You are on page 1of 182

TURBULENCE CONTROL BY PASSIVE MEANS

FLUID MECHANICS AND ITS APPLICATIONS


Volume 4

Series Editor: R. MOREAU


MADYLAM
Ecole Nationale Superieure d' Hydraulique de Grenoble
Boite Postale 95
38402 Saint Martin d'Heres Cedex, France

Aims and Scope of the Series


The purpose of this series is to focus on subjects in which fluid mechanics plays a
fundamental role.
As well as the more traditional applications of aeronautics, hydraulics, heat and
mass transfer etc., books will be published dealing with topics which are currently
in a state of rapid development, such as turbulence, suspensions and multiphase
fluids, super and hypersonic flows and numerical modelling techniques.
It is a widely held view that it is the interdisciplinary subjects that will receive
intense scientific attention, bringing them to the forefront of technological advance-
ment. Fluids have the ability to transport matter and its properties as well as
transmit force, therefore fluid mechanics is a subject that is particulary open to
cross fertilisation with other sciences and disciplines of engineering. The subject of
fluid mechanics will be highly relevant in domains such as chemical, metallurgical,
biological and ecological engineering. This series is particularly open to such new
multidisciplinary domains.
The median level of presentation is the first year graduate student. Some texts are
monographs defining the current state of a field; others are accessible to final year
undergraduates; but essentially the emphasis is on readability and clarity.

For a list of related mechanics titles, seefinal pages.


Turbulence Control
by Passive Means
Proceedings of the 4th European Drag Reduction Meeting

edited by

E. COUSTOLS
ONERAICERT,
Toulouse Cedex,
France

KLUWER ACADEMIC PUBLISHERS


DORDRECHT I BOSTON I LONDON
ISBN-13: e-ISBN-13:
DOl: 10.1007/

Library of Congress Cataloging in Publication Data


European Drag ReductIon Meeting (4th 1989 Lausanne. Switzerland)
Turbulence con t ro l by pass Iv e means proceedings of the Fo ur t h
European Dra g Reduction MeetI ng ! edI ted by E. Coustol s.
p. cm.
"T he ~ ourt h European Drag Reduction Meeting was held. July 24.
1989 i n Lausa nn e . Sw i tzerla nd . "
I n c luc es o l bl 1 ograp hi ca l ref e r ences an d index.
ISBN 0-7923 -1 020-9
I .F1 L l a dynam l cs--Congresses. 2. Tur bulent bound ar y l ayer -
-Co ngress es. I . Coustols. E. I I. T itle.
TA357.5 .T87E93 1989
620.1 '064--jc2 G 90-48907
ISBN-13: 978-94-010-7471-1 e-ISBN-13: 978-94-009-2159-7
001: 10.1007/978-94-009-2159-7

Published by Kluwer Academic Publishers,


P.O. Box 17, 3300 AA Dordrecht, The Netherlands.

Kluwer Academic Publishers incorporates


the publishing programmes of
D. Reidel, Martinus Nijhoff, Dr W. Junk and MTP Press.

Sold and distributed in the U.S.A. and Canada


by Kluwer Academic Publishers,
101 Philip Drive, Norwell , MA 02061, U.S .A.

In all other countries, sold and distributed


by Kluwer Academic Publishers Group,
P.O. Box 322, 3300 AH Dordrecht, The Netherlands.

Printed on acid-free paper

All Rights Reserved


1990 by Kluwer Academic Publishers
Softcover reprint of the hardcover 1st edition 1990
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.
Contents

Preface vii

Final programme ix

External Manipulators (LEEUs)


Balallce of turbulent kinetic energy downstream a single flat plate Inaniplllator :
comparisons betll'een detailed experiments and modelling. C. Tf 1/ a ud. I run ny.
IP. Bonnet C' I Dclville
.\Ianipulation and modelling of turbulent pipe flow: some parametric studies of
single and tandem ring devices. A. Pollard. H. Thoma n 11 C A .Jl. SUl'ili 23
Large eddy simulation of manipulated boundary layer and channel flows. 11.
1,'hin C' R. Friedrich -\1
The importance of LEBF de\'ice shape [or turbulent drag reduction.A. Butril"ud f,7

Internal Manipulators (Riblets)


Boundary layer flO\\' visualisation patterns on a riblet surface. JJ C. C/n Ik 7')
Simultaneous flow \'isualisation and LDA studies over longit uclillallllicro-grom'ecl
surfaces. C.I.l. Pulles. K.K. Prasad () F. T..U. "YiuUL'stailt LJ7
Effects of longitudinal pressure gradients on turbulent drag reduction with riLkts.
1,'.S. Choi IOl)
Synthesis of experimental riblet studies in transonic conditions. E. COI/sto/.' C
IT. Schmitt 12.\
Effect of ri blet s on either fully dewloped boundary layers or illt ernal flo\\"" ill
laminar regime. I Liandrat. E. COllslo/s. L. Djenidi. F. Ansc/})ut. )( de SO/III-
\'iclol'. F. Fioc [;' L. Fu/achier 1-\1

Internal and External 1\1anipulators


Turhulent houndar\' layer OHT a ribleted surfClce with tandelll l11i111ipulators 1),-
ing surfac(' drag balances. I'.D ..Ygllycn. I Dickinson. }'. .]U/II. L Cha/Irollr. "I.
Small!. A. Pogc C F. PaquEt 15')

List of referees In
Preface

In the last decades, a lot of effort has been directed towards manipulation of
turbulent boundary layers by passive devices such as external manipulators (thin
flat plates or aerofoil section devices embedded in the outer layer) and/or internal
manipulators (small streamwise grooves acting directly on the inner region) for the
purpose of reducing viscous drag. The former are commonly referred to as LEI3U s or
BLADEs and the laHer riblets or grooves. Though the details of the mechanisms are
not firmly understood, world-wide experimenta.! data are available and consistent
enough in order to assert the potential of such devices for turbulent drag reduction.
It should be noted that following on from recent and successful flight tests, the
concept of using grooved surfaces is rather close to finding industrial applications.
During the last few years, in Europe, there has been considerable interest in
lookillg at the behaviour of such passi,'e turbulence manipulators. A lot of intense
research, concerning both experimental and theoretical studies. has been carried out
in some European research centres. For the last fi\'e years. informal gatherings. called
,.\ \'orking Pi\l'ty i\Ieetings" , have been set up, once a year; the aim of these meetings
is not only to bring together European researchers acti,'e in the field of turbulent
drag reduction by passi"e means and to hear about recent de\'c!opments but also to
o u t1 ine sui tit ble directions for future research or collaborative programmes.
Thus, follO\\'ing on from previous meetings at EPF Lausanne (September 1st-
:3rd, 1986), ONEllA/CEllT Toulouse (September lOth, 1987) and ONEnA/Chalillon
(September 20th-:30th, 1988), the fourth European Drag Reduction WorLing Party
:'Ieeting \\'itS held at EPF Lausanne on July 2 Hh, 1989 under the auspices of the
c

11COFTAC Pilot Centre by invitation of Professor 1.1. 11yhming and Doctor T.V.
Truong. T\\'ellty six participants from fi \'e European count ries (Fra nce, CermallY,
S \\' i tzerl a nd, The N etherla nels and U ni ted Kingdom), together \\'il h collaborating
colleagues from Canada, presented 18 contributions of \\'ork which eilher had heen
recently completed or was in progress. The purpose of this Proceedings Book is
to pro\'ide papers '\'hich reflect the contents of the talks given at that conference.
It should lw lloled that "A report on the 4 th European Drag neduction \\'orking
Party T\lccting" \\"as published in the EHCOFTAC Bulletin IV. December 19S'}
011 a personal note, I would like to say how grateful I am to all the rc1'erees
for reviewing some of the dozen or so papers offered after the meeting. Thanks
to their assist ance. this Proceedings Book should be published in time for the :5th
\\'orking Parly i\Ieeting which will be held at I3i\IT Fluid i\Iechanics Limited. Ted-
dington. Uniled Kingdom. I would also like to thank Doctor Nigel Hollillg\\'orth,
frolll KIll wer academic ))ll blishers, for all his va 1uable commen b dming the prepa-
ration of this Proceedings Book.

Toulouse, August 1990

E. Coustols
vii
Final Programme

4th European Turbulent Drag Reduction


\\lorking Meeting 1989
S,viss Federal Institute of Technology
Lausanne EPFL - July 24th, 1989

Session 1 - External l\1anipulators in Internal and External Flows


(Chairman: E. Coustols)

High speed LEBU device drag measurements. J.P. Bonnet &; J. Delville
Complementary drag reduction measurements. L.C. Squire &; A.:'.I. Savill
I-Combination of riblets and manipulators at low Re ; 2-Some observations of tllC
static pressure distributions in pipe flow perturbed by a LEEU ring. Y.D. I\guyen
&; J. Dickinson
Turbulent pipe flow manipulation and modelling. A. Pollard, A.'\1. Sa,'ill &; H.
Thomann

Session :2 - Internal l\1anipulators : Riblets. (Chairman: A.::\I. Savill)

Effects of pressure gradients on riblets. K.S. Choi


Pressure grilc1icnts on riblcts in subsonic and transonic flm\"s. E. Coustols
Further supc'I".sonic riblet results. L. Gauclet (presented by A.:\f. Sa\'ill)
::\[ultiple light-plane flow visualisation of the structure \\"ithin riblcts. D. Clark
Rihlcts ill Cl difl"llser. Ph. Puh'in
Hiblets in a channel flow. ::\1.\1. Lm\'son
Flml" \'isllalisdtion ancl turbulence measurements moer micro-grooves. 1\:. Prasad c\.:
IT. Schwarz
Hiblds on a cylinder. D . .'\eumalln

Session j - Other Drag Reduction Approaches and l\lodelling.


(Chairman: 1.L. Hyhming)

Complltations of turbulent How in L-shaped riblets. R.E. Launder &; Li Shaopillg


Computations for L-shaped riblets in laminar boundary layers. E. Coustols
:further riblet computations in a laminar boundary layer. :r. Anselmet ,\.: L. Djenicli
CurY<ltme cll'ag reduction modelling. A.flI. Savill S: T.B. GatsLi
Video flow \Oisualisation of sparse d-type "roughness" including effect of an aerol"oil
LEBU. P. Filllourakis & A.M. Savill

Sessioll /[ - Round Table. (E. Coustols, A.M. Savill &; LL. Ryhming)

ix
Balance of Turbulent Kinetic Energy
Downstream a Single Flat Plate Manipulator
Comparisons Between Detailed Experiments and
Modelling.

c. Tenaud - J. Lemayt - J.P. Bonnet - J. Delville


Centre d'Etudes Aerorodynamiques et Thenniques,
1. E. A. - U. R. A. C.N.R.S. 191
43, Route de l'Aerodrome, 86000 Poitiers, FRAT\CE.

t Present address: Mechanical Engineering Department,


Laval University, Quebec, CA:r\TADA G1K 7P4

Summary
Previous numerical studies undertaken at ONERA/CERT /DERAT in Toulouse, ha\'e
shO\vn that usual closures, using three or five equations, were able to quite well repro-
duce the developpement of a manipulated boundary layer for the specific geometrical
configuration given from CERT. The goal sought after, in this study. is to check
the predictiw capabilities of the CERT code on an other specific configuration. given
from CEAT Poitiers, for which detailed experiments were available in the downstream
vicinity of an external manipulator of a rectangular section. In the .'\ avier-Stokes
equations, the unsteady and the diffusion terms in the streamwise direction are ne-
glcctecl. For the turbulent motion, a two layer model is adopted using a mixing length
scheme in the near wall region and a k, E, u'v' transport equations model everywhere
else. A brief description of the numerical procedure is first given. The experimental
apparatus and techniques are then described. The numerical results agree qualita-
tively \yell with the experimental mean velocity, turbulence and shear stress data
profiles. The experiments provide the terms of the energetic balance of the kinetic
energy equation ; they are, for further information, compared with the numerical
terms. Although the manipulated flow is strongly out of equilibrium, conventional
closure models appear to be able to predict rather correctly the measured quantities,
particularly the turbulent diffusion term and the dissipation behaviour.
2

1 Introduction
During the last decade, lot of experiments have been conducted in order to reduce skin
friction drag by the control of the eddy structures of turbulent boundary layers. For
airfim'>s, the techniques proposed can be divided into active and passive ones [2], [15].
One of the passive methods consists in immerging thin ribbons, having either rectan-
gular or aerofoil section, within the external part of a turbulent boundary layer. This
method, so called external turbulence manipulation, is studied in this article.
Experimental investigations have shown that the mechamisms leading to drag re-
duction are complex. Among several effects, one can notice that the externalmanipu-
lator induces large skin friction reductions and spectacular modifications on turbulent
profiles (excess of turbulence level in the device wake. important decreases belm\' the
mean axis of this wake, ... ). Furthermore, the manipulated boundary layer is highly
out of equilibrium O\'er a large dmvnstream extent.
Besides these experiments, numerical studies have been undertaken. The goal
was then to check usual turbulent closures, using several transport equations, for spe-
cific geometrical configurations. l"lost of the numerical codes, used up to now. '\Tre
based on boundary layer assumption and thereby started downstream of the device
with the initial data given from experiments. These numerical studies have been
reviewed by Coustols &: Savill [4]. Let us point out that, using appropriate initiali-
sations. authors were able to represent the mean and fiuctuating velocity profiles at
further stations downstream of the manipulator with a satisfactory agreement with
experiments. However, if one aims, for instance, at optimizing the device geometry,
it is necessary to start the computation upstream of the device and to follow the
whole of the fiow development. This second approach has been used in works pre-
viously undertaken at the Aerodynamic department of ONERA/CERT [3]. [13]. On
a given experimental configuration, check of several closure models have been done:
two, three and fiw transport equations, using cOlwentional turbulence models. The
conclusions ,'>as, though the k - E model was able to reproduce qualitati"ely the be-
haviour of such manipulated flow, better results have been obtained using three or
five transport equations. The mean and turbulence velocity profiles as well as the
skin friction coefficient have been calculated in better agreement with the experimen-
tal data [3] [13]. Let us point out that three or five equations models give rather
comparable results, except that the latter allows to get a better representation of the
C f evolution. A remaining question can be pointed out: Is these conclusions on the
CERT configuration stay valid for different device geometries?
The purpose of this study is then to check the predictive capabilities of the
C.E.R.T. numerical code on an other configuration given from the C.E.A.T. Poitiers,
for which detailed experiments were available in the downstream vicinity of an ex-
ternal manipulator having rectangular section [9], [10]. We decided to restrict the
present numerical approach by using only the three transport equations code. This
option gives a good compromise between computer time saving, previous prediction
3

quality and detailed description of the turbulent field.


A brief description of the numerical procedure is first given ; the experimental
apparatus and techniques are described, then a comparison between experiments and
modelling on the mean and fluctuating profiles at further locations downstream of
the manipulator is discussed. The experiments provide the energetic balance of the
kinetic energy equation. Then, for further information, we compare the numerical
terms of the k transport equation with their experimental estimations.

2 Numerical approach.
2.1 Basic equations.
The equations of the mean flow are the Reynolds averaged Navier-Stokes equations.
The usual two-dimensional assumptions are used; x represents the streamwise direc-
tion and y the axis normal to the wall. The unsteady terms as well as the diffusion
ones in the :r direction are neglected. It follows:

aU aV
-+-=0. (1)
a:r ay

a ([J[j)
aol'
+ a(UV)
ay
= _~ aP
pool'
+.!!.- [~ (aU)] + .!!.- (-U'2) + .!!.- (-177l)
ay p ay ax ay (2)

a(0'V)
ax
+ a(11V)
ay
= _~ of +.!!.- [~(a1l)] +.!!.- (-u'v') +.!!.- (_v'2)
p ay ay p ay a:r ay (3)
For the turbulent motion, a two-layer model is adopted:

1. Although the flow is out of equilibrium even close to the ,Yall, a mixing length
scheme is applied, for simplicity, in the near wall regions (up to y+ = 60),
including the lower and upper sides of the device. This scheme is based upon
an eddy \'iscosity assumption:

-,-, III (aU; aUj ) - 2


- uu- = -
I ] P
-
aXj + --
ax; -kfJ
3 IJ
(4)

The eddy viscosity coefficient {LI is expressed, using a mixing length formulation.
as follows:
(5)

where I is the mixing length and F the Van Driest type corrector function for
the viscous sublayer. The use of this scheme, close to the manipulator, can be
4

made if we assume that the boundary layers, developping on both sides of the
device, are turbulent. Let us notice that this assumption is not clearly evidenced
by the available experiments.

2. A three transport equations model is used everywhere else (y+ > GO). These
equations are :

t he kinetic energy equation in a general form :

---
It au: au:
p aXj aXj
(6)

C" and D" are respectively the convect.ion and molecular diffusion terms; they
do not require any modelisation. The turbulent diffusion term T" is modeled as
follows [8] :

(7)

Let 11S notice that this model is based upon a high turbllient Reynolds number
assumption (RI = pk 2 / w:). Close to the limit y+ = GO or in the deyice wake
the RI values could be moderate; therefore, we use a wall damping function
III :

[ -3.4 ] with
11' = exp (1 + Rt/50)2
The pressure-velocity correlation <Pk is neglected. c represents the dissipation
rate .

the transport equation for c is expressed as follows:

(8)

where 12 is a corrector function for low R t regions:

12 = 1. - 0.3exp (-Rn
5

Finally, the transport equation for the turbulent shear stress is used :

(9)

with
120U
P12 = -v -oU
D12 = -u,2-
-
oy oy
and
C 2 +8
a: = - - - -
11
f3 = 2 - 30C2
55 ,=- 8C112 - 2

The constants are the usual ones [8]


Col = 1.44 ; Co2 = 1.92 ; C/1 = 0.09 ; C l 1.44
(Jr = 0.9

In this three-equations model, one needs closure relationships for the diagonal
Reynolds stress components. From the transport equations for these components and
through a local equilibrium assumption, we can correlate the diagonal components
(u;u:, fori = j) to the turbulent kinetic energy; the ratios (U:llj / k, for i = j)
are function of constants C l , C 2 . The choice C l = 1.5 and C 2 = 0.4 leads to the
following values:

U,2 V'2 W,2


k = 0.925 k = 0.465 T = 0.610

However, it has been experimentally evidenced [10] that, in the imediate \icinity
of the manipulator, these three ratios are only slightly affected. Then, we keep the
assumption of local equilibrium for the three normal components of the Reynolds
stress tensor. Furthermore, the - u'v' / k ratio is strongly modified just downstream
of the deyice [10]' justifying the use of the u'v' transport equation.

2.2 Grid scheme and numerical procedure.


The resolution of the N avier-Stokes equations is carried out oyer a distance of about 70
Do, where Do denotes the boundary layer thickness at the device trailing edge location.
The calculation starts 15 Do upstream of the leading edge of the manipulator. The
domain is about 20 Do high. The mesh is tightened ncar the wall and in the vicinity
of the device, both in the streamwise (x) and in the normal to the wall (y) direction.
6

The grid size is then 110 x 160 (x x y) ; the manipulator is discretized on 13 x 4 (x


x y) mesh points.
The calculation scheme employs a finite volume approach, following the work
made by Patankar et al. [12]. The initial conditions are given from the experimen-
tal results for natural boundary layer, as regards to the mean velocity profiles and
the Reynolds stress components. The dissipation rate s at the upstream station. is
experimentally e\'aluated ; howe\'er, a mixing length scheme is still used for the c
estimation. as it was the case for the ONERAjCERT calculations. It should be no-
ticed that the choice of this initial E estimation is of secondary importance since the
calculation starts sufficiently upstream of the device, the turbulence profile being well
established before the flow reaches the manipulator. The resolution of the equations
of motion uses a X-marching technique. As the mean flow equations are elliptic in
the streamwise direction (x), the pressure must be imposed at the last station of the
domain. Convergence is then made upon the pressure field.

3 Experimental investigation.

3.1 Experimental apparatus.

Wind tunnel.

The experimental study is performed in a closed loop wind tunnel with a test section
2. III long and 300 x 300 mm 2 cross section. The settling chamber is equiped with
filters and grids: the contraction ratio is 10. Both the upper and lower walls are
adjusted in order to pre\'ent pressure gradient. The turbulence le\'el in the free stream
is less than 0.2%.

Flow and manipulator characteristics.

Natural boundary layer at the location of the manipulator trailing edge


External velocity Ue = 26 111.8 1
B.L. thickness (at UjUe = 0.995) 60 = 23.5 mm
Shape factor H = 1.35
:"Iomentulll thickness e= 2.3 mm
Reynold number based on Ue and e Reo = 4,000
Manipulator
Chord length c = 25 mm = 1.160
Thickness t = 0.1 111m = 0.00460
Height in the B.L. h = 10 mm = OA36 0
Chord Reynolds number Re c = 36,000
7

3.2 Data acquisition procedure.


General procedure.

The experimental data proceed from hot wire anemometry measurements (TSI 1750
CTA). Both single and X wires are used in order to measure the different quanti-
ties involved in the transport equation of the turbulent kinetic energy. The spatial
gradients included in this equation are properly estimated via a finite difference mea-
surement grid, tightened in the regions of strong gradients (near the wall and in the
vicinity of the trailing edge). The measurements are however perfonned far enough
from very strong gradients in order to limit integration effects arising from spatial
extent of the X wire probes [10]' [l1J.

Hot wire specifications.

Single normal wire: modified miniature probe TSI 1260 T1.5 operating with
tungsten wire of 0.5 Tl7m long and 2.5 Ilm in diameter.

X wires: TSI 1248 Tl.5 probe with tungsten wires of l.25771m long and 4 I'm.
in diameter.

Acquisition system.

The CA\IAC type data acquisition system is composed of programmable filters-


amplifiers and AID converters, coding the numerical data on 12 bits. The signals
collected from t\\'o anemometers are first filtered at 20 J( H z and then simultaneously
sampled at a rate of 50 J( Hz. For single and X wires measurements, a dynamic
memory allows to make sample size of respectively 256 J( and 128 l\" data. The
instantaneous digitized signals are stored on magnetic tape for further processing on
mini computer IvlD500. The whole acquisition and probe displacements are controlled
by a microcomputer.

3.3 Measured quantities.


The instantaneous signals used with the calibration laws [10], [11 J provide the instan-
taneous velocities. The following quantities are obtained from direct data processing
on the instantaneous velocities provided by different probe configuration [10J. [11 J :

1st order moments: U , V

2nd order moments: u,2 V'2 W,2 u'v'


'-----v----'
k

3rd order moments: '11/ 3 u'v'2 'I1.'w,2 u'2v' v,3 v'w,2


8

All the terms of the transport equation of k (high Reynolds number assumption)
are measured except 1>" and E. An estimation of 1>" + E is then obtained by balancing
this equation. If one assumes that the pressure-velocity correlation 1>" is negligeable,
the remainder of the k equation can be considered as an estimation of E.

4 Results and discussion.


The numerical code, using three transport equations, is checked on a specific geomet-
rical configuration corresponding to the above mention ned experiment.

4.1 Mean and fluctuating profiles.


Comparisons with CEAT experiments [9], [10] are presented at several locations down-
stream the device (~ = X/Do, X = 0 refers to the trailing edge of the manipula-
tor.), on the mean velocity profiles and the components of the Reynolds stress tensor
(Fig. 1,2,3) .

i\Iean velocity

vVith this three equations model, a rather good overall agreement with experi-
mental data can be observed (Fig. 1). Howeyer, at the first location (X/Do = 0.43),
the predicted wake is too large when compared with experiments. This discrepancies
might come from the use of a mixing length scheme in the treatment of the near wall
region on both sides of the device. Further downstream, one can notice that the evo-
lution of the deficit pocket is slower than the one observed in the experiments. Similar
conclusions have previously been pointed out on the CERT numerical results. At the
farthest location (X/Do = 20.), the mean velocity profile is slightly overestimated over
the whole thickness of the boundary layer (Fig. 1) .

Turbulent fields

At the first downstream station (X/Do = 0.43), the excess of longitudinal turbu-
lence intensity is quite well predicted (Fig. 2). On the other hand, farther downstream
(X/Do = 7.5), the well known effect of the manipulator inducing the important de-
creases of 1l'2 component below the wake axis, is well reproduced by the calculation.
Unlike the comparisons with the ONERA/CERT data [3], [13], the streamwise tur-
bulence intensity is well predicted in the region near the solid wall where a mixing
length assumption is made.
The results concerning the turbulent shear stress profiles are plotted on figure 3.
Close to the trailing edge of the manipulator, in the outer region located above the
mean axis of the wake, the calculation overpredicts the u'v' level. These discre-
pancies might be due, one more time, to the use of a mixing length scheme for the
9

calculation of the turbulent motion on both sides of the device. This overestima-
tion of u'v' decreases as the downstream distance increases, to fit with the data at
about 7.5 boundary layer thicknesses (Fig. 3). At contrary, below the device height,
the important decrease of the turbulent shear stress profile, experimentally observed
downstream of the device, is rather well predicted by the model (Fig. 3). However,
close to the wall, the numerical code provides with too high u'v' values. At the last
measured station (X/Do = 20.), the Reynolds stress tensor component v'v' tends to
be overestimated in a large part of the boundary layer thickness.

The performance of the present model can also be evaluated on its ability to predict
the evolution of the skin friction coefficient in manipulated flow. Indeed, this effect
is the most important for pratical purposes and its correct prediction is crucial for
optimizations. The values of C f / C fre! obtained both by calculations (C f = l/;~U;'
Tp = (p~U)
uy
)
y=o
and data acquisitions (Preston tube), are plotted \"Cl"SUS X /6 0 on
figure (Fig. 4) ; C f,.e! denotes the skin friction coefficient corresponding to the natural
flow.
The calculations show a more important local skin friction reduction than the
experiments (Fig. 4). Once again, the wall region treatment employed (mixing length
scheme) might explain these discrepancies. Furthermore, the location of the maximum
of local reduction is located slightly upstream of the measured one. This behaviour
can be explained by the fact that the calculated deficit pocket of the mean wlocity
spreads faster towards the wall than the experimental one. Then, the calculated
wake reaches the wall at a point located slightly upstream when compared to the
experiments. This interpretation is supported by previous experiments : actually,
visualizations and thermal marking of the wakc have experimentally revealed that the
location of maximum C f reduction corresponds to the point whcrc the wakc reaches
the solid wall [1], [6], [10]. Finally, one can notice that the predicted relaxation of the
skin friction coefficient toward natural values is in good agrecment with experimental
data (Fig. 4).
From the analysis of this preliminary step, one can observe rcsults on mean and
fluctuating quantities profiles in rather good qualitative agreement with the CEAT
experimcnts. j\Ioreovcr, some discrepancies occur on shear stress profilcs in the im-
mediate downstream vicinity of the manipulator and \"Cry close to the wall. In fact,
the turbulence model, used in the wall region of both sides of the device, can be
responsible of these discrepancies. This application of a three equation model to the
flat plate manipulator expcriments of CEAT [9], [10] confirms thc results prc\iously
obtaincd \\"ith the experimental data of ONERA/CERT [3]' [13].
In ordcr to more preciscly analysc the different turbulence models and tcrms
involvcd in the present closurc, the behaviour of several terms of the equations can
be studicd. On one hand, as previously described, one original aspect of the available
experiments [10]' [11] is the estimation of every term of the transport equation of the
turbulent kinetic energy k. On the other hand, the present code allows to calculate
10

all the terms involved in the k equation. The purpose of thc next section is then to
compare expcrimental and numerical valucs of the term balance of the k equation.

