs identification of anisotropic
topographical components. Analysis and filtering of anisotropic patterns by
morphological rose approach", Trans. 7th.Int. Con$ On Metrology & Properties of
Engng Surfaces, 222-230 (Goteborg, 1997)
CHAPTER 10

APPLICATIONS: CONTACT MECHANICS

Although it is clear that the existence of surface irregularities has profound effects
in numerous engineering situations, knowledge of these effects has until
comparatively recently been largely qualitative. Examples of the successful
quantitative relation of specific surface parameters to engineering function are
rather rare. The reasons for this are to be found in earlier chapters of this book.
To restate them briefly here, none of the conventional surface parameters is an
intrinsic property of a surface, hence its correspondence with any particular surface
phenomenon will be at best accidental. All surface parameters vary with the scale
over which they are measured. To apply a surface measureIlient to an engineering
problem it is essential that the scale of the problem and the scale of the
measurement be related.
As an illustration (Thomas & King 1977), imagine taking a 1:50 000
geographic map and progressively enlarging it by linear factors of 10. The
smallest feature we could resolve initially would be about l00m across. After only
one enlargement the topography starts to have an engineering effect; height
variations with a wavelength of 10 m will cause vibrations in the suspension of an
aircraft as it lands which will have to be allowed for in the design. After another
enlargement, to 1 m, a similar effect will be produced on the suspension of road
and rail vehicles. Amplitudes on this scale may vary from 10 to 100 mm.
At 10 cm we are down to the scale of surface features which influence tyre-
road interactions such as skidding. We are also for the first time within the range
of machined surfaces; the performance of a machine tool, for instance, will be
influenced by features on this scale which transmit vibrations through its joints.
Features of this size may also slow down ships by increasing hull friction
(amplitude 0.1 - 1 mm).
Below this we are firmly in the region of machined surfaces. Undulations of
wavelengths from 1 mm down to 1 pm may increase friction and wear and promote
noise, rough running and finally failure in bearings of all kinds. They may also
affect the performance of nuclear power stations and space satellites by increasing
their thermal resistance, or the functioning of telephone exchanges and other
switchgear by affecting electrical resistance (amplitude 0.0 1 -- 10pm).

199
200 Rough Surfaces

At 1 pm and below another set of properties becomes of engineering


importance, namely the reflection or diffraction of electromagnetic radiation.
When surface irregularities are present at wavelengths comparable with those of
visible light the appearance of the surface will alter, e.g. a painted surface such as a
car body may appear dull instead of glossy (amplitude 0-10 nm).
Thus the essential problem of the quantitative application of surface
measurements is the choice of scale.
In a previous chapter it was shown that the numerical values of height-
dependent parameters depend on the longest wavelengths measured, while those of
texture-dependent parameters such as slopes and curvatures depend on the shortest
wavelengths measured. If one attempts to include the effect of all the wavelengths
in the continuous spectrum present on most real surfaces one will obtain the trivial
result that all roughness parameters tend to infinity or zero.
To obtain finite numerical values for surface parameters it is necessary to
reject certain portions of the spectrum at both its short-wavelength and its long-
wavelength ends. A measuring instrument does this as a matter of course by virtue
of its design and construction; the finite dimensions of the probe remove certain
short wavelengths (the action of the so-called "footprint" described by Newland
1986) and the filter circuits remove certain long wavelengths. Unfortunately the
portions of the spectrum thus rejected are chosen quite arbitrarily, and there is no a
priori reason why the roughness values thus obtained should have any functional
significance.

Wavelengths tw long
L ffect interacbon

0
a

UL OH 1 I Wavelength

Figure 10.1. Functional filtering:only the pass-band of surface wavelengths between high-pass cutoff oHand
low-pass cutoff wL take part in the surface interaction

The obvious solution is to confine measurement to the portion of the spectrum


of wavelengths which actually take part in the phenomenon under investigation
(Fig. 10.1). The definition of this band of wavelengths is equivalent, in the
Contact Mechanics 20 1

terminology of communications engineering, to the application of a band-pass


filter, and we have termed the process "functional filtering" (Thomas & Sayles
1973). The pass-band is specific not to the surface but to the particular interaction
involved; the roughness and asperity density "seen" by reflected electromagnetic
radiation, for instance, will be quite different from those "seen" by the contact of,
say, a journal bearing on the Same surface. The problem then reduces to the choice
of the cut-offs which define the pass-band.
The cutoff which rejects the long wavelengths - the high-pass cutoff, in
electronics terminology - is the easier one to select, as it fairly clearly must be
related to the largest horizontal dimension of the surface interaction. In many
contact problems, for instance, the high-pass cutoff is set by the dimensions of the
nominal contact area. This is simply to say that wavelengths much longer than the
nominal contact area will not affect what goes on inside it. The dimensions of the
nominal contact area are not always the obvious ones; for instance, in a constrained
reciprocating contact the critical dimension may be the stroke.
Whether the numerical value of the cutoff and contact size should be the
same is not entirely clear. It has been suggested that a cutoff of twice the contact
size is more realistic (Leaver et al. 1974) because it allows for the lack of sharpness
(the roll-off) characteristic of all real filters. However, as the roughness often
increases only as the square root of the high-pass cutoff, the exact value chosen is
probably not critical. In the following pages we will discuss some practical
problems in which surface roughness is involved, in terms of functional filtering.

10.1. The Contact of Rough Surfaces

There are many situations, in engineering and other disciplines, where rough
surfaces are brought together and it is important to know the topography of the
contacting area. We will discuss the actual mechanics of contact a little later; for
the time being we can start from the observation of Bowden and Tabor (1950) that
the real area of contact is independent of the nominal area and is in most practical
situations only a tiny fraction of the nominal contact area. We will also simplify
the model to the contact of a rough surface with a smooth flat plane. Greenwood &
Tripp (1971) have discussed the contact of two rough surfaces and have shown that
this can be reduced to the contact of a single equivalent rough surface with a
smooth flat plane. In practice one contacting surface is often in any case so much
smoother than the other that no important information is lost by considering it as
perfectly smooth. What we are interested in for the moment is the relationship
202 Rough Surfaces

between the approach of the surfaces and the real area of contact, and the way in
which this real contact area is distributed over the nominal area of contact.
The simplest model is the intersection a of rough surface with a plane parallel
to the rough surface’s mean plane. In physical terms this is equivalent to contact
with complete loss of the displaced material, or an abrasive process which removes
all material above a given height without disturbing the material below. This
model may seem somewhat unrealistic, but in fact the resulting errors only become
appreciable when the surfaces approach each other very closely (Pullen &
Williamson 1972). The fractional area of contact at any height h from the mean
plane of the rough surface is then by definition the bearing area or bearing length
introduced in Section 7.4. If we normalize this height as a dimensionless
separation t = h / CT = h / Rq, then for a Gaussian random surface, in terms of the
dimensionless height s = z / 4

and the fractional real area of contact becomes

(10.2)

where @ (t) is the cumulative Gaussian probability function. Note that by this
definition of separation, the fractional area of contact is still only % at zero
separation; negative separations are possible; and complete contact is never
attained.
The next question to investigate is how the real area of contact is made up.
How many discrete areas of contact, or contact spots, are there at any separation,
and how big are they? By extending Nayak’s (1971) theory it can be shown
(Sayles & Thomas 1976, 1978) that the density (i.e. the number per unit nominal
area) of closed contours at any separation is

D, = ( 2 7 ~ ) ~ / ~ ( m 2 / m o ) q ( t ) (10.3)

and an upper bound for the mean contact spot radius is

(10.4)
Contact Mechanics 203

In other words, as the surfaces get closer, the numLer of closed contours
increases but their average size stays more or less the same (Fig. 10.2), so that the
increase in the real area of contact is almost entirely due to the increase in the
number of contacts. Of course a closed contour may be a hole inside a region of
contact, but at large separations the probability of this is small enough to be
ignored (Sayles & Thomas 1978). Note that up to this point the discussion has
been purely geometrical; we have made no assumptions about the actual
mechanism of contact or about the relationship between separation and load.

7/ ii* I
lo-.t
I /
/ // C’ I
l0-Y /
W*

Figure 10.2. Dimensionless mean contact spot radius, contact density and thermal conductance as a function
of separation and dimensionless load (Sayles & Thomas 1976)

Sayles & Thomas (1978) suggested that the distribution of contact spot sizes
was log-normal. Majumdar, however, (Majumdar & Tien 1990, Majumdar &
Bhushan 1990) observed that for fractal surfaces at any separation the number of
contact spots Nw larger than a given area % follows a power law:

(10.5)

where !RiL is the area of the largest spot at that separation (Fig. 10.3). He pointed
out that although the number of infinitesimally small spots is infinitely large, their
204 Rough Surfaces

contribution to the total area of contact is negligible, an observation confirmed by


Klimczak (1992).
--
MAGNETIC TAPE

0 38MR

10’ -
MAGNETIC RIGID DISK

.
.
too 101 Id Id Id
AREA. .[lUn’l

Figure 10.3. Size distribution of contact spots for a magnetic tape surface under two different loads, and a
magnetic disk (Majumdar & Bhushan 1990)

Such fractal spots are not generally circular; there is a power-law relationship
between their perimeter and their area, but the exponent is significantly larger than
% (Russ 1994) (Fig. 10.4). This has obvious implications for the calculation of
conductance or contact stress.

Log area

Figure 10.4. Total perimeter length as a function of total area of closed contours on a simulated isotropic
fractal surface (Russ 1994)
Contact Mechanics 205

10.2. Rough Contact Mechanics

Now that we have some idea how the real area of contact varies with separation, it
is time to consider how the separation might vary with the load. The following
discussion concentrates on those aspects of contact mechanics which depend on the
general topography of the surface; for a more general account of the principles of
contact mechanics and their application to rough surfaces, the reader is referred to
the books by Johnson (1985) and Hills et al. (1993). Rough contact is also
reviewed by Greenwood (1992).
Up to now we have considered rough surfaces as a continuum, but to
calculate their behaviour during contact it is necessary to model them as an array
of discrete physical objects distributed in some way both in the plane of the surface
and perpendicular to it. These objects are often termed asperities in the literature,
and we may think of them as related to the summits of the Nayak theory (though
remembering that Nayak’s summits are local maxima and so may occur at any
height, whereas contact must take place at least initially only at the highest parts of
the surface). The local load on an individual asperity will cause it to deform, at
first elastically. If the load continues to increase beyond its capacity to recover
elastically it will suffer irreversible plastic deformation.
We will consider later how to determine the mode of deformation. In general
some of the asperities may deform elastically and others plastically, depending on
their original geometry, the load and the physical properties of the bulk material.
The easiest cases to consider are those where the mode of deformation is
everywhere the same, that is where all the asperities deform plastically or where
they all deform elastically. We will deal with the plastic case first as the load P per
unit nominal contact area can be related directly to the separation through the real
area of contact and the hardness H by means of the bearing area:

A, = P/H = 1 ~ @(t) (10.6)

Knowing the separation, the variation of number of contacts and mean


contact spot size can be obtained from Eqns. 10.3 and 10.4. The number of
contacts increases almost linearly with load while the mean contact size remains
almost constant, a result confirmed by many experiments (e.g. Thomas & Probert
1970, Uppal et al. 1972, though not universally accepted: see Tian & Bhushan
1996). Note that these results are obtained without assuming any particular
geometrical model for an individual asperity. Care should be taken in selecting a
value for the hardness; the indentation hardness, commonly taken as 3 x the yield
206 Rough Surfaces

strength of the bulk material, is not appropriate for predicting the behaviour of an
unsupported asperity, which will deform under a normal pressure much closer to
the yield strength itself (Thomas et al. 1971).
At very high loads, as in drawing operations, one would expect behaviour to
change, as some of the simplifying assumptions made above no longer hold. Also
as the mean separation of individual contact spots approaches their mean size, the
asperities will begin to lend one another mutual support and the effective hardness
will increase. Pullen & Williamson (1972) maintained that as the load increased
the non-contacting parts of the surface moved uniformly upward, though they were
never able to suggest a mechanism for this arresting phenomenon. The well-
known figure which purports to show this effect (Fig. 10.5) in fact refers all heights
to their respective profile mean lines as separately calculated. But there was no
independent height datum for the series of experiments, hence no evidence that the
successive mean lines remained at the same height relative to the bulk material.
The load progressively flattening the tips of the asperities has the effect of
censoring the upper tails of the successive height distributions and so pushing their
recalculated means closer to the lower tails; if the successive means are then
constrained to line up, the rest of the distribution will of course appear to move
upwards.

Load (kN)

Figure 10.5. Spurious upward movement ofthe non-contacting surface under increasing load (Pullen &
Williamson 1972)
Contact Mechanics 207

Elastic contact is not quite so straightforward. To apply the theory of


elasticity it is necessary to have some geometrical model of an individual asperity
or at least of its tip, usually the model of an elastic sphere originally due to Hertz
(Timoshenko & Goodier 1951). For an elastic sphere of radius R in contact with a
semi-infinite elastic half-space, Hertz obtained the area of contact A and the load W
in terms of the compliance w (the distance by which points outside the deforming
zone approach):

A = nRw (10.7)

W = (413) E'd@w2) (10.8)

where the harmonic elastic modulus E' is given by

2 2
1 - 1-v,
+-1-v, (10.9)
E' E, E2

and v is Poisson's ratio. Archard (1957) modelled a rough surface as an array of


hemispherical asperities covered with smaller asperities, each of which is covered
by even smaller asperities, and so on (Fig. 10.6). By applying Hertz's relations
successively he showed that for such a surface the relationship between load and
real area of contact becomes closer and closer to proportionality as the number of
layers of asperities is increased. He explained that "the essential part of the
argument was not the choice of asperity model: it was whether an increase in load
creates new contact areas or increases the size of existing ones" (Greenwood &
Williamson 1966).

Figure 10.6. Multiple-scale asperity models for elastic contact (Archard 196 1): real area of contact is
proportional to the load raised to the power (a) 2/3 (b) 415 ( c ) 819 (d) 14/15 (e) 26/27 (t)44/45
208 Rough Sut$aces

Greenwood & Williamson combined Archard's model of a surface covered


with hemispherical elastic asperities with a statistical distribution of asperity
heights each of the same radlus, obtaining expressions for the density of contact
spts, real contact area and load per unit nominal area:

0,= Sds F o ( t ) (10.10)

A, = ZSdsR o F I ( t ) (10.1 1)

P = ( 413 ) Sds E' R'" d/2F3,2 ( t ) (10.12)

where
m

F,(t) = j ( ~ - t ) ~ p ( s ) d s (10.13)
t

This is a general expression for any probability distribution; if p(s) is Gaussian, the
area of contact is found proportional to the load.
The GW theory is in good qualitative agreement with experiment (Handzel-
Powierza et al. 1992), and many attempts have been made to obtain quantitative
agreement. They have mostly foundered on the usual reef, the difficulty of
obtaining unique values of summit density and curvature. In addition the
assumption that all summits have the same curvature and a Gaussian height
distribution has been superseded by the work of Nayak discussed previously. For
this reason Bush et al. (1975) attempted an "asperity-free" model of elastic contact
based on Nayak's approach and showed that at large separations,

e x p - t 2 I 2) (10.14)
A, =
2 t G

(10.15)

Eliminating the separation gives strict proportionality between area and load:

(10.16)
Contact Mechanics 209

The real area of elastic contact is just half the bearing area, that is half the
real area of plastic contact. Unfortunately the theory does not distinguish between
number and size of contact spots, and again suffers from the difficulty of defining a
unique second moment. McCool(1986) makes a detailed numerical comparison of
the predictions of Bush et al’s theory and Greenwood & Williamson’s theory with
those of an asperity simulation model and finds that they are in broad agreement.
Chang et al. (1987) modified the GW theory to allow for the plastic
deformation of the most highly loaded asperities, using a model based on volume
conservation of the plastically deformed region. Their model predicts significantly
higher separations at high loads but otherwise agrees with the GW model within a
few percent. The Chang-Etsion-Bogy (CEB) model is now widely used, though
there seems little practical gain to justify the extra complexity.
Bhushan and his co-workers have carried out extensive investigations of
rough contact (Oden et al. 1992, Ganti & Bhushan 1995, Tian & Bhushan 1996,
Poon & Bhushan 1996a, b, Yu & Bhushan 1996). As we saw in Section 9.4, for
fractal surfaces, all the moments of the power spectrum which define roughness,
slopes and curvatures can be expressed in terms of fractal parameters. It should be
possible, then, to recast theories of elastic contact in fractal terms. They began
(Majumdar & Bhushan 1990, 1991) by modelling a self-affine fractal surface as a
Weierstrass-Mandelbrot function, but it proved rather difficult to extract the
characteristic parameters from a real surface for purposes of experimental
comparison. However, they were able to show that the real area of contact, and the
relative proportions of it in elastic and plastic contact, are quite sensitive to the
fractal parameters. In an extension of this work to bifractal surfaces (Bhushan &
Majumdar 1992), the relationships could be quantified by invoking one extra
parameter, the size of the largest contact spot. Fractal parameters were found from
a profile structure function, and the largest contact spot was presumably measured
directly, though the paper does not make this clear. In any event the theory
underestimated the real area of contact by an order of magnitude, possibly due to
the uncertainty of statistical inference from an extreme value. A model of elastic-
plastic contact of self-af€ine fractal surfaces based on the Cantor set has been
proposed by Warren & Krajcinovic (1999, but only relates load to displacement.
Lee & Ren (1996) generated computer models of Gaussian rough surfaces
with varying degrees of weak anisotropy and simulated elastic-plastic contact
numerically. They obtained relationships between separation and real area of
contact in terms of dimensionless hardness and dimensionless load normalised by
the correlation length. They found that at high loads the variation of contact area
with load was quite non-linear (Fig. 10.7).
210 Rough Sudaces

E
.-0
Y
u
k

1 .o 1 ."