4.2 Energetic terms balance.


The comparison between experiments and modelling is presented in a dimensionless
form on figures (5) to (8) at three stations downstream of the manipulator (X /00 =
2. ; 1.5 ; 20.). The dashed zone, close to the solid wall, represents the region where
a mixing length schcme is employed in the calculations. Indced, no result could be
drawn hcre as regards to balance of k equation. The molecular diffusion term (DJe)
is negligeable when compared to the other terms, it then has not been reported on
thcsc fig1ll'es.
From an experimental point of \'iew, it is often difficult to propcrly c\'aluatc the
dissipation rate (E). Thc E is estimated as the remainder of this equation, assum-
ing that the pressure-\'eloci ty correlations (<I> d is negligiblc in the total thickness of
the boundary layer. This cxperimental assumption is consist ant ,vith the modelling
one (see 2.1).
Close to the trailing edge of the manipulator (X /00 = 2.), experiments ha\'e shown
important modifications of thc convection profile, especially both at and below thc
de\'ice height (y / 00 = 0.43) [11]. At the same location (X /00 = 2.), the predicted
conwction (C k ) (Fig. 5) is in rather good agrccment with the experimental results.
Howe,'er, in the immediate vicinity of the wake axis, a slight overestimation of calcu-
lated C k is obscrvcd (Fig. 5). As we proceed downstream of the manipulator (Fig. 5),
one can observe that the numerical results reprcsent correctly the \'ery rapid experi-
mental relaxation of the convection tcrm.
Thc t1ll'bulent diffusion (TJ.:) profiles are plotted on figure 6. At the first station
(X/oo = 2.), the experimental results show a change of sign undcr and upper the
manipulator hcight. Lemay [10] proposes that this sign change on Tk profilc con-
tributes to disrupt the energetic exchanges between the both sides of the wake axis.
This sign im'Cl'sion is still present at X /00 = 7.5, but occurs closcr to thc solid ,vall
than at the first location. From a numerical point of view, the usual first gradient
formulation (7) used for the modelisation of thc third ordcr correlation, can reproduce
rather wcll thc typical experimental behaviour of Tk at the first station X /00 = 2.
(Fig. 6). r-.loreover, the downstream evolution of the turbulent diffusion profile is in
a quite good agreement with experiments (Fig. 6). vVc must noticc that, though the
manipulated flow downstream of the device is greatly out of cquilibrium. thcrc is no
need to employ a more sofisticated modelisation for the t1ll'bulent diffusion as, for
instancc, a Daly-Harlow model [5] or a Hanjalic-Launder onc [7].
The cvolution of the production term Pk is plotted figure 7. Experiments show,
close to thc trailing edge of the manipulator (X /00 = 2.), a near-zero production level
just below the device height, while above, a peak on the Pk profile is obscrved [11].
The modelling gi\'es a rather good evolution of Pk in the normal to thc wall dircction
II

(Fig. 7) compared to the experimental one. Just below the manipulator posItIon
the association of the wake velocity defect and the boundary layer profile imposes
~u = 0 ; the production term then vanishes whatever the turbulent shear stress is.
J~st aboYe the manipulator height, the production of f.;; is numerically O\'erestimated,
due to theu'v' overprediction already mentionned (see 4.1).
For locations dO\vnstream of the device, one observes that the numerical relaxation
of h: is slower than the experimental one in the lower part of the boundary layer.
In this region, theu'v' level is well predicted, the discrepancies on PI,; can only be
attributed to the mean velocity prediction. Even at the last station (X/[,o = 20.)
(Fig. 7), the production term given from calculation has not recovered its natural
value while, in the experiments, there is no difference between manipulated and non-
manipulated PI,; profiles [10].
The dissipation rate E, calculated through the transport equation( 8), is plotted on
figure (8). Though this equation involves modelisations developped for high Reynolds
numbers, an overall rather good agreement is achieved on the E profiles compared to
the estimated experimental values (Fig. 8), even in the near wake where the Reynolds
number is relatively low; then the low R t corrector functions seems to be sufficient.
Howe\'Cl", at the first location (X / [,0 = 2.), the predicted width of the E increase
is larger than the experimental results. These discrepancies might be related to
the owrprediction of Pk , since the production term of the dissipation equation 8 is
modeled through the use of PI,;.

5 Conclusion.
After the numerical studies undertaken at ONERA/CERT /DERAT in Toulouse, it
\vas raised that usual closures, using three or five equations, were able to reproduce
quite well the clewloppement of a manipulated boundary layer for a specific geomet-
rical configuration given from CERT. The goal sought after, in this study, is to check
the predictive capabilities of the CERT code on other experimental configuration. As
detailed experiments were available at CEAT Poitiers, in the downstream vicinity of
an external manipulator having a rectangular section, we ha\'e decided to applied the
CERT code to this specific CEAT geometrical configuration. For computer time sav-
ing, we decided to restrict the present study; the turbulent motion is then calculated
using a three transport equations model (k, c, u'v').
Numerical results have shown a rather good agreement on the evolution of the
mean and fluctuating velocity profiles with experiments, except close to the trailing
edge of the device where some discrepancies occur on the shear stress profiles in
the manipulator wake. As regards to the downstream evolution of the skin friction
coefficient, one can notice that the maximum of C f reduction is overestimated by the
calculations. However, the location of this maximum as well as the relaxation of C f
are rather well predicted. One more time, a remainding difficulty in the calculation
concerns how to model the wall region, including both the upper and lower sides of
12

the device. Let us add that rather same observations have been pointed out when
discussing the numerical results on the CERT-type configuration. A solution might
be to use more sophisticated closures using transport equations in the wall layer.
As experimental balance of the turbulent kinetic energy equation was available at
three locations downstream the manipulator, we have decided to undertake a com-
parison between experiments and modelling on each term of the k equation.
The molecular diffusion is negligible in front of the other terms. The numerical
estimations of convection and turbulent diffusion terms agree quite well with the cor-
responding experimental profiles; the downstream relaxation of these terms is pretty
well reproduced. As regards to production and dissipation profiles, rather good agree-
ment with experiments could also be observed. However, some slight overpredictions
occur at the device height, in the downstream vicinity of the manipulator; this might
be related to the overprediction of U'V ' at the same location. Furthermore, the relax-
ation of Pk seems to be slower than in the experiments; this can be attributed to the
prediction of the downstream evolution of the mean velocity.
At last, let us say that, although the manipulated flow is strongly out of equilib-
rium, conventional closure models are able to predict rather correctly the measured
quantities, particularly the turbulent diffusion term and the dissipation behaviour.

Acknowledgements
This work has benefited from an active collaboration between ONERA/ CERT /DERAT
Toulouse and CEAT Poitiers. The detailed experiments has been recorded at the
CEAT Poitiers while the calculations have been performed at the Aerodynamic De-
partment of the CERT Toulouse.
C. Tenaud thanks specially Professor Jean Cousteix, head manager of the Aerody-
namic Department of CERT, and Doctor Eric Coustols, research engineer in the same
group, for their numerous and valuable comments and dicussions on the modelling of
disturbed boundary layers through outer layer devices.

References
[1] J.P. Bonnet - J. Delville - J. Lemay Study of LEBUs Modified Turbulent
BOllndary Layer by Use of Passive Temperatllre Contamination. International
Conference on Turbulent Drag Reduction by Passive Means, The Royal Aero-
nautical Society, London: RAeS 1 pp.45-68, 1987

[2] D.M. Bushnell AGARD R 723, pp. 5.1-5.26 (1985)

[3] E. Coutols - C. Tenaud - J. Cousteix Maniplllation of Tllrblllent BOllndary


LayeTs in Zero-Pressure Gradient Flows: Detailed Experiments and Modelling.
Turbulent Shear Flows 6, Springer Verlag, Berlin Heidelberg 1989, pp. 164-178
13

[4] E. Coustols - A.M. Savill Rcsume of important results presentcd at the


Third Turbulent Drag Reduction Working Party. Applied Scientific Research 46,
pp.183-196, 1989

[5] B.J. Daly - F .H. Harlow Transport equations in turbulence. Physics of fluids
Vol. 16, p. 157 (1973)

[6] J. Delville - J.P. Bonnet - J. Lemay Analysis of the Wake of an Outcr


Layer Manipulator. 2nd IUTAM Symp. on Structure of Turbulence and Drag
Reduction, Zurich, Springer Verlag, Berlin Heidelberg, 1990 (to appear)

[7] K. Hanjalic - B.E. Launder A Rcynolds strcss model of turb~tlencc and ds


application to thin shear flows. J. Fluid Mech. Vol. 52, part 4, pp. 609-638 (1972)

[8] B.E. Launder - G.J. Reece - W. Rodi Progress in the development of


Reynolds stress turbulence closure. J. Fluid Mech., Vol. 68, Part 3, pp. 537-566
(1975 )

[9] J. Lemay - A.M. Savill - J.P. Bonnet - J. Delville Some Similarities


Betwccn Turbulent Boundary Laycrs Manipulatcd by Thin and Thick Flat Plate
Manipulators. Turbulent Shear Flows 6, Springer Verlag, Berlin Heidelberg 1989,
pp. 179-193

[10] J. Lemay Etude Experimentale du Comportcment de la Turbulcnce dans une


Couchc Limite Imcompressible en Presence d'un Manipulatcur Externe. Ph.D.
Thesis, Faculte des Sciences et de Genic, Laval University, Quebec, ~day 1989.

[11] J. Lemay - J. Delville - J.P. Bonnet Turbulent Kinetic Energy Balance in


a LEBU Modified Turbulent Boundary Layer. Experiments in Fluids. 1990, (in
press)

[12] S.V. Patankar - D.B. Spalding A calculation procedure for heat, mass and
momcntum transfer in thrce dimensional parabolic flows. Int. J. Heat \1ass Trans-
fer, Vol. 15, p. 1787 (1972)

[13] C. Tenaud - E. Coustols - J. Cousteix Modelling of Turbulent Boundary


Layers Manipulated with Thin Outer Layer Devices. International Conference on
Turbulent Drag Reduction by Passive Means, The Royal Aeronautical Society,
London: RAeS 1, pp. 144-168, 1987

[14] C. Tenaud Simulation Numcriquc de l'Ecoulemcnt AUtOUT d'un Manip~tlatcuT'


Extcmc dc Couche Limite. Ph.D. Thesis, University of Toulouse, France, Sept.
1988

[15] A.S.W. Thomas AGARD R 723, pp. 1.1-1.20 (1985)


14

~---.~-~

2.
40 1------+---- ~- --
40 ~
----

4
~- - - --j------+----1

201-----+----, 20
')
0.4
o
o 0.4
lJ 0.8

y (mm)

t
7.5 20.
40 1- -~
401--~--~~--r_~

lL
20 20 f---f--1-----1

lJ~
o OL--L__~~__L_~
o 0.4 0.8 o 0.Lt 0.8

Figure 1: Streamwisc mean velocity profiles


(. experiments [9], [10] ; - three equatioll model)
15

X/oo =0.43 2.
f.O - -~---I

20
~ ~
~~
o L--_~_'-----'
o 5 10 5 10
y (mm)

t
~u'2/Ue(%)
-.-
7.5 20.
40 40 --f-----,----I

20 \
--i
\
~
5 10 10

Figure 2: StreClmwise turbulence intensity profiles


(. experirnents [9], [10] ; - three equation model)
16

X/5 0 =0.43 2.
LI 0 -~f-------. 40

201----il-----+ 20 ~
\~
~
O~ __L-~L_~_=~ o ~
-.002 0 .002 ~002 o .002

y(mm)
-u-,-,/U
t
2
V e
~

7.5 20.
40 40

- -_ . f
4

20
~. 20 \
o
-.002 o
" b
.002
o
f------ - -

-.002 o
" :~ .002

Figure 3: Turbulent shear stress profiles


(. experiments [9], [10] ; - three equation model)
Cf
1.2 tCfref

1.

0.8
XI[)o
o. 10 20 30 40

Figure 4: Evolution of the skin friction coefficient


(. experiments [9], [10] ; - three equation model)

-J
00

y/5
-
1.2 1.2 1.2
XI fJ o = 2, 7.5 20,
t I
0.8 0.8 0.8
I
!
, t !
i ,
~ :> ~ - -
0.4 0.4 0.4 I I

I I I

"-
Ii !
r--
t !
: I
o ~ o i<:LLLLc.L.LJ::.L..:L..L.
i o.
I-
-10 o 10 -10 0 10 -10 o 10
---I.~ Ck , 01 U~ x 1 0 4

Figure 5: Balance of the transport equation of thc turbulellt killetic encrgy :