0.5

.o
0.0
m b

Figure 10.7. Real contact area as a function of dimensionless hardness and dimensionless load (Lee & Ren
1996)

All the theories we have considered so far have assumed isotropy, but as we
saw earlier, most machined surfaces are anisotropic. Bush et al. (1978) obtained a
solution for elastic contact of strongly anisotropic rough surfaces in terms of
moments of the surface, rather than the profile, power spectrum. At low loads the
area of elastic contact was proportional to the load. So & Liu (1991) extended this
to include plastic contact. They found that the elastic portion of the contact area
was almost independent of anisotropy, but the proportion of plastic contact varied
significantly with the degree of anisotropy (Fig. 10.8). Lee & Ren, however,
concluded that for weak anisotropy, contact could be treated as two-dimensional for
asperity aspect ratios greater than 6. McCool (1986) compared the predictions of
the Bush et al. model with his own numerical model for an isotropic surface and a
surface with 10:1 anisotropy. The models agreed that at a given separation the real
area of contact in the anisotropic case was the same as in the isotropic case, but
real pressures were much lower.
Contact Mechanics 21 1

U
a
\
a
a

-2
16'
'Bp;
Figure 10.8. Proportion of plastic contact as a function of dimensionless load for various degrees of
anisotropy (So & Liu 1991)

As Chang et al. (1987) observed, theories of rough contact are extraordinarily


hard to verify experimentally. There are various reasons for this: it is difficult to
measure the area of elastic contact as the asperities will recover their original
geometry; displacements normal to the surface are very small, and the stiffness of
the asperities may be of the order of the stiffness of the whole experimental rig; the
height of the first point of contact and the absolute height of the rough surface
mean plane are difficult to determine; small contact spots may be below the limit of
resolution of the measuring system. Woo & Thomas (1979) reviewed the
published experimental results for plastic contact and found real area of contact
proportional to the 0.83 power of load up to dimensionless loads of 0.1, and
separation inversely proportional to log (load ) (Fig. 10.9). No correlation was
found between size or number of contact spots and load, as insufficient information
was available in the published literature to normalise the data by any of the
methods dscussed above.
212 Rough Surfaces

5 I

I 6

Bearing
AIM it2
Ralio

163

la4 1 1 I I I
IbC 113) 16 16' .l I0
Oimcnsionlcrs Lood

Figure 10.9. Collation of measurements of variation of rough plastic contact with dimensionless load from 7
sources (Woo & Thomas 1979): (above) real area of contact, solid line is best power law fit; (below)
separation, solid line is best log-linear fit

10.2.1. Contact of Curved Surfaces

This is a case of some practical importance as many, perhaps the majority, of


engineering contacts are curved. The problem here is complicated by the difficulty
Contact Mechanics 213

that because of the gross curvature the pressure distribution varies over the nominal
area of contact in an unknown way. An iterative approach to this problem was
suggested by Greenwood and Tripp (1967) for spherical contacts and investigations
were extended to cylindrical contacts by Lo (1969), Dobychin (1988) and
Merriman & Kannel (1989). Rather than a rough sphere on a flat plane, Poon &
Sayles (1994) modelled a smooth sphere on a rough plane. Greenwood and Tripp
found that the maximum pressure, and the area over which the load was
distributed, could differ significantly from the classical Hertzian solution for
smooth surfaces (Fig. 10.10).

Radial distance

Figure 10.10. Dimensionless pressure as a function of dimensionlessradial distance from the centre of an
elastic contact between a rough sphere and a plane (Greenwood & Tripp 1967). Broken line: classical Hertz
solution for a smooth surface; solid line: rough-surface solution

Again, to obtain numerical solutions a pass-band of surface wavelengths must


be defined. Such a calculation has been carried out for two practical cases
(Thomas 1979). The first case was the contact of a pushrod of cup radius 10 mm
with a rocker arm of ball radius 5 mm under a mean working load of 1 kN, taking
their combined roughness as 0.5 pm. The low-pass cut-off was then estimated as
33 pm from the plasticity approach, giving 1,100 asperities available for contact
per unit area with a mean tip radius of curvature of 310 pm. After iteration the
final rough contact area was found to be 20% greater in diameter than the
corresponding smooth contact area, and the maximum stress was 95 per cent of
214 Rough Su$aces

that for a smooth contact. It appears, therefore, that in this practical case the
roughness does not have a significant effect.
In the second case a journal bearing with a bronze liner was considered,
having a nominal radius of 25 mm, a clearance ratio of 1 in 1,000, an axial length
of 50 mm, a working load of 8kN and a combined roughness of 1 pm. Because of
the relative softness of the bronze liner the low-pass cut-off was as high as 0.6 mm,
i.e. all surface wavelengths shorter than this are immediately flattened plastically.
The density of asperities was correspondingly low at 3.7 mm-2and their radius of
curvature was no less than 11.3 mm. Convergence after three iterations yielded an
effective roughness of 6 pm. The nominal contact area was 2.2 times wider than
that for a smooth contact and the maximum stress was only 64 per cent that of the
smooth case. The implication would seem to be that it is worth making a journal
bearing as rough as is consistent with adequate lubrication.

10.2.2 Joint Stiffness

Machine-tools are not generally manufactured as a continuous casting or


fabrication; the reasons for this are functional, such as the necessity to incorporate
guideways, and also difficulties in manufacture and transportation. Most practical
designs for machine-tool structures, therefore, incorporate some form of connection
between the basic elements (Back et al. 1973). The stiffness or lack thereof of such
a joint is likely to affect the dynamic properties of the machine. It is known that
the dynamic stiffness of machine-tool joints is proportional to their preload, and
this result can be derived by considering the joint as an assembly of Gaussian
elastic asperities after Greenwood and Williamson's model (Thomas & Sayles
1977). The constant of proportionality was calculated for some experimental
stiffness measurements of Thornley and Lees (1971) on joints of various planforms,
and reasonable agreement was found between theory and experiment.
To find the numerical value of the constant it is necessary to know the density
and mean curvature of the asperities, both of which, as noted above, depend on the
pass-band of surface wavelengths chosen. A detailed calculation was carried out
from surface measurements of the contact between the bed of a lathe and its saddle
(Thomas & Sayles 1977). Here the high-pass cut-off was set at 61 cm by the
nominal length of the contact. The roughness measured at this cut-off was 3.3 pm.
The low-pass cut-off was found from the plasticity argument given above to be 19
pm. The mean plane separation could then be calculated by the method of
Contact Mechanics 215

Greenwood and Williamson to be 17 pm under a load of 1 kN supported by only 16


asperities, and the stiffness was calculated as 1.4 MN/mm.

10.3. The Plasticity Index

Greenwood & Williamson (1966) looked for a criterion which would determine the
mode of deformation of an array of asperities of varying heights. They found that
the mode of deformation of the highest asperities was almost independent of load.
Sharp asperities would deform plastically even under the lightest loads, while blunt
asperities would deform elastically even under the heaviest loads. The criterion of
sharpness was the ratio of the standard deviation of the height distribution, o,,to
the asperity radius of curvature R, and their result applied to any exponential
distribution of asperity heights. For the special case of a Gaussian distribution,
they showed that the top 2% of asperities would deform plastically under any load
for ry > 1, and elastically under any load for ly < 0.6, where

y~ = (E'/H)t/(o,/R) (10.17)

In the region 0.6 < ly < 1 the mode of deformation is dependent on the load. The
plasticity index ry is thus a dimensionless figure of merit which can predict the
dominant mode of deformation.
Many workers have used the plasticity index, with some success, as a
qualitative index of comparison witlun a series of tests, i.e. the higher the plasticity
index, the more likely a surface is to suffer wear or similar problems. There are
difficulties in comparing results between laboratories or in using it as an absolute
index, however, because of the difficulties in quantifying the surface parameters R
and o,. Because these are both properties of the peak distribution they depend on
the definition of a peak, and raise the problems which we have encountered
previously. For this reason various other formulations of the plasticity index have
been proposed, for instance one due to Mikic (1974) which replaces peak
parameters by the mean slope.
Bush et al. (1978) developed an expression for the plasticity index of a
strongly anisotropic surface in terms of Nayak's moments:

(10.18)
216 Rough Sudaces

where moo is the variance of surface heights, moa m2a mO4and m40 are the second-
and fourth moments of the power spectra of profiles parallel to and across the lay,
respectively, and

(10.19)

where moo 'mqo . - moo Om04


(10.20)
a,= 2 , a 2 - 2
17220 m02

According to Bush et al., deformation will be plastic for y > 0.7, elastic for ty <
0.5, and load-dependent in the intermedate region.
How anisotropic must a surface be to require the rather laborious anisotropic
treatment? According to Wu & Zheng (1988) the correction for anisotropy
increases very slowly with the degree of anisotropy y (Fig. 10.11). Their paper
does not appear to distinguish between strong and weak anisotropy, but this does
not affect the present argument. The degree of anisotropy is the ratio of the major
and the minor asperity radii of curvature, which is approximately the square of the
ratio of the major and minor axes of the ellipse projected when the asperity
intersects a plane parallel to the surface mean plane (Bush et al. 1978).

2.0

1.8

1.6

1.4

1.2

1 .o

0.8
1 1 . 5 2 3 4 5

Figure 10.11. Anisotropy correction factor for plasticity index as a function of degree of anisotropy y(Wu &
Zheng 1988)
Contact Mechanics 217

Remembering that the plasticity index only deals with the highest regions of
the surface, one may observe that the higher one looks on any surface, however
anisotropic, the less elliptical the tips of asperities appear (Fig. 10.12). If one
attempts to fit a best ellipse to some of these irregular shapes, the ratio of major to
minor axes is not more than about 3, corresponding to y = 0.1, in which case, from
Fig. 10.11, the anisotropy correction would be only about 10%. Bearing in mind
the large statistical uncertainties in measurement of somc of the other surface
parameters, it seems likely for many surfaces that isotropic calculations will
suffice, in which case Eqn. 10.18 simplifies to

(10.21)

Figure 10.12. "And we in dreams behold the Hebrides": contours more than 1 prn above the mean plane on a
plateau-honed surface do not look very elliptical
218 Rough Surfaces

Recalling that the second moment of the profile power spectrum is related to
the mean slope Bby Eqn. 9.15:

so
(10.21)

or
(10.22)

This formulation of the plasticity index simplifies the measurement problem


considerably. Instead of measuring asperity radii of curvature or power spectra
directly, we can replace these by the much easier measurement of the profile slope,
which merely requires computation of the standard deviation of the first differential
of the profile. But there are still some remaining practical obstacles. The slope is
a function of the sampling interval and increases as the sampling interval
decreases. This simply indicates that as asperities get smaller they also get
sharper. It follows that, by varying the sampling interval, we can obtain any
desired value of the slope, and hence of the plasticity index. To put this another
way, there always will be features on the surface so small and sharp that they will
deform plastically on contact.
In obtaining numerical solutions in all the above cases the basic problem is
the same: the choice of the low-pass cutoff. This amounts to asking the question:
What is the smallest surface feature which will affect contact? There does not
seem to be any general answer to this question. Many workers have implicitly, and
Ganti & Bhushan (1995) have explicitly, assumed that no features smaller than the
resolution of the particular instrument are important, but it is difficult to find
physical grounds to justify this.
One possible approach is to work back from the plasticity index itself
(Thomas & Sayles 1977). As asperities get smaller and sharper, a size will be
reached below which they will deform plastically during the very first cycle of
contact and so dlsappear (Archard 1961). During the subsequent lifetime of the
component, it will behave elastically as if the corresponding range of surface
wavelengths did not exist; in other words, the initial encounter of the surface will
act as a natural low-pass filter. The critical wavelength at which this filtering
Contact Mechanics 219

occurs may be found from the relationship between the second moment and the
plasticity index.
If vCis the critical value of the plasticity index above which deformation will
be plastic at any load, then the critical second moment

r n Z c = (71( 2 - nJ2 )I-” wc2 ( H I E ’ ) (10.23)

The exact form of Eqn. 9.22 is

m2 = B (3 + p).’ ( 271)-p ( e3+p


- m3+p) (10.24)

If p> 1 and the bandwidth of surface wavelengths is reasonably wide, i.e. oL)) oH,
then Eqn. 10.24 reduces to

m2 E B Q , ~ +1~( 3 + p ) ( 271)p (10.25)

The critical wavelength dc= 271/ CL)LC, so combining Eqns. 10.23 and 10.25,

(10.26)

From Bush et a1 1978, IY, , = 0.7. Combining this with the numerical constant,
replacing B and p by the corresponding fractal parameters from Eqns. 8.10 and
8.11 and rearranging, we have finally

20-1

(10.27)

where

(10.28)
f ( D ) = (-)@z)iO-LT(I-
1175 2D)cos- 2 - 0
1-20 2

This is a relationship between three dimensionless numbers: the critical


wavelength normalized by the topothesy, the fractal dimension and the material
property ratio. The dimensionless wavelength is highly sensitive to the other
parameters (Fig. 10.13), and for a given fractal dimension and material property
220 Rough Su?$aces

ratio the critical wavelength increases as the topothesy. Thus we can in principle
now find a unique short-wavelength cutoff, depending only on material properties
and intrinsic topography parameters, which we can use to determine the elastic
behaviour of the contact.

90
0
E-IH

Figure 10.13. Dimensionless critical wavelength as a function of material ratio and fiactal
dimension ( R o s h et al. 1997)

10.4. References

Archard, J. F., "Elastic deformation and the laws of friction", Proc. Royal
SOC.,A243, 190-205 (1957).
Archard, J. F., "Single contacts and multiple encounters", J. Appl. Phys., 32,
1420-1425 (1961).
Back, N., Burdekin, M, and Cowley, A., "Review of the research on fixed and
sliding joints", Proc. 13th Int. Machine Tool Des. & Rex Con$, 87-97 (1973).
Bhushan, B.; Majumdar, A., "Elastic-plastic contact model for bifractal
surfaces", Wear 153, 53-64 (1992)
Bush, A. W., Gibson, R. D., Keogh, G. P., "Strongly anisotropic rough
surfaces", Trans. ASME: J. Lub. Tech. 101, 15-20 (1979)
Contact Mechanics 22 1

Bush, A. W., Gibson, R. D. and Thomas, T. R., "The elastic contact of a


rough surface", Wear, 35, 87-111 (1975).
Dobychin, M. N., "Elastic contact of rough cylindrical bodies", Soviet
Journal ofFriction and Wear 9, 1-5 (1988)
Ganti, S., Bhushan, B., "Generalized fractal analysis and its applications to
engineering surfaces", Wear 180, 17-34 (1995)
Greenwood, J. A., "Contact of rough surfaces", 37-56 in I. L. Singer & H. M.
Pollock eds., Fundamentals of piction: macroscopic & microscopic processes,
(Kluwer, Dordrecht, 1992).
Greenwood, J. A. and Tripp, J. H., "The contact of two nominally flat rough
surfaces", Proc. 1. Mech. E., 186,625-633 (1970/71).
Greenwood, J. A. and Williamson, J. B. P., "Contact of nominally flat
surfaces", Proc. Royal SOC.A295, 300-319 (1966).
Handzel-Powiem, Z., Klimczak, T. and Polijaniuk, A., "On the
experimental verification of the Greenwood-Williamson model for the contact of
rough surfaces", Wear, 154, 115-124 (1992)
Hills, D. A., D. Nowell and A. Sacwield, Mechanics of elastic contacts
(Butterworth-Heineman, Oxford, 1993).
Johnson, K. L., Contact mechanics (Cambridge University Press, London,
1985).
Klimczak, T., "Predicting microcontact spots size distribution in contact
problems", Ann. CIRP 41,609-612 (1992)
Lee, S. C., and Ren, N., "Behavior of elastic-plastic rough surface contacts as
affected by surface topography, load, and material hardness", Trib. Trans. 39, 67-
74 (1996)
Lo, C. C., "Elastic contact of rough cylinders", Int. J. Mech. Sci., 11, 105-115
(1969).
Majumdar, A.; Bhushan, B., "Role of fractal geometry in roughness
characterization and contact mechanics of surfaces", Trans. ASME. Journal of
Tribology 112,205-216 (1990)
Majumdar, A.; Bhushan, B., "Fractal model of elastic-plastic contact between
rough surfaces", Trans. ASME. Journal ofTribology 113, 1-11 (1991)
Majumdar, A.; Tien, C. L., "Fractal characterization and simulation of rough
surfaces", Wear 136, 313-327 (1990)
McCool, J. J., "Comparison of models for the contact of rough surfaces",
Wear 107,3760 (1986)
222 Rough Surfaces

Merriman, T. and J. Kannel, "Analysis of the role of surface roughness on


contact stresses between elastic cylinders with and without soft surface coating."
Journal of Tribology, 111,87-94, (1989)
Mikic, B. B., "Thermal contact conductance: theoretical considerations", Int.
J. Heat Mass Transfer, 17,205-224 (1974).
Moalic; H., J. A. Fitzpatrick and A. A. Torrance, "A spectral approach to the
analysis of rough surfaces", Journal of Tribology, 111, 359-363, (1989)
Nayak, P. R., "Random process model of rough surfaces", Trans. A.S.M.E.
Ser. F. J. Lubr. Tech., 93, 398-407 (1971).
Newland, D. E., "The effect of a footprint on perceived surface roughness",
Proc. Roy. SOC.Lond. A405,303-327 (1986)
Men, P. I.; Majumdar, A.; Bhushan, B.; Padmanabhan, A,; Graham, J. J.,
"AFh4 imaging, roughness analysis and contact mechanics of magnetic tape and
head surfaces", Journal of Tribology, Transactions of the ASME 114, 666-674
(1992)
Poon, C. Y., and Bhushan, B., "Nano-asperity contact analysis and surface
optimisation for magnetic head slider/disk contact", Wear 202,83-98 (1996)
Poon, C. Y., and Bhushan, B., "Numerical contact and stiction analyses of
Gaussian isotropic surfaces for magnetic head slideddisk contact", Wear 202, 68-
82 (1996)
Poon, C. Y.; Sayles, R. S., "Numerical contact model of a smooth ball on an
anisotropic rough surface", Journal of Tribology, Transactions of the ASME 116,
194-200 (1994)
Pullen, J . and Williamson, J. B. P., "On the plastic contact of rough surfaces",
Proc. R. SOC.Lond. A327, 159-173 (1972).
Ro&, B.-G., R. Ohlsson and T. R. Thomas, "Nano metrology of cylinder
bore wear", Trans. 7th. Int. Con$ On Metrology h Properties of Engng Surfaces,
pp. 102-110 (Giiteborg, 1997)
Russ, J. C., Fractal surfaces (Plenum Press, New York, 1994).
Sayles, R. S. and T. R. Thomas, "Computer simulation of the contact of
rough surfaces", Wear, 49, 273-296 (1978).
Sayles, R. S. and Thomas, T. R, "Thermal conductance of a rough elastic
contact", Appl. Energy, 2, 249-267 (1976)
So, H.; Liu, D. C., "An elastic-plastic model for the contact of anisotropic
rough surfaces", Wear 146,201-218 (1991)
Thomas, T. R., "Calculation of elastic contact stresses for rough-curved
surfaces", ASLE Trans., 22, 184-189 (1979)
Contact Mechanics 223

Thomas, T. R., and King, M. J., Surface topography in engineering: a state


of the art review and bibliography (BHRA, Cranfeld, 1977)
Thomas, T. R. and Probert, S. D., "Establishment of contact parameters from
surface profiles", J. Phys. D: Appl. Phys., 3,277-289 (1970).
Thomas, T. R. and Sayles, R. S., discussion to Radhakrishnan, V., "Analysis
of some of the reference lines used for measuring surface roughness", Proc. 1.
Mech. E., 187, 575-582 (1973).
Thomas, T. R. and Sayles, R. S., "Random-process approach to the prediction
of joint stiffness", Trans. ASME: J. Eng. Znd. 99B, 250-256 (1977)
Thomas, T. R., Uppal, A. H. and Probert, S. D., "Hardness of rough
surfaces", Nature Physical Sci., 229, 86-87 (1971).
Thornley, R. H. and Lees, K., "The effect of planforni shape on the normal
dynamic characteristics of metal to metal joints", I. Mech. E., Paper C62/71,
(1972).
Tian, X., and Bhushan, B., "A numerical three-dimensional model for the
contact of rough surfaces by variational principle", Trans. ASME: J. Trib. 118, 33-
42 (1996)
Timoshenko, S., and Goodier, J. N., Theory of elasticity (McGraw-Hill, New
York, 1951)
Uppal, A. H., Probert, S. D. and Thomas, T. R., "The real area of contact
between a rough and a flat surface", Wear, 22, 163-183 (1972).
Warren, T. L.; Krajcinovic, D., "Fractal models of elastic-perfectly plastic
contact of rough surfaces based on the Cantor set", Internatanal Journal ofSolids
and Structures 32,2907- 2922 (1995)
Whitehouse, D. J. and Archard, J. F., "The properties of random surfaces in
contact", In: Surface Mechanics. Ling, F.F. (Ed.) 36-57 (A.S.M.E., New York,
1969).
Williamson, J. B. P., Pullen, J. and Hunt, R. T., "The shape of solid
surfaces", In: Surface Mechanics. Ling, F.F. (Ed.). 24-35 (A.S.M.E., New York.
1969).
Woo, K. L. and T. R. Thomas, "Contact of rough surfaces: A review of
experimental work", Wear, 58, 33 1-340 (1980).
Wu, C. and Zheng Linqing, "General expression for plasticity index." JVear,
121, 161-172, (1988)
Yu, M. M., and Bhushan, B., "Contact analysis of three-dimensional rough
surfaces under frictionless and frictional contact", Wear 200, 265-280 (1996)
CHAPTER 11

TRIBOLOGY

Discussion of the effect of roughness on contact mechanics leads naturally to a


discussion of the effect of roughness on friction, lubrication and wear. As we
would argue that all real contact is rough contact, it follows that roughness is a
complication in all tribological situations. There is no space here to do justice to
the full scale of the tribological implications of roughness, and we will confine
ourselves to discussing a few of the more interesting examples which illustrate
some of the previous topics. No comprehensive account of rough surface tribology
appears to exist, though Bhushan (1990) has a useful review.

11.1. Friction

It was recognised very early that friction was associated with surface roughness, to
the point where Coulomb suggested that friction was due to the effort required for
one surface to climb up the asperities of the other during translation (Bikerman
1944, Bowden & Tabor 1950). When it was pointed out that the energy thus
dissipated would be largely recovered when the surface slid down the reverse slopes
of the asperities, the Coulombic theory lost some of its popularity, but
microgeometry remains a important factor in friction. Manj workers have found a
correlation between roughness parameters and friction; friction increases with
average roughness (Furey 1963, Koura 1980) and also with mean profile slope
(Myers 1962, Ghabrial & Zaghlool 1974, Eiss & Warren 1975, Koura & Omar
1981, Moalic et al. 1987) (Fig. 11.1). Road roughness is llkely to influence friction
between a wet road and a tyre as asperities pierce the intervening film of water
(Taneerananon & Yandell 1981).
An opposite effect of roughness on friction has also been reported. Ogilvy
(1991, 1993) developed a statistical roughness model for adhesive friction which
predicted that friction would decrease with increasing roughness, tending to a
constant value independent of roughness. A wholly elastic and a mixed elastic-
plastic versionof the model were developed, dependin3 on two roughness
parameters, the RMS roughness height and the correlation length. In tests with a

225
226 Rough Surfaces

thin molybdenum disulphide film on a rough steel subtrate, the theory gave good
qualitative agreement, and reasonable quantitative agreement, with experiment
(Fig. 11.2).

Figure 11.1. Variation of dynamic friction coefficient with mean profile slope (Koura & Omar 198 1)

- dosli-plastic model

w b l l y elastic model

I ’ cipwimenlol volucs

0.00 0-10 0.20 0-30 0.40 0 50


rms heipht (microns)

Figure 11.2. Variation of fiction coefficient with roughness for steel coated with MoSz (Ogilvy 1993)

Chapman & Rizkallah-Ellis (1979) provide an interesting example of the


effect of scale of roughness on tribological interactions. They found a pronounced
directional effect in the coeficient of friction of automobile brake linings which,
after eliminating other possible causes, they concluded was due to the surface
topography. Reflectance measurements showed directional differences which could
Tribology 227

not always be detected by a stylus instrument. The implication is that at least some
of the surface features responsible for friction in this casc were smaller than a
stylus can resolve.
Proctor & Coleman (1988) and Harris and Shaw (1988) showed that the
roughness of floor surfaces was an important parameter in determining pedestrian
slip-resistance, at levels of floor roughness much lower than had been reported by
other researchers. It was well known that values of peak-to-trough surface
roughnesses (Rtm) of 15-600 pm had an effect (Jung and Riediger, 1982).
However, this level of roughness is many times higher than the Rtm values found
by Harris and Shaw to have an important influence on the safety of water wetted
floors. The significance of this finding was that it opened up new possibilities for
ensuring the safety of floors exposed to contamination by water. In essence, all
that is required is to incorporate a degree of roughness into the floor surface, that is
small enough not to detract from the aesthetic appearance or create problems for
cleaning. Since water is the most common floor contaminant, especially in public
areas exposed to wet footwear, this is a significant development.
Other researchers have confirmed the importance of surface roughness for
assessing both floor surfaces and footwear (Manning et al., 1991, Stevenson, 1989.
Gronqvist et al., 1990). Manning found that footwear soles of a granular
construction that maintain surface roughness during wear, retain their slip resistant
properties, whereas footwear soles that wear smooth, do not. He found a
correlation between sole roughness Rtm and coefficients of friction measured
during walking. He also found that the footwear ranked in approximately the same
order on seven floor/contaminant combinations; this suggest; that surface structure
determines the footwear ranking (Proctor 1993). Lloyd & Stevenson (1992)
obtained a correlation between slip resistance and a combination of Rq, skewness
and average wavelength.

11.2. Lubrication

A large number of papers have appeared describing or modelling the effect of


roughness on various lubrication regimes. Rough boundary lubrication has been
investigated by Hisakado (1978), Nivatvongs et al. (1991) and Denape et al. (1992,
1995). Christensen (1965/6, 1971, 1972), Rao & Mohanran: (1993), Wakuri et al.
(1995) and Liu et al. (1996) have reported work on rough mixed lubrication.
Rough hydrodynamic lubrication has been treated by Bush & Gibson (1980),
Kumar (1980), Nakahara et a1.(1984), Bayada & Chambat (1988), Cheng & Xie
228 Rough Sufaces

(1992), Sugimara & Yamamoto (1995) and Tonder and his co-workers
(Christensen & Tonder 1971, Tonder & Christensen 1972, Prakash et al. 1979,
Tonder 1980, 1987). Rough elastohydrodynamic lubrication (EHL) has been
investigated by Tallian et al. (1965/6),Coy & Sidik (1979), Johnson et al. (1972),
Kaneta & Cameron (1980), Karami et al. ( 1987), Lubrecht et al. (1988), Sadeghi &
Sui (1989), Sinha et al. (1987), Fan & Zheng (1991), Kaneta (1992), Chang et al.
(1993, 1994) and Ishibashi & Sonoda (1994). Chang (1995) has also reported on
rough partial EHL. Micro-EHL, which by definition deals with rough surfaces, has
been investigated by Baglin (1986), Kweh et al. (1989, 1992), Huang & Wen
(1993), and Chang & Zhao (1995); the topic has been reviewed by Chang (1995).
The specific application of lubrication in rough sliding has been dealt with by
Tzeng & Saibel (1967), Patir & Cheng (1979), Shukla & Kumar (1979), Hughes
(1981) and Anderson & Salas-Russo 1994. Other practical applications of rough
lubrication include compliant surfaces (Darbey et al. 1979), improvement of roller
bearing fatigue life (Akamatsu et al. 1991. 1992) (Fig. 11.3) and gear contacts
(Peeken et al. 1990). Unfortunately some of these researches have confined
themselves to simple deterministic roughness models and are thus of limited
applicability.

Skewness

Figure 11.3. Influence of skewness on relative fatigue life of rolling bearings (Akamatsu et al. 1991)
Tribology 229

An early application of functional filtering was to the case of a roller-bearing


operating in a regime of mixed lubrication (Leaver et a1 1974). The surface finish
of the bearings and races was measured before and after running-in, and a
pronounced difference was noted, though no scuffing or other damage was
apparent on the run-in surfaces. Measurements of the film thickness, however,
showed it to be much smaller than the combined roughnesses of the contacting
surfaces, and it was difficult to reconcile this with the absence of damage.
Application of the principle of functional filtering suggested that all wavelengths
longer than the Hertzian contact width should be ignored, as they took no part in
the interaction with the lubricant film. The effect of this was to reduce the
effective roughnesses to about a tenth of their measured valces, thus obtaining at a
stroke a far more plausible set of film-thickness ratios and allowing the no-contact
times to be predicted in fair agreement with experiment.
One suspects that if this principle were more generally employed it would
result in a substantial increase in the numerical values quoted in the literature for
the so-called lambda ratio, the ratio of oil film thickness to roughness. Cann et al.
(1994), in a review of the literature on lambda ratios, conclude that the behaviour
inside the Hertzian contact zone depends on the mean slope and the asperity
density as well as the roughness. They admit that many systems are known to run
successfully at low lambda ratios. They point out that as soon as the lubricated
surfaces are rough and the roughness heights are not negligible compared with the
mean oil film thickness, the local pressure fluctuations cause3 by the asperities will
have an influence on the elastic deformations of the surfaces.

11.3. Wear

The accommodation of two sliding surfaces over a period of time, variously known
as running-in, brealung-in or shakedown, causes changes in their initial
topography, and itself depends on that topography (Kapoor et al. 1994, Anderson
et al. 1996). Kang & Ludema (1986) reported an optimum initial roughness of
about 0.1 pm; smoother and rougher surfaces failed more quickly.
Chandrasekaran (1993) found a proportionality between reniprocal wear rate and
the reciprocal of initial roughness (Fig. 11.4). Summers-Smith (1969)
dstinguishes two basic types of running-in mechanism. On the one hand he
describes what he calls a “plastic squeezing” of the surface, that is a change in its
shape by redistribution of material due to plastic flow without net loss. On the
other hand there are the various wear mechanisms, adhesive or abrasive, all of
230 Rough &$aces

which do involve net material loss. These two types of running-in are associated
with quite different geometrical effects.

Figure 11.4. Wear rate of steel sliding against steel as a function of initial roughness (Chandrasekaran 1993)

The plastic redistribution of material during running-in is related to the


finishing process known as roller burnishing (Black & Kalen 1973). The unit
event is the compression of a single asperity by the roller until the plastic limit is
exceeded, when the asperity will deform plastically so as to redistribute its material
into the adjoining valleys. In general it will change its shape only until the new
contact area is large enough to support the stresses elastically. The zone of affected
asperities will therefore approximate to the nominal area of Hertzian contact
between the roller and the rough sphere. Surface features significantly larger than
this zone wilt not be affected by the burnishing process.
If the length of the nominal contact is small compared with that of the cut-off
of the measuring filter, very little change in roughness will be apparent from this
type of running-in, because the roughness measurement is weighted heavily by the
greater power associated with longer wavelengths. Whitehouse and Archard
(1969), for instance, found a maximum change in R a of less than 20% in these
Tn.bologv 23 1

conditions, although the changes in the actual profile were clearly visible to the
naked eye.
One way of overcoming this difficulty is to compare worn and unworn values
of roughness as the filter cut-off is progressively shortened (Fig. 11.5a). If
burnishing has occurred then power will have been lost by some band of
wavelengths and hence some cut-off will be reached where the roughnesses
diverge. The power spectrum is a more rational way of presenting the same
information (Fig. 11Sb).
(a)

I
i

2
6.(m
Cut-off wavelength (urn)

10 -6
f I
2
I
3 4
I 1
6
1
810
1 1
20
1 1 I l
3040 6 0 8 0 1 0 0
l
200
Frequency cyclelmrn :lmm
I

Figure 11.5. Topography of worn (circles) and unworn (crosses) inner races of a taper roller bearing
characterised as (a) roughness v. high-pass filter cutoff (b) power spectra (Laver et al. 1974)
232 Rough Surfaces

The simplest case of abrasive wear is a clean removal of the tops of the
asperities without smearing or tearing. This may be performed experimentally
under rather artificial conditions (Thomas 1972), but probably does not often occur
in practice. However, it is convenient to investigate because it is easily simulated
by computer (Thomas 1972; Willn 1972). As successive layers of the surface are
abraded, parameters such as roughness, mean slope and mean peak curvature
decrease in a systematic manner (Fig. 11.6).

I P

I T'i
-0 , I I I
IS 10 5 0 -5 - 10
Height of worn surface above original mean line (rm)
Figure 11.6. Effect of pure abrasion on (circles) RMS roughness (triangles) mean peak curvature (squares)
profile curvature standard deviation (lozenges) mean absolute slope. Open symbols are simulated results,
filled symbols are experimental. P = highest peak, V = lowest valley on unworn profile. Solid line is
Gaussian bearing-area curve (Thomas 1972)
Tribologv 233

These parameters are secondary rather than primary, as they can all be
represented by joint probability distributions of heights. A more fundamental
representation, then, is the height distribution. Hence it is important to describe
correctly what happens to the distribution. Suppose we have a distribution,
originally p{z), abraded until there are no heights higher than h above the mean
line. This is sometimes described as being truncated at h, so that the new
distribution

p{z) = p{z) for z 2 h


= Oforz>h

In fact the change is correctly described as being censored at h (Marcus 1967):

p(z) = p(z) for z < h


= p(h)forz>h

Figure 11.7. Proposed typology of running-in (Thomas 1978). From top to bottom: unworn profile with
power spectrum and height distribution; censoring without filtering; low-pass filtering without censoring;
high-pass filtering without censoring; high-pass filtering with censoring
234 Rough Surfaces

That is, for a truncated distribution the fraction of values of z greater than h
vanishes, but for a censored distribution it becomes equal to h, causing a Dirac
spike at h (Fig. 11.7). The distinction makes a considerable difference to the
moments of the distribution.
It is more characteristic of abrasive wear in general, and also of adhesive
wear, that changes in topography are due to the progressive removal of many small
particles over long periods of time. Golden (1976) made a mathematical
investigation of the effect on an initially Gaussian height distribution of a wear
mechanism such that the rate of loss in height of an individual asperity with time is
proportional to the depth to which the asperity has penetrated the opposing surface.
He concluded that the resulting topography is that of the original surface up to
some height representing the mean separation of the contacting surfaces. Above
this height is superimposed another distribution, also Gaussian but smoother; as
time progresses it will remain Gaussian but become smoother still (Fig. 11.8).
Topographies of this type have been found experimentally by Ostvik and
Christensen (1968/69) and also by Williamson et al. (1969) (see Section 7.4).