Convection term (Cd
(. experiments [10], [11] ; --three equation model)
ylB
~~~

1.2 ,--
1.2 1.2
X 10 0 =2. t 7.5 20.
0,8I 0.8 0.8

r. i ~
0.4It--f-d ~ 0,4 0.4
l
lr 't
)
/

o O. o '%
r "/// //hJ

-10 o 10 -10 0 10 -10 o 10


3
Tk.O/U e X10 4

Figure G: Balance of the transport equation of the turbulent kinetic energy :


Turbulent diffnsion term (Tk )
(. experilllcnts [10], [11] ; - - threc eqnation lllodel)

\0
N
o

t y 115 1.2 1.2


1.2
X1Bo= 2. 7.5 20.
4

0.8 0.8 0.8


t
.- \
... -..:- ~6
,~ \
0.4 ,...... 0.4 0.4
~ .~ ~\
"-.!
~ ~~ ~t---
i'e-t-- '/., '/., V//
o. o. o.
o. 10 20 o 10 20 o 10 20
- ~B I U~ X10 4

Figure 7: Balance of the transport equation of the turbulent kinetic ellergy :


Production term (Pk )
(. experiments [10]' [11] ; - - three equation model)
t y I~
1. 2 ,------~~_____r 1.2 1.2
X/Bo= 2.
I I I r
75 20.
0.8 0.8
OBi I I I I !
~.
~ 0.4 0.4 \
04 ~.
- ...oc
i !li ~
~
o.~ o. o --
o 10 20 o 10 20 o 10 20
---tI__ e.o / U~ X 1 0 4

Figure 8: Balance of the transport equation of the turbulent kinetic energy:


Dissipatioll term (c)
(. experiments [10]' [11] ; - three equation model)

N
Manipulation and Modelling of Turbulent Pipe Flow:
Some Parametric Studies of Single and Tandem Ring Devices

A.Pollard H.Thomann
Professor, Professur fiir Str6mungslehre,
Department of Mechanical Engineering, Eidgen6ssiche Technische Hochschule,
Queen's University at Kingston, Sonnegstrasse 3,
Ontario K7L-3N6 Zurich CH-8092
CANADA SWITZERLAND
A.M.Savill
Rolls-Royce Senior Research Associate,
Engineering Department,
University of Cambridge,
Cambridge CB2 IPZ
ENGLAND

SUMMARY

The effect of both single and tandem plate manipulators on developing pipe flow
has been investigated both experimentally and computationally. Numerical calcula-
tions have been performed for nearly developed flow using a finite- volume, elliptic,
low-Re turbulence model code. The results are compared with measured wall pres-
sure, mean velocity, and turbulence intensity profiles, as well as earlier computational
results for similar devices mounted in the initial developing flow region. The numer-
ical results agree closely with experiment and both suggest that nett drag reduction
may only be possible in the fully developed case and over a longer development
length behind a more optimum tandem configuration than could be considered in
the present studies. Further calculations are now being performed to examine this
latter possibility prior to any further experimental investigations.

INTRODUCTION

An extensi\-e research program carried out by a large number of research groups


throughout the world has demonstrated that turbulent skin friction drag of external
boundary layers can be reduced through the use of passive devices such as surface
modifications, including both riblets and other "roughnesses", and by various ma-
nipulators introduced into the flow itself. [e.g. see Wilkinson et al. (1987), Savill et
al. (1988), Savill (1989)]. In both cases apparent optimum device parameters (scaled
on wall units or {j respectively) have been determined and appear to be valid for a
very wide range of flow velocities and conditions. Indeed it is now accepted that ri-
blets offer 6- 7% drag reduction for both marine and flight applications up to at least

23
24

M=1.2 [e.g. see Savill (1989)]. Whether similar benefits may be attainable through
the use of aerofoil manipulators is still a matter for debate and further study, but it
has been shown that these can be constructively combined with riblets and indeed
both devices have been successfully tested in conjunction with other established flow
control techniques such as polymers and blowing [e.g. see Savill (1989), Sellin &
Moses (1989)]. At the same time considerable interest has arisen in the possibili-
ties of using manipulators, and also riblets, for other purposes e.g. to reduce wall
pressure flnctuations, and hence noise, and to control heat transfer and mixing.
Initial attempts to model manipulated boundary layers with simple mixing length
and k-E models have led rapidly to more sophisticated treatments incorporating stress
transport closures [Savill (1989)] and even, most recently, some Large Eddy Simula-
tion results [Friedrich & Kline 1989,1990]. Parabolic stress transport computations
ha\'e been used successfully; first to post-dict experimental results for thin single
and tandem flat plate devices in zero and adverse pressure gradient boundary layers,
including the influence of free-stream turbulence; then to predict the effect on such
manipulated layers of larger free-stream turbulence intensities and length scales rep-
resentative of conditions within aero-engines [Savill 1987], and streamline curvature
[Savill (1988)], for which there are as yet no experimental data. At the same time
an elliptic scheme, starting upstream of the manipulators and employing k-E, or 3-
or 5-equation turbulence models has been developed for ONERA-CERT [Coustols
et al. (1987b)] and used, again successfully, to predict the optimum height and
spacing of thin tandem plate manipulators for low Reynolds number operating con-
ditions. The same approach is now being extended to curvilinear co-ordinates in
order to make parametric optimisation computations for the more robust aerofoil
devices which will be required in any practical application. Some extensive results
have already been obtained for N ACA 0009 and ONERA D profile devices in laminar
flow conditions [Djenidi et al.(1990a)], for which there is experimental evidence that
Tollmien-Schlicbting wave growth can be delayed by manipulators [Bardakhanov et
al. (1989)], and these are presently being compared with the only other available
'data' from simulations performed by Friedrich & Klein [1990]. Some initial tur-
bulent computations are also now also being performed by Djenidiet al. (1990b)
for comparison with the experimental results of Coustols et al. (1987a). It is clear
that there are still some deficiencies in the model approximations nsed, particularly
for the near-wall, near-device, and device-wake regions. To overcome these waIl
function or mixing length models, low-Re closures are being used. Indeed initial
low-Re k-E results have already been obtained by Johnston (1987) and Shenxi &
Zhou (1989), using the model of Chien (1982), as well as by Tenaud (1988) using
the Launder-Sharma (1974) model, and in future it seems likely the introduction of
newly-developed 2-dimensional, 2-component limit models [e.g. see Fu et ai. (1987)]
will lead to a further improvement in predictions, particularly since it has already
been demonstrated that these can also improve predictions for wakes.
The concept of using such passive devices to beneficially control external turbu-
25

lent boundary layers is thus well established, but surprisingly, particularly in view
of the possible applications to pipelines and ducting, considerably less attention has
been paid to internal flow manipulation. In fact riblet performance levels similar to
those found in external flows have recently been recorded in a fully developed duct
Lowson et al. (1989) and in a range of pipe flow experiments, Dinkelacker et al.
(1987), Reidy and Anderson (1988), Rohr et al. (1989), Liu et al. (1989); even for
the case of two-phase flow [Kelman et al. (1989)). By contrast, manipulator experi-
ments which have been conducted in fully developed channel flow at NAL Bangalore
[by Prabhu et al. (1987), Vasudevan et al. (1990)), and in developing pipe flow, at
ETH Zurich [Pompeo & Matievic (1987a,b), Lineton (1988), Samec (1989)) and at
Laval University [by Dickinson & Nguyen - see Coustols et al. (1990)), have revealed
no nett benefits. However only a limited number of device configuration parameters
have so far been investigated, and it is not clear what should be the optimum values
for these, or even how they should be scaled in the case of fully developed flow when
fj can no longer be used as a reference. So far it has generally been assumed that
this can simply be replaced by the duct/channel half-height or the pipe radius, but
the NAL group's finding of a best result for a device mounted at the duct centre-
line casts some doubts on this. Despite the fact that further improvements are still
required to near-wall/near-device modelling, there is clearly some advantage to be
gained from applying current computational schemes to internal flows in an attempt
to resolve this scaling question and to help circumvent the need for very extensive
and time-consuming parametric experimental optimisation studies. Indeed, prelimi-
nary low-Re k-e elliptic computations, performed previously by Pollard et al. (1989),
have already successfully post-dicted some of the ETH experimental results for single
ring manipulators mounted in the developing flow region of the same pipe.
The purpose of this paper is to present new low-Re k-e predictions for both
single and tandem ring manipulator configurations in nearly developed pipe flow
and to compare these with the earlier computational results for developing flow
and a wider range of experimental data obtained at ETH Zurich. The aim of the
present study has been to try to establish the criteria for optimising devices in pipe
flows, before conducting any further experimental studies, in order to minimise the
parametric measurements required to determine the maximum possible benefit that
can be achieved by such manipulation. As such, this work forms part of a much
wider collaborative program on turbulent drag reduction and modelling involving a
far larger number of research groups both within the European Research Community
on Flow Turbulence and Combustion (ERCOFTAC) and in Canada [see Lemay et
al. (1990)].

EXPERIMENTS

The ETHZ experiments were conducted in a pipe 100mm in diameter (D=2R)


and 6000mm long at a mean velocity of 24m/s (ReD = 180,000). Various ring
manipulators were mounted in the pipe at a number of locations (X/D=5,18 & 28)
26

downstream from the inlet. The manipulator rings were of different radii (r) so as to
vary the distance (h=R-r) of these from the pipe wall, and were of metal construction,
with a cord length (1) of 20mm and a separation for the tandem arrangements of
135 or 35 mm; each ring being supported from the internal surface of the pipe
by three, thin, equally-spaced struts. The basic geometry considered in both the
experiments and the computations was thus as indicated by Figure 1, although in
the experiments the manipulators were tapered with their thickness (t) falling from
0.3mm at the leading (rounded) edge to 0.15mm at their trailing (sharp) edge.
The pipe walls were instrumented with pressure taps and mean velocity and
turbulence data were recorded by means of Pitot (employing a Macmillan correction
for shear and wall inflence effects) and hot-wire probes respectively. Data collected
included axial wall pressure distributions, and radial mean velocity and turbulence
intensity profiles at a circumfirential location mid-way between two of the support
vanes.
The range of parameters covered by the experiments are indicated in Table 1.
Note that for the case of manipulators introduced into the initial developing flow
region. where a wall layer thickness 8 could still be identified, these included the
apparent optimum values of: h;::::O.78, 1;::::1.58, and for the tandem case: s;::::108 sug-
gested for external flow applications. The corresponding device parameters for the
similar fully developed pipe flow experiments conducted subsequently at Laval, and
channel flow investigations performed at N AL, are also included in Table 1, together
with those adopted for the Large Eddy Simulations performed at Technische Uni-
versitiit l\hinchen. It can be seen that, a1though similar Reynolds numbers and wall
separations have been investigated in each case, the ETHZ manipulators were rather
shorter and more closely spaced than the devices studied at either Laval or N AL
(when all are scaled in terms of pipe or duct diameter for fully developed flow con-
ditions). The total development length of 120R, and hence the maximum available
reCO\'ery length, was also rather less for the ETHZ experiments than for the other
experimental studies, and the flow only attained a fully developed state at X=55D
(although the initial wall layers had merged by x;::::30D).

TURBULENCE MODELLING

The computations were performed using a low-Reynolds number k-t model, in-
corporated into a finite-volume code with a non-uniformly distributed staggered grid,
which has previously been validated for turbulent pipe flow by Pollard and Mart-
inuzzi (1989a,b). In view of their recommendations, based on comparative tests of
various k-t and stress transport models, the Lam & Bremhorst (1981) low-Re k-t
model was adopted. Since this low- Re model accounts for molecular viscosity effects
by introducing a damping function II' for the eddy viscosity which varies between
o at a solid surface and 1 in the fully turbulent region, the distance from any such
surface must be specified uniquely. For the manipulated flow configurations this dis-
27

tance was evaluated by taking the smallest distance between any point in the flow
and the nearest solid boundary. (The support struts were ignored).
The code utilised hybrid-differencing for the convective terms since Martinuzzi
and Pollard, and many others have shown that for pipe flow this results in negligible
false-diffusion. However placing a manipulator within the flow introduces a wake, the
spreading of which may be over-estimated by the use of such a one-sided differencing
scheme [e.g. see Pollard and Siu (1982)]. Fortunately this effect can be reduced by
using a grid sufficiently fine enough to ensure that central differencing is recovered
when the grid Peclet number is less than 2. Thus for the present computations the
number of grid points was increased to 297 (in x) by 98 (in 1') from the 147x68 mesh
used for the earlier developing pipe flow calculations. The grid \vas further refined
near the wall, 40% of the points being below y+=50 with the first at y+ :::;1. In
addition, for some of the present computations the number of iterations was about
1000 resulting in a factor 20 improvement in normalised residuals for axial and radial
momentum reported earlier. While these additional 400 iterations produced only
negligible differences in other flow quantities (much less than 1%), the pressure field,
particularly in the region downstream of the manipulators, was altered by about 5%.
This change occurred early in the extra iteration count, thus it is believed a fully
converged solution for all variables was obtained.
The manipulators were delineated by 6 control volumes across their thickness,
which was assumed constant at 0.3mm, and 20 control volumes along their length,
so that at both the leading and trailing edge the first grid point away from the device
was located at approximately one wall unit from its surface. In addition care was
taken to ensure that the control volumes bordering the manipulators were suitably
modified to account for the presence of the blockage. (Note that all estimates in wall
units are based on the value of U T for the undisturbed fully developed pipe flow).
The wall boundary conditions used in the present calculations were the same
as those defined by Martinuzzi and Pollard (1989a,b). The inlet conditions required
some modification in order to achieve conformity with the mean axial velocity profiles
measured for the initial plane, un-manipulated pipe flow. The "best fit" agreement
was found when a standard log-law (constants, A=2.,5, B=5.5) was specified from
the wall to y+=250, with a linear sublayer below y+=l1.5, and a constant velocity
between y+=250 and the pipe centre-line. In the absence of any data, the inlet
conditions for the turbulence kinetic energy and its dissipation rate were assigned
uniform values across the pipe radius: k=0.002 U~verage and f=Cl'k 3 / 2 /O.02R. Typical
computations for 1000 iterations on the 297x98 grid required 137 CPU minutes on
an IBM 3081G.

COMPUTATIONAL RESULTS
Increasing the grid resolution resulted in a clear improvement in the model per-
formance, even for the plane or unmanipulated pipe flow, as illustrated by the com-
parison of predicted and experimental Cp (== 1 f'j2- P , ) distributions for the initial
2 P average
28

developing flow region presented in Figure 2. The resulting Cp development, with and
without a single manipulator introduced into the nearly developed flow at x=28D, is
shown in Figure 3. These computations, performed over 600 iterations, reveal some
discrepancies with the experimental data for the more developed region of the plane
pipe flow. This was still evident, but somewhat diminished, after the additional
400 iterations were performed to ensure more complete convergence on the pressure
field. A further indication of the performance improvement with grid refinement
for the un-manipulated case is provided by the kinetic energy profiles at X=30D
plotted in Figure 4. It can be seen that the higher resolution computations are in
good agreement with experiment bearing in mind the uncertainty involved in both
of these.
'-Vith the same refined mesh the computed mean velocity defect (manipulated mi-
nus un-manipulated U) close behind a single manipulator introduced into the nearly
developed flow at X=28D from the pipe inlet (leading edge of device at X=27.3D)
were also in considerably better agreement with the experimental data than previous
lower-resolution computations of Pollard et al. (1989) - see Figure 5. The lack of
significant improvement in the calculations for X=10.4D is quite probably due to a
too rapid spreding of the wake; however, the error in the experimental data, obtained
with a Pi tot tube, can also be exagerating the discrepancy. The corresponding delta
Cp (manipulated minus un-manipulated Cp ) distribution is plotted on Figure 6 and
is also in respectable agreement with the experimental results (initial 147x68 grid
computations overpredicted the delta Cp to the extent that the results would lie just
above the domain covered by the figure). The dashed curve in Figure 6 indicates the
somewhat improved results obtained after 1000 iterations as compared with those
after only 600 iterations (solid curve). Again both computed curves fit the exper-
imental data within the experimental and numerical uncertainties (note that the
experiments also indicated that the Cp development was unchanged when ReD was
increased by a factor of 2-3). In addition the computations reproduced the trend of
all the experimental results up to X=50D, except in the initial region immediately
downstream of the device. A similar discrepancy is evident in high-Re k-E computa-
tions for external boundary layers performed by Coustols et al. (1987), and must be
attributed to defficiencies in the modelling of this region with such closure schemes.
The quality of the predictions obtained with the highest grid resolution/number
of iterations is better appreciated when these are replotted on a different scale for
comparison with similar experimental and computational results for an equivalent
tandem maipulator configuration (one with the first plate at the same location as
for the single device just considered and with the same parameters for both plates,
but s=2.7R), as in Figure 7. However it is clear that the tandem configuration is not
so well predicted, and that the computations indicate the effect of introducing the
second plate is simply to double the pressure drop due to the single plate manipula-
tor, presumably due to the increased blockage (in both cases only about half of the
pressure drop due to the device appears to be recovered), whereas the experiments
29

show that the penalty associated with the second device is less than for the first. In
fact the pressure was clearly still relaxing even at the end of the test section in the
tandem case (65R downstream of the device) where the .6.C p had fallen to the level
attained behind the single plate [see Pollard et al. (1989b)]. For the latter it ap-
peared recovery was completed within 40R, a similar recovery length to that found by
the N AL researchers. Unfortunately, storage limitations (8Mb) restricted the extent
of the flow development that could be considered in the computations and so it was
not possible to discover if the trend of the experimental results in the fully developed
section of the flow could be reproduced by these although the indication from the
shape of the curves presented in Figure 7 is that this was unlikely. The reason for
such a discrepancy is unclear at present, but is probably partly related to the well-
known defficiency of current transport models, based on local equilibrium closure
approximations, that they tend to recover too rapidly to equilibrium conditions and
thus underpredict the overall effect of disturbances imposed on flows [c.g. see Savill
(1987)]. However the fact that, for devices introduced into the initial developing flow
region, recovery seems to have been completed within 60R, or approximately 2008
(a distance comparable with that found for external boundary layer manipulation
by similar device configurations and reproduced by model computations - see Fig-
ure 8), together with fact that the expected shape of the turbulence profiles in the
immediate vicinity of the device were captured well by the calculations (see Figure
9), suggests there is some additional problem with the physical modelling of the
manipulator influence in the fully developed flow situation. It has previously been
generally thought that the lack of an entrainment surface under these conditions
might preclude any nett drag reduction (and also perhaps simplify the modelling),
because the important suppression of such outer layer effects seen in external bound-
ary layers would be missing. However, at least for fully developed duct flow, the
experimental studies of Antonia and Teitel (1989) and others have shown that this
external influence is replaced by an equally significant interaction between the flow
in the two halves of the duct; large scale motions from one side moving across the
centre-line to disturb the flow on the opposite side wall.
It is evident from an analysis of the initial low resolution 32x16x16 Large Eddy
Simulation results obtained at T.U. Miinchen by Friedrich & Klein (1989a) that ma-
nipulators not only influence the flow structures in the near wall region, but also
in the core flow; both reducing the scale of these and suppressing any cross-flow
interactions for some distance behind the devices [see Savill & Djenidi (1990)]. This
might help explain why the N AL researchers obtained their best result for a device
mounted on the centre-line of their duct. However it is significant that their next best
result was for two single plates mounted at h=0.38R. Considering all of the experi-
mental/LES data and computational results obtained so far it would seem that the
optimum height for essentially developed flow conditions may be of order h=0.25R.
It is also evident, from the different best results obtained at Laval (whole device
drag recovered), N AL (device drag almost recovered) and ETHZ (approximately half
30

device drag recovered) using similar height single plate devices, with progressively
smaller l/R, that a manipulator chord length of at least I.25R (the value adopted
for the most detailed LES study) is required for optimum performance, and further
model computations are now being performed to evaluate the benefits of such longer
devices [Pollard ct al. (1990)]. Regarding the question of the optimum spacing it
seems clear from the comparison of the tandem results from N AL (nett drag twice
that for single device) and ETHZ (minimum nett drag approximately less than or
equal to single device for s=2.7R, but greater than single for s=0.7R), that simply
re-scaling the optimum value found for external applications by the initial devel-
oping wall layer thickness, or by half the duct or pipe diameter, is insufficient to
maximise the benefits of tandem manipulators for internal flows. Instead a much
closer non-dimensional spacing of around 3R or 8 appears to be much more bene-
ficial. Howe\u if one studies the predicted C f (== 1 ~~'all ) distributions for both
2 P average
single and tandem de\'ices (which were not measured in the experiments) as plotted
in Figure 10 severa1 points emerge. First considering the whole C f development for
the case of a single manipulator it is evident that, while the downstream C f assymp-
totes to the Blasius profile va1ue as expected, this is actually a higher ("overshoot")
va1ue than that predicted immediately upstream of the device. Furthermore when
the C f distribution in the immediate downstream vicinity of the single manipulator
is examined in greater detail and the equivalent distribution computed when the
second plate of the tandem with s=I.35D is superimposed (see (0) curve on Figure
10, lower figure) it is clear that there is an upstream influence of this second plate
which pre\'ents the minimum C f due to the first plate alone from being attained
before the effect of the second plate (leading edge at X=28.65D) is felt by the flow.
Clearly the tandem configutaion could be improved if the second plate were shifted
downstream. To test this suggestion some additional computations have now been
performed for a similar plate configuration, but with the leading edge of the second
plate at first X=28.9D, and then X=29.4D. The results are presented in Figure II.
Assuming that minimising the overall C f is the most important criteria for achieving
optimum nett drag performance, these suggest that a spacing of s=2D (or 4R) can
be recommended. In the present case this corresponds to 101 which is similar to the
external flow value, but is not clear whether this is significant since this condition
was nearly satisfied for the N AL tandem experiments also. It seems more likely that
s should also be scaled on D or R to maintain a close physical presence of the two
plates particularly as the immediate effect of the device on Cp persists for only 0.6R
(as opposed to approximately 58 for an equivalent device: 1=0.48, h=0.28 t=0 ..5mm
in external flow) and the maximum C f reduction of 5% occurs only about 3R from
the de\'ice (as opposed to 1.5% at 88), although the recovery lengths of 40R and 508
are comparahle.

CONCLUDING REMARKS
The results of the present work can be summarised as follows.
31

New computational and experimental results have been presented for turbulent
pipe flow manipulated by both single and tandem ring devices. It appears that
the low-Re k-c model employed is sufficient to produce predictions in reasonable ac-
cord with the experiments or at least equivalent to the level of agreement achieved
previously in external flows. The results, when considered together with earlier com-
putational and experimental results including data obtained elsewhere, suggest it is
unlikely that any nett benefit could be obtained in developing pipe flow, but that
this may still be possible in fully developed flow given an optimum tandem manip-
ulator geometry and sufficient development length in which to recover the device
drag. In particular it would appear that (when scaled in terms of R rather than
8) the non-dimensional height and spacing of the devices both need to be reduced
relative to their optimum values for external flow operation, although a similar,
non-dimensional, chord length may be appropriate. For the best present case the
development length both upstream and downstream of the manipulators was in-
sufficient to determine whether any nett benefit could be obtained, but there was
little difference in the slope of either the measured or computed longitudinal pres-
sure gradients indicating that any drag reduction was less than the experimental or
numerical uncertainty. It is hoped that further improvements to the understanding,
modelling and optimisation of internal flow manipulation will come from a more
detailed analysis of higher resolution 64x32x32 and 192x32x32 LES 'data' currently
being performed by Savill it et al. (1990).

ACKNOWLEDGEMENTS
Professor Pollard acknowledges the financial support of the Natural Science and
Engineering Research Council of Canada, Queen's University Computing Centre and
The School of Graduate Studies of Queen's University. Dr.Savill acknowledges the
continuing support and interest of Rolls-Royce pIc. The experimental work was
unfunded and performed by students under the supervision of Professor Thomann
at ETH, Zurich. Additional support and computing resources were provided by the
EPFL ERCOFTAC Pilot Centre in Lausanne, Switzerland as part of a collaborative
ERCOFTAC project, and these are also gratefully acknowledged.

REFERENCES
R.A. Antonia and M. Teitel (1989): Organised motion in a fully developed tur-
bulent duct flow, Proceedings, 10th Australasian Fluid Mechanics Conference, Mel-
bourne, Decemder.
S.P. Bardakhanov, V.V. Kozlov and V.V. Larichkin (1989): Influence of an acous-
tic field on the flow structure behind a LEBU in a turbulent boundary layer. Poster
Paper at IUTAM Symposium on Turbulence Structure and Drag Reduction, Zurich.
F. Chen, Y-P. Tang and M-Z. Chen (1990): An experimental investigation of loss
reduction with riblets on cascade blade surfaces and isolated aerofoils. Proceedings
ASME Turbo Expo - Land, Sea & Air, Brussels.
32

K.-Y. Chien (1982): Predictions of channel and boundary layer flows with a low
Reynolds number turbulence model. AIAA J. 20, p.33-38.
E. Coustols, J. Cousteix and J. Belanger (1987a): Drag reduction performance on
riblet surfaces and through outer layer manipulators. Proceeding RAeS International
Conference on Turbulent Drag Reduction By Passive Means. 2., pp.250-289.
E. Coustols, C. Tenaud and J. Cousteix (1987b): Manipulation of turbulent
boundary layers in zero pressure-gradient flows: detailed experiments and modelling.
Proceedings 6th Symposium on Turbulent Shear Flows, Toulouse, Paper 11-5:1-5.
E. Coustols and A.M. Savill (1988): Resume of important new results presented at
the Third Turbulent Drag Reduction Working Party, ONERA/Chatillon, September
29-30. Appl.Sci.Res. 46, pp. 183-196.
E. Coustols, A.M. Savill and T.V.Truong (1990): Report on the 4th European
Drag Reduction Working Party, Lausanne PC. ERCOFTAC Bulletin IV, p.8-12.
J. Dickinson, V.D. Nguyen, Y. Chalifour, A. Smaili, A. Page and F. Paquet
(1989): Boundary layers over a ribleted surface using tandem turbulence manipula-
tors. Proceedings 1st Canadian Symposium on Aerodynamics, Ottawa.
A. Dinkelacker, P. Nitsche-Kowsky and W.-E Reif (1987): On the possibility of
drag reduction with the help of longitudinal ridges in the walls, Proceedings IUTAM
Symposium on Turbulence Management and Relaminarisation, Bangalore, (Springer-
Verlag) pp. 109-120.
L. Djenidi, A.lVI. Savill and L.C. Squire (1990): Parametric computational study
of aerofoil manipulator performance in laminar flow. For Proceedings Volume 5th
European Drag Reduction Working Party Meeting, BMT London (Kluwer Academic
Publishers; Ed: K.-S.Choi).
1. Djenidi, C. Tenaud, E. Coustols and A.M. Savill (1990): Comparison of com-
putations with experiments for aerofoil manipulators in turbulent flow. For Proceed-
ings Volume 5th European Drag Reduction Working Party Meeting, BMT London
(Kluwer Academic Publishers; Ed: K.-S.Choi).
R. Friedrich and H. Klein (1989a): Large scale turbulence structures in a manip-
ulated channel flow. Proceedings IUTAM Symposium on Turbulence Structure and
Drag Reduction, Zurich. Springer-Verlag.
R. Friedrich and H. Klein (1990): Manipulating large scale turbulence in a chan-
nel. Private communication.
S. Fu, B.E. Launder and D.P. Tselepidakis (1987): Accommodating the effects
of high strain rates in modelling the pressure-strain correlation. U11IST Mech. Eng.
Dept. Report TFD /87/5.
A. Johnston (1987): Private communication.
J.S. Kelman, 1. Shearer and J. Hall (1989): Riblet surface pressure losses in single
and two phase flow. See Sellin & Moses (1989).
H. Klein and R. Friedrich (1989b): Manipulating large-scale turbulence in a
channel and a boundary layer. Proceedings 7th Turbulent Shear Flows Symposium,
Stanford p.25.4.1-6.
33

C.K.G. Lam and K. Bremhorst (1981): A modified form of the k-E model for
predicting wall turbulence, ASME, Journal of Fluids Engineering, 103, pp. 456-460.
B.E. Launder and B.I. Sharma (1974): Application of the energy dissipation
model of turbulence to the calculation of flow near a spinning disc. Letters in Heat
and mass Transfer, 1, p.131-138.
J. Lemay, A. Pollard, J.P. Bonnet and A.M. Savill (1990): International co-
operative research programme in the area of turbulent boundary layer manipulation.
7th Canandian Conference on Engineering Education, Toronto, June 27-28.
B. Lineton (1988): Unpublished ETHZ data.
K.N. Liu, C. Christodoulou, O. Riccius and D.D. Joseph (1989): Drag reduction
in pipes lined with riblets. Proceedings, 2nd IUTAM Symposium on Structure of
Turbulence and Drag Reduction. Zurich, Switzerland, July (Springer- Verlag).
M.V. Lowson, A.D. Jewson and J.H.J. Bates (1989): Drag reduction by riblets
in a duct flow. See Sellin & Moses (1989).
R. Martinuzzi and A. Pollard (1989): A comparative study of turbulence models
in predicting turbulent pipe flow. Part I: algebraic stress and k-E models, Part II:
Reynolds stress and k-E models. AIAA Journal, 27, No.1, pp. 29-:36 and 27, No.
12, pp. 1714-172l.
A. Pollard, A.IVI. Savill and H. Thomann (1989a): Turbulent pipe flow manip-
ulation: some experimental and computational results for single manipulator rings,
Journal of Applied Scientific Research, 46, No.3, July, pp. 281- 290.
A. Pollard, A ..M. Savill and H. Thomann (1989b): Turbulent pipe flow manipu-
lation: some experimental and computational results for tandem manipulator rings.
Proceedings 10th Australasian Fluid Mechanics Conference, Melbourne. Vol. II paper
9C.3
A. Pollard and A.L-W Siu (1982): The calculation of some laminar flows us-
ing various discretisation schemes, Computer Methods in Applied l'vIechanics and
Engineering, 37i, pp. 293-313.
A. Pollard, H. Thomann and A.M. Savill (1990): Further computational results
for manipulated pipe flow. Proceedings Volume 5th European Drag Reduction \Nork-
ing Party Meeting, BMT London (Kluwer Academic Publishers; Ed: K.-S.Choi).
1. Pompeo and T. Matievic (1987a): Turbulenzbeeinflussung, Semesterarbeit in
Fluiddynamik, ETH Zurich, Institut fur Aerodynamik Report WS 1986/87.
L. Pompeo and T. Matievic (1987b): Investigation of drag reduction in manipu-
lated turbulent boundary layers. ETHZ Institut fur Aerodynamik Internal Report.
A. Prabhu, P. Kailas Nath, R.S. Kulkarni and R. Narasimha (1987): Blade ma-
nipulators in channel flows. Proceedings IUTAM Symposium on Turbulence Man-
agement and Relaminarisation. Bangalore, Springer-Verlag, pp. 97- 108.
L.W. Reidy and G.W. Anderson (1988): Drag reduction for external and internal
boudary layers using riblets and polymers, AIAA-88-0138, 26 Aerospace Sciences
Meeting, Reno, Nevada.
J.J. Rohr, L.W. Reidy and G.W. Anderson (1989): An experimental investigation
of the drag reducing benefits of riblets and polymers in pipes. See Sellin & Moses
(1989).
K. Samec (1989): Investigation of drag reduction in manipulated developing pipe
flow. ETHZ Institut fur Fluiddynamik Internal Report.
A.M. Savill (1987): Algebraic and Reynolds stress modelling of manipulated
boundary layers including effect of free-stream turbulence. RAeS International Con-
ference on Turbulent Drag Reduction by Passive Means, 1, pp. 89-143.
A.lIt Savill (1988): Structural refinements to Reynolds stress closures for plane,
manipulated and wake-boundary layer flows. Proceedings, 3rd International Sympo-
sium on Refined Flow 110delling and Turbulence Measurements, Tokyo (Universal
Academy Press Inc.).
A.:\I. Savill (1989): Recent developments in turbulent drag reduction by pas-
sive means. Rcview lecture paper for the 2nd. IUTAM Symposium on Structure
of Turbulence and Drag Reduction, Zurich, Switzerland, July, to be published by
Springer- Verlag.
A.M. Savill and L. Djenidi (1990): Lausanne ERCOFTAC Pilot Centre Visitors
Report.
A.:\1. Savill, T.V. Truong and 1.L. Rhyming (1988): Turbulent drag reduction by
passive means: a review and report on the first European drag reduction meeting.
Journal de lIIecanique 1 (4), pp. 353-378.
A.:\1. Savill, H. Klein and R. Friedrich (1990): Pattern recognition analysis of
structure in manipulated channel flow simulations. For Proceedings Volume 5th
European Drag Reduction Working Party Meeting, BMT London (Kluwer Academic
Publishers; Ed: K.-S. Choi).
R.J.Scllin and R.T. rvloses (1989): Drag Reduction In Fluid Flows: techniques for
friction control. Proceeding, IAHR International Conference DRAG REDUCTION
89 on the Fundamentals and Applications of Drag Reduction Techniques: Additives
and Passive Devices. Davos, Switzerland, (Ellis Horwood).
S. Shengxi and Zhou. 1!J.D. (1989): The influence of plate manipulator wake on
skin friction. (Unpublished Report from Nanjing Aeronautical Institute, China)
C. Tenaud (1988): Unpublished ONERAjCERT computational results.
B. Vasudevan, A. Prabhu and R. Narasimha (1990): Blade manipulators do not
reduce pressure losses in turbulent duct flow. To appear.
S.P. Wilkinson, J.B. Anders, B.S. Lazos and D.M. Bushnell (1987): Turbulent
drag reduction research at NASA-Langley: Progress and plans. In: Turbulent Drag
Reduction by Passive Means, Royal Aero. Society, Sept.
35

ETHZ ReD=180,000 D=lOOmm l=20mm t=0.3~0.15mm t max =1.5% 1


x/D 8(mm) 81R 1/8 h/8 s18,
5 14.0 0.28 l.4 l.07, 0.71, 0.36 9.6
18 28.7 0.57 0.7 0.54, 0.36, 0.18 4.9 (or 1.25)
28 40.5 0.81 0.49 0.37, 0.24, 0.12' 3.3
30 50.0 1.0 0.4
55 Fully Dev.
60 End of Pipe
Laval ReD=140,000 D=55.5mm l=25mm t=0.3 & 0.15mm t max = 1.2%1
x/D 8(mm) 81R 1/8 h/8 s/8
Fully Dev. 1.0 1.2 0.18 & 0.57
+75 End of Pipe
NAL ReD=145,000 D=50mm 1=22mm t=0.26mm t max =1.2% 1
x/D 8(mm) 8/R 1/8 h/8 s/8
73 25.0 1.0 0.8 1.00, 0.68, 0.38 -, 10, 10
+50 End of Duct
TU Mu. ReD=150,000 t=Omm
x/D 8(mm) 81R 1/8 h/8 sl8
Fully Dev. 1.0 0.8 0.75
Fully Dev. 1.0 1.25 0.5
Fully Dev. 1.0 1.5 0.31, 0.5, 0.75
+lOmax End of Duct
Also at ReD= 370,000 and 450,000.

Table 1: Manipulator Parameters Investigated.


36

u R r

Figure 1: Basic geometry, not to scale.

0t--------------------------
I,
G ';
-0,1

-0.2 \
\
.,
-0,

-0] :al::~aet~~:~~tal Data


-0'1 _-0 __ 68 147 \, \ I
I --- 98' 267
-r'--.--,-,,,'-----,,----,-,-r,
-O.6't',----" ----"-,
--.----c--,'~,----,-,

12 16 20 24 28
X/D

Figure 2: Axial wall pressure distribution, comparison of experimental data and


calculations for two grid distributions.
37

0.1
0-
C>

-0.1

-0.2

-0.3 98 * 297 grid


-0.4

-0.5 o Experiment
-0.6

-0.7
+ Calculation Pipe Flow

-o.B o Calculation Single Manipulator


-0.9
20 40
X/D

Figure 3: Cp vs. X/D for pipe flow with single manipulator with leading edge at
X/D=27.3, r/R=O.8, Re=180,OOO.

+ +
"+
"+
"
"
o Experimental data (Laufer) (Fully developed)

o
o
<>
Calculation 68 * 147

+ Calculation 98 267

20 40 60 BO 100

y+ (=y Utau / nul

Figure 4: Comparison of calculated turbulence kinetic energy profiles at X/D=30


to Laufer's fully developed pipe flow data.
38

0.03 , - - - - - - - - - - - - - - - - - - - - - - ,
I~ ~:~i ::t=:--+=::-=:::-:==-+-----~-_-_-
~ -OO~ +--'-------=--:...------~-\---.,~---.::j.

<l =~:~5
-0.04
-0.05
-0.06 X"/D~O.4
-0.07
-0.08
-0.09 +------::-'C,----~--~-~-~-+-_--1
0.2 0.4 0.6 0.8

0.025
0.02
0.015

-~--
0.01
0.005 f--- ----
0

~
-0.005
-0.01 X"/D~2.4
-0.015
-0.02
0.2 0.4 O.S 0.8

0.02
0.018
0.016
0.014
0.012
0.01
0.008
0.006
0.004
.
0.002
-0.002
-0.004
0
X"/D=10A
---- ~
-0.006
0.2 0.4 0.6 0.8
r/R

Figure 5: Normalised velocity difference (with minus without manipulator) down-


stream (X" /D) of single manipulator with leading edge at X/D=27.3, r/R=O.S,
Re= ISO ,000. ( .) Experimental data, (-) Calculations, 9S* 297 grid, (- -) Calcula-
tions, 6S*147 grid.
0.022

~
0.021
"'-
. ,,+
0.02 Experimental Data
0.019
~ 0.018 1 0.7
0.017 -+- l ' O.B
5 0.016
0.015 +
'\ c o l ' 0.9
~ 0.014 c \
0- 0.013
u 0.012
.3 0.011
0.01
'"" 0.009
0.008
+---r--r:--
l + .B3 Ii
0.007
0.006 Calculations (1 O.B)
0.005
0.004
0.003 600 Iterations
0.002
0.001 1000 Iterations
0
30 31 32.5 35.5 38.5 40.5 42.5 44.5 46.5 48.550.7554.75 57 60
X/D
Figure 6: Cp difference downstream of single manipulator with leading edge at
X/D=27.3, r /R=O. 7, O.S and 0.9. (-) Calculations 600 iterations, (- -) Calculations,
1000 iterations.
39

0.07

ExpL Data Calculations

o +-
0.05
Single at X/D = 27.3 -
~
:5
0.05 !:. Tandem at X/D = 27.3 & 28.65 - 0 -
;;;
0.04-
:5
~ 0.03
A
u
0.02
-<l

0.01

28 32 36 40
44 X/D 48

Figure 7: Cp difference downstream of single and tandem manipulators, compar-


ison of experimental data and calculations.

0.0< , - - - - - - - - - - -_ _ _ _ _,
rlpHlmental
0.035 o hO.7
olxO.8
~
, 0.03 -+hO.9
1.21:0.8
~ 0.025 _ _ 1:rO.7Ca!cs

:5 0.02

GO.015

~ 0.01

0.005
1-1-00C25

0.00 20.00 40.00 60.00


x/ 0

Figure 8: Cp difference for single manipulator at either r/R=0.7, 0.8 or 0.9 and
tandem manipulators at r/R=0.8 (axial separation of 1.35D). Comparison of calcu-
lations with data for single manipulator at r/R=0.7

o X/D=4.900

o XjD::5.100

6. X/D=5.201

X X/D:::5.250
I
~ I -=i=i I ,I

1

j
0.6 0.7 0.8 0.9
r/R
Figure 9: Radial turbulence kinetic energy distribution prior to, on and after a
single manipulator with leading edge at X/D=5, r/R=0.7, Re=180,000.
40

4.50E-03
Cf 4.40E-03
4.30E-03
4.20E-03
4.10E-03
4.00E-03 Blasius Cf=0.00384
3.90E-03
3.80E-03
3.70E-03
3.60E-03
3.50E-03
HOE-03
20 40
4.50E-03
Cf 4.40E-03
4.30E-03
4.20E-03
4.10E-03
4.00E-03
3.90E-03
3.80E-03
3.70E-03
3.60E-03
3.50E-03
HOE-03
27 27.4 27.8 28.2 28.6 29 29.4 29.8

X/D
Figure 10: C J distribution vs. X/D. Top figure - single manipulator at X/D=27.3.
Bottom figure - Window of 27 ::; X/D ::; 30 (D) single manipulator, (< tandem
manipulator. Leading edges at X/D=27.3 and 28.65, r/R=0.8, Re=180,000.
0.0043

Cf
0.0042 I.
0.0041

0.004

0.0039

0.0038

0.0037

0.0036

Ii
0.0035
20 24 28 32 36 40 44 48
X/D
Figure 11: C J predictions for single and two different tandem arrangements. (-
+-) single, leading edge at X/D=27.3; (< Tandem, leading edges at X/D=27.3
and 28.9; (- - -) Tandem, leading edges at X/D=27.3 and 29.4. All for r /R=0.8,
Re=180,000.
Large-Eddy Simulation of Manipulated Boundary Layer and Channel
Flows

H.Klein & R.Friedrich


Institute of Fluid Mechanics
Technical University of l11unich
Federal Republic of Germany

Abstract
The large-eddy simulation (LES) technique is used to study boundary layer and
channel turbulence at high Reynolds numbers altered with flat plate manipulators.
The simulations demonstrate the effects of such manipulators on the instanta-
neous turbulence structures as well as on the statistically averaged flow field. The
time-dependent and three-dimensional filtered :\a,"ier-Stokes equations for an in-
compressible fluid are solved. A two-part eddy-viscosity model takes into account
the non-resolved subgrid-scale effects. Characteristics of the numerical method are
a leapfrog-scheme for explicit time integration, the projective method and the ca-
pacitance matrix technique to solve the Poisson equation for the pressure. First,
the results of a zero-pressure-gradient boundary layer flow are discussed. The ma-
nipulated boundary layer (1\1BL) is compared with the standard boundary layer
(SBL) without any manipulator under the same inflow conditions. Secondly, the
comparison of a manipulated and standard channel flow, abbre,"iated :\ICF and
SCF, respecti,"ely, is presented. Both the results of the MBL and l\ICF exhibit a
decrease in the turbulence intensity. Correspondingly, coherent structures of the
instantaneous flow field are suppressed in both fiow types by the manipulators.
The results of the IvIBL and MeF indicate no fundamental differences concerning
the effects of fiat plate manipulators. This beha,"iour can be explained by the fact
that the manipulators act in regions where the turbulent structures in both flows
are wry similar.
NOlnenclature
Symbols
~A.j grid cell surface
cf = 1 ;;;:~ , skin-friction coefficient
Ci convection term
Di diffusion term
Dij deformation tensor
E;.o SGS kinetic energy (isotropic part)
Fr total drag force

41
42

h channel half-width
h characteristic mesh-width
hm wall-distance of manipulator
1m length of manipulator
linh characteristic length of inhomogeneous SGS model
lmix mixing length
p pressure
Re = v,,! I re ! Reynolds number
VreJ ' ...
t. !:::"t time, time step
bulk flow velocity
= ( <T~> ) ~ , friction velocity
free-stream velocity
grid volume
Vi,U,l'.W velocity components
Xi, X, y. Z coordinate directions (longitudinal, spanwise, vertical)
Zwall nearest wall-distance in z-direction
5 boundary layer thickness
52 momentum thickness
5ij Kronecker symbol
Tij = ~e D ij , shear-stress tensor
Tw \vall shear-stress
J' J.1iso isotropic SGS eddy-viscosity
ji J.1inh inhomogeneous SGS eddy-viscosity
~ distance downstream of manipulator trailing edge
Definitions
<9> statistical mean value
rjy" (Reynolds') fluctuating value
F-
9 grid volume average
j,
CJ surface average, taken oYer !:::"Aj
9' deviation from GS quantity
Subscripts
inh inhomogeneous part of the SGS model
~so isotropic part of the SGS model
Introduction
The dynamics of turbulent flows is not at all dominated by random motions.
In definition of the phenomenon "turbulence" coherent structures and coherent
motions play an important role. \Vhile studying wall bounded flows so-called wall
events can be observed. There is for example the burst-process involYed in the
breakup of a single longitudinal streak structure (Bogard et al. 1987). As a con-
sequence of a burst, low-speed fluid is carried away from the \\'all region \vhich is
called an ejection. In contrast, sweeps are characterized by the rapid transport of
high-speed fluid towards the wall. Ejections as well as sweeps contribute about
43

70% to the total Reynolds stress (Landahl et a1. 1986) and represent an impor-
tant mechanism of producing surface drag in turbulent boundary layers (Corke et
a1. 1980). Furthermore, free or wall bounded turbulent shear flows contain large-
scale coherent structures (Fiedler 1987). They dominate the turbulent transport,
contain most of the turbulence energy and consequently control the mixing rate,
noise-production and -emission and vibration phenomena. As far as boundary
layer flows are concerned, some of these large-scale structures propagate into the
outer part of the layer and cause the entrainment process (Coustols et a1. 1986.
Tennekes et a1. 19(2). The speed, at which the interface between the turbulent
and nonturbulent fluid with the large-scale structures moves into the irrotational
region, controls the entrainment (Tennekes et a1. 19(2). In addition, fluid from
the irrotational region penetrates into the boundary layer (Praturi et a1. 19(8),
which explains the intermittent character of the outer part of a boundary layer
flow. Coustols et a1. (1986) summarize that turbulent structures are related to a
cycle of movements featuring near wall events (e.g. ejections, sweeps) and outer
flow e,'ents (e.g. entrainment).
It is pointed out in several experimental studies (Plesniak et a1. 1985, Savill et
a1. 19S5, Anders et a1. 19S5) that flat plate manipulators are a means of reducing
the drag in turbulent boundary layers at low Reynolds numbers. A computational
investigation with various statistical turbulence models (Coustols et a1. 1989) pro-
duces similar results but is, of course, not able to show the modification of instan-
taneous flow structures. High Reynolds number experiments (Reo, >S, 000) on an
axisymmetric body in the NASA Langley Tow Tank resulted in no net drag reduc-
tion by positioning tandem airfoil manipulators in the outer part of the boundary
layer (\Valsh et a1. 19S9). Concerning the manipulation of turbulent internal flows
drag reduction is not found at all. Prabhu et a1. (19SS) found such results in the
case of a turbulent channel flow at low Reynolds number (Reh =24,500). Equally,
recent experimental and computational studies of Pollard et a1. (19S9) give the
conclusion that single ring manipulators do not reduce the drag in pipe flows at
high Reynolds number (ReD=IS0, 000).
Finally, the results of experimental studies (Savill et a1. 19S5, Savill 1987,
Guezennec et a1. 1990, Lemay et a1. 19S9) in boundary layers show that flat plate
manipulators change the flow field in several ways. The mechanisms of how ma-
nipulators work are classified into "plate" and "wake" effects, compare table l.
The "plate" effects include the alteration of the flow field in the immediate "icin-
ity of a plate. Due to the presence of the manipulator all velocity fluctuations
are suppressed, in particular wall normal fluctuations (pI). The destruction of
the large-scale structures is understood as the break-up of large-eddies (p2). In
consequence of the flow displacement in a boundary layer the circulation around
the plate produces lift and induces downward velocities in the wake according to
Biot-Savart's law. This phenomenon is called the "downwashing" of the plate \vake
(p3). The "wake" effects determine the flow modifications due to the wake of a ma-
nipulator. New energetic small scales are introduced into the base-flow immediate
downstream of the manipulators. These small scales promote an enhanced energy
cascade from existing larger scales (wI) (Savill et a1. 19S5). Furthermore, the wake
acts as a shield that prevents incursions of high-speed potential fluid and blocks
44

the interaction between the outer and near wall part of the flow (w2) (Coustols et
al. 1986, Savill et al. 1988). Both the "plate" and "wake" effects are responsible
for the skin-friction reduction (Savill 1987). With reference to the maximum drag
reduction, it is important to point out that the minimum local skin-friction occurs
close to the position where the wake reaches the wall. This explains why the dis-
tance between manipulator and wall is an important parameter. In a boundary
layer it should vary inversely with the Reynolds number (Savill et al. 1988).

"plate" effects "wake" effects

suppressIOn (pI) small scales (wI)


large-eddy break-up (p2) shield (\\'2)
downwashing (p3)

Table 1: Effects of flat plate manipulators.


In studying the effects of manipulators in boundary layer and channel flows
the large-eddy simulation is used in this work. The starting point is a model of
Schumann (1975) which was developed to simulate turbulent flows in plain chan-
nels and annuli. Richter et al. (1987) have successfully applied this procedure to
boundary layer flows with an external pressure gradient. As the model allows for
flow inhomogeneities, it is also suited for complexe geometries as in the present
case. The numerical procedure has also been used successfully to simulate the
backward facing step flow (Schmitt et al. 1987, Friedrich et al. 1990).
Numerical Model
The time-dependent and three-dimensional!\' ayier-Stokes equations for an in-
compressible fluid are the starting points of the present LESs. Integrating these
equations over a finite volume according to Schumann's (1975) "volume balance
procedure", the high-frequency part of the spectrum is filtered out. Grid-scale
v-
(GS) quantities are defined as cell-volume averages, 9, or cell-surface ayerages,
J-;-
9:
t:.<p~:= Al { IjJ dip . (1)
Uip } t:.<p

The subgrid-scale (SGS) quantities represent the deviations from the GS quantities:

(2)

Using a cartesian grid the filtered equations haye the following non-dimensional
form:
(3)

aF'iji .. j-_.
-at- + oJ (J'ijJ'ij
J t
+ v.lv'
J t
+Jp-C -)~ .. ) -
U)l I JI -
0 (4)
45

where
j- j-
.
JJ'r- -Re
- D 1"
1 j- jD .. _ OVj + OVi
( 5a, b)
J' -
OXi OXj

The volume integration of divergence terms leads directly to differences of the


cell-surface averaged quantities, defined by

(6)

By decomposing velocity components into a GS part and a SGS part according


to (2) SGS stresses are produced,

j- j-,-,
Tji,SGS = - VjVi , (7)

which have to be expressed in terms of GS quantities. To handle this closure


problem we adopt the two-part eddy-viscosity concept of Schumann (1975) and
Grotzbach et al. (1977):

(8)

j-.. - ji . j D" (9a)


T)l,iso - /-LIS0 ji'

ji f1iso = C2c V.6.A. j j C5 v ' isa ji C , (9b)


v- C2c C20 1 j-- 2
' iso = ---min( C31c [mix, h )h- l ' c (D'J;) , (9c)
C3c 2
j-II _ j - j-
Dji - Dji - < D ji > , (9d)
1
h=(.6.V)" , (ge)
C2
C2c =- (91)
11
j=-c - ji j-D..
'J',inh - f1inh< J'> (only for ji=xz,zx) (lOa)

ji _ (Z. )21 o<xli> O<=w> 1


f1inh - tnh oz + ox ' (lOb)

111
linh
--~+--
C2.inh h Imix
, (10c)

(11 )
This model contains a fluctuating part (9) which considers the locally isotropic
SGS effects and a statistical part (10) that takes into account inhomogeneities
46

near walls and manipulators. The isotropic SGS energy of relation (9c) results
from equating the production and dissipation terms in the transport equation for
v-
E'isa. For more details of the SGS model see Schumann (1975) and Friedrich
(1988). The factors of the isotropic model (-Yl=0.93, 12=1.0, C31=0.74) are the
same as those used by Schmitt et al. (1987) to simulate the fiow O\-er a backward
facing step. The grid dependent constants are computed by assuming the the-
ory of locally isotropic turbulence and the validity of the Kolmogorov spectrum
(Schumann 1975, Schmitt 1988). For the grid used in this study the following
values are obtained: c2=0.069, c2o=0.83, c3=0.595, xC5 =Y C5 =0.785, zC5=0.939,
xX c=YY c=1.589, zZc=1.435, xY c=0.981, xZ c=yz c=0.747. The constant of the inho-
mogeneous model is chosen as, C2,inh=0.1.
The isotropic SGS energy amounts about 18% to the total turbulence energy
for the used grid and Reynolds number. The statistical part of the model becomes
effective near walls and manipulators and contributes to an enhanced momentum
transport.
The system of equations (3),(4) is solved on a staggered grid so that one may
assume vVi ~;v; and jp ~vp. All GS quantities in equation (4), which are not
defined, are approximated by algebraic averages, i.e.

(12)

The time-integration of (4) proceeds in cycles of explicit steps consisting of 49


leapfrog steps (h=I, 12=0, h=2) and an averaging step (h=~, 12=~, h=~) in
order to avoid 26t-oscillations:

(13 )

(14)
Fortin's version of the projection method is used in (13) which first neglects the
pressure-gradient. To update the velocity V;+l, P has to be determined from a
Poisson equation which results from equation (14) and (3):

h6t(6;6;p) = 6iVi (15)

Since the fiowfield is homogenous in spanwise direction, a fast Fourier transfor-


mation can be applied to convert the 3D Poisson problem into a set of decoupled
2D Poisson problems. A cyclic reduction algorithm serves to solve the latter in
the case of a box-like computational domain. In order to apply this direct solver
in multiply connected fiow domains as produced by fiat plate manipulators, the
so-called capacitance matrix technique (CMT) is needed (Schumann 1980, Schmitt
et al. 1988).
In matrix notation the C~IT procedure can be formulated as follows. Given is
the set of equations A.p=q, the so-called" A-problem". A is such that no efficient
direct soh-er exists. The equation Bx=y defines the " B-problem". which can be
47

solved directly. B is "similar" to A in the sense that these two n X n-matrices are
equal except for a few mn rows. These m equations belong to the mesh cells
with internal boundaries. Correspondingly, A, Band q are decomposed into

q= GJ
(AI,B I : mxn-matrices, A 2 : (n-m)xn-matrix, qI vector of length m, q2 vector of
length (n-m)). We get the solution of the A-problem with the help of the m x m-
capacitance matrix C and the nxm-'selection'-matrix lV

C = .h B- 1 W W= (~)
(I: mxm-unit matrix, 0: (n - m)xm-zero matrix) by applying the steps:

Bp=q (160)

Cv = w q=q+Wv (16b, c. d. e)
Bp=q (16f)
(U). w. v: ,'ectors of length m, p. q: vectors of length n). That means we obtain the
desired solution at the cost of solving two B-problems (16a,f) and a m x m algebraic
system of equations (16d). The capacity matrices (one for each decoupled plane)
depend only on the irregular geometry and are pre-computed at the start of a
simulation by solving m B-problems to determine one matrix.
Boundary Conditions
The specification of proper inflow boundary conditions is very costly in the
present case of spatially developing flows. At each grid point of the entrance plane
the LESs of ~dBL and MCF need the instantaneous GS velocity vector and the
SGS turbulence energy at each time step of the simulation process. Separate LESs
provide this data and allow storing of all relevant quantities from a plane normal to
the flow direction at each time step. In the case of a channel a fully dewloped flow
(FDCF) ,,ith periodic boundary conditions in the main flow direction and in the
case of a boundary layer a zero-pres sure-gradient flow (ZPGBL) with inflow/out-
flow boundary conditions, described in Richter et al. (1987), have to be simulated.
The Reynolds number, grid spacing, time step and SGS model are the same as in
the cases of the manipulated flow.
A special treatment of the "elocity wctor in the outflow plane has prowd suc-
cessfully. It consists in a linear extrapolation of the mean streamwise component
and a constant extrapolation of the remaining two mean components, but in soh'-
ing a conwction equation for all the velocity fluctuations using < u > as proper
transport velocity. For more details, see Richter et al. (1987).
The velocity component normal to solid walls is set to zero. Resolution restric-
tions require approximate boundary conditions tangential to walls as Piomelli et
48

al. (1989) propose for flows of engineering interest. By means of Schumann's (1975)
condition the instantaneous wall shear-stress is in phase with the instantaneous
velocity in the logarithmic layer. This assumption is supported by experimental
investigations (Piomelli et al. 1989). The same concept is applied at surfaces of
manipulators. The authors are aware that the validity of the logarithmic law is
questionable there. However, the inviscid damping of vertical velocity fluctuations
as the primary effect of manipulators is considered correct.
In the case of the boundary layer flow we specify Dirichlet conditions at the top
of the computational domain for the tangential velocity components and the SGS
turbulence energy. The normal velocity is obtained from the continuity equation.
'Ve use periodic boundary conditions in both flow types in the spanwise direc-
tion.
Results

Three different LESs were performed for boundary layer as well as for channel
flows.
l.) Two different LESs serve to produce inflow boundary condi tions for both flows.
a) Boundary layer flow:
LES of a zero-pres sure-gradient boundary layer flow (ZPGBL) with inflow /
outflow boundary conditions. Some sort of structural periodicity (Richter
et al. 1987) is assumed in this case to generate inflow conditions.
b) Channel flow:
LES of a fully developed channel flow (FDCF) with periodic boundary
conditions in the main flow direction.
2.) LESs of the manipulated boundary layer (MBL) and manipulated channel flow
(MCF) with inflow boundary conditions from l.)
3.) LESs of the standard boundary layer (SBL) and standard channel flow (SCF)
with the same inflow conditions as in 2.). "Standard" refers to unmanipulated
flows.

Boundary layer flow


Two computational domains of different length haye been chosen. In terms of
boundary layer thickness 5 at the inflow plane the long domain measures 24 x 4 x 2
(x,y,z-directions) and the short one 8x4x2. A 192x32x32 and 64x32x32 equidis-
tant grid have been used. The results are independent of the length of the com-
putational domain except in the outflow-region. Based on Ur,l and 5 the Reynolds
number is 3240, which corresponds to a value of 8700 based on free-stream velocity
U oo and momentum thickness 02' All quantities are non-dimensionalized with Ur,l
and 5. An infinitely thin manipulator is fixed at a distance 25 downstream of the
inflow plane. Since the gap h m between the manipulator and the wall should de-
crease with increasing Reynolds number in order to get the maximum skin-friction
reduction (Sa\'ill et al. 1988), we have set h m = 0.31255. The length 1m of the de-
vice is l.55. This is in accordance with Savill et al. (1988), who obtain the largest
integrated skin-friction reduction when the length of a manipulator is greater than
the boundary layer thickness.
49

I
o
50

The ZPGBL was computed with 36,000 time steps. Inflow data of 12,000 time
steps for the MBL and SBL were stored on magnetic tape. Consequently the
simulations of the MBL and SBL covered 12,000 time steps out of which 6,000
are needed for statistics. Besides time averaging spatial averaging is used in the
homogeneous y-direction.
Figures 1a,b give a qualitative impression of how a single flat plate manipulator
affects a boundary layer flow. The contour surfaces are plotted for high \'alues of
the instantaneous (-u" w" )-correlation. The Reynolds stress mainly results from
these structures. In the SBL the (-u" w")-structures are uniformly distributed
over the whole computational domain (fig.1a). Figure 1b shows the suppressing
effect of the manipulator. Unlike the portion of the flow field upstream of the
manipulator, there is no similarity in the MBL and SBL downstream of the plate.
The manipulator destroys larger structures and breaks them up into smaller ones.

8
I I I10 12

Fig.2 Mean longitudinal velocity component.


1~ 16 1B 20 22 X 2~

The effect of the manipulator on the mean longitudinal velocity component is


demonstrated in fig.2. The plate produces a wake, which develops downstream and
is nearly indistinguishable in the last profile. In this context, it has to be mentioned
that the grid-resolution is to low to get detailed information concerning the wake
and the developing boundary layers on the manipulator surfaces. However. the
dewlopment of the wake is more obvious from plotting the momentum deficit
6<U>=<U>MBL-<U>SBL downstream of the manipulator, compare fig.3. It
is shown that the maximum deficit shifts towards the wall. This phenomenon
coincides with the" downwashing" of the wake.

SC"LEI - - - 8 -4.000E+OO

I
z I I
I
I I
)
e
'='
)
J
=
tc;
J
El
,
.::l
= ~
,

8 10 12 I ~ 16 18 20 22 X 24

Fig.3 \fomcntum deficit 6<U>=<U>MBL-<U>SBL downstream


of the manipulator.
51

The rms-values of the longitudinal velocity fluctuations clearly reflect the sup-
pressing-effect of the manipulator, compare fig.4. The solid line refers to the ma-
nipCllated and the dashed line to the standard boundary layer, respectively. \Vhile
a slight influence is found immediate downstream of the manipulator at x=48
(~=0.58), the turbulence intensity is considerably reduced at x=98 (~=5.58). It
is of interest that the velocity fluctuations are reduced mainly between z=hm and
the wall. The measurements of Coustols et al. (19S6,19S9) confirm these results.
Further downstream the urms-values recover gradually and reach almost the level
of the SBL at x=198 (~=15.58).

SCALE' - - - 6.000[1'00

2 10 12 14 16 1B 20 22 X 24

Fig.4 rrns-Hlues of longitudinal velocity fluctuations.


- lIIBL, - - SBL.
Figures 5a,b demonstrate the effect of suppressing and destruction of large-scale
structures in a ~\'IBL by comparing the fluctuating wlocity-vectors in a (x,z)-plane
with those in a SBL. Fig.5a gives an impression of the turbulence dynamics in a
SBL at two different instants. The flow field is characterized by ejection and sweep
type events. A typical turbulence structure, marked with arrows, travels more than
28 downstream in the course of 100 time steps. \;Vhat happens to this structure
when a manipulator obstructs the flow? Fig.5b shows the flow field in a :vIBL at
the same moments. At time step 6S00 the vectors which form the same structure
just described are aligned with the plate. Still more importanL the structure is
clearly weakened 'while trawlling downstream (lower plot of fig.5b). Furthermore,
the plots for nt=6S00 show how the sweep and ejection type ewnts at the end of
the computational domain are influenced by the manipulator.
The instantaneous flow fields of the MBL and SBL are compared in plate 1 by
means of cartographic representations of velocity fluctuations and the SGS kinetic
energy. The upper panel of each figure corresponds to the SBL and the 100\'er panel
to the \IBL. The blue, green and red regions are characterized by strong negative.
nearly zero and strong positive \'alues, respectively. In the 100\'er panels the flow
field upstream of the manipulator is still the same as in the SBL. As the manipulator
breaks up the oncoming structures, the flow field downstream is changed. \Yhile
the longitudinal velocity fluctuations are scarcely affected (compare plate la), the
changes are more obvious in the other wlocity components. The figures of the
spanwise (plate Ib) and the vertical velocity fluctuations (plate lc) illustrate the
wake of the manipulator, which divides the flow field into an upper and lower
52

Q .... .. ..........
.. . . . . .. . .
. . . . . . . . .......... ..
. . . .. .. .. .. .. .. .. .. .. .
.... . ...... .
....
.. .. .. .. .. .. .. .. .. ...
,
...... . .........