-
10-5

-
10-6

lo-'-

10-8 -

10-91 I I I I I I I I
-10.00-R.00-6.00 -4.00 -2.00 0.00 2.00 4.00 6.00
Height, units of standard deviation

Figure 11.8. Transitional double Gaussian height distribution of a surface at progressive stages of wear
(Golden 1976)
Tribologv 235

Censoring a random process reduces the amplitude of the autocovariance


h c t i o n but has little effect on its form. When the autocovariance function is
normalized to the autocorrelation function by the profile variance, the change due
to censoring is barely perceptible even when abrasion has progressed as far as the
mean line, as shown theoretically by Marcus (1967) and King et at. (1978) and
confirmed by experiment (Willn 1972). Radhakrishnan (1977) divided the profile
along the mean line and cross-correlated worn peaks and unworn valleys, but
without any greater success. It follows that no change will be seen in the
normalized power spectrum as this merely presents the same information in a
m e r e n t form.
A tentative classfication of run-in topographies was proposed by Thomas
(1978) (Fig. 11.7). The original surface may be censored without filtering, or it
may be high-pass filtered or low-pass filtered or both with or without censoring. In
principle, then, it should be possible to define a run-in topography fairly closely by
three or four parameters: the cut-off wavelength or wavelengths, the censoring
level of the height distribution and, if a transitional topography, the roughness of
the superimposed finish.

11.4. Seals

Roughness is a factor in determining the rate of leakage through seals, both static
w t c h e l l & Rowe 1967/8, 1969, Wallach et al. 1968, Chivers & Hunt 1978, Hehn
1970, Kazamaki 1974, Otto 1974, Warren et al. 1988, Matsuzaki et al. 1992, 1993,
Etsion & Front 1994) and dynamic (Lucas et al. 1994, 1995). Vacuum seals are a
special case of static seals where the performance criterion in terms of leakage rate
is particularly rigorous (Roth 1966, 1971, Yanagisawa et al. 1991).
The performance of radial lip seals is known to depend on roughness, but the
actual mechanism is not clear (Horve 1991, van Bavel et al. 1995). In the absence
of a satisfactory physical model, a discriminant analysis approach has been applied
(Thomas et al. 1975a, b) to distinguish between a set of rubber lip seals, individual
members of which had either sealed or leaked. Worn and unworn profiles on the
seals were measured and first 9 and then 11 parameters were computed. Two basic
and relatively straightforward procedures were implemented in the analysis of the
data. In the first, each of the nine features was examined individually. The
procedure of evaluation was based on simple analysis of variance. The seals were
first sorted into sealed or leaked categories. Then one of the surface measurements
(e.g. zero-crossing density) was examined. An average value for this parameter
236 Rough Su$aces

was computed for each performance category, and the overall or grand average
value for all data was calculated. Then the pooled sum of squares was computed
for the difference between each individual measurement and its corresponding
group mean. This quantity is called the within sum of squares. Next, a sum of
squares was computed by observing the difference between each group mean and
the grand mean. This sum of squares, expressed on a per-observation basis, is
called the between sum of squares.
The ratio of the latter to the former sum provides a measure of the
discriminating information available from the measurement in question. It is
reasonable to consider that the greater this ratio the better the ability to
discriminate between good and bad seals on the basis of a single parameter. In
particular, if the ratio is large, it connotes a wide separation between the two
groups of measurements but a relatively close clustering of individual
measurements within each group. Accordingly, the several measurements can be
ranked according to their (individual) ability to differentiate between effective and
ineffective seals. It turns out that this ranking is different for the worn and unworn
cases.

Table 1 1 . 1 . Discrimination between sealed and leaked in terms of 1 1 surface parameters measured from
worn and unworn profiles of 15 lip seals (Thomas et al. 1975a)

Discriminants Eigenvector x 10

Unworn WOA.n
Rq -6.9 -0.5
Ra 7.2 0.4
DO 0.8 -0 6
Peak height 5.8 -2.6
Valley depth 3.4 -0 1
Peak curvature -0.8 10.0
Valley curvature 1.4 -3.0
Profile curvature 0.1 -7 1
Slope -6.4 3.9
Sk 1 .o 07
............K
................. .*/................. .................. 0.5 .-1
................................_..
3 ........
...............__

Rather than attempt a classification of seals on the basis of a single 'best'


parameter, one may make much more effective use of the infomation contained in
Tribologv 237

the data by means of linear discriminant functions. In this approach, the various
parameters are combined as a weighted sum and the weights are adjusted so as to
maximize the ratio of between to within sum of squares. The construction of the
transformations appropriate to multiple linear discriminant functions is
straightfonvard and well documented. In essence, the approach defines a sum of
squares ratio in terms of the coefficients of a linear expansion. The ratio is then
differentiated with respect to the coefficients and solved for the values which
maximize the ratio. The problem reduces to finding the eigenvectors of a non-
symmetric matrix, and can be solved in a straightforward manner by any of a
number of standard eigenvectodeigenvalue routines. The eigenvector is composed
of the nine coefficients in the linear expansions. The associated eigenvalue
provides a measure of the amount of variance accounted for by the discriminant
function.
Separate analyses for worn and unworn surfaces ezch separated the two
categories and showed that the order of importance in which the parameters were
ranked for the worn surfaces was quite different from that of the unworn (Table
11.1). The hypothesis advanced to explain this was that the geometry of the worn
surfaces directly affects the sealing process, whereas that of the unworn surfaces
affects it indirectly only in so far as it influences the production of the final
geometry of the worn surfaces. As a result of the analysis it is possible to
reconstruct ideal models of successfully and unsuccessfully sealing surfaces (Fig.
11.9).

Figure 11.9. Reconstructions from pattem-recOgnition analysis of profiles of the contacting surfaces of (a) an
ideally good @) an ideally bad lip seal (discussion to Thomas et al. 1975b)
238 Rough Surfaces

Although this form of analysis cannot replace a proper understanding of a


particular problem it can aid it in several ways. By its assignation of degrees of
importance to particular parameters it can offer the theoretician a useful initial
guide for the formulation of ideas. In some practical cases it may be necessary to
go no further, and the discriminant functions themselves may serve as part of the
manufacturer's armoury of quality controls. The technique is also well suited to
interactive computing work, where successive parameters can be dropped from the
analysis until the engineer's subjective judgement decides that separation is no
longer adequate.

11.5. References

Akamatsu, Y., Tsushima, N., Goto, T. and Hibi, K., "Influence of surface
roughness skewness on rolling contact fatigue life", Trib. Trans. 35, 745-750
(1992)
Akamatsu, Y., Tsushima, N., Goto, T., Hibi, K., Itoh, K., "Improvement of
roller bearing fatigue life by surface roughness modification", SAE Trans. 100, 44-
49 (1991)
Anderson, P., Juhanko, J., Nikkila, A.-P., Lintula, P., "Influence of
topography on the running-in of water-lubricated silicon carbide journal bearings",
Wear 201, 1-9 (1996)
Anderson, S., Salas-Russo, E., "Influence of surface roughness and oil
viscosity on the transition in mixed lubricated sliding stee: contacts", Wear 174,
71-79 (1994)
Archard, J. F., "Elastic deformation and the laws of friction", Proc. Royal
SOC.,A243, 190-205 (1957).
Archard, J. F., "Single contacts and multiple encounters", J. Appl. Phys., 32,
1420-1425 (1961).
Baglin, K. P., "Micro-elastohydrodynamic lubrication and its relationship
with running-in", Proc. I. Mech. E, 200, 415-424 (1986)
Bayada, G. and M. Chambat, "New models In the theory of the hydrodynamic
lubrication of rough surfaces."Journal of Tribology, 110, 402-407, (1988)
Bhushan, B., Tribology and mechanics of magnetic storage devices
(Springer-Verlag, New York, 1990)
Bikerman, J. J., "Surface roughness and sliding friction", Rev. Mod. Phys.,
16, 53-68 (1944).
Tribology 239

Black, J. T. and Kalen, S. E., "The anatomy of a rolier burnished surface",


Proc. Int. Conf: on Surface Technol., 507-526 (SME, Dearborn, 1973)
Bowden, F. P. and Tabor, D., The friction and lubrication of solids Part 1
(Oxford University Press, 1950).
Bush, A. W.; Gibson, R. D., "Effect of surface roughness and elastic
deformation in hydrodynamic lubrication - A perturbation approach", in Surface
roughness effects in hydrodynamic and mixed lubrication, 173-191 (ASME, New
York, 1980).
Cann, P.; Ioannides, E.; Jacobson, E.; Lubrecht, A. A,, "Lambda ratio - a
critical re-examination", Wear 175, 177-188 (1994)
Chandrasekaran, T., "On the roughness dependence of wear of steels: a new
approach", J. Mat. Sci. Lett. 12,952-954 (1993)
Chang, L., "Deterministic model for line-contact partial elasto-hydrodynamic
lubrication", Tribology International 28,75-84 (1995)
Chang, L., "Deterministic modeling and numerical simulation of lubrication
between rough surfaces - a review of recent developments", Wear 184, 155-160
(1995)
Chang, L.; Jackson, A,; Webster, M. N., "Effects of 3-D surface topography
on the EHL film thickness and film breakdown", Tribology Transactions 37, 435-
444 (1994)
Chang, L.; Webster, M. N.; Jackson, A,, "On the pressure rippling and
roughness deformation in elastohydrodynamic lubrication of rough surfaces",
Journal of Tribology, Transactions of the ASME 115,439-444 (1993)
Chang, L. ; Zhao, W., "Fundamental differences between Newtonian and non-
Newtonian micro-EHL results", Journal of Tribology, Transactions of the ASME
117,29-35 (1995)
Chapman, R. J. and A. A. Rizkallah-Ellis, "Effect of the surface finish of
brake rotors on the performance of brakes", Wear, 57, 345-356 (1979).
Cheng, Y., Xie, Y., "Study of squeeze film between non-normal rough
surfaces under partial hydrodynamic lubrication", Journal of Xi 'an Jiaotong
University 26,59-65 (1992)
Chivers, R. C. and Hunt, R. P., "The achievement of minimum leakage from
elastomeric seals", 8th Znt. ConJ on Fluid Sealing, Paper 24 (BHRA, Cranfeld,
1978)
Christensen, H. and Tonder, K., "The hydrodynamic lubrication of rough
bearing surfaces of finite width", Trans. A.S.M.E. Ser.F. J.Lubr.Tech., 93, 324-330
(1971).
240 Rough Surfaces

Christensen, H., "A theory of mixed lubrication", Proc. I. Mech. E., 186,
421-430 (1972).
Christensen, H., "Nature of metallic contact in mixed lubrication", Proc.
Instn.Mech. Engrs., 180, Pt.3B (1965/66)
Christensen, H., Yome aspects of the functional influence of surface
roughness in lubrication", Wear, 17, 149-162 (1971).
Coy, J. J., S. M. Sidik, "Two-dimensional random surface model for asperity
contact in elastohydrodynamic lubrication", Wear, 57, 293-3 11 (1979).
Darbey, P. L.; Higginson, G. R.; Townend, D. J., "Lubrication of rough
compliant solids", , Proc. 5th. Leeds-Lyon Trib. Symp. 398-403 (MEP. London,
1979)
Denape, J.; Marzinotto, A,; Petit, J. A., "Roughness effect of silicon nitride
slidmg on steel under boundary lubrication", Wear 159, 173-184 (1992)
Denape, J.; Masri, T.; Petit, J.-A,, "Influence of surface roughness and oil
ageing on various ceramic-steel contacts under boundary lubrication", Proc. I.
Mech. E: J. Eng. Trib. 2095, 173-182 (1995)
Eiss, N. S. and Warren, J. H., "The effect of surface finish on the friction and
wear of PCTFE plastic on mild steel", S.M.E. Paper IQ75-125 (1975).
Etsion, I.; Front, I., "Model for static sealing performance of end face seals",
Tribology Transactions 37, 1 1 1-1 19 (1994)
Fan, Y., Zheng, L., "A study on the limit criterion of full and partial
lubrication", Wear 143,22 1-229 (199 1)
Furey, M. J., "Surface roughness effects on metallic contact and friction",
A.S.L.E.Trans., 6,49-59 (1963).
Ghabrial, S. R. and Zaghlool, S. A,, "The effect of surface roughness on static
friction", Int. J. Mach. Tool Des. Res., 14, 299-309 (1974).
Golden, J. H., "The actual contact area of moving surfaces", Wear, 42, 157-
162 (1977).
Gronqvist, R., Roine, J., Korhonen, E., Rahikainen, A,, "Slip resistance
versus surface roughness of deck and other underfoot surfaces on ships", J. Occup.
Accid. 13, 291-302 (1990)
Harris, G. W. and Shaw, S. R., "Slip resistance of floors: users' opinions,
Tortus instrument readings and roughness measurements", J. Occup. Accid. 9,
287-298 (1988)
Hehn, A. H., "Effects of friction and wear on a sealing interface", Lubr.
Engng. 26, 206-212 (1970).
Hisakado, T., "Influence of surface roughness on friction and wear in
boundary lubrication", J. Mech. Eng. Sci. 20,247-254 (1978).
Tribology 24 1

Home, L., "Correlation of rotary shaft radial lip seal service reliability and
pumping ability to wear track roughness and microasperity formation", SAE Trans.
100,620-627 (1991)
Huang, P.; Wen, S. Z., "Sectional microelastohydrodynamic lubrication",
Journal ofTribology, Transactions of the ASME 115, 148-151 (1993)
Hughes, W. F., "A cell theory of rough surface lubrication", Wear, 67, 31-53
(1981).
Ishibashi, A., Sonoda, K., "Mirrorlike finishing of precision rollers and
changes on the roller surfaces caused by loaded running", JSME International
Journal, Series 3 35,286-293 (1992)
Johnson, K. L., Contact mechanics (Cambridge University Press, London,
1985).
Johnson, K. L., Greenwood, J. A. and Poon, S. Y., "A simple theory of
asperity contact in elastohydrodynamiclubrication", Wear, 19, 91-108 (1972).
Jung, K. and Riediger, G., "Recent developments regarding the inspection of
non-slip floor coverings", Die Berufgenossenschaft 6,l-7 (1982)
Kaneta, M.; Cameron, A., "Effects of asperities in elastohydrodynamic
lubrication", ASME Papern 79-Lub-6 (1979).
Kaneta, M., "Effects of surface roughness in elastohydrodynamic
lubrication", JSME International Journal, Series 3: 35,535-546 (1992)
Kang, S. C., and Ludema, K. C., "The breaking-in of lubricated surfaces",
Wear 108,375-384 (1986)
Kapoor, A.; Williams, J. A.; Johnson, K. L., Steady state sliding of rough
'I

surfaces", Wear 175,81-92 (1994)


Karami; G., H. P. Evans and R. W. Snide, "Elastohydrodynamic lubrication
of circumferentially finished rollers having sinusoidal roughness." Proc. I. Mech.
E. Part C, Mechanical Engineering Science, 201,29-36, (1987)
Kazamaki, T., "An investigation of air leakage between contact surfaces.
(3rd report, in which iron and brass were used as specimen)", Bull. J.S.M.E., 17,
1321-1331 (1974).
King, T. G., Watson, W., Stout, K. J., "Modelling the microgeometry of
lubricated wear", Proc. 4th. Leeds-Lyon Symp., 333-343 (MEP, London, 1978)
Koura, M. M., "The effect of surface texture on friction mechanisms", Wear,
63, 1-12 (1980).
Koura, M. M. and M. A. Omar, "The effect of surface parameters on
friction", Wear, 73, 235-246 (1981)
Kumar, S., "Stochastic models with variable viscosity for hydrodynamic
lubrication of rough surfaces", Wear, 62, 329-336 (1980).
242 Rough Sudaces

Kweh; C. C., H. P. Evans and R. W. Snidle, "Micro-elastohydrodynamlc


lubrication of an elliptical contact with transverse and three-dimensional sinusoidal
roughness." Journal of Tribology, 111,577-584, (1989)
Kweh, C. C., Patching, M. J., Evans, H. P. and Snidle, R. W., "Simulation of
elastohydrodynamic contacts between rough surfaces", J. Trib., ASME, 114, 412-
419 (1992)
Leaver, R. H., Sayles, R. S. and Thomas, T. R., "Mixed lubrication and
surface topography of rolling contacts", Proc. I. Mech. E., 188,461-469 (1974).
Lee, S. C., and Ren, N., "Behavior of elastic-plastic rough surface contacts as
affected by surface topography, load, and material hardness", Trib. Trans. 39, 67-
74 (1996)
Liu, K., Liu, X. J., Xie, Y. B., "Definition of the transition from fluid
lubrication to mixed lubrication for rough surface", Lubrication Science 8, 287-295
(1996)
Lloyd, D. G., and Stevenson, M. G., "An investigation of floor surface profile
characteristics that will reduce the incidence of slips and falls", Mech. Eng. Trans:
Inst. Engrs. Australia ME17, 99-105 (1992)
Lubrecht; A. A., W. E. ten Nape1 and R. Bosma, "The influence of
longitudinal and transverse roughness on the elastohydrodynamic lubrication of
circular contacts."Journal of Tribology, 110, 421-426, (1988)
Lucas, V., Bonneau, O., Frene, J., "Roughness influence on turbulent flow
through annular seals including inertia effects", ASME Paper 95-TRIB-11 (1995)
Lucas, V., Danaila, S., BOMeaU, O., Frene, J., "Roughness influence on
turbulent flow through annular seals", Journal of Tribology, Transactions of the
ASME 116,321-329 (1994)
Manning, D. P., Jones, C., Bruce, M., "A method of ranking the grip of
industrial footwear on water wet, oily and icy surfaces", Safety Science 14, 1-12
(1991)
Marcus, A. H., "Statistical model of a flooded random surface and
applications to lunar terrain", J. Geophysical Rex, 72, 1721-1726 (1967).
Matsuzaki, Y., Funabashi, K., Hosokawa, K., "Effect of surface roughness on
contact pressure of static seals. (Effect of tangential force on conical inside-seal
surface)", JSME International Journal, Series 3: 36, 119-124 (1993)
Matsuzaki, Y., Hosokawa, K., Funabashi, K., "Effect of surface roughness on
contact pressure of static seals", JSME International Journal, Series 3: 35,470-476
(1992)
Tribologv 243

Merriman, T. and J. Kannel, "Analysis of the role of surface roughness on


contact stresses between elastic cylinders with and without soft surface coating. "
Journal ofTribology, 111, 87-94, (1989)
Mitchell, L. A. and Rowe, M. D., "An assessment of face seal performance
based on the parameters of a statistical representation of surface roughness", Proc.
I. Mech. E., 182, Part 3K, 101-107 (1967/68).
Moalic; H., J. A. Fitzpatrick and A. A. Torrance, "A spectral approach to the
analysis of rough surfaces", Journal of Tribology, 111, 359-363 (1989)
Myers, N. O., "Characterization of surface roughness". W7ear, 5 , 182-189
(1962).
Nakahara, T., M. Takesue and H. Aoki, "Effects of surface roughness and
bearing slenderness ratio on hydrodynamic lubrication." Journal JSLE Int Ed,
No.5, 65-70, (1984)
Nivatvongs, K., Cheng, H. S., Ovaert, T. C. and Wilson, W. R. D., "Influence
of surface topography on low-speed asperity lubrication breakdown and scuffing",
Wear, 143, 137-148 (1991)
Ogilvy, J. A,, "Predicting the friction and durability of MoS2 coatings using a
numerical contact model", Wear 160,171-180 (1993)
Ogilvy, J. A,, "Numerical simulation of friction between contacting rough
surfaces", Journal of Physics D (Applied Physics) 24, 2098-2 109 (1991)
Ostvik, R. and Christensen, H., "Changes in surface topography with
running-in", Proc. I. Mech. E., 183, part 3P, 57-65 (1968/69).
Otto, D. L., "Triangular asperities control seal leakage and lubrication",
S.A.E. Paper No. 740201, (1974).
Patir, M., Cheng, H. S., "Application of average flow model to lubrication
between rough sliding surfaces", J. Lubr. Technol. Trans. ASUE 101, 220-230
(1979).
Peeken, H.; Ayanoglu, P.; Knoll, G.; Welsch, G., "Measurement of
lubricating film thickness, temperature and pressure in gear contacts with surface
topography as a parameter", Lubrication Science 3, 33-42 (1990)
Poon, C. Y., and Bhushan, B., "Numerical contact and stiction analyses of
Gaussian isotropic surfaces for magnetic head slideddisk contact", Wear 202, 68-
82 (1996)
Poon, C. Y., and Bhushan, B., "Nano-asperity contact analysis and surface
optimisation for magnetic head slideddisk contact", Wear 202, 83-98 (1996)
Prakash, J.; Tonder, K.; Christensen, H., "Micropolarity roughness
interaction in hydrodynamic lubrication", ASME Paper 79-Lub-8 (1979).
244 Rough SurJaces

Proctor, T. D., "Slipping accidents in Great Britain - an update", Safety


Science 16, 367-377 (1993)
Proctor, T. D., and Coleman, V., "Slipping, tripping and falling accidents in
Great Britain - present and future", J. Occup. Accid. 9, 269-285 (1988)
Fbdhakrishnan, V., "The application of correlation functions in wear
measurements", Wear 41, 169-177 (1977).
Rao, A. Ramamohana; Mohanram, P. V., "Study of mixed lubrication
parameters of journal bearings", Wear 160, 111-118 (1993)
Rosen, B.-G., R. Ohlsson and T. R. Thomas, "Nano metrology of cylinder
bore wear", Trans. 7th. Int. ConJ On Metrology & Properties of Engng Surfaces,
102-110 (Goteborg, 1997)
Roth, A., "The interface-contact vacuum sealing processes", J. Vacuum Sci.
& Technol. 9, 14-23 (1971).
Roth, A,, Vacuum sealing techniques (Pergamon, Oxford, 1966)
Sadeghi, F. and P. C. Sui, "Compressible elastohydrodynamlc lubrication of
rough surfaces." Journal of Tribology, 111, 56-62, (1989)
Shukla, J. B. and S. Kumar, "Effects of viscosity variation and surface
roughness in the lubrication of a slider bearing", Wear, 52, 235-242 (1979)
Sinha; P., J. S. Kennedy and C. M. Rodkiewicz, "Effects of surface
roughness- and additives in lubrication: generalised Reynolds equation and its
application to elastohydrodynamic film." Proc. I. Mech. E. Part C, Mech. Eng.
Sci., 201, 1-9, (1987)
Sugimara, J., and Yamamoto, Y., "Hydrodynamic lubrication of self-affhe
fractal surfaces", Trans. JSME 61C, 475 1-4756 (1 995)
Summers-Smith, D., An introduction to tribology in industry (Machinery
Publishing Co., London, 1969)
Tallian, T. E., McCool, J. I. and Sibley, L. B., "Partial elastohydrodynamic
lubrication in rolling contact", Proc. I. Mech. E., 180, Part 3 8 , 169-84 (1965/66).
Taneeranonon, P. and W. 0. Yandell, "Microtexture roughness effect on
predicted road-tyre friction in wet conditions", Wear, 69, 321-337( 1981).
Thomas, T. R., "Computer simulation of wear", Wear, 22, 83-90 (1972).
Thomas, T. R., "The characterisation of changes in surface topography
during running-in", Proc. 4th Leeds-Lyon Symp., 99-108 (MEP, London, 1978)
Thomas, T. R., Holmes, C. F., McAdams, H. T. and Bernard, J. C., "Surface
features influencing the effectiveness of lip seals: a pattern - recognition approach",
S.M.E. Paper 1475-128 (1975a).
Tribologv 245

Thomas, T. R., Holmes, C. F., McAdams, H. T. andBernard, J. C., "Surface


microgeometry of lip seals related to their performance", Proc. 7th. Int. Con$ on
Fluid Sealing, Paper J32 (BHRA,Cranfield, 1975b)
Tonder, K. and Christensen, H., "Waviness and roughness in hydrodynamic
lubrication", Proc. I. Mech. E., 186, 807-812 (1972).
Tonder, K., "Effects of skew unidirectional striated roughness on
hydrodynamic lubrication. Part 2- Moving roughness." Journal of Tribology, 109,
671-678, (1987)
Tonder, K., "Numerical investigation of the lubrication of doubly periodic
unit roughness", Wear, 64, 1-14 (1 980).
Tzeng, S. T. and Saibel, E., "Surface roughness effect on slider bearing
lubrication", A.S.L.E. Trans., 10,334-338 (1967)
Van Bavel, P. G. M.; Ruijl, T. A. M.; Van Leeuwen, H. J.; Muijdennan, E.
A., "Upstream pumping of radial lip seals by tangentially deforming, rough seal
surfaces", ASME Paper 95-TRIB-13 (1995)
Wakuri, Y., Hamatake, T., Soejima, M., Kitahara, T., "Study on the mixed
lubrication of piston rings in internal combustion engine", Nippon Kikai Gakkai
Ronbunshu, 61C, 1123-1128(1995)
Wallach, J., Hawley, J. K., Moore, H. B., Rathben, F. V. and Gitzendanner,
L. G., "Calculation of leakage between metallic sealing surfaces", A.S.M.E. Paper
68-LUB-15, (1968).
Warren; W. E., J. G. Curro and D. E. Amos, "On the nature of O-rings in
contact with rough surfaces. Journal ofTribology, 110,632-637, (1988)
Is

Willn, J. E., "Characterisation of cylinder bore surface finish - A review of


profile analysis", Wear, 19, 143-62 (1972).
Yanagisawa, T., Sanada, M., Koga, T., Hirabayashi, H., "Influence of
designing factors on the sealing performance of C-seal", ME Trans. 100, 651-657
(1991)
Yu, M. M., and Bhushan, B., "Contact analysis of threedmensional rough
surfaces under frictionless and frictional contact", Wear 200,265-280 (1996)
CHAPTER 12

SOME OTHER APPLICATIONS

So far we have had the opportunity to discuss the application of surface roughness
studies to contact mechanics in some detail, and we have also been able to outline
briefly some of the more important areas of application in tribology. Regrettably, it
is not possible within the scope of tlus book to cover the whole range of the other
engineering and scientific effects of roughness. In this concluding chapter we will
select a few of these applications, chosen at least partly to illustrate our earlier
arguments about functional filtering. Contact resistance is interesting to electrical
engineers and also important in heat transfer studies. The effect of roughness on
fluid flow is appreciated in many technical fields, from aero- and hydrodynamics to
chemical engineering, and also to workers in the earth sciences. Machining
problems are the domain of manufacturing and production engineers, and
Qmension and tolerance are also important to quality control personnel and
inspectors. Finally we conclude by touching briefly on two topics of interest to
their eponymous communities, bioengineering and geomorphology.

12.1. Contact Resistance

In many engineering applications involving contact it is sufTicient to know the real


contact area. However, an important practical class of problems where a more
detailed knowledge is necessary is that of thermal or electrical contact resistance.
When two surfaces are in contact and electricity or conductive heat passes between
them, it does so through the discrete areas of contact. The lines of equal potential,
parallel at a distance from the interface, become increasingly distorted as the
contact zone is approached, and the flow lines bunch together to pass through the
individual contacts (Fig. 12.1). Holm (1958) has shown that the conductance of a
single contact spot is proportional not to its area but to its radius. The conductance
per unit nominal area of the interface is

c, = 2Kn t/a (12.1)

247
248 Rough Sursaces

where a is the mean contact spot area, n is their number per unit area and K is the
thermal or electrical conductivity. It is therefore necessary to determine the
variation with load both of the number of contact spots and of their individual size.

Figurel2.1. Contact resistance due to constriction of flow lines

Electrical conductivity is of practical importance in the electronics and


electrical power industries, and the effect of roughness has received some attention
(Barkan & Tuohy 1965, Harada & Mano 1968, Hisakado 1977, Lanchon et al.
1986, Bryant 1994, Clausen & Leistiko 1995). The economic importance of
thermal contact resistance is also considerable (Thomas & Probert 1972). In the
form of the resistance between the fuel element and its container, it affects the
economics of nuclear power (Boeschoten & van der Held 1$57, Tillack & Abelson
1995). In space technology also, the electronic equipment inside a satellite or
space vehicle generates heat which can only be dissipated by solid conduction to
the outside (Chung et al. 1993, Chung 1995, Chung & Shefield 1995).
The property of the contiguous solids most difficult to define quantitatively is
the topography of their contacting surfaces (Yip & Venart 1967/68, Cooper et al.
1969, Mikic 1974, Thomas 1982, Majumdar & Tien 1991, McWaid & Marschal
1992). Not only the surface roughness is involved but its waviness also, and a
number of workers have considered the combined resistances of both (OCallaghan
et al. 1989, Lambert & Fletcher 1995, Torii & Nishino 1995). It seems generally
to be felt that the resistances are additive, so that the total resistance of the
interface is the sum of a macroscopic resistance due to wavkiess and a microscopic
resistance due to roughness. The microscopic resistance is due to the convergence
of lines of flow through individual microscopic contact areas which are due to the
Other Applications 249

surface roughness. Clusters of these spots are held to be contained within larger
macroscopic contour areas due to waviness.
The effective of waviness on thermal contact resistance was investigated
(Thomas & Sayles 1975) by considering the behaviour of the interface at very light
loads W when it could be treated as a three-point Hertzian elastic contact. Here the
high-pass cut-off was again a hnction of the diameter of the bar, and a low-pass
cut-off was sought such that the enclosed bandwidth would contain only three
asperities. This was found by substituting the appropriate asperity density into the
integrated moment equation and solving the resulting transcendental equation
iteratively.
The strategy was as follows. One starts by assuming a form for the spectrum
of wavelengths and goes on to define a pass-band representing waviness. The
long-wavelength limit of this band must be set by the size A of the nominal contact
area; the low-pass cut-off is found in terms of the contact size from the initial
contour area density. Knowledge of the bandwidth permits calculation of the mean
height and radius of curvature of the contacting asperities. The usual Hertzian
formula will then give the area of real contact in terms of the nominal contact size
and the total roughness at that size. This total roughness may be found from the
measured roughness, leading to a figure of merit, sensitive to the effect of waviness
on resistance, which is a function of readily measurable parameters.

Figure 12.2. Thermal contact resistance:experimental results of Fried (1965) replotted as a fhnction of
waviness number <(Thomas & Sayles 1975)

The waviness number < = W/E’cr& so defined has a number of properties


whlch make it suitable as a figure of merit for determining contact mechanism.
Although it cannot at this stage be used to describe the lower limit of microscopic
250 Rough Sudaces

resistance, it has been argued, with some experimental sup?ort, that the effect of
waviness can be neglected for 6 > 1. Its predictions are qualitatively reasonable;
for a high load on a small smooth contact of low Hertzian modulus, jwill be large
and the waviness will flatten to let the microscopic resistance predominate; for a
low load on a large rough contact of high elastic modulus, cwill be small and the
macroscopic resistance will predominate. One set of experimental results replotted
in terms of waviness number can be seen in Fig. 12.2. The sharp dog-leg
indicating a change of regime occurs near <= 1 as predicted.
It turns out to be very difficult to find experimental data in the literature for
which this condition is satisfied. The implication is that for the range of conditions
of engineering interest the effect of waviness can never be neglected. There is
some support for thw conclusion. When resistance is entitely due to waviness it
can be inferred that the total resistance of the interface is proportional to W o 3 3 .
From results on surfaces specially prepared with roughness but no waviness, it is
known that when resistance is entirely due to roughness the exponent of W is
between -0.95 and -0.99 (Thomas & Probert 1970). However, for the vast majority
of results in the literature the exponent lies between these extreme values at about -
0.73 (Thomas & Probert 1972). It seems likely, therefore, that in most practical
situations thermal contact resistance is due to the combined effects of waviness and
roughness.
The case where microscopic resistance only is present is rather more difficult
to tackle if plastic contact is assumed. The reason is that there are then no grounds
for truncating the power spectrum at a short-wavelength limit and consequently the
higher moments are infinite. Lf elastic contact is assumed then we can again use
the argument of repeated contact to impose a low-pass filter based on the plasticity-
index criterion. This allows us to find an upper-bound solution for the average
contact spot size and an exact solution for the number of contacts, and hence an
upper-bound solution for the conductance (Sayles & Thomas 1976a). All these are
in terms of the separation of the contacting surfaces, but as we also know the
variation of load with separation we can plot all three in terms of load (Fig. 10.2).

12.2. Noise and Vibration

Small surface irregularities can give rise to noise and vibration in rolling contact
(Nayak 1972, Ananthapadmanaban & Radhakrishnan 1982, Hess & Soom 1991).
This causes problems in rolling bearings (Yhland 1967/8, Kanai et al. 1987),
railroad operation (Gray & Johnson 1971, Thompson 1993a, b) and transmission
Other Applicatiom 25 I

design (Mengen & Weck 1992, Aziz & Seireg 1994). The dynamic characteristics
of rolling element bearings (Poon & Wardle 1978) are such that an extremely high
number of resonances exist just in the bearing itself. When housing alignment,
geometry and the general bearing environment are considered the problem
becomes more complex, and a bearing design which would avoid or attenuate
vibration seems very unlikely. The problem is best solved by removing or reducing
the source of vibration, namely the wavelengths of the specific surface features
which are responsible for the input of vibration energy. In achieving this it is first
necessaq to identi@ these wavelengths and under what circumstances they can
produce vibration (Su et al. 1993), and secondly to establish how they can be
effectively monitored and removed in a production environmat.
The input of vibration energy is due to random surface interaction occurring
predominantly at the rolling-element interfaces. In some cases a significant energy
input can be supplied by ball-cage interaction, macroslip and spin, but at
conventional speeds and reasonable loads these effects are of secondary
importance. By considering the possible partial and full-film lubrication
conditions, and the physical interpretation of these mechanisms in terms of surface
contact and the interchange of energy, the surface interaction can be split into three
distinct mechanisms: rolling response due to form and waviness effects, shock
noise due to local elastic deformation, and impulsive shocks due to asperity
collisions and debris.
Having established the size of asperities which can Aastically conform, it
becomes possible to determine their density and therefore frequency of interaction
(Sayles & Poon 1981). This can be easily accomplished through Eqn. 9.17,
modfied to take account of anstropy. It is achieved by a circumferential and an
across-lay measurement from the raceway components. From such measurements
on all the conventional finishing processes it appears that this form of vibration
generation predominates over the others and can affect all frequencies, although
the most severe effects are apparent in the range of about 300-10,000 Hz.

12.3. Fluid Flow

Roughness is of interest and importance in a number of applications involving


fluid flow. Flow through a rough gap or between rough planes, for instance, is a
phenomenon affecting the performance of air bearings (Lau & Harman 1975) and
of static seals such as vacuum seals (Thomas 1973a). The elemental mass flow
252 Rough Sui$aces

rate dQ through a length dx of a onedimensional rough gap of mean width h is


given by (Thomas & Olszowski 1974)

dQ K (h -z)"dx (12.2)

where z is the instantaneous height of the rough surface above its mean plane and n
depends on the flow regime. As z is not an analytic hnction of x we cannot
integrate this directly over the length of the gap. However, we can transform the
integration into one over the height probability distribution p(z). The fraction of
the gap length at height z is p(z)dz, hence the mass flow per unit length is

dQ K (h - z)" p(z)dz (12.3)

and the total mass flow rate will be

(12.4)

I
-2 -1 0 I 3
2 t
Figure 12.3. Variation ofdimensionless effective gap width with dimensionless mean plane separation,
accordingto Eqn. 12.5:(1) laminar, n = 3; (2) smooth turbulent, n = 12/7; (3) rough turbulent, n = 3/2
(Thomas& Olszowski 1974)
Other Applications 253

The values of n for laminar and turbulent flow are 3 and 3/2 respectively. If the
height distribution is Gaussian, the integral is closed-form for n integer and can be
evaluated numerically for other n. Jt is convenient to present the result as an
equivalent gap 6 defined such that

(12.5)

The results indicate (Fig. 12.3) that the mean gap "seen" by the fluid diverges
markedly from the real gap as the surfaces approach. This has implications for, for
instance, the design of pneumatic gauges (see Section 4.2.3).