~~~~~~~~ll~;ll~~~~~j~~

tttfHU1HU~llU}lll
2 3 5 6 7

x
t
o 2 3

x
t
2 3

Fig ..5a.b Fluctuating velocity-vectors in a (x.z)-plane. (a) SBL. (b) '\lBL.


53

< < <

1
< <

I
i
S'l s
SIH-Z

S I H-Z SiX\,-Z

Fig.6a.b Vertical velocity fluctuations and fluctuating velocity-vectors


in a (y,z)-plane (x=3.81258. ~=O.31258). (a) SBL, (b) !'IIBL.
54

part. The SGS kinetic energy is a measure of the absolute value of the fluctuating
deformation tensor. It shows that the edge of the boundary layer is not affected
by the manipulator, compare plate Id.
Figures 6a,b compare fluctuating quantities in a plane perpendicular to the main
flow direction. The plane lies slightly downstream of the manipulator trailing edge.
The contour-lines of the vertical velocity fluctuations show the break-up of large-
scale structures into smaller ones. The wake of the plate can be made out from
those fluctuating "elocity-vectors which are aligned with the y-direction. They are
located at the height of the manipulator, h m =O.31258.
The development of the momentum thickness 82 for the MBL compared to that
for the SBL in fig.7a reveals the typical characteristics. The increase in the range
of 2.2::; x::; 3.7 is induced by the presence of the manipulator. Further downstream.
for ~ ;:: 2.58, the slightly smaller slope for 82 in the Iv1BL indicates the tendency
to-wards a skin-friction reduction.

f
:':
~
;i.1
>=<
I
:::;
I
:: I'"
I~ ~
- SIlL

- ' ,illL
I

10 15 20 X

Fig. ia Comparison of momentum thickness in the SBL and MBL.

~ M I

b ~~~~~~_~
i-~L ~
i - . 1illL I
I

o :1 ========================================
10 15 20 x
Fig.ib Comparison of skin-friction coefficient in the SBL and MBL.

A true skin-friction reduction for ~;:: 2.58 is reflected in the shape of the skin-
friction coefficient C f in fig.7b. The reduction of C f immediately upstream of the
plate is a consequence of flow deceleration. Consequently, the following increase at
the manipulator location is due to flow acceleration. The maximum skin-friction
reduction of about 11.5% is found at ~ = 138. The cf-distribution shows a skin-
friction reduction until the end of the computational domain. The total drag
force F,. results from integrating the shear stress oYer the boundary layer wall and
the two manipulator surfaces. F,. does not giYe a net drag reduction by using a
computational length of 248 (~=20.58). The ratio of F,. in the ~1BL to that in the
55

SBL drops from the yalue 1.66 for the short computational domain down to 1.16
for the longer domain:
FTMBL = 1.66 ---+ 1.1 6 .
F TSBL

::
- SDL
- MDL
.---. MUL 2

X 8

Fig.8a Comparison of momentum thickness in the SBL, MBL and MBL2.

M
o

~ /~ I
~ I
- SDL
- MDL
<>-----<> MOL 2

X 8

Fig.8b Comparison of skin-friction coefficient in the SBL, MBL and MBL2.


Figures 8a,b show the influence of different manipulator heights on the momen-
tum thickness and the skin-friction coefficient. It is found that a plate positioned
at h m =0. 758 (MBL2) provides no skin-friction reduction within the short compu-
tational domain of 88 length. The higher level of 82 downstream of the manipulator
is due to the higher fluid velocity in the environment of the manipulator. The ratio
of FT for the MBL2 to FT for the SBL is greater than that in the case above:

This fact gives support to the conclusion of Savill et a1. (1988) that the optimum
position of a manipulator should be close to the wall at high Reynolds numbers.
For comparison with the experimental data of Coustols et a1. (1988) and Tenaud
(1988) a manipulated boundary layer at low Reynolds number REb2 ::::2500 was
simulated. The length of the manipulator is 0.758 for the simulation and 0.78 in the
experiment. The downstream development of the skin-friction coefficient related
to the coefficient of the nonmanipulated boundary layer is in good agreement with
the measurements, compare fig.9. However, the maximum skin-friction reduction is
56

1. 05

1. 00 _._._.-._._._._._.-
-'-'-'-'

o
.95 o 0
o
o
.90

.e5~----------------------------~------------
-5 10 15 20 ~ 25

Fig.9 Downstream development of the skin-friction coefficient.


- simulation, 0 experiment.

somewhat larger in the simulation. In addition, the simulation reflects the increase
of the skin-friction at the location of the manipulator.
Channel flow
The size of the computational domain in a (x,y,z )-coordinate system is agam
24x4x2 in the long channel case and 8x4x2 in the short channel case. It is
measured in terms of the channel half-width h. Just as for the boundary layer flow
a 192 x 32 x 32 and 64 x 32 x 32 equidistant grid has been choosen. The Reynolds
number based on h and the friction velocity Ur,l at the inflow plane is 3240, based
on the channel width 2h and the bulk velocity Ub is 150,000. All quantities are
non-dimensionalized with hand Ur,l' By imaging the channel flow as the result
of two merged boundary layer flows, it is natural to group two manipulator plates
symmetrically. They are fixed at a distance 2h downstream of the inflow plane.
The geometrical data of one plate are the same as in the boundary layer case. Thus
the wall-distance h m of such an infinitely thin plate amounts to 0.3125h and the
length lm to 1.5h.
The LES for the FDCF was performed with 38,000 time steps in order to
generate inflow data. The ~\'fCF and SCF were simulated with 17,000 time steps
out of which 12,000 were used for statistical averaging besides averagmg in the
homogeneous y-direction.

~ ---~ fI 4.U-4E+01

I Ii I I I
"
z

0 6 8 10 12 14 16 18 20 22 X 24

Fig.l0 ~fean longitudinal velocity component.

Fig.10 gi\'es an impression of the manipulated flow field in terms of the mean
longitudinal velocity component. One can make out the wakes of the manipulators.
57

which spread and weaken similarly to those of the MBL while developing down-
stream. The effect of the manipulators becomes more obvious from plotting the
momentum deficit 6.<U>=<U>MCF-<U>SCF, compare fig.lI. The momentum
deficit of the wakes shifts in the channel as well as in the boundary layer flow to-
wards the wall. This leads to the conclusion that the" downwashing" -phenomenon
exists also in the channel. As a consequence of continuity the flow is accelerated
in the core region. Thus, the velocity profile becomes peakier in the manipulated
case and leads to a reduced wall shear-stress.

;t1 JJJJ11l ))))J) ) }


SCALE I - - - 8 4. OODE+OO

o
11
8 10 12 1~ 16 18 20 22 X 24

Fig.111Iomentum deficit 6.<U>=<U>MCF-<U>SCF downstream


of manipulators.
The manipulators affect the urms-values in a channel in a way comparable to
the boundary layer, compare fig.12. The turbulence intensity is reduced mainly in
the region near to the wall. This is shown most clearly at x=9h (~=5.5h). The
influence of the manipulators on the urms-values has almost completely disappeared
at x=19h (~=15.5h).

~.

(
/

\, \,
~
o
~> '~
o 6 B 10 12 14 16 18 20 22 X 24

Fig.12 rms-values of longitudinal velocity fluctuations.


- MCF, - - SCF.
Figures 13a,b compare the fluctuating (u, w)-velocity \'ectors in a l--.ICF and a
SCF in their time-development. The plots of fig.13a demonstrate how two large-
scale structures in the upper and lower part of the SCF (marked with arrows) trawl
downstream in the course of 2xlOO time steps. Fig.13b presents the vector fields
of the :tIfCF at the same moments. Remember, that the inflow conditions for both
flows are the same. Therefore at time step 5800 the same large-scale structure is
found in front of the upper manipulator. 100 time steps later the structure has
58

~::"'-""' ,,:?~. i~?~:' .':::~~~::;i


I.::~:":-::::;;;::'::~~! i::~::: ::~::::::::::: ;::::-::: i
.".,- .. " , .... "':: .. ~::;..;:

:.:::::1 ~U:::":~<J)~I

~r"::;f~'
I:~. ~~I
Ihl~~~:1
N N

Fig.]3a.b Fluctuating yelocity-yectors In a (x.z)-plane. (a) SCF. (b) .\1CF.


59

..0

Fig.14a,b Spanwise velocity fluctuations. (a) SCF, (b) .\lCF.


60

passed the manipulator while the fluctuating velocity-vectors have been aligned in
the direction of the plate. Again 100 time steps later the structure has travelled
further downstream, but the strong fluctuations in the vicinity of the ,vall have
quite completely disappeared. The effects of suppression and destruction of large-
scale structures by the manipulators become also evident from the structure present
in the lower channel half. \Vhile travelling downstream it is split into two parts by
the plate and remains devided.
The time-development of spanwise velocity fluctuations in terms of contour-
lines is shown in figures 14a,b. Among the three velocity components, v" most
perfectly reveals the typical inclination of large-scale structures with respect to
the walls. In fig.14a a structure which travells downstream in the SCF is marked
with arrows. By contrast, in the MCF the same structure has been deformed and
broken up by the lower manipulator (fig.14b).