-
2-
1.1

1.0
v
-
=n.Y

-2 n . u
0.7

0.6

0.5

0.4

0.3
0.2
2.6 2.X 3.1) 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4 6 4.8 5.0 5.2 5.4 5.6 5.8 6.0

1% 10 Re
Figure 12.4. Skin fiction as a function of Reynolds number in pipes coated with increasinglylarge grains of
sand; each curve represents grains twice as large as those for the curve below (Hunsaker & Rightmire 1947
after Nikuradse)

An application of even greater economic importance involves the effect of


roughness on skin friction. This has been well known since Nikuradse's classic
experiments on flow in rough pipes (see, for instance, Hunsaker & Rightmire
1947). He roughened a pipe artificialiyby cementing sand grains of similar size on
its interior surface and repeated this with a number of pipes for various sizes of
sand grain. For a smooth pipe the friction factor .ro/pu2,where is the wall shear
and p and tl are the fluid density and velocity respectively, is inversely proportional
254 Rough Surfaces

to the logarithm of the Reynolds number Re, the constant of proportionality


depending on whether flow is laminar or turbulent. For a rough pipe, however, the
friction factor departs from its logarithmic dependence at some value of Re and
tends to a constant value; that is, the shear stress increases as the square of the
velocity. This constant friction factor increases, and the critical Reynolds number
decreases, as the roughness is increased (Fig. 12.4).
This has a profound effect on the performance of ships, where skin friction
can amount to 80-90% of the total resistance to motion (Lackenby 1962). Ships
tend to become rougher in service due to paint breakdown, corrosion and fouling,
and this can necessitate an increase of power of as much as 40 per cent to maintain
the same speed. An increase in roughness of 25 pm increases resistance by about
2%%; to put this in perspective, the roughness of a new hull is on average about
175 pm, at a high-pass cut-off of 50 mm (Lackenby 1962). Measurements on
replicas of ships’ hulls (King et al. 1981) have shown that their height distributions
are reasonably close to Gaussian and their power spectra are of the familiar
inverse-square form; in other words they are mathematically very similar, on a
larger scale, to the machined surfaces discussed throughout this book.
Surprisingly little fundamental research has been carried out on this
phenomenon since Nikuradse’s original work. The qllalitative explanation
generally offered for the roughness resistance is that extra turbulence is caused by
the projection of the roughness elements through the laminar sublayer immediately
adjacent to the solid surface; if the height of the roughness elements is small
relative to the thickness of the sublayer then the surface will behave as though it
were hydraulically smooth, and the transition to rough behaviour is due to the
decrease of thickness of the laminar sublayer with increasing Reynolds number. It
is usual to express the results of any fluid measurement involving a rough surface
in terms of an equivalent sand-grain number, a descriptive approach which is not
altogether satisfactory in modem terms.
Two points are apparent in the light of today‘s knowledge of surfaces. The
first is that most surfaces generally have much gentler slopes than abrasive
surfaces, and the abrupt changes in height between adjacent grains may themselves
be likely to induce a special sort of turbulence not present for surfaces of a different
texture. The second is that real surfaces are not composed of roughness elements
of the same geometric form and at the same height. A Gaussian height distribution
would imply a gradual increase in the number of summits penetrating the laminar
sublayer as its thickness decreased, resulting in a transition region between “fully
smooth and “fully rough regimes, and indeed such a transition region is apparent
(Fig. 12.4). This transition regon is much wider for “natural” surfaces than for
Other Applications 255

Nikuradse’s pipes where no doubt an attempt was made to level the grains to the
same height.
The actual process of momentum transfer from a fluid to a rough surface does
not seem to have been at all widely studied. A theoretical treatment in terms of the
power spectrum of surface wavelengths (Singh & Lumley 1971) yielded predictions
at variance with experiment. One might expect a priori that both the roughness
and the slope would exert an important effect, and also thaf as usual it should be
possible to isolate a pass-band of surface wavelengths which affect skin friction.
What is the shortest wavelength, for instance, to which the laminar sublayer will
conform? Clearly, wavelengths longer than thw will play no part in roughness
effects. A correlation is reported between increased power requirement and mean
apparent amplitude measured at 50 mm cut-off, but this cut-off is admittedly
arbitrary (Lackenby 1962). An apparent peak in the surface power spectrum
around 50 mm wavelength (Chaplin 1967) is probably an artefact of computation.

12.4. Dimension and Tolerance

A relationship between surface finish and dimensional tolerance has long been
suggested (Schlesinger 1944, Ber & Yarnitzky 1968, Bryan & Lindberg 1973,
Osanna 1979). It has also been well known for many years that the tolerance of
machned components must be increased with component size. The reasons for
k s are not clearly defined, although in BS 4500: 1969 it is stated that to some
extent this can be explained by an increased difficulty in manufacture. Thermal
expansion effects must also be considered, as these increase with component size
but are generally small in relation to tolerance with components below 500 mm
diameter. BS 4500 gives empirical rules which relate tolerance to diameter which
are supposedly derived from “extensive practical investigations”, but the
fundamental reasons why this should be so are still somewhat obscure. As an
example, a 3 cm diameter shaft, produced on say a lathe, would be associated with
tolerance under the IT10 grade of 40 pm, whereas the same machine and tolerance
grade allows a 25 cm shaft 84 pm.
Section 1.6 of BS 4500 suggests that geometry, form and surface texture must
also be considered in some circumstances. In fact this is the principal reason why
tolerance is necessary. The increase in Wiculties in manufacture with component
size stated as a reason by BS 4500 is simply another way of stating the arguments
leading up to the non-stationarity result of Section 8.4, in that as the size increases
the number of potential surface geometric errors also increases.
256 Rough Surfaces

Fig. 12.5 shows the fractal topography-length law (Sayles & Thomas 1978)
plotted against the empirical equations governing tolerance grades IT6-16 of BS
4500 up to 500 mm. The parameters of the fractal equation are chosen so that the
curves overlap; however, the values are typical of an average ground surface. The
fact that the curves do overlap is unimportant other than it provides the best way of
comparing the trends. The agreement in trend is surprisingly good, even to the
extent of anticipating tolerance requirements above 500 mm, and given by a
separate equation in BS 4500 which is shown by the dashed line. The figure
demonstrates that for a given tolerance grade, to maintain a constant grade of fit
we are increasing the tolerance in the same way as the surface topography is
changing. In other words we are maintaining the same relative surface interface
conditions.

Figure 12.5. Tolerance Ias a function ofworkpiece diameter D. Dashed line:I = 0.4501'3+ 0.0010 (D 5
500 m)(BS 4500: 1969); dotted line: I = 0.0040 + 2.1 (D > 500 m)(BS 4500: 1969); solid line: I =
0.4 + $10" x D ) (Sayles &Thomas 1978)

From a production engineering point of view the fractal relationship can be


considered at two different levels. Firstly it gives us an insight into the way in
which the increase in tolerance is linked to size; an empirical fact whch has long
been established and taken for granted. Secondly, and on a more practical level, it
can provide the production engineer with a means of monitoring the condition of a
machine. We know that machines producing the same components can generate
differing values of fractal parameters, a good measure of machine condition and
environment. We have shown how the fractal parameters can be related to a class
of tolerance; thus it seems possible to classlfy a machme quality in terms of its
Other Applications 257

ability to produce components within a given tolerance grade. Periodic checks on


the fractal parameters of surfaces being produced would also act as a good
indication of the potential useful life of the machine.

12.5. Abrasive Machining

Ground surfaces have a welldefined lay owing to the parallel orientation of


grinding scratches, which are much longer in the direction of rotation of the
grinding wheel than they are wide. It is of interest in the study of grinding
processes to determine the average dimensions of a grinding scratch. Thts is not
readily done by direct observation, as the scratches are superimposed on each other
to a confusing degree. A possible solution is to examine profiles taken at various
angles to the lay; it can plausibly be shown that the correlation length should
represent the average dimension of a scratch in that direction (Fig. 12.6). There
can be seen a marked peak in the direction of the lay; the average length and
breadth of a grinding scratch in this case were deduced a$ 252 pm and 34 pm
respectively (Thomas 1973b).

-E
120 '
0
0
0
0
3 IOU 'B
000

I I I I I 1 I I I I I
0 80 7 0 6 0 5 0 4 0 M 20 10 0 - 1 0

Angle to lay (degrees)

Figure 12.6.Correlation length as a function of angle of stylus traverse kom the lay of a ground surface
(Thomas1973b)

The height distribution of a ground surface is often cited as Gaussian and a


practical example of the central limit theorem. However, Sayles & Thomas (1979)
found a highly sigruficant negative skewness in samples of several hundred
thousand height readings from a number of ground surfaces (Fig. 7.11). The
258 Rough Sur$aces

distributions were truncated at their upper ends, in effect possessing fewer high
peaks than a normal distribution. A study of the literature revealed that a similar
skewness is in fact almost always present in height distributions from simple
abrasive processes, and is often present to a lesser extent in ground surfaces.
Kapteyn (as quoted in Hald 1960) developed a statistical theory of skew
distributions in terms of a function which can be interpreted as the effect of the
abrasion or grinding mechanism at the interface. This mechanism has been the
subject of much dxussion, but it has generally been established that a combination
of conventional cutting, ploughing and plastic deformation exists. Such a
mechanism suggests that the geometry of the abrasive surface imposes itself on the
machined surface, irrespective of the metal removal or deformation mechanism
involved at each individual grain. If this is so, then where light cuts are involved
and several passes are made without increasing the feed, Kapteyn's theory would
predict a Gaussian height distribution; conversely, with heavy cuts and single
passes a truncated distribution would result.
Distributions of both types are reported in the litenture; it is suggested,
however, that the strict Gaussian distribution, heretofore accepted as the norm, is
an artefact of the care taken to obtain a specimen, and that on the production line,
where single-pass and plunge grinding are extensively employed, the height
distributions created on many engineering components are negatively skewed. If
so, this may have important practical implications: for instance, the degree of
truncation of the height distribution has been found to influence the running-in of
cylinder bores (Campbell 1972), and it might also be supposed that in a seal with a
ground surface the sealing action would be affected by the number of high peaks on
the surface.
A detailed theoretical consideration of the grinding process (Sayles &
Thomas 1976b) enabled a prediction of an effective height distribution for a
grinding wheel which showed a strong positive skewness. This will produce its
mirror image, a negatively skewed height dwtribution, on the ground surface. The
exact form of the distribution depends on the number of effective profiles n on the
grinding wheel which intersect with the surface; for small n the distribution is
closely Gaussian, but becomes increasing skewed as n increases (Fig. 12.7a). The
parameter n is a function of wheel and work speed, number of passes and wheel
specification; the ratio of workpiece roughness to grit size can thus be calculated
explicitly (Fig. 12.7b). Roughness is approximately inversely proportional to the
square of the logarithm of n, in accordance with everyday experience that changing
to a finer wheel is an easier way to improve the finish than prolonged machining
with a coarser wheel.
Other Applications 259

-
N
d
c,.
1.0
0.8
0.6
0.4

0.2
0.0
-3.0 -2.0 -1.0 0.0 1.0 2.0 3.0 0.4 5.0
z/R9

0.04

: : : 0.m
0.0010' 102 103 104 105 106

Figure 12.7.(above) Height distributionsofthe envelopes of n effective abrasive profiles, assuming a normal
distribution for a single profile; (below) variation of workpiece roughness, normalised with respect to grain
size r, with number of effective abrasiveprofiles: circles are experimental, solid line is theory (Sayles &
Thomas 1976b)

12.6. Bioengineering

In human joints the bone ends are separated by a soft porous tissue known as
articular cartilage. The entire joint is enclosed in a sealed capsule containing a
lubricating medium called synovial fluid (Fig. 12.8). Considered as a bearing,
then, the joint behaves as two porous compliant surfaces backed by rigid solids and
separated by a fluid. Roughness of artificial hip joints is known to be associate
with increased wear perbyshire et al. 1994).
The mechanism of lubrication of this bearing is far from clear, but its
roughness is believed to be an important factor (Tandon & Rakesh 1981).
260 Rough Sur$aces

Hydrodynamic, elastohydrodynamic and boundary regimes have all been


postulated, but there are experimental objections to all of them. One theory of
"boosted lubrication" (Longfield et al. 1969) relied on the observation that articular
cartilage is rough. It was suggested that lubrication under light or moderate loads
is normally hydrodynamic. Under high-impact loads, as when walking or
jumping, the sudden approach of the surfaces squeezes the watery components of
the synovial fluid into the pores of the cartilage. The larger molecules which are
left are trapped in the pools formed by the interlocking asperities and act as a
boundary lubricant

Spmd membne

Artmrlar cartilage

Bone

Figure 12.8. Schematic of a human joint (Longtield et al. 1969)

The problem of direct measurement of a yielding surface with a stylus


instrument was discussed in Section 2.3.3. It turns out that the surface is highly
irregular and that the distribution of heights is quite closely Gaussian (Fig. 12.9).
A hip joint is basically a ball-and-socket joint in engineering terms, and
calculations can be carried out on it in exactly the same way as for other
engineering contacts (Thomas et al. 1980). Assuming a 40 mm diameter for the
femoral head the high-pass cut-off becomes 20 mm and from the cartilage
measurements the effective roughness was 20 pm. The plasticity approach gave a
low-pass cut-off of 4.1 pm. It is of some medical interest to know the
configuration of the joint under a high transient load, for instance at heel-strike in
the normal walking cycle. This can be deduced in some detail from the surface
measurements and material properties, using now the elastic theory of Eqns. 10.14-
10.16 which requires only the first two even moments of the profile power
spectrum-
Other Applications 26 1

It turns out that the real area of elastic contact calculated under these
conditions is 104 mm2, the mean plane separation is 58 pm, the real contact
pressure is 24 N/mm2 and the stlffness is 185 kN/mm, more than two orders of
magmtude less than that of the machine-tool interface quoted previously. None of
these predictions is in serious conflict with existing experimental measurements.
The volume enclosed between the cartilage surfaces at heel strike can be calculated
to be about two-thirds of the volume when standing still, thus lending support to
the theory of boosted lubrication of human joints.

lo p m
I H
I mm

Height above mean line ( g m )

Figure 12.9. (a) Profile of surface of human articularcartilage: (b) distribution of profile heights; broken line
is Gaussian distribution with the same standard deviation (Thomas et al. 1980)

Roughness also affects the assimilation to the body of surgical implants and
prostheses of various kinds. Where these are inserted into bone, it is essential that
the bond between the implant and the living tissue should possess mechanical
strength. To ensure this, individual cells must adhere to the surface of the implant
itself or to its fasteners, and the local roughness is an important determinant of this
adhesion. Wennerberg (1996) reviews the extensive literature on the effects of
implant roughness with more than 200 references.

12.7. Geomorphometry

Geomorphometry has been defined as the science which treats of the geometry of
landscape, and plays a vital role in both military and civil engineering (Bekker
262 Rough Su$aces

1969, Mitchell 1991). Geomorphometry is a recognized sub-field within geology,


geography, geomorphology, hydrology and digital cartography (Thorn 1988,
Richards 1990, Clarke 1990). Its specific applications range from measuring
highway roughness (Hegmon 1979, Xu et al. 1992, al-Mansour et al. 1994),
mapping sea-floor terrain (Hennings et al. 1994), and assessing soil erosion (Hagen
& Armbrust 1992, 1994, Govindaraju & Kawas 1994) to meteorology (Wieringa
1992, Roberts et al. 1994, Hignett & Hopwood 1994), and analysing wildfire
propagation (Kasischke et al. 1994). Recent technical advances have presented
information on geographical relief as Cartesian data sets similar to those acquired
by scanning microscopes and similar instruments, though of course on a much
larger scale. Such data sets are referred to in the literature as digital elevation
models OEMs). The field is reviewed by Plke (1995a), who has also published a
bibliography of the geomorphometric literature (Pike 1993) which with
supplements (Pike 1995b, 1996) runs to more than 3000 entries.