J0 20 x
Fig.lS Comparison of wall shear-stress in the SCF and MCF.
Finally, the downstream development of the wall shear-stress Twin the l\1CF
is compared with that in the SCF, see fig.15. The reduced Tw slightly upstream
of the manipulators is a consequence of flow deceleration, as the following increase
in T w is due to the flow acceleration between manipulator and ,vall. The following
decrease in T w for ~ ;::: 2.5h leads to a maximum wall shear-stress reduction of
11 % at ~ = 13.5h. Although T w remains below its SCF -value until the end of the
computational domain (~= 20.5h) no net drag reduction is found. The ratio of the
total drag force Fr in the MCF to that in the SCF decreases from a value of 1.67
for the short to 1.16 for the long computational domain:

F rMCF = 1.6-( ----t 1.16 .


FrscF

By comparing this result with that in the MBL no difference can be made out
concerning the drag behaviour of both manipulated flow types.
Summarizing discussion
The presented results show that flat plate manipulators act in similar ways in
boundary layers and in channels though there are differences between both flow
types. Table 2 contains correspondences and differences between these flows. Since
both are bounded by walls they produce wall ewnts, such as ejections and sweeps.
61

Both flows also contain large-scale coherent structures which are characteristic of
any shear layer. The entrainment process does not exist in a channel. The reason
is the missing free-stream boundary. Last not least the fully developed channel
flow exhibits no mean flow displacement.

boundary layer channel

yes wall events yes


yes large-scale coherent structures yes
yes entrainment no
yes mean flow displacement no

Table 2: Comparison between boundary layer and channel flow.

The following arguments aim at underlining the similarities between manipu-


lated high Reynolds number boundary layer and channel flows. The suppression
of velocity fluctuations due to the primary "plate" effect is present in both cases.
Since at high Reynolds numbers the distance of the manipulators from the wall
has to be small, the plates act in regions where both flows are very similar. This
fact may also explain the "downwashing" behind the plates in a channel, where one
would not expect it, because of the missing flow displacement. As far as the "wake"
effects are concerned, new energetic small scales which would be introduced in the
real flow, cannot be resolved in the simulation. The frequently quoted shield-effect
of the wake is irrelevant to the channel as there is no incursion of high-speed po-
tential fluid. Howeyer, as manipulators are mounted close to the walls, their wakes
should act in a similar way in both flows. Namely, they inhibit the interaction
between large-scale structures in the outer region and smaller scale structures near
the wall.
Conclusions
LESs were performed in order to investigate the manipulation of boundary layer
and channel flows at high Reynolds number with infinitely thin flat plate devices.
The results rewal no fundamental differences between the two flows concerning the
effects of such manipulators. The suppressing of the velocity fluctuations due to the
presence of the plates is shown in terms of statistical as well as in terms of instan-
taneous field variables. Furthermore, large-scale coherent structures are destroyed
in both flow types by the manipulators. \Vith regard to the" dO"wnwashing", the
~1BL and MCF indicate a similar behaviour. The spatial resolution of the grid
chosen does not suffice to resolve details of the manipulator-wake. Consequently,
the simulations cannot predict the new small-scale turbulence generated immedi-
ate downstream of the manipulators. However, the wake is demonstrated in both
flow types in the mean velocity component. The resolvable part of the turbulence
intensity is clearly reduced in the region near the wall. The maximum skin-friction
reduction amounts about 11.5% in the MBL and about 11% in the MCF. But.
in the present boundary layer as well as in the channel simulations, no net drag
reduction could be found up to a distance of 20.58 (resp. 20.5h) downstream of
62

the manipulator trailing edge. Larger computational domains in the streamwise


direction would be needed to show a possible net drag reduction.
Acknowledgement
The work is supported by the Stiftung Volkswagenwerk
Literature
Anders,J.B., vVatson,R.D., Airfoil Large-Eddy Breakup Devices for Turbulent
Drag Reduction. AIAA-85-0520, AIAA Shear Flow Conference, Boulder, Colorado,
March 12-14, 1985.
Bogard,D.G., Coughran,M.T., Bursts and Ejections in a LEBU-Modified
Boundary Layer. Proc. of the 6th Int. Symp. on Turb. Shear Flows, Toulouse,
France, Sept.7-9, 1987.
Corke,T.C., Guezennec,Y.G., Nagib,H.M., Modification in Drag of Turbulent
Boundary Layers Resulting from Manipulation of Large-Scale Structures. - In:
Hough, G.R. (ed.): Prog. in Astronautics and Aeronautics Vol. 72, pp. 128, 1980.
Coustols,E., Cousteix)., Reduction of turbulent skin friction: Turbulence \Iod-
erators. La Recherche Aerospatiale, N2, pp. 63-78, 1986.
Coustols,E., Cousteix)., Turbulent Boundary Layer Manipulation in Zero Pres-
sure Gradient. 16th ICAS Congres of the International Council of the Aeronautical
Sciences, Jerusalem, 1988.
Coustols,E., Tenaud,C., Cousteix,J., Manipulation of Turbulent Boundary Lay-
ers in Zero-Pressure Gradient Flows: Detailed Experiments and Modelling. - In:
Andre,J.-C., Cousteix,J., Durst,F., Launder,B.E., Schmidt,F.\V., vVhitelaw, J.H.
(eds.): Turbulent Shear Flows 6, Selected Papers from the 6th Int. Symp. on Turb.
Shear Flows, Toulouse, France, Sept. 7-9, 1987, Springer Verlag, pp. 164-178, 1989.
Fiedler,H.E., Coherent Structures. - In: Comte-Bellot,G., Mathieu,J. (eds.):
Advances in Turbulence. Proc. of the First European Turbulence Conference, Lyon,
France, 1-4 July 1986. Springer-Verlag, pp. 320-336, 1987.
Friedrich,R., On Large-Eddy Simulation of Turbulent Flows. - In: Kiki.H.,
Kawahara,M. (eds.): Computational Methods in Flow Analysis, Vol.2. Okayama
Univ. Press, Japan, pp. 833, 1988.
Friedrich,R., Arnal,M., Analyzing Turbulent Backward-Facing Step Flow with
the Low-Pass Filtered Navier-Stokes Equations. - In: Murakami,S. (ed.): Journal
of \Vind Engineering and Industrial Aerodynamics, Vol. 36, nos. 1-2, Special Issue
on Computational Fluid Dynamics, 1990.
Grotzbach,G., Schumann, U., Direct numerical simulation of turbulent velocity-,
pressure-, and temperature fields in channel flows. Proc. of the Symp. on Turb.
Shear Flows. Penn. State Univ., Apr.18-20, 1977.
Guezennec,Y.G., Nagib,H.M., Mechanisms Leading to ~et Drag Reduction in
Manipulated Turbulent Boundary Layers. AIAA Journal, Vol. 28, No.2, Feb. 1990.
Landahl,1\LT., Mollo-Christensen,E., Turbulence and random process in fluid
mechanics. Cambridge University Press, 1986.
Lemay)., SaviILA.M., Bonnet,J.-P., Delville,J., Some Similarities Between Tur-
bulent Boundary Layers Manipulated by Thin and Thick Flat Plate Manipula-
tors. - In: Andre.J.-C., Cousteix,J., Durst.F., Launder.B.E .. Schmidt. F.W ..
63

Whitelaw,J.H. (eds.): Turbulent Shear Flows 6, Selected Papers from the 6th
Int. Symp. on Turb. Shear Flows, Toulouse, France, Sept. 7-9, 1987, Springer Ver-
lag, pp. 179-193, 1989.
Plesniak,M.W., Nagib,H.M., Net Drag Reduction in Turbulent Boundary Lay-
ers Resulting from Optimized Manipulation. AIAA-85-0518, AlA A Shear Flow
Control Conference, Boulder, Colorado, March 12-14, 1985.
Pollard,A., Savill,A.M., Thomann,H., Turbulent pipe flow manipulation: some
experimental and computational results for single manipulator rings. - In: Sav-
ill,A.M. (ed.): Applied Scientific Research, Vo1.46, No.3. Special issue: Drag Re-
duction Applications of Riblets and Manipulators, pp. 281-290, 1989.
Prabhu,A., Vasudevan,B., Kailasnath,P., Kulkarni,R.S., Narasimha,R., Blade
manipulators in channel flow. - In: Liepmann,H.W., Narasimha,R. (eds.): Turbu-
lence Management and Relaminarisation, IUTAM Symp. Bangalore, India, 1987,
Springer-Verlag, pp. 97-107, 1988.
Praturi,A.K., Brodkey,R.S., A stereoscopic visual study of coherent structures
in turbulent shear flow. J.Fluid Mech., vo1.89, part 2, pp. 251-272, 1978.
Richter,K., Friedrich,R., Schmitt,L., Large-eddy simulation of turbulent wall
boundary layers with pressure gradient. Proc. of the 6th Symp. on Turb. Shear
Flows, Toulouse, France, Sept. 7-9, 1987.
Savill,A.M., On the Manner in which Outer Layer Disturbances Affect Tur-
bulent Boundary Layer Skin Friction. - In: Comte-Bellot,G., Mathieu,J. (eds.):
Advances in Turbulence. Proc.of the First European Turbulence Conference, Lyon,
France, 1-4 July 1986. Springer-Verlag, pp. 533-545, 1987.
Savill,A.M., Mumford,J.C., Manipulation of turbulent boundary layers by outer-
layer devices: skin-friction and flow-visualization results. J.Fluid Mech., vol. 191,
pp. 389-418, 1988.
Schmitt,L., Grobstruktursimulation turbulenter Grenzschicht-, Kanal- und Stu-
fenstromungen. Diss., Lehrstuhl fur Stromungsmechanik, TU Munchen, 1988.
Schmitt,L., Friedrich,R., Application of the large eddy simulation technique to
turbulent backward facing step flow. Proc. of the 6th Symp. on Turb. Shear Flows,
Toulouse, France, Sept. 7-9, 1987.
Schmitt,L., Friedrich,R., Large-eddy simulation of turbulent backward facing
step flow. - In: Deville,M. (ed.): Proc. of the 7th GAMM-Conf. on Numerical
Methods in Fluid Mechanics, Louvain-La-Neuve, Belgium, Sept. 9-11, 1987. Notes
on Numerical Fluid Mechanics, Vol. 20, Vieweg, pp. 355-362, 1988.
Schumann,U., Subgrid scale model for finite difference simulations of turbulent
flows in plain channels and annuli. J.Comp. Phys.18, pp. 376-404, 1975.
Schumann,U., Fast elliptic solvers and their application in fluid dynamics. -
In: Kollmann,W. (ed.): Compo Fluid Dynamics. Hemisphere, pp. 402-430, 1980.
Tenaud,C., Simulation Numerique de l'Ecoulement autour d'un Manipulateur
Externe de Couche Limite. These Docteur, l'Ecole Nationale Superieure de l'Aero-
nautique et de l'Espace, Toulouse, 1988.
Tennekes,H., Lumley,J.L., A First Course in Turbulence. The MIT Press, 1972.
\Valsh,M.J., Anders,J.B., Jr., Riblet/LEBU research at NASA Langley. - In:
Savill,A.M. (ed.): Applied Scientific Research, Vo1.46 , No.3. Special issue: Drag
Reduction Applications of Riblets and Manipulators, pp. 255-263, 1989.
64

P late Ie Cartographic representation of vertical velocity fluctual.Lons

Plate Id Cartographic representation of SGS kinetic energy


65

Plate 1a Cartographic representation of longitudinal velocity fluctuations

Plate Ib Cartographic representation of spanwise velocity fluctuations


THE IMPORTANCE OF LEBU DEVICE SHAPE
FOR TURBULENT DRAG REDUCTION

by
Arild Bertelrud
High Technology Corporation, Hampton, Virginia, U.S.A.
and
FFA, The Aeronautical Research Institute of Sweden, Bromma,
Sweden

SUMMARY
It is proposed that a well organized vortex street behind the device is a necessary
requirement for turbulent boundary layer drag reduction with outer-layer
manipulators.
The communication from the outer region to the wall is inhibited through the
existence of the vortex layer. The necessary wake structure is obtained with an
open laminar separation on the front device. At low Reynolds numbers most flat
plates and airfoils can provide this. At high Reynolds numbers different shapes are
required, since conventional airfoils with sharp trailing edges produce a turbulent
wake when the Reynolds numbers are increased. For tandem devices, the two
airfoils shed vortices at slightly different frequencies, causing development of a
strong, low-frequency component in the transitional wake.

OUTER-LAYER MANIPULATORS
At low Reynolds number several experiments have shown that local and overall
skin friction reduction is possible for turbulent boundary layers manipulated by
so-called LEBU's. These are flat plates or airfoils in single or tandem arrangement
in the outer part of the boundary layer. Low Reynolds number experiments
performed at 'host' boundary layer Rea of below 8000 gave 8-15 % net drag
reduction for an optimum tandem device at y/8=0.8, L 2 8, gap25 8. Due to the
high drag of a plate with finite thickness and its tendency to flutter, it was assumed
that an efficient device would be a minimum drag airfoil at practical Reynolds
numbers.
At high Reynolds numbers the limited experiments performed with airfoils so far,
have shown small local and no overall drag reduction 1. This has been attributed to
a scale mismatch between device and turbulence structure and/or a difference in
turbulence structure at high Reynolds number, Rea , of the 'host' boundary layer. A
sensitivity to minute geoemetric changes has also been noted by several authors.
In the present paper flow changes over the airfoil and in the wake due to Reynolds
number variation and device shape are emphasized.

67
68

ANALYSIS
Several mechanisms have been proposed for the influence outer-layer
manipulators 26 : for example a plate effect, i.e. reducing the overturning of the
large vortices in the outer part of the boundary layer, and a wake effect where the
downstream wake has beneficial effects on the turbulence structure. In the present
paper we observe that an increase in Reynolds number based on the 'host'
boundary layer momentum thickness also leads to a completely different flow
pattern ON the device AND IN ITS WAKE. The wake structure may have a large
influence on the downstream development of the turbulence structure.

Figure 1 shows a schematic of vortex development behind a single and tandem


device. On the airfoils open laminar separations change into closed bubbles as the
Reynolds number is increased. The well organized vortex shedding of flat plates
and airfoils under low Reynolds number conditions, changes into a complex
combination of effects as the same airfoils are used at higher Reynolds numbers.

There are several important aspects in the discussion of how and why a particular
outer-layer manipulator is suitable for high Reynolds number conditions:
- 'Host' boundary layer structure
- Organized, laminar vortex shedding
- Slightly different shedding frequency from the rear device
- Moderate device drag.

'Host' boundary layer structure


According to Bandyopadhyay and Hussain 2 there is a distinct difference in the role
of large and small scale structures in the inner and outer part of a boundary layer.
The two scales are in phase close to the wall while the are in opposite phase in the
outer layer. Here we propose that the drag reducing effects of the outer layer
manipulators are associated with the disruption of communication between the
inner and outer layer through the existence of an organized vortex structure in
between. Bradshaw3 showed how the turbulence close to the wall can be split into
an 'active' and a 'passive' part. The vortiCity in the outer part of the boundary layer
contributes to the wall shear, but only in a 'passive', indirect fashion. Spectra taken
in the inner part of the boundary layer could be collapsed on one curve for wave
numbers below 2 1C f . Y / U = 2 , where U is the local mean velocity where the
spectra were taken.
A controversy exists concerning how the turbulence structure changes with
Reynolds numbers. At low Reynolds numbers Badri Narayanan et. al. 4 found that
wall shear fluctuations scaled on inner variables, while the zero-crossings scaled
on outer variables. Bursts appeared to scale on inner or outer variables,
depending on where and how they are detected, and Badri Narayanan suggested
a possible pairing process. Andreopolous et.al. 5 did experiments at intermediate
Reynolds numbers, without being able to draw any firm conclusions concerning
the scaling.
69

If a direct interaction between the LEBU wake and the boundary layer turbulence
is responsible for the drag reduction, the devices must be mounted close to the
wall at high Reynolds numbers. If the communication loss is the responsible
mechanism, the required changes in location inside the boundary layer with
Reynolds are very small. Up to y+=100 Andreopolous' data showed a bursting
period T B U / 8 independent of Reynolds number (i.e. outer scaling), while the
burst period more than doubled for y+> 100 and scaling with outer variables was
not possible.
To have the drag reducing effect lasting over any useful distance, the wake must
have a structure allowing it to survive far downstream. It is turbulent over the major
part of the downstream region, and longitudinal vortices dominate the structure of
the far wake. The role of the manipulator is to set up this vortex system with a
minimum of device drag.
A modified laminar flow airfoil would be suitable, but it is possible to obtain the
open laminar separation with a traditional NACA profile turned backwards, as
suggested in Ref. 6. The drag as function of angle of attack? for a NACA 0012
airfoil mounted traditionally and reversed are shown in Figure 2 at Re.c=321 ,000 ,
i.e. at Reynolds numbers appropriate for flight. The drag coefficient is higher for
the reversed airfoil, but similar to the drag coefficient at a somewhat lower
Reynolds number for the traditionally mounted airfoil. This has been illustrated in
Figure 3, where drag measurements with NACA0012 airfoils have been compared
with data from the literature 8 ,9. The NACA0012 reversed is comparable to a flat
plate of 2% thickness. There is a consistent shift in drag coefficient with free stream
turbulence level, and one may expect that locating an airfoil inside a boundary
layer has an effect also on the drag. Also an increase in velocity will have an effect
- simple compressibility rules yield that a NACA0012 in incompressible flow has a
pressure distribution equivalent to a NACA0009 in high subsonic flow (Mach = 0.6
- 0.7). At transonic Mach numbers one may expect that open laminar separations
can create sharp drag increases.

Vortex street behind a single device


An airfoil or flat plate located in the fluid will shed vortices. A two-dimensional
vortex street will form downstream over almost any velocity range. The frequency f
of vortex formation behind a bluff body is proportional to velocity U. Taneda 10
found that bodies with a sharp trailing edge have f ex U1.5. The Strouhal number
is often used to characterize the non-dimensional frequency of vortex shedding.
It is defined as
S=f . d /U where d is a typical transverse length scale. For convenience we use a
Strouhal number based on the geometrical thickness for a plate or the maximum
thickness for an airfoil. Over a Reynolds number region the Strouhal number
changes - each change is a sign that a different local flow pattern is developing on
the body and that another law governs the Reynolds number dependence.
70

For LEBU applications flat plates and airfoils are of interest. The Strouhal number
selected by the flow depends on the thickness, t, of the fluid layer11 involved in
the vortex street formation, and will vary depending on the type of device flow:

-Flat plate with square trailing edge: t = d + 2 . 8*.


For a flat plate of given length L, 8*/L = 1.7208 / -V Re~. For extremely long
manipulator plates transition occurs on the plate, and local imperfections also
cause premature transition to a turbulent wake with loss in coherence. For a flat
plate with round trailing edge, the displacement thickness , 8*, is modified by a
separated zone. The local separation will depend on the microgeometry, but the
wake is laminar for low and moderate Reynolds numbers.

-Airfoil with open laminar separation: t = 2 . ( h + 8*) if we neglect the trailing edge
thickness. The 'effective thickness' is determined exclusively by the displacement
thickness, 8*, and the separation height, h, at the trailing edge.The trailing edge
thickness is negligible. However local geometrical changes on the airfoil can
cause a change in separation position or transition pattern for the wake.

Figure 4 gives a variety of values when plates of rounded leading and trailing
edges 12 -14 are included. For airfoils the data falls in two categories: laminar and
turbulent as will be discussed later. Both Bauer16 and Liandrat 16 have found
Strouhal numbers in excess of 0.6 along the laminar curve when the maximum
thickness is used as basis. Similar numbers are found by Paterson et.aI. 18 , Arbey
and Bataille 19 and others along the turbulent curve.

Flow on the airfoil - Open laminar separation


For manipulators laminar attached flow or separation must exist on at least on
one side of the device for the distinct vortex shedding to occur.
The laminar flow on the airfoil is estimated using Thwaites' method 20 for laminar
boundary layers:

e2 = ( 0.45 . v /U6) f U5 dx = L2 W / ReL (1)

where e= momentum thickness, v = kinematic viscosity, U =Iocal free stream


velocity, (U/U.ref)sep = Velocity at separation/Reference velocity, Re.L = Uref' L /
v = Reynolds number based on device chord
71

_---->00=.4""'-5_ _ Sf (U/ Ure f)5 dS (2)


and W is the integral: W = (U/ Uref ).sep 0

S = xsepfL

The integral W is a result of the inciscid flow set up around the airfoil, and since it
depends on the pressure distribution and separation location, it is to a first
approximation independent of Reynolds number.

The shedding frequency fsh for flow around bluff bodies 11 is fsh t / U = 0.16. For
an airfoil the frequency of shedding fsh from each side is given by21:

fs h=0.08 . U / t (3)

A rough estimate of the separation height h, is given by the separation angle:

h = L . (1-S) * tan(y) (4)

For a closed separation bubble y is related to the momentum thickness Reynolds


number at separation by the empirical equation. If we assume that this relation
holds also for an open laminar separation at low angles of attack:

Y= tan- 1( B / Res.s ep ) where we use the fit B= - 5.736 + 3.704 In(ReS.s ep ) (5)

The parameter B is usually assumed constant with a value of 15 to 20. The fit
chosen here is based on results by Bandyopadhyay17 and Liandrat 16 with two
different NACA0009 airfoils. Choice of a constant B would change the computed
curves slightly.
The displacement thickness is obtained directly from the momentum thickness at
separation using the shape parameter at separation, Hsep = 3.55 and assuming
that it does not change over the separated area. This is a reasonably assumption
since the pressure gradient is almost zero over the region back to the trailing edge.

The 'equivalent thickness' of an airfoil with open laminar separation becomes:

(6)
j(
_ 2 1 - ~sep) B + Hsep )
- t- -
8 sep (-.lL).
Uinf sep
W
72

The Strouhal number for an airfoil with open laminar separation is :

0.16 . H
<p= 3LL
(1 - ~sep) 8 H iW (7)
(-.1L)
Uinf. sep
YW + sep

Transitional/turbulent flow
From Figure 5, it is clear that the cases with open laminar separation are well
described with the simple equations. As the Reynolds number increases, the
shedding frequency drops and in a region it is possible to obtain several values of
the Strouhal number. The existence of a bifurcation point due to an absolute
acoustic instability in the wake has been predicted by Koch22 for flow behind flat
plates with blunt trailing edges. In the present case, it is reasonable to regard the
separated airfoil case as similar, since the separated area acts as a bluff body at
the trailing egde. This corresponds to the sudden jump in Bauer's result for the
NACA 0012 airfoil, and the knee in the drag polar of the airfoil.

As the Reynolds number is increased further, one obtains Strouhal numbers

*
following a correlation curve given by Paterson et.aL 18, which for our purpose is
expressed:
s = 0.011 YRed (8)

As discussed by Arbey and Bataille 19 , the actual shape of the curve is governed
by an acoustic feedback loop producing a ladder-type behaviour associated with
laminar flow on the pressure side of airfoils.

APPLICATION TO MANIPULATORS
If the manipulator geometry is sharp edged plates or ribbons, a laminar wake will
be formed immediately behind the front (or single) device as long as the Reynolds
number based on device length is well below the transition Reynolds number.
Airfoils tested so far have been limited to NACA0009 and related shapes. They
have an open laminar separation only at low Reynolds numbers, and will not work
as the Reynolds number is increased or if they are moved inward in the boundary
layer where the turbulence intensity and structure can cause premature transition
in the free shear layer of the airfoil.
73

Although the wake behind the front device initially is laminar, the first phase of
transition will develop rather rapidly as described by Sato and Kuriku 12 and
Miksad et.aI. 23 . In the first, short region, the linear instability region, higher
harmonics to the fundamental start to develop. The maximum amplitude is initially
concentrated to the vertical positions corresponding to the highest mean velocity
gradient. Rather soon the amplitude spreads more evenly through the wake, but
this does not mean the flow is turbulent. The transition region is rather long 12 , and
is still dominated by distinct 'carrier' frequencies and their sidebands. It is essential
that the rear airfoil is located in the deterministic, transitional part of the wake.

Flow on the rear device


In visualizations, the vortices of an airfoil wake follow two separate paths 26 , all
vortices of the same sign forming a sheet. The impingement of these sheets on the
rear airfoil is critical, since it may cause inhibition or enhancement of the vortex
street behind the rear device 24 .
The flow on the rear device is important for the formation of the vortex street. The
boundary layer is initially laminar, but depending on the 'free stream' turbulence, in
this case the turbulence structure in the outer part of the 'host' boundary layer, the
flow may go through transition and the airfoil mayor may not have laminar
separation bubbles. The largest structures of the 'host' boundary layer, up to 5 0
will have limited influence due to the scale mismatch. The vortex street from the
front device will have a much larger influence, since it contains a fundamental
closer to the maximum unstable frequency of the separated layer. The main role of
the rear device is to create a vortex shedding that can modulate the downstream
wake. The device always has a vortex shedding of a frequency f2 detuned from
the front device frequency f1' since its drag and general flow characteristics are
close to, but not exactly equal to, that of the front device.

The wake behind the rear device


In the wake close behind the rear device, the two basic frequencies of the vortex
streets f1 and f2 will exist, but more importantly development at the difference
frequency f 2-f 1 develop. Sato and Sait0 25 gave the following three rules for the
non-linear interaction in a transitional wake:
i) The growth of a spectral component is suppressed by the presence of another
strong component.
ii) Mutual interaction is more effective when the amplitudes of interacting
components are closer.
iii) Stronger interaction takes place between components of closer wavenumbers.
The wake downstream of the rear device thus develops under the influence of a
growing, low frequency component created by the two basic harmonics f1 and f2.
74

While the higher harmonics soon dissipate and lose their identity in the turbulent
wake, the low difference frequency, f2-f 1, is capable of setting up the strong vortex
pattern that is essential for the turbulence manipulation.

So far only limited documentation exists to support the model described.


Liandrat 16 has confirmed the existence of the two frequencies f1 and f2 behind
NACA 0009 airfoils at low Reynolds numbers, and also the development of the
f 2-f 1 frequency downstream of the rear airfoil. This has been done both outside the
boundary layer and in the outer part of it with airfoils of chord L=38.1 mm, made of
polished stainless steel. For high Reynolds numbers, however, no data is yet
available to assess the model.

CONCLUSIONS AND SUGGESTIONS


A model for the development of a wake structure suitable for manipulation of a
turbulent boundary layer has been described. Some of the important points in the
analysis are:
- The main effect in reducing the skin friction is production of a vorticity structure
between the inner and outer part of the boundary layer, blocking the phase transfer
of the small scale motions.
-Airfoils have a vortex shedding distinctly different from that of flat plates, and to
yield a unique, laminar vortex development, the airfoil must have laminar or
separated flow at the trailing edge.
-A laminar or transitional wake must exist at the location of the rear device in a
tandem configuration.
-The rear device sheds vortices at a frequency slightly different from the incoming
vorticity
-Interaction between the two harmonics produce a low frequency fluctuation that is
capable of delaying the wake transition process and give it the structure required
for its blocking role.