PUR
10 1

0 I
0 50 100 150 200
Roughness (dkm)

Figurel2.10. Ride quality (Pavement User Rating) as a function ofzero crossing density (Potter et al. 1992)

The measurement of terrain roughness is an essential preliminary to the study


of vehicle dynamics (Van Deusen 1967). Levels of vibration in truck shipments
(Marcondes & Singh 1992), dynamic pavement loadings (Gyenes et al. 1992, Lin
et al. 1994) and bridge loadings (Hung et a1 1992) can be predicted from road
roughness. Subjective perception of ride quality also appears to depend on
pavement roughness (Potter et al. 1992, Gerardi & Schmerl 1995); comfort
decreases as the average wavelength shortens (Fig. 12.10). Wambold & Henry
(1982) review techniques for measuring road roughness with nearly 40 references.
Other Applications 263

12.8. References

Ananthapadmanaban, T. and Radhakrishnan, V., "An investigation on the


role of surface irregularities in the noise spectrum of rolling and sliding contacts",
Wear 83, 399-409 (1982)
Aziz, S. M. A,; Seireg, A,, "Parametric study of frictional noise in gears",
Wear 176,2528 (1994)
Barkan, P. and Tuohy, E. J., "A contact resistance theory for rough
hemispherical silver contacts in air and in vacuum", Trans. I.E.E.E. PAS-84,
1132-1143 (1965).
Bauer, T. W., Taylor, S. K., Jiang, M., Medendorp, S. V., "An indirect
comparison of third-body wear in retrieved hydroxyapatite-coated, porous and
cemented femoral components", Clinical Orthopaedics & Related Research 298,
11-18 (1994)
Bekker, M. O., Introduction to Terrain-Vehicle Systems (Univ. Michigan
Press, Ann Arbor. 1969)
Ber, A. and Yarnitzky, Y., "Functional relationship between tolerances and
surface finish", Microtecnic, 22, 449-45 1 (1968).
Boeschoten, F. and Van der Held, E. F. M., "The thermal conductance of
contacts between aluminium and other metals", Physica 23, 37-44 (1957)
Bryan, J. and Lindberg, E., "Relationship of surface finish to dimensional
tolerance", Proc. Int. Con$ on Surface Technol., Pittsburgh, 117-130, (S.M.E.,
1973).
Bryant, M. D., "Resistance buildup in electrical connectors due to fretting
corrosion of rough surfaces", IEEE Transactions on Components, Packaging, and
Manufacturing Technology 17A, 86-95 (1994)
Campbell, J. C., "Cylinder bore surface roughness in internal combustion
engines: its appreciation and control", Wear, 19, 163-168 (1972).
Chaplin, P. D., "The analysis of hull surface roughness records", European
Shipbuilding, 16, 40-47 (1967).
Chung, K. C.; SheEeld, J. W.; Sauer, H. J. Jr.,; O'Keefe, T. J.; Williams, A,,
"Thermal contact conductance of a phase-mixed coating layer by transitional
buffering interface", Journal of Thermophysics and Hear Transfer 7, 326-333
(1993)
Chung, K. -C., "Experimental study on the effect of metallic-coated junctions
on thermal contact conductance", JSME International Journal, Series B: Fluids
and Thermal Engineering 38, 100-107 (1995)
264 Rough Sursaces

Chung, K.-C., Sheffield, J. W., "Enhancement of thermal contact


conductance of coated junctions", Journal of Thermophysics and Heat Transfer 9,
329-334 (1995)
Clausen, T., Leistiko, O., "Surface roughness and specific contact resistance
of AuGeNflnP ohmic contacts", Materials Research Society Symposium
Proceedings 355,389-394 (Materials Research Society, Pittsburgh, 1995)
Cooper, M. G., Mikic, B. B., and Yovanovich, M. H., "Thermal contact
conductance", Int. J. Heat Mass Transfer 12, 279-300 (1969).
Derbyshire, B.; Fisher, J.; Dowson, D.; Hardaker, C.; Brummitt, K.,
"Comparative study of the wear of UHMWPE with zirconia ceramic and stainless
steel femoral heads in artificial hip joints", Medical Engineering & Physics 16,
229-236 (1994)
Dowson, D., El-Hady Diab, M M., Gillis, B. J., Atkinson, J. R., %fluence of
counterface topography on the wear of ultra high molecular weight polyethylene
under wet or dry conditions", in Lee ed., Polymer wear and its control 171-187
(Amer. Chem. Soc.,St. Louis, 1985)
Fried, E., "Study of interface thermal contact conductance", General Electric
Co. Report 65SD4395 (1965)
Gerardi, T., Schmerl, H., "Ride quality as a part of airport pavement
management systems", Transportation Congress, Proceedings 1, 588-599 (ASCE,
New York, 1995)
Gray, G. G. and Johnson, K. L., "The dynamic response of elastic bodies in
rolling contact to random roughness of their surfaces", J. Sound & Vibration, 22,
323-42 (1972).
Gyenes, L.; Mitchell, C. G. B.; Phlipps, S. D., "Dynamic pavement loads and
tests of road-friendliness for heavy vehicle suspensions", SAE Technical Paper
922464 (1992)
Hald, A., Statistical theory with engineering applications (Wiley, New York,
1960)
Harada, S. and Mano, K., "The effects of surface roughness on contact
resistance of sphere-plane contact", Proc. 4th Int. Res. Svmp. Electric Contact
Phenomena, Chicago, 25-28 (Illinois Inst. of Technol., Chicago, 1968).
Hess, D. P.; Soom, A,, "Normal vibrations and friction under harmonic
loads. 11. Rough planar contacts", Transactions of the ASME. Journal of Tribology
113,87-92 (1991)
Hisakado, T., "Effects of surface roughness and surface films on contact
resistance", Wear, 4, 345-359 (1977).
Holm, R., Electric contacts handbook 3e (Springer-Verlag, Berlin, 1958)
Other Applications 265

Huang, D., Wang, T.-L., Shahawy, M., "Impact analysis of continuous


multigirder bridges due to moving vehicles", Journal of Structural Engineering
118,3427-3443 (1992)
Hunsaker, J. C. and Rightmire, B. G., Engineering applications of fluid
mechanics (McGraw-Hill, New York, 1974)
Kanai; H., M. Abe and K. Kido, "Estimation of the surface roughness on the
race of bails of ball bearings of vibration analysis." Journal of Vibration, Acoustics,
Stress, and Reliability in Design, 109,60-68 (1987)
King, M. J., Chuah, K. B., Olszowski, S. T. and Thomas, T. R., "Roughness
characteristics of plane surfaces based on velocity similarity laws", ASME Paper
81-FE-34 (1981)
Lackenby, H., "The resistance of ships, with special reference to skin friction
and surface condition", Proc. I. Mech. E., 176, 981-1014 (1962).
Lambert, M.A.; Fletcher, L.S., Thermal contact conductance of nonflat,
rough metals in vacuum", ASMEIJSME Thermal Engineering Joint Conference -
Proceedings 1, 3 1-42 (ASME, New York, 1995)
Lanchon; H., B. Makaya; J. Saint Jean Paulin; E. Kroener and A. Mirgaux,
"Mathematical study to obtain qualitative effects of roughness in technical
problems." Wear, 109, 99-111, (1986)
Lau, H. and Harman, C.M., "Externally pressurized compliant air bearing
operating on a rough moving surface", Trans. A.S.M.E.,Ser. F., J. Lubr. Tech., 97,
63-68 (1975).
Lin, W.-K., Chen, Y.-C., Kulakowski, B. T.; Streit, D. A., "Dynamic
wheeVpavement force sensitivity to variations in heavy vehicle parameters, speed
and road roughness", Heavy Vehicle Systems 1, 139-155 (1994)
Lon@ield, M. D., Dowson, D., Walker, P. S., Wright, V., "Boosted
lubrication of human joints by fluid enrichment and entrapment", Biomed. Engng.
4 , 5 17-522 (1969)
Majumdar, A.; Tien, C. L., "Fractal network model for contact conductance",
Transactions of the ASME. Journal ofHeat Transfer 113, 516-525 (1991)
Marcondes, J., Singh, S. P., "Use of road roughness to predict vertical
acceleration in truck shipments", Advances in Electronic Packaging 2, 999-1004
(ASME, New York, 1992)
McWaid, T. H.; Marschall, E., "Application of the modified Greenwood and
Williamson contact model for the prediction of thermal contact resistance", Wear
152,263-277 (1992)
266 Rough Surfaces

Mengen, D.; Weck, M., "How to ensure precise analysis of gear surfaces and
diagnosis of changes during operation", Proc 92 Int Power Transm Gearing Conf
DE43,605-612( ASME, New York, 1992)
Mikic, B. B., "Thermal contact conductance: theoretical considerations", Int.
J. HeatMass Transfer, 17,205-214 (1974).
Nayak, P. R., "Contact vibrations", J. Sound & Vibration, 22, 297-322
(1972).
O'Callaghan, P., Babus'Haq, R. and Probert, S., "Predictions of contact
parameters for thermally-distorted pressed joints", Am. Inst. Aeron. & Astron., 24th
Thermophysics Con$, Paper AIAA-89-1659 (1989)
Osanna, H. P., "Surface roughness and size tolerance", Wear, 57, 227-236
(1979).
Plke, R. J., A bibliography of geomorphometry, with a topical key to the
literature and an introduction to the numerical characterization of topographic
form. - U.S. Geol. Survey Open-file Rept. 93-262-A (1993)
Pike, R. J., "Geomorphometry - progress, practice and prospect", in Pike &
Dikau eds., Advances in geomorphometry, 2. Geomorph. Supplementband 101,
221-238 (1995a)
Pike, R. J., A bibliography ofgeomorphometry, Supplement 1.0. - U.S. Geol.
Survey Open-file Rept. 95-046 (1995b).
Pike, R. J., A bibliography ofgeomorphometry, Supplement 2.0. - U.S. Geol.
Survey Open-file Rept. 96-726 (1996).
Poon, S. Y. and Wardle, F. P., "Running quality of rolling bearings assessed",
Chartered Mechanical Engineer (April 1978)
Potter, D., Hannay, R., Cairney, P., Makarov, A,, "Investigation of car users'
perceptions of the ride quality of roads", Road and Transport Research 1, 6-27
(1992)
Sayles, R. S. & S. Y. Poon, "Surface topography and rolling element
vibration", Precis. Eng 3, 137-144 (1981).
Sayles, R. S. and Thomas, T. R., "Thermal conductance of a rough elastic
contact", Appl. Energy, 2, 249-267 (1976a)
Sayles, R. S., Thomas, T. R., "Stochastic explanation of some structural
properties of a ground surface", Int J Prod Res 14,64 1-655 (1976b)
Sayles, R. S. and Thomas, T. R., "Surface topography as a nonstationary
random process", Nature, 271,43 1-434 (1978)
Sayles, R. S.; Thomas, T. R., "Measurements of the statistical microgeometry
of engineering surfaces", J Lubr Techno1 TransASME 101,409-417 (1979).
Other Applications 267

Schlesinger, G. "Surface finish and the function of parts", Proc. I. Mech. E.,
151, 153-158 (1944)
Singh, K. and Lumley, J. L., "Effect of roughness on the velocity profile of a
laminar boundary layer", Appl. Sci. Res. 24, 168-186 (1971).
Strahler, A. N., "Quantitative geomorphology of drainage basins and channel
networks", in: Chow, V. (ed.)Handbook of applied hydrology 4, 39-76 (McGraw-
Hill, New York, 1964).
Su, Y.-T.; Lin, M.-H.; Lee, M.-S., "Effects of surface irregularities on roller
bearing vibrations", Journal of Sound and Ebration 165,455-466 (1993)
Tandon, P. N. and L. Rakesh, "Effects of cartilage roughness on the
lubrication of human joints", Wear, 70, 29-36 (1981).
Thomas, T. R., "Influence of roughness on the deformation of metal surfaces
in static contact", Proc. 6th Int. ConJ on Fluid Sealing, B3, 33-48 (BHRA,
Cranfeld, 1973a).
Thomas, T. R., "Correlation analysis of the structure of a ground surface",
Proc. 13th Int. Machine Tool Des. &Re x Con$, Manchester, 303-308 (1973b).
Thomas, T. R., "Defining the microtopography of surfaces in thermal
contact", Wear 79, 73-82 (1982).
Thomas, T. R. and Olszowski, S. T., "Theory, design and performance of a
porous-diaphragm hoverpallet", Proc. 6th. Int Gas Bearing Symp. D6, 73-92
(BHRA, Cranfeld, 1974)
Thomas, T. R. and Probert, S. D., "Thermal contact resistance: The
directional effect and other problems", Int. J. Heat Mass Tran.sfir, 13, 789-807
(1970).
Thomas, T. R. and Probert, S. D., "Correlations for thermal contact
conductance in vacuo", Trans. Am. Soc. Mech. Engrs., 94C, 176-180 (1972)
Thomas, T. R. and Sayles, R. S., "Random-process analysis of the effect of
waviness on thermal contact resistance", A.I.A.A. Paper No. 74-691, (1974).
Thomas, T. R., Sayles, R. S. and Haslock, I., "Human joint performance and
the roughness of articular cartilage", Trans. Am. SOC.Mech. Engrs: J. Biomech.
Eng., 1026,50-57 (1980)
Thompson, D. J., "Wheel-rail noise generation, Part I: Introduction and
interaction model", Journal of Sound and Vibration 161,387-400 (1993)
Thompson, D. J., "Wheel-rail noise generation, Part V: Inclusion of wheel
rotation", Jourvlal of Sound and Vibration 161,467-482 (1993)
Tillack, M. S.; Abelson, R. D., "Interface conductance between roughened Be
and steel under thermal deformation", Fusion Engineering and Design 27,232-239
(1995)
268 Rough Surfaces

Toni, K., Nishino, K., "Thermal contact resistance of wavy surfaces", Revista
Brasileira de Ciencias Mecanicas 17,56-76 (1995)
Van Deusen, B. D., "A statistical technique for the dynamic analysis of
vehicles traversing rough yielding and non-yielding surfaces", NASA Report CR-
659 (1967).
Wambold, J. C., and Henry, J. J., "Evaluation of pavement surface texture
significance and measurement techniques", Wear 83, 351-368 (1982)
Wennerberg, A., On surface roughness and implant incorporation, PhD
thesis, Goteborg University (1996)
Yhland, E., "Waviness measurement - an instrument for quality control in
rolling bearing industry", Proc. I. Mech. E., 182, Part 3K, 438-445 (1967/68).
Yip, F. C. and Venart, 3. E. S., "Surface topography effects in the estimation
of thermal and electrical contact resistance", Proc. I. Mech. E., 182, Part 3K, 81-91
(1967/68).
INDEX

Abrasive composites, 15 Asperities, 225,229,230,232,241,243.


Abrasive machining, 156 251,260
ACF. See autocorrelation function collisions, 25 1
Acoustic density, 74,229,249
interferometer, 84 Astigmatic focussing, 103
speckle, 84 Asymmetry, 143
waves, 83,84 Atomic force microscope, 70
ACVF. See autocovariance fimction Autocollimation, 8, 39
ADC.See analogue-todigital conversion Autocorrelation, 133, 151, 153, 155,
Adhesive Giction, 225 157, 161, 176-178
AFM. See atomic force microscope Autocovariance fimction, 153,235
Air Autofocussing, 65
bearings, 251 Automobile brake linings, 226
gap, 17 Autoregressive moving-average, 156
jet, 101 Average
Aircraft, 1,43,44 slope, 50, 162
engineers, 3 wavelength, 152, 170,227,262
Aliasing, 115 Averaging circuitry, 20
Alignements, 2
Allison system, 136, 137
Amplifier, 14
Amplitude parameters, 1 33, 151, 161, Back-scattered electrons, 65
173,186 Backscattering, 83, 84, 85
Analogue methods, 113 Bad data, 95
Analogue-todigital conversion, 1 13, 117 Ball filter, 173
Angle of extinction, 47 Ball-cage interaction, 25 1
Bandwidth, 181, 184,185,249
Angular Beamsplitter, 40
distribution, 47, 50, 5 1, 52
reflectance function, 103 Bearing area, 144, 145, 146, 153, 167,
uncertainties, 7 176
Anisotropy,48, 133, 169, 171, 177, 190, Gaction, 144, 176
1 92- 1 98 ratio, 144
correction factor, 2 16 Beta distribution, 143
Anode, 17 Bifractal, 167, 168
Area of contact, 23 Birmingham 14, 173, 174,177, 186
Area of real contact, 5 Boosted lubrication, 260,261
ARMA. See autoregressive moving- Boundary lubrication, 227,240
average Boustrophedon scanning, 94
Armature, 17 Breaking-in, 229,241
Articular cartilage, 84,259-261,267 Bridge circuit, 17

269
270 Rough Surfaces

Bridge loadings, 262 Conductance, 247,250,263-267


Bulldozer, 24 Conducting probe, 68
Burnishing, 230,23 1 Confocal microscope, 42
Constriction of flow lines, 248
Contact spots, 248
Contacting envelope, 2 1
Calcium fluoride films, 65 Convolution, 118, 119, 161
Cali-block, 2 1,28, 34 Coordinate measuring machines, 8
Calibration, 28,29, 30, 32, 33, 133 Core fluid retention index, 176
specimens, 28,29, 33, 99, 106 Comer frequency, 167
Cantilever, 70 Correlation length, 11, 133, 153, 159,
Cantor set, 11,25 161, 174,257
Capacitance, 17, 66-68, 77,78, 85, 87, Correlogram, 152, 170
88,100 Coulombic theory of friction, 225
Carbon replicas, 65 Crankshaft, 97, 103, 107
Carnac, 2 Crops, 9, 10
Carrier, 17, 19 Cross-correlation, 180,235
Cartographers, 3 Cross-covariance function, 190
Cartridge, 16 Crystal faces, 53
Casting, 16, 167 Cutoff, 74, 76, 79, 80, 83, 118-125, 130,
Catchment areas, 194 138,161, 183,184,231
Cathode ray tubes, 92 cutting
Causal filters, 119 fluids, 100
Censoring, 233-235 speed, 102
Central limit theorem, 141 Cylinder bores, 258
Characteristic depth, 129
Charge force microscope, 72
Charge-coupled diode, 45
Chart recorder, 19,22 Damping effects, 70
Chatter, 167 Data acquisition, 94, 100
Chemical balances, 9 Datum, 16, 18,26,28
Chisel-shaped stylus, 23 Default settings, 95
Circular DEM. See digital elevation models
lay, 190, 192 Demodulation, 17
profiler, 44 Density of extrema, 181
Closed contours, 4, 5,6 Dental cement, 98
Clusters of contacts, 249 Diamond styli, 15
Coefficient of variation, 133 Diamond-turned, 50,51, 57
Coherence radar, 46 Dielectric surfaces, 5 1
Coherent light, 45, 50, 54, 57, 103 Differential thermocouple, 75
Compact disc player, 40 Diffracted radiation, 47
Compliant Diffuse reflection, 36, 39, 60
seal, 80 Digital
surfaces, 27,98,228,259 cartography, 262
Compressible fluids, 8 1 elevation models, 262
Concrete surfaces, 84 filter, 117
Index 27 1