REFERENCES

1. Anders,J.B. 'LEBU drag reduction in high Reynolds number boundary


layers.' AIAA Paper 89-1011 (1989)

2. Bandyopadhyay,P.R. and Hussain, A.K.M.F.: ' The coupling between scales


in shear flows.' Phys. Fluids 27(9), Sept 1984

3. Bradshaw, P.: ' 'Inactive' motion and pressure fluctuations in turbulent


boundary
layers.' J Fluid Mech, Vol 30, part 2, pp. 241-258 (1967)

4. Badri Narayanan, M.A., Raghu, K. and Poddar, K.: 'Wall shear fluctuations in
a turbulent boundary layer.' AIAA Journal ,Vo122 No, pp. 1336-1337
75

5. Andreopoulos, J., Durst, F. and Jovanovic,Z.Z.: ' Influence of Reynolds


number on characteristics of turbulent wall boundary layers.' Experiments in
Fluids, 2, pp. 7-16, (1984)

6. Bertelrud,A.: ' A profile family for use in boundary layers with large-eddy
breakup devices." DRAG REDUCTION '89, Davos, Switzerland, July 31-August 3,
1989

7. He,Dexin and Zhang, Weizhi: ' Aerodynamic performance of 30 types of


airfoils at high angles of attack.' (In Chinese) CARDC Low Speed Aerodynamics
Institute & China Wind Energy Development Center. (1988)

8. Anders,J.B. and Watson,R.D.: ' Airfoil large-eddy breakup devices for


turbulent drag reduction.' AIAA Paper 85-0520 (1985)

9. Miley,S.J.:' A catalog of low Reynolds number data for wind turbine


applications. ' Department of Aerospace Engineering, Texas A&M University,
College Station, Texas, Report No. RFP-3387 (1982)

10. Taneda, S.: ' Oscillations of the wake behind a flat plate parallel to the flow. '
J. Phys. Soc. Japan, 13 (4) ,418-425 (1958)

11. Levi, E.: ' A universal Strouhal law.' J. Engineering Mechanics, ASCE, Vol.
109, No.3, June 1983, pp. 718-727.

12. Sato,H. & Kuriki,K.: ' The mechanism of transition in the wake of a thin flat
plate placed parallel to a uniform flow.' J. Fluid Mech, 11, 321-352 (1961)

13. Lawazceck,O. and Kreplin, H.-P.: ' Experimental investigations on the


unsteady behaviour of the flow in the wakes of flat plate profiles.' DFVLR IB
29001-83 A 03 (1983)

14. Heinemann,H.J., Lawaczeck, 0. and BOtefisch,K.A.:' V. Karman vortices and


their frequency determination in the wakes of profiles in the sub- and transonic
regimes,' In Symposium Transsonicum II, Gottingen, September 1975. Springer
Verlag (1976)

15. Bauer,A.B.: ' Vortex shedding from thin flat plates parallel to the free stream. '
J. Aero. Sci., 28, 340-341 (1961)

16. Liandrat,Marie-Pierre. Private communication.


76

17. Bandyopadhyay, P.R.: Resonant flow in small cavities submerged in a


1

boundary layer.' Proc. R. Soc. London A. 420,219-245 (1988)

18. Paterson, R.W., Vogt, P.G., Fink, M.R. and Munch, C.L.: ' Vortex noise of
isolated airfoils.' J. Aircraft, May 1973, pp. 296-302.

19. Arbey, H. and Bataille, J.: ' Noise generated by airfoil profiles placed in a
uniform laminar flow.' J. Fluid Mech, (1983), vol 134, pp. 34-37

20. White,F.M. 'Viscous fluid flow.' McGraw-Hili (1974)

21. Sigurdson,L.W. and Roshko,A.: 'The structure and control of a turbulent


reattaching flow.' In Liepmann,H. and Narasimha, (Eds.): 'Turbulence
Management and Relaminarisation.' IUTAM Symposium, Bangalore, India (1987)

22. Koch, W.: 'Local instability characteristics and frequency determination of


self-excited wake flows.' J. Sound and Vibration (1985) 99(1), 53-83

23. Miksad,R.W., Jones,F.L. and Powers,E.J.:' Measurements of nonlinear


interactions during natural transition of a symmetric wake. Phys. Fluids (1983) Vol
26 (6) , June 1983, pp. 1402-1409.

24. Meier, G.E.A. and Timm, R.: 'Unsteady airfoil vortex interaction.' AGARD
CP-386. (1985)

25. Sato,H. and Saito,H.: ' Fine-structure of energy spectra of velocity


fluctuations in the transition region of a two-dimensional wake.' J. Fluid Mech
(1975), vol. 67, part 3, pp. 539-559.

26. Savill,A.M. and Mumford,J.C. : ' Manipulation of turbulent boundary layers by


outer-layer devices: skin-friction and flow-visualization results.' J. Fluid Mech
(1988), vol. 191, pp. 389-418.
77

Figure 1 Schetch of vortex development behind a tandem device.

Forced
ShC<lr~Layer Possible ~airing l::1
C____ --=-- ? ~? ~ ? ~ ? ~ C_____
----
~
????~$~~=- FT
~~--.:>-~~v_=_ ~

Figure 2 Drag coefficient for NACA0012 at Re.c=321,OOO. (Ref. 7)

0.06
C NACA0012
NACA 0012 Reversed
0.05

l/
0.04
vt;
0.03 1/
CD l/

0.02
V
t--- ji....
"
1,..-1,..-
L.--

0.01

0.00
-5-4-3-2-1012345

Alpha [deg]
78

Figure 3 Drag of NACA 0009 airfoil at Alpha = 0 deg


as function of Reynolds number.
0.04
--- Computations, Tu=O.l%
I:. experiment, Tu=2%
experiment, Tu=O.08%

0.03
-... ,

CI
NACAOO09reversed(Llandrat)
Flat plate, IIL=O.Ol
NACA0009 (Anders)

0.02 0
0 NACA0009 (Llandrat)

CD
1:..
0.01 o~
0.00
1 03

RE.c

Figure 4 5trouhal number 5=f dIU for NACA 0009 & 0012 airfoils.
Flat plate results are shown for comparison.
1.0
Comp.-Laminar
Comp.turbulent
0.8 I. D Bandyopadhyay
o Liandrat
I
Bauer
) Arbey & Batailie
0.6 I Lawaczeck & Kreplin
5TROUHAL

"
I,. Sharp T.E.
NUMBER ti

0.4
J1 .. !,: a.
RoundT.E.

I:.~ '&~
0.2
~~
./

0.0
100 1000 10000 100000 1000000

RE.d
BOUNDARY LAYER FLOW VISUALISATION PATTERNS ON A RIBLET
SURFACE

D. G. CLARK, Department of Aeronautical Engineering


Queen Mary and Westfield College, University
of London, Mile end Road, London E1 4AS U.K.

Abstract
Boundary layer flow visualisation methods, developed at
Queen Mary and Westfield College, have been applied to a
riblet surface. The results reveal cellular cross flows
developing in the grooves between the riblets. These local
flow regimes appear to have little direct effect on the flow
in the wall layers immediately adjacent to them.
Qualitatively, the behavior of the wall layers appears to be
that which would be expected if a virtual surface existed at
a level slightly above the riblet tops, but a tendency for
the origin of longitudinal eddy pairs to become anchored to
the top of a riblet is noted.

1. Introduction
Experiments with flow visualization in a turbulent
boundary layer have been in progress in the Department of
Aeronautical Engineering at Queen Mary and Westfield College
for a number of years (Refs. 1 & 2). Oil droplet smoke is
used, with plane beams to give cross sections of the
resulting smoke patterns. Such techniques are now familiar,
but the light beams used at Q.M.W. are not produced by the
more usual method of spreading a laser beam, but by shining
a focused beam, from either tungsten halogen projector lamps
or standard photographic flash units, through narrow slit
masks. The original aim of these experiments was to find a
less expensive alternative to laser light.
As a preliminary to a proposed programme of studies,
tests were undertaken to establish whether these lighting
methods could be successfully applied to a riblet surface.
Though the tests were only exploratory, the photographs
obtained were unexpectedly informative.

79
80

Central to the technique is the use of low free stream


speeds to obtain representative turbulent features on a
large physical scale, developing relatively slowly. This
has the disadvantage that without facilities of inordinate
size, the boundary layers studied will always be in the
transitional range and must be tripped for consistent
results. The element of artificiality introduced by the
presence of the trip, and the low Reynolds number of the
boundary layer must always be borne in mind when applying in
a wider context any conclusions reached .
A detailed explanation for the effectiveness of
streamwise riblets in reducing drag is closely dependent on
explanations of the mechanisms of turbulence production
within the wall layers. The picture of activity presented
by Head and Bandyopadhyay (Ref 3) now seems over-simplified,
in that it placed too much reliance on one flow feature, the
vortex loop. Though such features can be found, and though
their initiation, growth, and decay may well represent an
important component of turbulent activity, the work of Falco
(Ref 4) suggests that a wider variety of features needs to
be invoked for a complete account.
Whatever the precise mechanisms are, it is clear that
lateral movements of the flow close to the wall are an
essential feature, and it seems likely that it is in
interfering with these cross-flows that the riblets produce
their effect. Falco discusses their role in inhibiting the
formation of the "pocket" features which figure prominently
in his descriptions of flow mechanisms in the wall layer.
Bacher and smith (Ref. 5) suggest that an important effect
may be the action of riblets in weakening the activity of of
the streamwise vortices or eddies which feature in most
descriptions of wall layer activity.
The optimum height and spacing for drag reducing
riblets of triangular cross-section has been given as 15
wall units by Walsh and Lindemann (Ref. 6). Wilkinson and
Lazos (Ref. 7), using thin L-shaped or "blade" riblets,
found an optimum height of 8 wall units, with a spacing of
10 wall units, but spacings of more than 100 wall units
appeared to result in drag reduction, provide that the
riblet height was not increased. In associated hot wire
studies, the same authors detected a well defined change in
the distribution of turbulence in the vicinity of the
riblets when the spacing was raised beyond a value of
approximately 50 wall units.
For the proposed flow visualisation, it was desirable
to use riblets of as large a size as would be consistent
with representative results. Apart from the indication of
81

an abrupt change when the spacing exceeds 50 wall units,


noted above, it seemed likely that the essential features of
the interaction between the riblets and the flow features in
the wall layers would be observable on riblet heights and
spacings somewhat larger than the optimum values for drag
reduction. An upper limit of 15 wall units for riblet
height, and 30 for spacing was therefore decided on. The
dimensions achieved, in terms of wall units, were somewhat
smaller, as explained below.
2. The Wind Tunnel and Lighting Methods
The floor boundary layer of a simple, purpose built, open
return wind tunnel was used for these tests. The working
section floor has a recessed opening 1. 83m (6ft) long by
0.61m (2ft) wide into which removable panels can be
inserted. For the primary flow visualisation technique,
plate glass panels are used. These are coated with an
opaque layer of matt black paint on one face. The paint is
scraped away along strips 1.0-2.0mm wide leaving a series of
transparent slits. An incandescent or discharge tube light
source, with a suitable reflector, placed about 20cm below
the floor, then produces approximately plane beams emerging
through the tunnel floor. The resulting beams are somewhat
divergent, but they are well defined in the region of
interest close to the surface.
The method is less versatile in some ways than the use
of a laser, but it has the advantage that simple multiple
beam configurations present no problem, and colour can be
used to discriminate between the different beams if
required.

The general arrangement of the tunnel and the riblet


surface is shown in Fig. 1. The tunnel has a length of some
4.5m to the working region, and a cross-section 1.2m wide by
0.6m high. The results were obtained at a free stream speed
of 0.8 m/s. At this speed, the floor boundary layer was
just unstable, and, in absence of a trip, turbulent spots
appeared sporadically in the working region. with a
cylindrical trip 6mm (1/4 inch) in diameter, positioned on
the tunnel floor about 1.4m downstream of the entry gauze, a
steady turbulent boundary layer was present in the working
region, though it was at the lower limit of Reynolds number
for sustained turbulence.
At the above tunnel speed, velocity measurements in the
boundary layer at the location used for the present
photographs, made using both a hot wire probe and combs of
2mm diameter pitot tubes, give a boundary layer momentum
thickness of 14mm, and a corresponding momentum thickness
82

Reynolds number of 770. These measurements have also been


fitted to a modified Clauser plot, using the inner wall law
proposed by Van Driest (Ref. 8), with constants recommended
by Coles (Ref 9). The result is shown in Fig 2. The values
for shear velocity given by these plots are 0.0364 m/s for
the pitot tube readings, and 0.0374 m/s for the hot wire
measurements. The difference is attributed to the
difficulty associated with with obtaining accurate and
consistent readings of flow velocity and dynamic pressure at
the low speeds involved. it is felt however, that the
agreement is sufficiently close for for a value of shear
velocity of 0.037 to be assumed, with an associated scale of
2.5 wall units/mm.
3. The Riblet Surface
In order to undertake the proposed programme, a
difficulty to be overcome was that of forming a suitably
transparent riblet surface. Machining from solid perspex
and subsequent polishing would probably produce a very
effective lateral diffuser for the emerging beam. A plane
beam normal to the riblets would not be unduly affected, but
beams at other angles would be seriously degraded. The
solution finally adopted was simply to glue riblets, in the
form of strips of matt black solid polystyrene, directly on
to the surface of the glass plate, on which the slit masks
had already been formed. The household adhesive "UHU" was
found to be very effective for this purpose. The
pOlystyrene strips were left approximately square topped for
these prel iminary experiments, but it was found that they
were sufficiently firmly attached for it to be possible to
shape them to some desired profile by scraping, if required.

A decision on riblet height was made on the basis of an


estimated scale of 3 wall units/mm. Using the upper limit
of 15 wall units, discussed in the introduction, a height of
5mm was chosen. A lateral spacing of 10mm was decided
partly by the need to allow for the insertion of a standard
hot film element between the riblets if required. At the
estimated scale, this corresponds to 30 wall units but also
conforms to the limits discussed earlier. The determination
of local surface shear stress given above indicates that the
actual height and spacing achieved are approximately 12.5
and 25 wall units respectively.
It should perhaps be emphasised here that the method
of construction employed dictated the use of L-shaped
riblets, and, as noted above, they are square topped. Any
conclusions reached may not be valid for riblets of other
cross-sections.
83

The riblets were mounted on a single glass plate, 1.2m


(4ft) long. The riblet region itself was 1m long by 28cm
wide, offset somewhat from the centreline so that an
adjacent region of plain surface, well clear of the tunnel
walls, was available for purposes of comparison. A series
of transverse slits in the paint surface was provided
towards the rear end of the riblet region, extending
laterally over the adjacent plain surface. The slits were
spaced at 2cm intervals.
4. The experimental arrangement
The location of the riblets in the tunnel can be seen
in Fig. 1. Smoke was introduced either through a tangential
slot in the surface, 7 cm upstream of the leading edge of
the riblets, or from a smoke wire close to the riblet tops,
immediately upstream of the region photographed (Fig. 1).
A standard 35mm camera was used, with a 200mm lens on
an extension mount. The flash units are a standard portable
professional type, used at 1/4 and 1/2 power, giving
effective exposure times of the order of 0.5 milliseconds.
The view chosen for the tests is directed upstream and
towards the surface at approximately 45 degrees. This view
has been found to give the most generally revealing and
easily interpreted smoke patterns. Since a single light
source is used, the beam array obtained with a multiple slit
mask is divergent. The actual geometry achieved is shown in
Fig. 3.
5. Results and discussion
The orientation diagram of Fig. 4 depicts a wide angle
view of the scene visible in the photographs. The camera
was positioned above the tunnel looking through a glass roof
panel. The 200mm lens used gives a partial telephoto
effect, which suppresses perspective, and yields an image
which can be confusing at first sight. The slits appear as
bright horizontal lines, whilst the riblets appear as a set
of near-vertical, slightly convergent, dark lines, brightly
outlined on each side.
5.1 Smoke input through surface slot
5.1.1 Multiple cross sections on the riblet surface
The first results, using the multiple beam array with
smoke input through the surface slot, are illustrated by the
examples shown in Figs. 5 to 6. These reveal, as might be
expected, smoke-filled fluid clinging to the internal
84

corners and floor of each groove. In many places however


there are also features which can only be interpreted as
signs of cellular cross-flows in the grooves between the
riblets. In some cases these cross-flows appear to have the
form of single rolls, with, in places, an accompanying
suggestion of lateral movement in the exterior flow, across
the top of the riblets. This is illustrated by the two
scroll features on the lower left of Fig. 5, and the single
feature a little right of centre in Fig. 6. In other cases,
the cross flow pattern appears to consist of an outflow along
both sides of one riblet with some sign of balancing inflow
along its neighbours. This is indicated slightly right of
centre, in Fig. 7, in the third cross-section from the
bottom. In some cases careful scrutiny of the negatives
reveals patterns which have a streamwise extent of six to
eight riblet spacings (150-200 wall units). some indication
of this elongation can be seen in both figures, though two
levels of reproduction tends to obscure the image.
These features could be interpreted as lending support
to the mechanisms suggested by Bacher and Smith, and
mentioned in the introduction. It is felt however that on
the evidence of the present results, such speculations are
not justified. The suggestion of cellular rotating flows
between the riblets is strong, but in the photographs
presented here, with the exception of the "mushroom"
features and "pockets" described below, there is little to
indicate the nature of the overlying motions.
5.1.2 Comparison with the plain surface
Moving the view across to include both the riblets and
part of the plain surface, illustrates the difference
between the two regions. A typical example is shown in
Fig. 7. In this view some low angle overall lighting was
added, and the lines of low speed streaks are just visible,
superimposed on the patterns of the cross-sections.
Apparent in this view, and in others of the series, are
broad regions on the plain surface which are cleared of
smoke, presumably by fluid from the outer wall layers moving
towards the surface, and spreading out across it. This type
of motion, responsible for the formation of a "pocket"
(Ref. 4) in a newly injected smoke sheet, is very
characteristic of the wall layer behaviour. Over the riblet
surface, in the layer visible with the illumination used
here, similar broad sweeps are not apparent.
In these multiple cross sections, viewed without the
use of colour, only the inner part of each cross section is
clearly visible. It has to be borne in mind that though the
images are bright, the smoke is in fact very tenuous; the
85

cross sections are transparent, and both the intensity of


the illumination and the density of the smoke, decrease with
increasing distance from the surface. Consequently, in most
cases, the outer portions of each section are not apparent
where they overlap the more vivid image from the inner part
of the next section.
5.1.3 Single cross sections
Restricting the lighting to just one slit, shows the
extent of smoke penetration into the outer layers, and
indicates something of the nature of the activity there, as
illustrated by Figures 8 to 11.
Fig. 8 shows, just right of centre, a feature which
appears to consist of a layer of relatively passive smoke
laden fluid which has been lifted away from the surface by
the action of a typical double eddy, with its characteristic
"mushroom" cross section, erupting below. It is noticeable
that this whole feature seems largely independent of the
cellular activity taking place between the riblets beneath
it. Some overall low angle illumination was also used in
this view, and it appears to reveal that the central feature
spans a low speed streak. Fig. 9 was obtained with the same
lighting, and carries a very clear example of the ubiquitous
"mushroom" feature, this time in isolation. This feature
also appears to span a low speed streak and again, the
activity which formed it seems to have been mainly
independent of the cross-low pattern between the riblets,
though in this and subsequent examples, there is some
indication that features of this particular form tend to be
"anchored" to a riblet immediately beneath them. It appears
that these typical wall layer patterns have their origin in
the layer of fluid just clear of the riblet tops.
In Figs. 10 & 11, single sections at the edge of the
riblet surface, this time without overall illumination, show
the general differences between the patterns over the plain
and riblet surfaces particularly vividly. The marked
difference in thickness of the surface smoke layer on the
two types of surface and the apparent origin of the
wall-layer features are both particularly clearly
illustrated.
5.2 Smoke input from the smoke wire
Though the impressions of activity given by the
photographs so far described are very persuasive, they are
not really firm evidence that cross-flow motion is actually
present. They could be patterns formed far upstream, say at
the leading edge of the riblets, and carried downstream in
86

relatively quiescent flow. The use of a smoke wire


immediately upstream of the region of observation can
provide direct evidence of activity since the distortion of
the smoke sheet can only be caused by activity in the region
itself.
Fig. 12 shows a typical example of multiple cross
sections, obtained with a smoke wire positioned 2mm above
the riblet tops, just upstream of the range of the camera
view. It becomes clear that the cross-flows indicated in
the earlier views are indeed active. Movement into and out
of the spaces between the riblets is taking place, and some
of the features of the cross-flow patterns noted earlier
develop in a comparatively short streamwise distance. A
clear example occurs on the third section from the top of
the figure, just left of centre. There, the smoke has moved
down to the floor of the groove on each side of a riblet,
and has moved round far enough to form one complete turn of
a scroll. This movement has taken place in a streamwise
distance of about 150 wall units.
At the top of Fig. 12, just right of centre, a feature
appears which is typical of those suggesting movement of a
sizable body of fluid toward the wall, between two counter
rotating scroll-like features. The pattern appears to be
similar to those seen over the plain surface, and remarked
on above (para. 4.1.2), except that here the flow does not
extend to the surface, but spreads out just above the riblet
tops, rather as if a virtual surface exists there.
In the views obtained with the smoke wire, clearly
identifiable patterns, such as those noted above, are quite
common, but, nevertheless, the majority of the field of view
tends to contain less easily interpreted shapes.
with the beam spacing used in these views, since the
smoke sheet is so rapidly distorted, the images in different
planes soon begin to overlap and the picture becomes
confused. Doubling the distance between the light sheets,
by blanking out alternate slits, gives more easily
interpreted images. An example of the results obtained,
wi th the beam spacing doubled in this way, is shown in
Fig. 13. A steady migration of the smoke toward the surface
over most of the regionis noticeable, with movement away
from the wall apparently confined to narrow regions. A
particularly clear double scroll formation is visible in
this view, just to the left of centre. The flow feature
responsible appears to be similar in form and scale to those
visible in the centre of Fig. 9 and on the right hand side
of Fig. 10. In this example, the origin of the feature in
the outward flow from either side of a riblet seems
87

particularly clear.
Raising the smoke wire to 7mm above the riblet tops
produces a marked change in the type of pattern produced.
An example is shown in Fig. 14. The smoke tends to remain
clear of the riblets for some considerable distance. At
that level the direct effect of the rib lets is apparently
less important than the larger scale movements taking place
in the wall layers immediately above. The uneven nature of
the smoke sheet in this view is due to the oil on the smoke
wire collecting into globules, an effect which is difficult
to avoid with certainty.
Moving the view across, as before, to include part of
the plain surface, again shows the difference between the
two flow regimes. The patterns which form on the plain
surface have a wavelike appearance, rolling over to form
streamwise scrolls in some instances but by no means always.
Figs. 15 & 16 show a typical examples. A comparison with
the general form of the pattern in Fig. 14, in which the
smoke wire was approximately one riblet height above the
rib let tops reveals similarities , giving further support to
the idea that the normal flow patterns of the wall layers
are not greatly altered by the presence of the riblets, but
that they are simply displaced from the surface.
6. Conclusions
Considerable activity directly associated with the
presence of the riblets, and consisting of cellular
cross-flows, exists in the region within about 1.2-1.5
riblet heights from the surface. These movements, wi thin
what might be termed a "riblet sublayer", appear to have
little direct effect on the activity in the adjacent wall
layers, but they are strongly influenced by the nature of
the motions in the adjoining flow.
The patterns which appear tend to have the same general
form as those which can be seen over a plain surface, but
the level from which recognisable features appear to
originate, is displaced from the layer immediately adjacent
to the surface, to a region just above the riblet tops.
These results are consistent with the view that one of
the primary effects of the riblets is to prevent inrushes of
high speed fluid from spreading out laterally close the
surface. Such movement is displaced away from the wall by a
distance somewhat greater than the height of the riblets,
and the ability of the high speed fluid to penetrate close
to the basic surface is strongly inhibited.
88

7. Acknowledgements

The author is indebted to Dr H.P.Horton for advice on


the method used to obtain the Clauser plot from which the
effective size of the riblets has been deduced, and for the
use of a computer program to perform the calculations.
The work was supported by a grant from British
Aerospace (Commercial Aircraft) Ltd., Airlines Division,
Hatfield.
References

1. Clark, D.G. "Visualisation of Flow features in Wall Bound


Flows" Contribution to Lecture Series 11, "Flow
Visualisation and Digital Image Processing",
Von Karman Institute, Brussels, June 1986.
2. Clark, D. G. "Organised Flow Features in Boundary Layer
Turbulence" Royal Aeronautical Society Conference,
"Turbulent Drag Reduction by Passive Means",
London, U.K., September 1987.
3. Head, M.R. & Bandyopadhyay, P. "New aspects of turbulent
boundary layer structure" J.Fluid Mech. (1981),
vol. 107, pp.297-338.

4. Falco, R.E. "New Results, A Review and Synthesis of the


Mechanism of Turbulence Production in Boundary
Layers and its Modification"
AIAA Paper No. 83-0377 (1983).
5. Bacher, E.V. and Smith, C.R. "A Combined Visualisation-
Anemometry Study of the Turbulent Drag Reducing
Mechanisms of Triangular Micro-Groove Surface
Modifications" AIAA Paper No. 85-0548 (1985)
6. Walsh, M.J. and Lindemann, A.M. "Optimisation and
Application of Riblets for Drag Reduction"
AIAA Paper No. 84-0347 (1984).

7. Wilkinson, S.P. and Lazos, B.S. "Direct Drag and Hot Wire
Measurements on Thin Element Riblet Arrays"
Turbulence Management and Relaminarisation, IUTAM
symposium Bangalor 1987 (Springer-Verlag 1988)
8. Van Driest, E.R. "On turbulent flow near a wall"
J Aeronaut. Sci. 23, pp.1007-1011 & 1036.
9. Coles, D. and Hirst, E.A. "Computation of turbulent
boundary layers" AFOSR-IFP-Stanford Conf. 2 (1968)
89

Smofte '-.titre

/X/-COYr'\era

/'
~

Ri.blet Geometr-!:I

Fig. 1 Experimental arrangement

20~----------------~----------------~----------~------,

Wall Law
u* (Van Driest)

10~----------------~~--------------~----------------~

Pi tot tube A
Hot Wire +

oL----------------L________________L -______________ ~

o 1 2 3
Log(y*)

Fig. 2 Tunnel boundary layer; Clauser plot


90

~'
, Fla.=e",,-_,.

Fig. 3 Geometry of light sheet array

Carne"'a vtew frame"

Fig. 4 The camera view


91

Fig. 5 Mul tiple Cross s,ections over the riblet surface

Fig. 6 Multiple Cross sections over the riblet surface


92

Fig. 7 Plain surface included for comparison

Fig. 8 Single cross section with overall lighting added


93

Fig. 9 Single cross section with overall lighting added

Fig. 10 Plain surface included for comparison


94

Fig. 11 Plain surface included for comparison

Fig. 12 Multiple sections, smoke wire 2mm above riblet tops


95

Fig. 13 Smoke wire 2mm above riblets, beam spacing doubled

Fig. 14 Smoke wire 7mm above riblets, beam spacing 2cm.


96

Fig.15 Smoke wire at 2mm, view including plain surface

Fig.16 Smoke wire at 2mm, view including plain surface


Simultaneous Flow Visualization and LDA Studies over
longitudinal micro-grooved surfaces

by C.J.A.Pulles 1, K.Krishna Prasad 1 and F.T.M.Nieuwstadt 2


1 Laboratory for Fluid Dynamics and Heat Transfer, Faculty of
Physics, Eindhoven University of Technology, Eindhoven
2 Laboratory for Fluid Dynamics, Faculty of Mechanical
Engineering and Maritime Technology, Technical University of
Delft, Delft, The Netherlands

Abstract
In this paper we present some results obtained from simultaneous LDA and flow
visualization measurements over several types of longitudinal, microgrooved
surfaces. No new structures were detected above the grooved walls apart from
those that exist over the smooth walls. The measurements indicate subtle
differences in flow over the smooth and grooved surfaces. In particular the shear
stress distribution in the low speed streak seems affected. These differences are of
great interest because the shape of the riblets used effects a decrease in drag
compared to the smooth wall drag (Walsh 1982).

1. Introduction.

In an earlier paper (Pulles et al. 1989) the authors presented measurements of


drag and local velocity statistics over a smooth plate, a spanwise rough wall and
several longitudinal microgrooved walls. The major conclusion of that work is
that the current theories are inadequate to explain the drag reduction observed
over grooved walls. In particular the VITA burst detection procedure and the
quadrant analysis did not lead to any evidence for a change in specific events that
clearly results in drag reduction. For example the streamwise fluctuating signal
showed according to the VITA procedure an increased bursting rate above the
grooved plate, while no significant difference in the conditional averages could be
detected. This not only questions the significance of the bursting process (as it
follows from the VITA technique) as an indicator of turbulent shear production
but more importantly it does not agree with the observed reduction of turbulent
intensity in the streamwise direction above a grooved plate. Applying the VITA
burst detection technique on the velocity component perpendicular to the wall we
found not only a decrease in frequency of triggering events ("vertical velocity
bursts"), but also a less pronounced average "burst "-shape, when one compares
the grooved wall measurements with the smooth wall measurements. Contrary to
the naive assumption that a decrease in "vertical velocity burst" frequency and
amplitude should be accompanied by a decrease in turbulent intensity normal to
the wall, no such decrease was evident in the measurements. The quadrant
analysis clearly shows a reduction in ejection activity and an increase in sweep

97
98

activity above the grooved plate. One would expect from these results that the
turbulence intensity normal to the wall should decrease and this is not supported
by measurements of turbulent intensity mentioned earlier. However the effect of
reduced sweep activity on the turbulence intensity is not that clear.

Since the work presented in this paper was done, a very detailed study on the
structure of a turbulent boundary layer over a riblet surface was published by
Choi (1989). The work included measurements of velocity, turbulence intensities,
skin friction apart from fairly extensive flow visualization photographs over
smooth and riblet surfaces (Choi also provides an up-to-date list of references on
the subject). The VITA technique applied to the wall friction signals showed that
the frequency of the near-wall bursts over the riblet surface was nearly eight
times that over the smooth plate. However the duration of the burst was reduced
by a factor of two. In addition the turbulent statistics for the two surfaces were
different. The skewness and kurtosis of the skin-friction signal over the riblet
surface was 1.31 and 6.82 respectively compared to 0.42 and 3.12 over the smooth
surface.

Choi also presented a conceptual model based on pairs of counter-rotating


longitudinal vortices. Flow visualization studies showed that the average spanwise
distance between these vortices over a riblet surface was nearly twice that over
the smooth surface. The conditionally averaged vertical velocity field of the
near-wall bursts indicate that the flow direction during the event was negative
and the extent of this region for the riblet surface was twice as large as that for
the smooth surface. The near-wall bursts were also observed to coincide with the
formation of these vortices. Taken together these results suggest that the near
wall bursts are dominated by sweep-like motions.
The discussion above suggests that there is a difficulty in reconciling the
"picture" suggested by the results of burst measurements by the VITA technique
and the results of quadrant analysis. The model suggested by Choi is inspired by
the VITA results, though it is supported by visualization studies. However these
results are not consistent with the measured turbulent intensities. It seems thus
the resolution of this conflict requires more detailed statistical analysis. We bear
in mind that much of the work on VITA-burst detection and quadrant analysis
was also initiated and stimulated by the results of flow visualization (see ego Kline
et al. 1967). In this paper we present some results of simultaneous flow
visualization and LDA measurement of flow velocities. We also provide the
results of a first attempt to associate in a statistical sense the measured Reynolds
shear stress with the visually observed structures. AS yet no accurate theory
predicting the shape, frequency and intensity of coherent structures in a turbulent
boundary layer exists. Therefore for purposes of comparison we include
measurements over smooth wall and the transverse grooved wall apart from those
on the longitudinal grooved wall.
2. The experimental arrangements.
The experimental results reported here were all acquired in a water channel.
Details of the system with its associated instrumentation are given by Pulles
(1988).
99

In the water channel a pulsed hydrogen bubble flow visualization facility is used.
Images were recorded with a standard video recording system. Processing these
video images with picture processing techniques enables us to obtain quantitative
streamwise velocity measurements, i.e. the velocity of the hydrogen bubbles.
Additionally measurements of the instantaneous streamwise and normal velocity
components were taken with a laser doppler anemometer (LDA). The LDA
measurement volume was located at 0.7 0.1 mm below the hydrogen bubble
generating wire and 1.0 0.2 mm upstream of it. As the LDA, the wire and the
video camera were mounted on the same traversing mechanism their relative
positions remained fixed during the experiments.

Four surfaces were studied during this investigation: one smooth (UU), two
longitudinal grooved (SA and SS) and one transverse rough (GG) surface. The
dimensions and the shape of the surfaces are shown in fig 1. The measurements

~!MA ~ll
~j ~
. ,.
~~G ---:-::-15.7=-----
ss 5.0

FIGURE 1. Geometry and nomenclature of grooves. SA & SS: longitudinal grooves;


GG: spanwise grooves.
were carried out 4.2 m downstream of the leading edge of the measurement plate
at two free stream speeds of 95 mmls and 140 mm/s. The u* values for the
smooth plate were 4.7 and 6.7 mmls respectively. In the case of the grooved wall
experiment aIm section of the smooth plate is removed and replaced by the test
surface. In these cases the measurement position was located at a distance of O. 7m
downstream of the surface roughness change.

The visualization experiments were made with the wire at right angles to the free
stream and parallel to the wall. The video images are obtained by photographing
a single hydrogen bubble line a few hundred milliseconds after its release by the
bubble generating wire. The position of the line is measured automatically by
computer processing and provides a measurement of the streamwise velocity
component as a function of the spanwise position. Comparison with the LDA data
showed a, good correlation between the flow velocity calculated from the position
of the bubble line and the flow velocity measured with the LDA (Pulles 1989). No
correlation exists between the local turbulent intensity or the local shear stress as
measured with the LDA and the velocity calculated from the position of the
bubble line.

The method is especially suited for measuring the spanwise correlation of the
streamwise velocity component and provides an estimate of the average spanwise
extent of flow structures. Since four released bubble lines are visible in the video
pictures one can observe the structure of the flow. The velocity measurements are
100

made on the bubble line closest to the wire (or the line that was released last in
the sequence of four). The signal of the LDA during this period gives additional
information, in particular on the vertical transport and the instantaneous
Reynolds shear stress which are not readily visible in the video frames. With this
setup two sets of data were obtained. One set consists of 500 velocity profile
measurements, recorded at four different heights above the four surfaces and two
main flow speeds. The total set thus comprises of 16000 velocity profiles.
Simultaneously with the profile measurements the two velocity components as
registered by the LDA were recorded. A second set consists of six times 500
velocity profile measurements, with the LDA measurements and six times 500
fragments of video film showing the movement of the bubble lines, during the
measurements. These were obtained on three different surfaces (the smooth, one
longitudinal grooved (SS) and the spanwise grooved surface) at two different
heights.

3. Results.

First the results derived from the first data set are presented and discussed. An
obvious quantity that can be calculated from the velocity profiles is the average
50~-------- __________________ ~

40

30

20

10

O~~-r~~4--~~6-.~~8--r-1~0-'~12
if
SO,---------------------------~b

40

30

20

10

OO~--c-~-4r-~-6r-~~8--~~10--~12
u+
FIGURE 2. Correlation lengths obtained by jitting an exponential curve to the
observerved average correlation function. a:Urn = 95 mm/s,. b: Urn = 140 mm/s. 0 :
smooth plate, UU,. + : longitudinal grooved plate, SA,. 0 : longitudinal grooved
plate, 55,. t, : spanwise grooved plate, GG.
101

spanwise velocity correlation. From this correlation function we derive a


correlation length. This can be done in several ways. The length presented in
figure 2 was obtained by fitting the function C(x) = exp(-x/L) with the
correlation function averaged over the 500 profiles at each height for the plates
studied. The L shown in the figure was obtained by a least-squares fit. Since the
definition of the zero reference plane for different types of rough surfaces is quite
arbitrary, we have chosen to plot the length scale against average velocities
obtained from the same data set at the concerned height. Thus we interpret an
increase in average velocity as an increase in the distance from the wall.

The results suggest a decrease in length scale near the wall . This is in agreement
with the measurement of Gallagher & Thomas (1984). However an increase in
length scale is observed in the region of the higher velocities, presumably above
the buffer layer. This is to be expected since the lengthscale near the wall must be