Dimensional tolerance, 255,263 Field coils, 65


Directional properties, 5 1, 188, 190 Film-thwkness ratio, 229
Discretisation, 113 Filtering, 116, 117, 121, 122, 126, 134,
Discriminant analysis, 235 159, 161, 170, 172, 176, 186, 197,
Dry dock, 13 198,229,233
Dynamic Finger, 11
range, 14,31 Fingemail, 71, 73
response, 24,31 Finite impulse response, 119
stiffness, 16 FIR. See finite impulse response
Flagging, 95
Flash gun, 37
Flatness, 1, 100
Effective F l e m i n g integrator, 11
hardness, 8 Flexible diaphragm, 78
profile, 2 1 Floor surfaces, 227
Elastic contact, 250,261 Flow pressure, 23
Elastohydrodynamic lubrication, 228, Fluid dynamics, 162
238-244 Fluid flow, 247,25 1
Elastomers, 27 Focus-detection, 40,41, 103
Electrical contact resistance, 1,247,268 Footprint, 2,24, 92
Electroforming, 99 Footwear, 227,242
Electromagnetic radiation, 35,46 Forest canopies, 9
Electron Form error, 68
beam, 92-94 Formprofil, 125
microscopy, 7,64, 86, 93, 97, 106 Foucault knife-edge, 37
Electrostatic forces, 70 Fouling, 254
Ellipsometry, 53,60,61, 104, 106 Fourier transform, 119, 121, 154, 160,
Ensemble averaging, 161, 179 161
Entropy, 2 , 3 Fourier transforming lens, 52
Error trap, 159 Fractal, 162, 164, 165, 166, 167, 168,
Errors of form, 39,94, 116, 125 169, 170, 185, 194, 195, 196,244,
E-system, 21, 125, 126, 131 256
Evaluation length, 18, 135, 138 dimension, 165
Excel, 94 Fracture surfaces, 162, 168, 169
Extreme-value parameters, 133, 136, Fraunhofer diffraction, 103, 109, 110
137,139 Frequency response, 119, 121
Friction, 225-22'7,238-245,253,254,
264,265
Face turning, 190 Fringe-field technique, 66
Facets, 51 Fringes, 45,46, 93
Fast Fourier transform, 160 FSD. See full-scale deflection
Fatigue life, 228,238 Fuel element, 248
Feed rate, 50 Full-scale deflection, 114
Feedback circuit, 70 Functional
Femoral head, 260,264 filtering, 3, 229
Fibre optics, 103, 105 parameters, 176
272 Rough Surfaces

Gamma function, 164 Highway roughness, 262


Gauge block, 29, 133 Hip joints, 259,264
Gauging nozzle, 8 1 Honing scratches, 195
Gaussian Horizontal
distribution, 114, 139-147, 157, 180, compression, 22
183,234,254,261, magnification, 19
filter, 123, 173 range, 6
weighting function, 120 resolution, 14,65,69
Gear Housing aligumnt, 251
contacts, 228,243 HSC. See high-spot count
teeth, 97 H-system, 136
General surface texture, 136 Hubble, 8, 10
Geography, 262 Hull fnction, 1
Geology, 262 Hiillprofl, 2 1, 125
Geomorphology, 194,247,262,267 Human teeth, 97
Geomorphometry, 261,266 Huygen’s principle, 54
Glass transition, 27,28 Hybrid parameters, 173, 175, 176
Glossmeter, 36,47,48, 91 Hydrodynamic lubrication, 227,238-245
Goodness-of-fit test, 142 Hydrodynamicists, 3
Gramophone, 11 Hydrology, 262,267
Grinding, 134, 142, 149, 169, 172, 195, Hyperbola, 147, 167, 168
257,258 Hysteresis, 68
scratch, 257
wheel, 257,258
Gritblasting, 185, 194, 195
Ground glass, 56, 160 W. See infinite impulse response
GST. See general surface texture Image clarity, 47
Impact wear, 97,107
Implant, 261,268
Impulse response, 1 18, 1 19, 122
Hairbrush, 8 Inclined plane, 74
Half-width, 104, 105 Indenter, 23
Heavy electrical engineering, 3 Index of refraction, 53
Heel-strike, 260 Inductance, 17,78
Height distribution, 10, 133, 140-149, Infrared laser diode, 4 1
155, 176,234,235,253,254,257, In-process measurement, 100, 102, 103
258 Inspection, 16, 91, 100, 111, 172,241
Hemispherical parts, 105 Instrument
Hertzian elastic modulus, 74 performance, 6 , 7
Heterodyne laser, 43 reference plane, 7
High-fidelity, 16 Intensity
High-pass filter, 117, 118, 121, 130, equations, 45
155,161 signal, 65
High-spot Interferometry, 42-46, 91, 93, 94, 107,
count, 151 108,111
spacing, 151 Internal combustion engine, 4
Index 273

Interrupted finishes, 145 Mathernatica, 94


Inverse Fourier transform, 161 Matlab, 94
Isometric display, 94, 95 Mean
absolute slope, 105, 158, 159
contact spot radius, 4, 5
hydraulic radius, 80
Kurtosis, 143, 173 line crossing, 152
peak curvatwe, 157
peak height, 157,158,182
Lambda ratio, 229 peak radius of curvature, 159
Lambert's law, 35 plane, 172, 176, 185,252,261
Laminar Mechanical
flow, 80 noise, 68
sublayer, 254,255 vibrations, 14
Landscape, 9 I , 166 Mecrin tester, 73
Laser Megalithic, 2
force microscope, 72 Menhir, 2
interferometer, 18 Meteorology, 262
scanning analyser, 103, 104, 106 Method divergence, 123
Lateral MHR. See mean hydraulic radius
deflection, 24,25 Michelson interferometer, 44
resolution, 66,68,70,71, 84, 101 Microdensitometer, 65, 88
stiffness, 24, 33 Microscopic resistance, 248,250
Lathe, 255 Milling, 190, 193
bed, 16 Mireau interferometer, 45
saddle, 16 Mixed lubrication, 227,229,239-245
Leakage rate, 235 Molten asphalt, 76
Length of traverse, 11 Molybdenum disulphide film, 226
Leptokurtic, 143 Moment of distribution, 133, 140, 143,
Levelling, 16, 17 155,162,183,249
Light-section microscope, 36, 37, 125, Monochromatic illumination, 55
134 Motif
Linear discriminant functions, 237 analysis, 126, 130, 193
Lip seals, 235,236,244,245 combination, 128, 129, 131
Log-normal distribution, 143 filter, 173
Long-crestedness, 192 Moving-coil transducer, 17
Loudspeaker, 19 MPRC. See mean peak radius of
Low-pass filter, 66, 117, 118, 123, 194 curvature
LVDT. See linear variable differential M-system, 120, 125, 130
transformer Multifractal, 166, 167
Multiprocess surface, 133, 145, 147

Macroscopic resistance, 248,250


Magnetic force microscope, 72 Nearest neighbows, 186, 187
Material ratio curve, 144 Negative slopes, 19
274 Rough Surfaces

Nickel replicas, 100 roughness, 262


Noise, 116, 156, 168,250,251,263,267 user rating, 262
Nomarslu microscopy, 39 PC. See phase-corrected filter
Nominal contact area, 74 Peak height, 73, 137,146,157,158,181
Non-causal filters, 1 19 182
Non-contacting stylus, 37 Peak-to-trough roughness, 227
Non-recursive filters, 119 Peak-to-valley height, 64, 74, 128, 129
Nonstationarity, 133 Pendulum, 74
Nozzle cross-sections, 81 Perspective view, 94
Nuclear power, 248 Phase
power stations, 1 diirerence, 40,42,43, 54
Nyquist distortion, 119, 120
frequency, 1 15 lag, 27
sampling theorem, 1 15 shift, 43
Phase-corrected filter, 120, 130
Phonograph, 11, 15, 16
Photodetector, 103
Ocean Photographic film, 52
surfaces, 180 Photographic negative, 37
waves, 43 Photon scanning tunnelling microscope,
Oil-droplet method, 83
72
On-the-fly sampling, 1 15 Photoresistors, 39
Operating envelope, 6, 7
Pickup, 14, 16, 18-20
Optical Piezo drives, 68
absorption microscope, 72 Piezoelectric
flat, 83, 94, 98 crystal, 16
instruments, 9 transducer, 45
lever, 11 Pmhole, 42
path, 45,46,52,98,103 Piston term, 172
probe, 37-39
Plaster of paris, 98
profilers, 37, 50 Plastic deformation, 258
sections, 36
Plateau honed surface, 95, 177, 178
stylus, 40, 103, 108
Platykurtic, 143
Oscillator, 17
Pneumatic
Oscilloscope, 19, 103, 104
gauging, 81,86, 89, 101, 108,253
Outflow meter, 80, 81, 85
profiler, 101
Out-of-balance signal, 39,41
Poisson’s ratio, 9
Polar scans, 92
Polarization, 40,42, 43,47, 52, 54, 103
Paper tape, 74 Polarizing interferometer, 42,43
Parameter rash, 133, 150 Polygonal mirror, 103
Partial EHL, 228 Polynomial
Pass-band, 155, 183 filter, 122
Pattern recognition, 128, 193 fitting, 172
Pavement Portable data format, 94
loadings, 262 Positional error?, 93
Index 275

Positive replica, 98, 99 Reflected beam, 35,40,43, 54


Power spectra, 11, 18,20,27, 51, 52, 99, Refractive index, 52, 98
115, 117, 129, 130, 133, 153-155, Relocation, 24,27, 28, 96, 97
160, 163, 173, 177, 178, 182, 192, Rendering of image, 94
231,233,235,250,254,255 Replica, 52,98,99, 100, 108, 109, 110,
Pressure transducers, 101 254
Pnmary standards, 4 Repulsive forces, 70
Probability density, 140, 141, 155, 181, Resolution, 4, 6, 8, 9
182, 197 Resonances, 25 1
Production engineers, 3 Reverse slopes, 24
Production time, 3 Reynolds number, 253,254
Profile &de quality, 262,264
curvature, 157,232 Rk,145,146, 148,150
length ratio, 152 RI, 152
Profilometer, 13, 30, 32 Rm,135
Projector, 12 Road roughness, 225,262,265
Prostheses, 261 Road surfaces, 77, 80
PSDF. See power spectrum Rocker arm, 15
Pt, 134, 135, 148 Roller burnishing, 230
PTB, 28,29 Rolling
Pushrod, 15 ball, 173
F'yramid, 15 bearings, 229,250,266
circle, 2 1
contact, 250,264
ellipse, 126
Quality Rolling of sheet metal, 97
control, 20, 100, 105, 145,238, 247, Roll-off, 118, 119, 121, 130
268 Rough
system, 5 pipe, 253
Quantisation, 113, 114, 137 sliding, 228, 243
Quantum tunnelling, 68 Roughness
of floor surfa.:es, 227
regime, 51
R3z, 136 standards, 5, 15,21, 28
Ru, 133, 138, 139, 143-149,230 Rp, 74,76,83, 135, 145, 146, 148
Rru'Rq, 139 Rq, 138, 139, 142, 144, 145, 147, 158,
Radar engineers, 3 162,167,236
Railroad operation, 250 Running average, 123
Raster scan, 68, 92, 93 Running-in, 96, 97, 106, 110,229, 230,
Rayleigh distnbution, 143 233,238,243,244
Razor blade, 36,76 Rm-out, 19
Real contact area, 4, 10 Rvk, 145, 146, 148
Reciprocating contact, 3 Ry, 135,136
Recording barometer, 24 Rz, 136
Recursive filters, 119
276 Rough Surfaces

SAq, 174,175 Shock noise, 251


Safety of floors, 227 Shortest correlation length, 174
Sample length, 121,122,135 Shot-blasting, 141
Sampling, 113, 115, 135, 137, 152, 158, Silicon wafers, 103
159, 176, 183-187 Sinusoidal surface, 6 , 7
interval, 1, 137, 158, 159, 183-185 Skewness, 143, 173, 177, 186,227,228,
Sand grain roughness, 253 238,257
Sand-patch, 76 Skid, 126
Satellite, 1, 248 Skid, 16, 18,26-32, 126
ranging techniques, 8 Skin
Sbi, 174, 176 friction, 253
scanning resistance, 79
chemical potential microscope, 72 Sliding contact, 97
ion conductance microscope, 72 Slip gauges, 4
near-field acoustic microscope, 70 Slope, 6,7,20,65,73,74,76,83, 104,
near-field optical microscope, 72 105, 123, 146, 149, 152, 158, 159,
thermal profiler, 72 162,164,166,175, 181, 182, 188,
tunnelling microscope, 68, 72 225,226,229,232,255
Scattering, 46,84 distribution, 19
Scattering, 84 measurement, 39
angle, 51 Sm,151,155,169,170
theory, 46 Smoked-glass plate, 11
Scatterometers, 36 Smoothing, 118, 161
Sci, 174, 176, 198 SNAM. See scanning near-field acoustic
Scraping technique, 76 microscope
Scuffing, 229,243 Snell's law, 35
Sds, 174, 182 Soil erosion, 262
Sea-floorterrain, 262 Sonar, 84
Seals, 235-251 Space vehcle, 248
Second law of thermodynamics, 2 Spark erosion, 166
Secondary Spatial coherence, 55
electrons, 65 Speckle, 47,54-59, 84, 85
length standards, 4 contrast, 47, 54, 55
Second-order effect, 3 pattern decorrelation, 54,55
Sectional measurement, 5 Specular reflection, 36, 37,47-50, 58,
Seismograph, 14 84, 147
Self-afine fractals, 117, 164, 165, 170 Speed of traverse, 19,24
Self-similarity, 162 Spherical contacts, 15
SEM. See scanning electron microscopy Spindle runout, 167
Separable filter, 173 Spiral scans, 92
Servo control, 68 Spline filters, 123
Sewing needle, 15 Spot diameter, 103
Shakedown, 229 Sq, 173,174,176,177,188
Shaping, 66, 190 Stagnant-layer method, 83
Sharpest curvature, 6 Static
Shps' hulls, 12, 97,254 friction, 73,74
Index 277

seals, 235,242,251 Taper


Stationarity, 116, 155, 161, 169, 180, roller bearing, 23 1
255 sectioning, 63,64
Std, 174 Telemetry, 102
Stedman diagram, 7, 8, 37 Television, 92
Steepest slope, 6,65, 123, 159 TEM. See transmission electron
Stereo pairs, 65 microscopy
Stereophotogrammetry, 9 1 Temporal coherence, 55
Stick-slip, 19 Ten-point height, 136, 173, 174, 186
Stm,68 Terrain roughness, 262
Str, 174, 196 Textile roughness, 37
Straddling skids, 16 Texture, 1, 3,6
Straightness, 1,2 aspect ratio, 174, 175, 192
Strain state, 53 direction, 174, 176, 177
Stratified surface, 145, 146, 148, 190, Thermal
195 comparator, 75, 76
Strong anisotropy, 192, 193,194 conductance, 75
Structure h c t i o n , 155, 156, 164-169, expansion effects, 255
177,179,180 noise, 14
Stylus resistance, 1
dimensions, 12, 14 Thermoelectric, 72
geometry, 24,28 Thetameter, 74, 75
instrument, 4, 7, 9-21,28-32, 96-104, Three-dimensional filtering, 171
167,227,260 Time series, 153, 161
load, 19,23,24 Tip radius, 15,21
tip, 20,21,23,28 TIS. See total integrated scatter
transducer, 14, 16 Tolerance, 255,256
width, 66 Tool
S h t , 10, 142, 171-190, 197 radius, 106
definition, 186 replacement, 100
Surface Topothesy, 162, 164, 185, 194, 195
bearing index, 176 Total integrated scatter, 49, 57
damage, 15,23,3 1 T,’ 139, 144, 145
discontinuities, 41 Traceability, 29, 32
slope, 182 Traceable uncertainty, 100
vibration, 54, 105 Traced profile, 20
Surgical implants, 26 1 Transducer, 113, 114
Surveying, 91 Transducer, 13-18,29, 32, 113, 114
Svi, 174, 176 Transition region, 254
Swad, 103 Transitional topography, 235
Switchgear, 1 Translation, 115, 125
Synovial fluid, 259,260 speeds, 67,102
stage, 37,39
tables, 93
Transmission
Tactile test, 71
characteristic, 17
278 Rough Surfaces

coefficient, 1 18, I 19, 121 Watersheds, 194


design, 251 Waviness, 12, 83, 103, 116, 125, 126,
Transmittance, 52,65 130, 167, 170,248-251,267
Transparent replica, 52 number, 250
Traverse length, 19 Weak anisotropy, 192, 193
Trend removal, 172, 182 Wear, 227-244,259,263,264
Tribology, 225,244 scars, 97
Triode, 17 Weighting
Truck shipments, 262,265 function, 66, 120-126
Tuning fork, 70 sequence, 1 19
Tunnelling current, 68,69 terms, 117
Turbulent flow, 253 Wheatstone bridge, 82, 89
Tyre-road interactions, 1 White light, 40
Wiener-Khinclune relation, 154
Wildfire propagation, 262
Wire-frame view, 94
Ultrasonic back-scattering, 102 Wollaston prism, 39,43
Ultrasound, 84, 85,88 Wood, 15, 17,24, 30, 31, 33
Uncertainty, 11, 133 Wristwatch. 70
Underwater surfaces, 84,97

Zero crossing density, 262


Vacuum Zonal filter, 173
chamber, 69 Zoom, 94
seals, 235
Valley
depth, 137,145,146
fluid retention index, 176
suppression filter, 122, 123
Van der waal's forces, 70
Vehicle dynamics, 262
Vertical
range, 4, 9, 14, 37-46,65-69, 103
resolution, 14, 18, 3748, 64-69,74,
75, 82-84
scanning interferometry, 46
Vibration, 250,251,262-267
Vibrator, 29, 30
Video camera, 37
Viscosity, 81,238,241,244
Visibility of fringes, 46

Wall gauge, 12
Walsh functions, 155, 169

You might also like