50

-20 -10 o I. 10 20 30 mm

50

~~~
Jt;-~ss 0 J\~A A~~
___
v '\r'-JV :v SA
'---::-2~0---~,0'---O~z-.~'0-+~2L-O----L30:-m-'m -20 -,0 o z.'O 20 30mm

FIGURE 3.Average velpcity versus distance to the nearest low speed streak. U00 =
95 mm/s, Ylda = 1.0 mm ~ 5 y*. SS andSA: longitudinal grooves; UU: smooth,'
GG: transverse grooves.
strongly determined by the distance between the grooves (12 and 24 viscous units
at 95 mmls main flow speed, 17 and 34 at 140 mmls main flow speed), which is
smaller than the distance between the low speed streaks that determine the
lengthscale higher up in the boundary layer (100 viscous units).

Figure 3 shows the average streamwise and vertical velocity components versus
102

the distance to the nearest low speed streak. These can be extracted from
thepresent measurements as follows. Firstly the position of the most prominent
low speed streak in the velocity profile as measured by the hydrogen bubble line is
determined. Then the velocity components measured with the LDA at its fixed
position are filed against the distance from the low speed streak identified earlier.
This operation was carried out for the 500 profiles over the four surfaces for one
vertical position of the bubble-wire LDA assembly. The figure was obtained by
averaging over the 500 measurements. The figure shows that above the smooth
wall there exists an average upwards velocity component at the position of a low
speed streak, as is to be expected and as is also found by other authors (Kline et
al 1967). This feature of positive vertical velocity is absent above the longitudinal
grooved wall SA but does persist over the plate SS. However this is in consistent
with the observation that at the speed at which these measurements were done
plate SS shows very little drag reduction (Pulles et a1.1989)

With the data obtained in the second set of measurements we attempt to


correlate the periods of high Reynolds stress with characteristic flow patterns.
The Reynolds stress T is calculated with:

T
T = fJ (U - U) . (V - V) dt (A)
o

Here U and V are the average velocity components averaged over the 500
sequences measured. The averaging time T is 240 ms, which is the period during
which video recordings of the flow were made. In figure 4 the 500 values of T
calculated from the expression (A) are ranked according to the value of their
10~-------------------------- ____________~

"j

..
s. -.: ... ,"",-
"' '. c
b . . . . ----..
.. ~ 20 sequences
min. stress \

max.. stress

-5
'.

-10L-------~------~------~------~-------"
o 100 200 J~O 400 500
N

FIGURE 4.Distribution of measured Reynolds stress over the 500 sequences.


Vertical axis arbitrarily scaled. a: smooth plate UU; b: longitudinal grooved plate
55; c: spanwise grooved plate GG.
103

Reynolds stress. The twenty sequences of the highest and the twenty sequences
with the lowest (most negative) Reynolds stress were selected for a more detailed
study. Figure 4 shows that these account for nearly 30% of the Reynolds stress on
the positive as well as on the negative side. Thus this set of sequences is highly
relevant, despite its small sample size. For comparison twenty random selected
sequences were also analyzed.

On basis of close examination of the total set of sequences we decided that we


could separate most of the observed flow patterns into 5 categories (I to V).
Considering this categorization we divided the set of sequences in twelve parts (1
to 12), mainly on basis of the position of the LDA measurement volume (in which
the shear stress is measured) in the flow patterns.

The five categories and the twelve parts are:


I The longitudinal vortex
1. The vortex is recognizable as loops in the hydrogen bubble lines. A
typical example is shown in the photograph in Fig.5a.
2. Longitudinal vortex, clearly associated with a low speed streak.
II The low speed streak
3. The end of a low speed streak. This was recognized when all the
four bubble lines showed a minimum and the minimum was clearly
disappearing. While this might just mean that the streak got lifted
and disappeared from the field of vision there was no way of
identifying this with the present technique. Figure 5b is a typical
example of this category.
4. Centre of a low speed streak. If all four hydrogen bubble lines show
a progressive V-shaped minimum the structure is identified as a
low speed streak (Fig. 5c).
5. Side of a low speed streak.
III The narrow high speed region
6. Side of a high speed region. The distinction between 5 and 6 is
difficult. The decision is made on the basis of whether the local
velocity is higher than the the average between the maximum
velocity in the high speed region and the minimum velocity in the
low speed streak. No distinction is made among the sides of the
different types of high speed regions.
7. Centre of the narrow high speed region. This is a sort of high speed
streak, but mostly flat topped in contrast with the V-shaped
minimum observed in a low speed streak. This is frequently
observed when two low speed streaks approach one another in their
meandering.
IV The Wide high speed region
8. Wide high speed region is characterized as a region where the
velocity is higher than the average and is more or less constant in
the z-direction. Fig. 5d is a typical example.
9. End of a high speed region. This is deemed to occur whenever a new
minimum appears in a wide high speed region or when the region
evolves into a less pronounced feature.
V Accelerating flow
10. If the last released bubble line moves faster or overtakes the
104

previous bubble lines. Fig. 5e is a typical example.

c f
FIGURE 5.Examples of structures. In each figure the flow direction is from bottom-
to top. a: longitudinal vortex; b: end of a low speed streak; c: a low speed streak; d:
a wide high speed region; e: an accelerating region; f: a pattern outside
classification.
11. Miscellaneous patterns. Includes patterns like vertical vortices,
105

peaked high speed streaks, or violent ejections from the wall.


12. Outside classification. Patterns that occurred once, but defy efforts
at classification. Fig. 5f is one such example.

fe~---------------r-----'----~---r------,
Maximum
I
'1+ =8 I
I
II IV IV
I
I
I
I
I
I
I

f~;:~~::::~::~~~~~::::~~~::::~
random
'1+ =8 I
I
II III IV IV
I
I
I
I
I
I
I

f~~~::~~~::~~~~~::::~~~~~~
mlnlrn..m
'1+ = 8
I
I II III IV
I
I
I
I
I
I
I

2 :I 4 e e 7 e It 10 1f 12

Category rn.mber

spanwise ,,"ooves o smooth


~ longi tlddlnal grooves

FIGURE 6.Distribution of structures. a: y+ =8


106

In figure 6 we present the frequency distribution of the events as occurring in the


selected sets of video sequences. These sets were taken at two different heights (y+

1e~---------------r----~------r--r------'
Maxirrun I
y" = 17 I
I I
II III IV IV
I
I

random
y" = 17
I I
I II III IV IV
I I
I I
I

minlmun I
y" = 17 I
I
, 10 :
II III IV IV
I I
I
I
I
I

2 3 4 e II 7 g 10 11 12

Category runber

spanwise grooves D smooth

~ longituQinal grooves

FIGURE 6.Distribution of structures. b: y+ = 17.


107

= 8 and y+ = 17), three different surfaces (smooth, longitudinal grooved and


transverse grooved) and using the three selection criteria already mentioned.
The figure shows that at y+ = 8 the events with the highest shear stress are
mainly concentrated in category 7: the centre of the narrow high speed region.
This fact is very pronounced for the case of the transverse grooved wall. The
counts for the two other types of wall are not so clear, but even so the majority of
the high shear stress events are associated with structure III: the narrow high
speed region. Figure 6 also shows that at y+ = 8 the low speed streak can be
associated with the events of negative shear stress on the smooth and transverse
grooved plates. In the case of the longitudinal grooved plate these events are
found to occur in the high speed region.

Higher in the boundary layer (y+ = 17) the classification becomes more difficult,
as is evident by the increasing numbers of events assigned to categories 11 and 12.
Above the smooth plate the high speed region is a much observed phenomena at
this height: it occurs in all three selections. So it has apparently no net
contribution to the shear stress. Above the transverse grooved wall the events
with negative shear stress are still the low speed streaks. The longitudinal grooved
wall shows no clear trend, but the events of high and low shear stress seem to be
more evenly distributed over the low speed streak and the high speed regions.

The authors made some further observations looking at the video sequences,
which were difficult to quantify but nevertheless interesting enough to mention
here.
i) The flow over the transverse grooved wall the low speed streaks are visually
dominating. It is tempting to relate this to the high shear stress in this
boundary layer.
ii) The low speed streaks appear to be shorter in length above the longitudinal
plate compared to the streaks above the smooth wall, indicating some rather
subtle interaction between the grooves and the flow.
iii) As the low speed streaks were clearly seen to meander across the crests of
the grooves any straight forward resonance effect is ruled out.

4. Discussion.

Visual observation of the flow revealed no new structures above the two grooved
walls in addition to the ones already found above the smooth wall. An important
question that is not answered during the course of the present work is the
statistical reliability of the results in figures 2 and 6. It is not at all clear that
merely increasing the sample size would provide the requisite information. However
if we accept the differences in counts in figure 6 as indicative of the possible
changes one can expect, i.t follows that the grooves affect both the vertical position
and/or size of the structures and their contribution to the turbulent shear stress.
The absence of vertical flow in the low speed streak above the longitudinal grooved
plate (fig. 3) also points to a different shear stress distribution in a low speed
streak above the grooved wall compared to the same feature above the smooth
wall. This mix of effects, together with the absence of great qualitative changes in
the inner boundary layer could explain the difficulties in reconciling the mean
velocity measurements with the quadrant analysis. In this context it is instructive
to quote Choi: liAs regards possible mechanisms of turbulent drag reduction with
riblets, there may be more than one involved" (1989). Our results show that it is
108

worthwhile to look at the shear stress distribution in the different flow structures
in greater detail.

Acknowledgements.

The research reported in this paper was partially supported by the Netherlands
Foundation for Technical Sciences (STW) as part of the programme of the
Foundation for Fundamental Research on Matter (FOM). Two of the authors
(CJAP & KKP) are grateful for the assistance of Ing. G. Trines, MI. J.C.
Stouthart, II. C. Nieuwvelt and Drs. A.M. Koppius during the course of the
investigation at Eindhoven.

References.

Choi,K.-S. 1989. Near-wall structure of a turbulent boundary layer with riblets.


J.Fluid Mech. vol.208, pp417-458
J.A.Gallagher and A.S.W.Thomas. 1984. Turbulent boundary layer characteristics
over streamwise grooves. AIAA paper no.84-2185.
S.J.Kline , W.C. Reynolds, F.A. Schraub and P.W. RundstadleI. 1967. The
structure of turbulent boundary layers. J.Fluid. Mech. 30 p 741.
C.J.A.Pulles. 1988. Drag reduction of turbulent boundary layers by means of
grooved surfaces. Ph.D. Dissertation,Eindhoven University of Technology.
C.J.A.Pulles, K.Krishna Prasad. and F.T.M.Nieuwstadt.Turbulence Measurements
over longitudinal microgrooved surfaces. Accepted for publication in Applied
Scientific Research
M.J.Walsh. 1982. Riblets as a viscous drag reduction technique. American Institute
of Aeronautics and Astronautics, Aerospace Sciences Meeting, 20 th, Orlando, FI,
Jan 11-14AIAA paper 82-0169.
Effects of Longitudinal Pressure Gradients on Turbulent Drag Reduction
with Riblets

KWING-SO CHOI
British Maritime Technology
1 Walde grave Road
Teddington
Middlesex TWll 8LZ, UK.

Summary

A study of turbulent boundary layers was carried out using the hot-\vire
and film anemometry over smooth and riblet surfaces under different
pressure gradient conditions. Detailed measurements into the near-wall
turbulence structure indicated that the changes in time averaged
turbulence quantities, produced by the addition of riblets, were not
significantly altered by the presence of a longitudinal pressure
gradient. This is in line with the previously reported results on the
near-wall structure obtained by a conditional sampling technique.

1. Introduction

Investigation of turbulent drag reduction using riblets has started at


NASA Langley Research Center (Wilkinson et al., 1987) in the late
1970's, and many investigations have shown (Choi, 1984; Bandyopadhyay,
1986; Savill et al., 1988) that up to 8% of turbulent dra~ reduction
can be obtained from the passive device. Although most of the
experiments so far were conducted in laboratories using flat plates
under zero pressure gradient condition, efforts were recently made to
confirm the results in the field (McLean et al., 1987; Choi, 1990) as
well as using three dimensional objects with double curvatures (Choi et
al., 1987; Nieuwstadt et al., 1989). Compressibility effects of riblets
were also investigated in a high-speed wind tunnel (Squire & Savill,
1987) .

However, a study of longitudinal pressure gradients on the performance


of riblets has not yet been conducted comprehensively, although it is
one of the critical aspects in design and application of riblets to
aircraft and underwater vehicles. It is in this context that we have
carried out an investigation of the turbulent drag reduction using
riblets. An emphasis of the present investigation was placed on the
structural differences of the near-wall turbulence under different
pressure gradient conditions, rather than direct measurements of drag.

109
110

In this paper, only the time-averaged quantities such as mean velocity


and turbulence intensity profiles and associated turbulence statistics
are reported. Conditional sampled and ensemble averaged data on the
near-wall turbulence structure are already given in Choi & Johnson
(1989). Detailed account of the mechanisms of turbulent drag reduction
by riblets is found in the recent paper by Choi (1989).

2. Experimental Set-Up

The boundary layer wind-tunnel (Sawyer & Winter, 1987) at Royal


Aerospace Establishment (RAE) at Bedford was used throughout the
present investigation. This tunnel was an open-return blower type with
a speed range of up to SOmis. The test section was S.4m long and 1.2m
wide, with its test surface at the roof made of ground plates of
aluminium alloy. The lower surface of the test section was constructed
with a flexible sheet of fibre glass whose height could be adjusted to
give a range of longitudinal pressure gradient conditions. The present
experiment was carried out under three different conditions - adverse,
zero and favourable pressure gradients. The boundary layer parameters
corresponding to these three conditions are shown in Table 1. The
spanwise variation of the momentum thickness over the test plate was
3% within one half plate width (200rnrn).

TABLE 1

Flow parameters of the present experiment.

Favourable Zero Adverse

Uw (rnls) 3.03 3.00 2.97

u * (rnls) 0.136 0.124 0.093

H l. 36 l. 38 l. 65

(rnrn) 7.5 13 23

RB 1600 2700 4700

v illi. 0.00208 0 0 0196


0: - ----;3 dx
pu

f3 -
Lilli. 0.16 0 3.1
TW dx

dU w
v 0.20 x 10. 6 0.57 x 10. 6
K - Vco 2 ~ 0
III

Figure 1 Cross-section of riblet.

The experiments were carried out using two interchangeable test


surfaces - a riblet plate and smooth plate, 400mm long and 400mm wide,
which were located 4.70m from the beginning of the test section. Cross-
sectional shape and dimensions of the riblets are given in Figure 1,
which were similar to the ones used by Choi (1989). The free-stream
velocity of the tunnel was set to 3m/s at the test plates. In the zero
pressure gradient condition, the non-dimensional height of the riblets
was approximately 12 wall units. The transition point of the boundary
layer was fixed by placing a 6mm square rod at the beginning of the
test section.

Velocity measurements were carried out using hot-wire anemometry with


DISA SSm bridge units. Most of the data were obtained using DISA SSP14
single wire probe with an overheat ratio of 1.8. The skin-friction
fluctuations were also measured during the present investigation using
TSI 1471 hot-film sensors. They were mounted flush with the surrounding
surface for the smooth plate, and in the valleys of microgrooves for
the riblet plate. Overheat ratio of these surface sensors were kept low
at 1.2 in order to minimise the effect of free convection from the
sensors. The virtual origin of the velocity profiles over the riblet
surface was one fifth of the riblet height (O.3mm) below the peak.

Data acquisition was conducted digitally using Zenith Z-248 PC at a


sampling rate of 800Hz with low-pass filters set at a cut-off frequency
of 400Hz. All the data were recorded onto magnetic cassette tapes which
were later replayed and analysed on the same computer.

3. Results and Discussions

The mean velocity profiles over the smooth and riblet surface are shown
in Figure 2 in log-law format for three different pressure gradient
conditions (see Table 1). The top two profiles correspond to the
112

30,----~--~--~--
dpJdx < 0

3D
dp/dx:: 0

u
~" 30
dpldx. > 0

1
J
15

Present study !
Smooth socta"
" Albie! surface
i preVi.OUS dato (ChoI, 1989)-
\\\Smooth surface
IIIAlbie! surfan>
o --~---~~-----
o , 2 3 4
Log y.

Figure 2 Mean velocity profiles in semi-logarithmic form of U/u * vs.


yu* Iv. The sequence of figures from top to bottom
correspond to the favourable, zero and adverse pressure
gradient cases.

boundary layers in a favourable pressure gradient (dp/dx < 0), the


middle ones in zero pressure gradient (dp/dx ~ 0) and the bottom ones
in an adverse pressure gradient (dp/dx > 0). Over the smooth surface,
the mean velocity profiles were fitted to the log-law obtained by Patel
(1965):

+ +
u 5.5 log Y + 5.45 (1)

from which the friction velocity u * was derived. The velocity profiles
over the riblet surface were, however, fitted to the following log-law
obtained by Choi (1989) for the experiment using similar riblets:

u+ ~ 5.5 log Y+ + 6.89 (2 )

where u + and y + are defined by

u
+ U/u * (3)

and y+ y u * Iv (4)
113

It is clear from this figure that the general fit to the log-law is
good for both smooth and riblet cases in all the three pressure
gradient cases tested. It is also noticed that there is an upward
(positive) shift in the intercept of the log-law for the riblet cases,
indicating a drag reduction due to thickening of the viscous sublayer.
Similar shifts are also reported by Hooshmand et al. (1983) over a
riblet surface and with Large Eddy Breakup (LEBU) devices (see
Bandyopadhyay, 1986). This is also a common phenomenon for the
turbulent boundary layer with drag-reducing polymers (see Lumley,
1973). The extent of logarithmic region is, however, smaller in the
adverse pressure gradient case, with a larger extent of wake region.
Strictly speaking, the log-law is only valid in zero pressure gradient
conditions. As Mellor (1966) and McDonald (1969) pointed out, however,
the effects of longitudinal pressure gradients on the log-law are small
as long as the pressure gradient parameters a defined by

*3
a ~ (v/p.u ) . dp/dx (5 )

is small enough. Our value for this parameter in the adverse pressure
gradient case is 0.0196 (see Table 1), which is small enough not to
affect the constants of the log-law. The value for a is even smaller in
the favourable pressure gradient case. A small deviation from the log-
law is, however, expected at large y+ even with a small value of a

1.00 -------,--------,

1.00 dp/dx < 0

100
dp/dx =0
11I1f/1I 11 "/l1f
/IIIIII~", '\ ,,,,, \ \\ \\ '\ \
.!.! I,ll, \ '\ '\ \ \ '\ \
U'" 0.75 II,..!.!'\\\\

"
""
050 dp/dx>O

Present study
Smooth surface
025 )( Riblet surface
Previous data (ChOI, 1989)
\\\ Smooth surface
/1/ Riblet surface
0 I
0 25 50 75

Figure 3 Mean velocity profiles in linear form. Figure sequence and


symbols as for Figure 2.
114

0.15,-----....,----r------,

0.15

~-;':::::::i!::=;~~ dp/dx < 0

~t?:;;s;;:;:?===7==;:~==~ dp/dx > 0

Pre-sent study
0.05 Smooth surface-
)( Rjbl~t surface-
Previous dato(Choi.1989):
\\\Smooth surface
II/Riblet surface
OL-______l -_ _ _ _ _ _ ~ ______ ~

o 25 50 75
y'

Figure 4 Profiles of the u-component turbulence intensity. Figure


sequence and symbols as for Figure 2.

(Mellor, 1966). The comparison with the previous results in the ze:o
pressure gradient condition by Choi (1989) is very good as shown ~n
Figure 2, although the extent of logarithmic region is slightly shorter
in the present case owing to its smaller Reynolds number of Re ~ 2.7 x
10 3 compared with Re ~ 4.6 x 10 3 in Choi (1989).

The corresponding linear profiles of the mean velocity are given in


Figure 3, together with the data by Choi (1989) in the zero pressure
gradient condition. The general shape of the present profiles as well
as the differences in profiles between the smooth and rib let cases are
in good agreement with those by Choi (1989). The difference in the
"fullness" of the linear profiles is again due to the effect of the
Reynolds number, as mentioned above, through the change in the skin
friction coefficient.

The turbulence intensity profiles for the u-component velocity are


given in Figure 4. In this figure, a reduction of between 5 to 13% in
turbulence intensity is observed over the riblet surface at the maximum
point of intensity of y+~14. It is also observed that the reduction is
slightly greater in the favourable pressure gradient case compared with
the adverse pressure condition case. At zero pressure gradient, the
present result of 8% reduction compares reasonably well with the
liS

previous results by Choi (1989), who obtained 10% reduction at y +=18.


The mean velocity and turbulence intensity profiles in the adverse
pressure gradient case were obtained using a cross-type hot wire, which
has limited some measurements close to the wall surface. The rest of
the turbulence statistics were obtained, however, using a single-wire
sensor.

Figures 5 and 6 show the skewness andkurtosis of the u-component


+
velocity very close to the wall surface (y <16). The corresponding
values of the wall-skin friction signal are also shown at y+=O in these
figures. In general, both the skewness and kurtosis are greater in the
adverse pressure gradient case compared with zero and favourable
pressure gradient cases at the edge of the viscous sublayer (y+=ll and
16), but they become similar to the others in the viscous sublayer
+
(y =0 and 7). The comparison with the data of Kreplin & Eckelmann
(1979) and Ueda & Hinze (1975) in zero pressure gradient condition is
good, with the skewness becoming zero and the kurtosis reaching minimum
at y+~ll. Approximately at this distance from the wall surface, the
viscous and turbulent stresses become equal (Kreplin & Eckelmann,
1979). However, the skewness and kurtosis of the wall-skin friction
signal, which is linear to the u-component velocity at small y+ within
"linear" velocity profile region, does not seem to be extrapolated very
well from the data of Kreplin & Eckelmann (1979) or Ueda and Hinze
(1975). This may be due to the lower frequency response of the hot-film

1. 0 ,---r---,-----,,-----,
Present study
Smooth surface
)( Riblet surface
0.5)(--- __ x dp/dx< 0
L----o~x
1) A , ~o~
~
Xl-~'A ..............
~I
PrevIous dotQ(d. p/dx = 0)
III 0 6.)(~ .0. Kreplin & Eclo:elmann
~ 0.5 0 Ueda & Hmze
.---- ...... a
~
lfl I ~X dp/dx=O
1.0 x I o~x
I"~ A 0

0.5 ~ ' .... , ..... x


dp/dx > 0
.-----.~x----

oI I 10--;'
o 5 10 15 20

Figure 5 Skewness of the u-component velocity and the wall-skin


friction in the favourable, zero and adverse pressure
gradient cases.
116

5 0 l---..----.- Pr~st'nt study


i

Smooth surfac~
1
4 0~ )( A,ble-t surface-

I" . . , "- dp/dx < 0

30r-, "
I ' _ ..... l(",---

i
\ -e ____! = : - _

5 0' ~,---',-~---'---

~ 4.0 L',
o \ 6 Kreplin & Eckeolmann
DUedo & Hinze

.s ~ \ "- dpldx =0
~ 30
I -""0',D~ 1
0
",

~o ox~

;ol\~ ~-1
40
1\ j
J
dp/dx> 0
I "
o "-

3 0 r'~-~'::~-j-~
20~-~ I
o 10 15 20
y'

Figure 6 Kurtosis of the u-component velocity and the wall-skin


friction in the favourable, zero and adverse pressure
gradient cases.

sensors to measure the wall-skin friction fluctuation because of the


heat conduction to the substrate. Positioning of hot-film sensors in
the valleys of riblets may also have some effects on the turbulence
statistics over the riblet surface. These should not, however,
seriously affect the changes in these values due to pressure gradients.

The effects of riblets on the turbulence statistics in the adverse


pressure gradient case, the difference between riblet and smooth data
in Figures 5 and 6, become smaller with y+ faster than the other two
pressure gradient cases. This suggests that the effects of riblets on
the higher order statistics of turbulence in the adverse pressure
gradient case does not extend further away from the wall surface
compared with the zero or favourable pressure gradient cases.

Figures 7, 8 and 9 show the probability densities of the u-component


velocity near the wall surface of the boundary layer as well as that of
wall-skin friction (y+~O). The general trends of the probability
117

Deviation

Figure 7 Figure 9

00 ------------.---~-

dp/dx = 0
Smooth 5urfac~
y' 0 16 RlbI~t 5Urfac~
Probability densities of the
u-component velocity and the
I) 6 wall-skin friction in the
favourable (Figure 7) , zero
(Figure 8) and adverse (Figure 9)
pressure gradient cases. A circle
symbol denotes smooth surface, a
cross denotes riblet surface.
.is
Dashed line represents Gaussian
E
o probability density.
a:

oJ r' u,_~,,.. <-


Wall-skin
friction

'1 l""_",,,.. 20 40
Deviation

Figure 8
lIS

density are such that the skewness changes its sign from positive to
negative with an increase in the distance from the wall surface. This
is also shown in the skewness profiles (see Figure 5). The positive-
tail probability is largest at y+~7 (closest position to the wall
surface), indicating that large positive excursions of the u-component
velocity signal are the dominant feature in the near-wall region of the
turbulent boundary layer. This is in good agreement with the results of
the quadrant analysis by Wallace et al. (1972) and Willmarth and Lu
(1972). They both showed that the "sweeps", events associated with a
positive u-signal and a negative v-signal, are more important closer to
the wall surface than the "ejections" (u<O and v>O). It is believed
that "near-wall bursts", quasi-periodic events associated with a sharp
increase in wall-skin friction (Choi, 1989), are responsible for this
large positive-tail probability since they are caused by the downwash
of high-momentum fluid between pairs of counter-rotating longitudinal
vortices close to the wall surface. The comparison of the present
results with the probability densities by Choi (1989) for the wall-skin
friction signal in zero pressure gradient condition is excellent for
both smooth and riblet cases. As far as the effects of riblets on the
probability densities in the adverse pressure gradient case are
concerned, they are only noticeable within a near-wall region of the
boundary layer (say, y +<11) although the effects are noticeable up to
+
y ~16 for the other two pressure gradient cases. This is in line with
the results of the skewness and kurtosis (see Figures 5 and 6) as
discussed above.

The energy spectra of the u-component velocity at y+~7, 11 and 16 and


of wall-skin friction (y+~O) are shown in Figures 10, 11 and 12. The
results obtained by Choi (1989) in zero pressure gradient condition are
also given in Figure 11, showing an excellent agreement with the
present spectra of wall-skin friction over both smooth and riblet
surfaces. It is observed from these figures that there is a reduction
in a low to middle part of the turbulence energy spectra over the
riblet surface. The reduction is particularly significant in the
spectra of wall-skin friction. Apart from the differences in general
shape, there is no noticeable changes due to pressure gradients in the
effects of riblets on the energy spectra.

4. Concluding Remarks

A detailed study of the near-wall turbulence structure over smooth and


riblet surfaces was carried out to investigate the effects of
longitudinal pressure gradients. The results on the u-component
turbulence intensity indicated a reduction of between 5 to 13% by
riblets, which was slightly greater in the favourable pressure gradient
case compared with the adverse pressure gradient case. The reduction of
the turbulence intensity, the second order moment of turbulence
statistics, was only observed within the near-wall region of the
+
boundary layer, say y <75. The higher-order moments, e.g. the skewness
(the third moment) and kurtosis (fourth moment), gave a similar result
119

100 ~ y' ~ 11

! y' ~ 7
~ 10 ~ __

I10~'fi
if}
Wall- skin
fnctlOn

~4
10~4
,0

~6L
I
~6

10 10 - - - Smooth sur/eel'"
- Riblet surface

1(),L __ ~ ____ -----L ______ ~~._J


10- 1 100 10 1 10 2 10]
10' 10'
Frequency (H z) FreqLJ'ncy (H z!

Figure 10 Figure 12

dpld, ~ a
I Spectral densities of the

l
u-component velocity and the
wall-skin friction in the

~
favourable (Figure 10) , zero
(Figure 11) and adverse (Figure
I 12) pressure gradient cases.
f Solid line indicates riblet data.
l Dashed line indicates smooth
I surface data. The sequence of
~
[
figures from
correspond
top to
respectively
bottom
to
+
y ~ 16, 11 and 7 with the bottom
figure being the wall-skin
10.'~! I

l
-- -
--
Smooth surface
Rlblet sur/acE' friction, y+ ~ O.
Drev,ous dato (ChOl.1989)
I \\\ Smooth surfaCE'
10~8L. __ ~ _/~~~~ SUr_f_Q~ __ _ _ ---1-___ _
~
~I
10 10' 10'
~'requency (Hz)

Figure 11
120

to the turbulence intensity in that there were no significant


differences in the degree of changes due to riblets for different
pressure gradient cases. It was, however, found that the effects of
riblets on these higher turbulence statistics were confined within much
closer to the wall surface, say y+<lS, in the adverse pressure gradient
case. The difference in mean velocity profiles (the first moment) due
to riblets extended further out to the edge of the boundary layer.

These results seem to indicate that the effects of riblets on the near-
wall turbulence structure would not, in general, be altered by the
longitudinal pressure gradients. This conclusion is also supported by
the previously reported results (Choi & Johnson, 1989) using a
conditional sampling technique that the wall-shear stress signatures
during "near-wall bursts" are essentially unaffected by the pressure
gradients. This does not necessarily mean, however, that the net drag
reduction by riblets is unaffected. This is because the performance of
riblets is determined by the balance between the turbulent skin
friction reduction and viscous skin-friction increase. What the present
results seem to suggest is that the turbulent skin friction reduction
due to riblets would not be altered by the pressure gradients.

Acknowledgements

The author would like to express his gratitude to Dr. R. Johnson for
his help in the execution of the experiment and data analysis. He would
also like to extend his acknowledgements to Prof. M. Gaster of
Cambridge University, Mr. M. Firmin and Mr. W. Sawyer of RAE, and
Dr. J.A.B. Wills and Dr. M.E. Davies of BMT for their encouragement and
valuable discussions. This work has been supported by British Aerospace
plc and the Department of Trade and Industry, U.K.

References

Bandyopadhyay, P.R. 1986. Review Mean flow in turbulent boundary


layer disturbed to alter skin friction. J. Fluid Engr., 108, 127.

Choi, K.-S. 1984. A survey of the turbulent drag reduction using


passive devices. NMI Report R-193. NMI Ltd., Teddington, Middlesex.

Choi, K.-S., Pearcey, H.H. and Savill, A.M. 1987. Test of drag reducing
riblets on a one-third scale racing yacht. Proc. International
Conference on Turbulent Drag Reduction by Passive Means, London.

Choi, K.-S. 1989. Near-wall structure of turbulent boundary layer with


riblets. J. Fluid Mech. 208, 417.

Choi, K.-S. and Johnson R. 1989. Pressure gradient effect on the


turbulence structure over riblet surface. In Advances in Turbulence II,
Eds. by Fernholz, H.-H. and Fiedler, H.E., Springer-Verlag.

Choi, K.-S. 1990. Drag-reduction test of riblets using ARE's high speed
buoyancy propelled vehicle - MOBY-D. Aeron J., March, 79.
121

Hooshmand, D. et al. 1983. An experimental study of changes in the


structure of a turbulent boundary layer due to surface geometry
changes. AlAA Paper 83-0230.

Kreplin, H.-P. and Eckelmann, H. 1979. Behaviour of the three


fluctuating velocity components on the wall region of a turbulent
channel flow. Physics of Fluids, 21, 1233.

Lumley, J.L. 1973. Drag reduction in turbulent flow by polymer


additives. J. Polymer Sci. : Macromolecular Reviews, 2, 263.

McDonald, H. 1969. The effect of pressure gradient on the law of the


wall in turbulent flow. J. Fluid Mech., }2, 311.

McLean, J.D. et al. 1987. Flight-test of turbulent skin-friction


reduction by riblets. Proc. Int. Conf. Turbulent Drag Reduction by
Passive Means, London.

Mellor, L.G. 1966. The effects of pressure gradients on turbulent flow


near a smooth wall. J. Fluid Mech., 24, 255.

Nieuwstadt, F.T.M. et al. 1989. Some experiments on riblet surfaces in


a towing tank. In Drag reduction in fluid flows Techniques for
friction control, eds. by Sellin, R.H.J. and Moses, R.T., Ellis
Horwood.

Patel, V.C. 1965. Calibration of the Preston tube and limitations of


its use in pressure gradients. J. Fluid Mech., 1, 185.

Savill, A.M. et al. 1988. Turbulent drag reduction by passive means : a


review and report on the first European drag reduction meeting.
J. Theoretical and Applied Mechanics, 2, 353.

Sawyer, W.G. and Winter, K.G. 1987. An investigation of the effect on


turbulent skin friction of surfaces with streamwise grooves. Proc. Int.
Conf. on Turbulent Drag Reduction by Passive Means, London.

Squire, L.C. and Savill, A.M. 1987. Some experiences of riblets at


transonic speeds. Proc. Int. Conf. Turbulent Drag Reduction by Passive
Means, London.

Ueda, H. and Hinze, J.O. 1975. Fine structure turbulence in the wall
region of a turbulent boundary layer, J. Fluid Mech., 67, 125.

Wallace, J.M., Eckelmann, H. and Brodkey, R.S. 1972. The wall region in
turbulent shear flow. J. Fluid Mech., 54, 39.

Wilkinson, S.P. et al. 1987. Turbulent drag reduction research at NASA


Langley progress and plans. Proc. Int. Conf. on Turbulent Drag
Reduction by Passive Means, London.

Willmarth, W.W. and Lu, S.S. 1972. Structure of the Reynolds stress
near the wall. J. Fluid Mech., 55, 65.
Synthesis of ExperiInental Riblet Studies in
Transonic Conditions

E. COUSTOLS 1 & V. SCHMITT 2


1 OIVEflA-CERT, Aerothe7'11todynomics Department, Tou1011se, France:
2 O;VERA, Aerodynamics DivisioH, Chatillon, France.

SUl\1MARY - The present paper summarizes the status of the experimental research car-
ried ant both at O~ERA/CERT and ONERA/1Iodane, as regards internal manipulators.
commonly named, by a lot of researchers, riblets. That turbulent boundary layer ma-
nipulation program was begun at CERT in mid-1986. Emphasized will be drag reduction
performances of such passive devices, tested at transonic conditions, on a cylindrical body,
a CAST I aerofoil for two-dimensional turbulent boundary layers, and, at last. a complete
l/11th scale Airbus A320 model.

1 Introduction
The structure of a turbulent boundary layer has been studied by numerous investi-
gators and its description in terms of coherent structures is subject to much debate.
Though turbulence seems to be a very complex phenomenon, a specific set of fea-
tures ill the turbulent boundary layer has been identified: streamwise ,'ortices, low
speed (walllilyer) streaks, ejections, break-up, sweeps, outer layer motions and their
interactions with the walL entrainment of irrotational flows, ... The production stage
of l urhulcnce has usually been referred to as the "bursting phenomenon", [5]. One
approach to illterpreting turbulence-control experiments is "to view them as attempts
to interfere with some component of the turbulence production cycle", [6] : thus. one
could think of modifying either the large outer eddies or the wall layer events. or.
consequently, of changing the communication bet ween the ,'arious scales.
III the last decade, a lot of effort has been devoted towards manipulation of
tl11'bulelll boullClary layers by passi\'e devices such as internal manipulators (small
slrealllwise grooYE's acting straightly on the inner region), commonly referred to as
riblets. for tile purpose of reducing viscous drag. Though the details of the mecha-
nisms are not firmly understood, "such a concept of llsing grooved surfaces to obtain
skin-friction reduction is rather close to industrial applications", [7]. In fact. the
question which arises is to address the problem of how and to what extent the riblets
alter the turbulent 11m\l structure in reducing viscous drag. Though it is not the
purpose of this paper to explore the eJIects of these passive devices on the mean and
fluctuating quantities of a turbulent boundary layer, [16], one must recognize that
some studies showed up that grooves would restrict the spanwise motion of the lon-

123
124

gitudinal \'orLices [7], [17], would not lock the low speed streaks in a fixed spanwise
location [5] but \Vould increase their spanwise spacing [4],
At NASA Langley Research Center, a complete experimental investigation of
such passi\'e drag-reducing devices was performed by Walsh [24], and \\lalsh and
colleagues [2,j], [26], [27], They have tested many different groove geometries and
found that the optimum shapes for drag reduction have a sharp peak protruding
into t he flo\\' and have a height and spanwise spacing of the order of the viscous
sublayer thickness, l\ett drag reductions up to S% have been recorded on the most
popular symmetric V-shaped riblet. 1Ioreover, performances of grom'ed surfaces were
unaffected by the presence of moderate ach'erse and/or favorable pressure gradients.
and by misalignment up to 1.5 0
The first experiments at CERT, as regards performances of ribbed surfaces. were
carried out in mid-19S6, Thus. initial tests performed in zero- and acl\'erse pressure
gradient conditions for low subsonic speeds, confirmed the earlier findings of nett drag
reduct ions. The efficiency of grooved surfaces was essentially determined through
eiilier momentum balance technique or wake surveys, [9], [12], [15]. These preceding
"fundamental" experiments have allowed to verify, with laboratory measurements,
that carefully designed internal devices could provide nett drag reductions. In order
to go closer to flight applications, the effect of small streamwise grom'es on a slender
body (1 j:38 th scale Airbus-type fuselage) was checked in the F2-0NERA/Fauga wind
tunnel. [11]. Nett skin-friction drag reductions \yere measured by means of an intern al
one-componellt drag balance.
Tile following step was to pursue the evaluation of such a passive drag reduction
at transonic speeds, in order to provide data on riblet performances applicable to
transport aircraft fuselage and/or wings, This paper summarizes the experimental
research carried out in transonic wind tunnels at O:\ERA,

2 Experilllents in the T2-wind tunnel at CERT


Two sets of experiments haye been conducted in the T2-transonic cryogenic wind
tunnel of OI\ETIA/CETIT, That \\'ind tunnel is a closed circuit pressurised (up to -
bar) facility. Operational runs can be carried out in cryogenic conditions, with self-
adaptiye walls, The flow is driven by an injector system (air) and can be cooled down
to 110 K by a liquid nitrogen injection, The aim of upper and lower adaptive walls
is to create an unconfined flow around the model, in a limited section, by controlling
bounda ry condi tions, [1], The dimensions of the test section are; 1:320mm length,
:390111m width and :170mm height.

2.1 Cylindrical body - Ogive


First of all, the efIiciency of grooved surfaces has been checked, for transonic condi-
tions, on an axisymmetric body, The turbulent boundary layer, which is going to
be manipulated by altering the wall geometry, deyelops along a cylinder-type body.
under zero-pressure gradient conditions, For transport aircraft applications. this ar-
rangement deals with a rather important part of the fuselage.
125

LcarhOrnndnIn hand

r 1'~halaJ-lce g -~-;==----===------+-)---'9) j
-,-----S-.----(_in

L : = J 5 0 L' - 1 - - - - - 550 '


Figure 1: Model geomeh'Y

2.1.1 Experimental apparatus - Drag measurements


The model is schematized on figure 1 ; it is a cylindrical body of 600mm length,
80mm diameter (<,6), with an upstream elliptical nose of 150mm length. The model
axis has bcen aligned with the external free-stream flow direction. The boundary
layer which develops along that ogive has been tripped with a carborundum band
(ayerage height close to 0.06mm) at 30111111 from the ogive nose; this location is far
upstream so that the turbulent boundary layer will be fully turbulent when it reaches
the leading edge of the rib let model. The last 550111m length of the cylindrical body
section were only covered \\'ith thin self-adhesive grooved surfaces. In order to judge
of ribs performances, the diameter of the cylinders has been reduced to 79.8mm : then,
whaten.. r ribld model is considered, the ribs trough plane is approximately mounted
Hush with the adjacent upstream smooth part, [9], [1:3]. This arrangement avoids
the facing sicp efIect. Four symmetric V-groove riblets, manufactured by the 3:\1
Company in an adhesive backed film, with an aspect ratio (s/h) of one. have been
tested: h = 0.023, 0.033, 0.051 and 0.076mm. The turbulent boundary layer length.
manipulated by such models, is then: L=550111m.
Drag measurements ha\-e been completed using an internal five-component bal-
ance. The observed accuracy on the drag coefficient is equal to 5 10- 4 for Cd=O.l.
That accuracy became two times larger when taking into account the measurement
uncertainties (variations of free-stream Mach number, stagnation pressure, ... ) but
is small ellough to define, precisely, drag reductions or increases due to grooved sur-
faces. [1:3].
The tests have been performed at ambiant stagnation temperature: the Reynolds
lllllnber H L , based on the manipulated length L, varies from :3.8 10 6 up to 1S.7 lOG.
for a free-stream 1\1ach number range, J\LX! : 0.:3 - O.Sl.5 and a stagnation pressure
range: l.1 - :3.0 bar. Lower and upper \Yalls were only adapted for expf'riments per-
formed at 1\1.::0 :2: 0.6 . For low-speed tests, it has been more difficult to reproduce
faithfully the drag measurements. However, at 1100 ~ 0.7 (resp. 0.4) the clata scatter
corresponds to ~Cd/Cd~l.3% (resp.:3.5%), which is quite acceptable. [1:3].

2.1.2 Local Mach number distributions


The local :\Iach number distributions, 1Iz, are plotted on figure 2 for [1'"C values of the
free-stream :\1ach number 1LX!=0.42, 0.55, 0.60, 0.70 and 0.80. The groO\'cel surfaces
"'ill be applied between x=200mm and 750mm, i.e x/c=0.27 and 1.00 if c denotes
the total modcllcngth. One can observe that the external Mach number is const.ant
all along tllis manipulated length; this could be achieved because of the presence of
adaptive \Yalls. In the vicinity of the elliptical nose, the higher the value of M,:xJ 1S,
126

M[

1.Or t Moo=
',I /~~- _ _ _ _ _ _ _ _ _ _ _ 0.80
I(~ 0.70
IV
" -------- - 0.60
0.5j~ -0.55
~~ 0.42
II

- X/c (c=0.75m)

o 0.5 1.0
Figure 2: Local .lIach number distribvtions (cylinder)

the greater the amplitude of the excess of velocity is. Knowing these distrihutions.
boundary Inyer computations were performed in order to estimate the contribution of
the friction drag to the total measured drag. The code. developped at CERT by Arnal
et aL [:3], soh'es the local boundary layer equations in three-dimensional compressible
flows. By integrating the local shear stress along the curvilinear abscissa. one ends up
wi th t he computed skin-friction drag coefficient, Cd J : the chosen reference surface
has been based on the cylinder diameter (7r9 2 / -'). Boundary layer tripping induces
inevitably the problem of over-thickening due to roughness clrag. Computing the
step in momentum thickness, 0, at the transition location, [2], and including it in tlw
bounda ry byer code, one could deduce the change in 0 at the end of the cylindrical
body. The btler is rather small (0.6%), [13], which means that the penalty roughness
drag can be lleglected. Thus, comparing Cd J to the measured total drag coefficient
Cel. rewalecl that the contribution of the pressure drag was rather weak (:::::: :3-''5'0)
and almost independent of the Reynold" number, [1:3].

2.1.3 Performances of grooved surfaces


The results fur the four riblet models collapse best when non-dimellsionalisecl by ht.
where the rib height is scaled with respect to the inner variables of the turbulent
bOllJldary Llyer taken at the wall : h~= (h/iJ w ) . jTw/ p"" Of course, h~ decreases
\\'i th respect tot iJe streamwisc coordinate, beca use of the "Reynolds effeci". IIowever,
its stream\\'ise \'cll'iation is smooth so that it is possible to define an a\'erage \'ahle.
ht, from intcgration of h~ all along the riblet modcllength, [9]. Then. that value is
represcntati\'c of the rib geometry for given test conditions.
In order to estimate as carefully as possi ble, the effect of streamwise groo\'Cs on
the skin-frict.ion drag coefficient, bvo main features have to be recalled:
the friction drag force contributes to about 96% of the total drag of the model:
the length of the manip1llated boundary layer is L=YiOmm. \\'hich implies that
tIte ratio: manipulated area / total \vetted area is close to 77%.
127

20

~~d If ("10) 0 .
10-
t B
..
0

0 h (mm) = 0.023
.r. 0.033

.. ~
0.051

.. . , 0.076
coo
0
08.
o&,' OP/Ox =0
8 I 0
9~. ~
,0 0

,",
-10 --' _h~

0 50 100

Figure 3: Synthesis of drag data (- - - low speed results)

For the [our considered riblet models, the variations of the friction drag coefficient,
over 1, arc plotted versus h~ on figure 3. For the explored range of free-stream 1Iach
number and stagnation pressure or Reynolds number, this choice of grom'ed surfaces
allow to go through h~=7.7 to 77. There is some scatter on figure 3 due to the fact
that drag decrease is very sensitive to cross-section uniformity, surface finition, ... It
can be noticed that the biggest rib size (h=76pm) constitutes the lower branch of the
data bane!. For this specific groove height, the lower the value of 1C"J is. the smaller
the non-dimensionalised height is, but also the weaker the consistency of results is,
according to the afore-mentioned explanations (i.e. non adapted lower and upper
walls of the test section, for low values of ?lIex,). That scatter might also come from
the recorded dependency of percentage of drag variations on the free-stream i\Iach
number, for a given Reynolds number. Indeed, for a given manipulated Reynolds
number, IlL of about 13 10 6 , it was noticed that the greater :\lx is, the larger the
percentage of drag reduction is and/or the smaller the percentage of drag increase
is. [13] .
.\'evertllcless, maximum turbulent skin-friction reductions close to 7 - 8% have
been recorded for h~ close to 10-15 : however, when h~ :s: 20 (res]). 2': 30) drag reduc-
tions (resp. ill creases ) exist whatever riblet model is concerned. Let us mention that
on this same figure :3, the dotted line represents results obtained in incompressible
flows for the same kind of symmetric V-groove riblet, manufactured by the 3M Com-
pany ; aspect ratio equal to one, but h=0.152mm, [9]. These data were obtained from
meaSllremellts of the change in momentum thickness. There are a few difficulties in
using that teclmique for evaluating the efficiency of rib let models; in fact, the drag
128

o McLean et al 19
1.08 o Walsh et al 26
/':, Coustols 9
1.04

1.00
til
0 Squire et al 22
'-.....
0 0.96

Fit of transonic data


0.92
o 4%band
s+
0.88 0
10 20 30 40

figure ,1: Transonic drag data (from Walsh and Anders)

reduction of riblets is small and the performance, i.e the le\'el of reduction, might be
affected either by the scatter in the momentum thickness measurements or in the (J
de\'elopment along the acti\'e riblet area exposed. From the reported experiments. it
seems that the "ariable h;; allows to gather data from high and low subsonic ]0\'"5.
Some authors argue that the upper limit of the reduction domain (the zero-drag
reduction crosSO\"Cr point, as quoted by Walsh et aI, [27]) be slightly higher in the
transonic regime. lIowe\"er, one has to be aware that for transonic conditions, the rib
height could be scaled with either the wall Yariables : h~= (hlv w ) . JTwl p"" or the
free-stream OWeS: h-:-= (hiVe) . JTwl Pe' The latter corresponds to the incompressible
\'alue and i~ used by some authors, for instance Squire et aI, [22]. Depending upon
the Reynolds and .'-Iach numbers, the difference between h~ and h-:- can be more
or less important. Thus, considering this cylinder-type body, when RL ~ 1:3 lOG
the ratio (h-:- I h~) is close to 7.2% (resp. 12.2%) for a free-stream Mach number
of 0.6 (resp. 0.82). This might be the reason of the differences between high and
low subsonic data since it is not always ob\'ious to guess what variable is used on
different plots ... Anyway, in their last paper. \\'alsh et aI, [11], summarized and cor-
related the <l,"ailable riblet film data. It appears that "significant riblet data are now
available tv firmly establish their drag reduction performance"" Collecting data from
Coustols, [9], ?dcLean et aI, [19], Squire et aI, [22] and Walsh et aI, [26], Walsh et
aL [27], found that all the transonic data, as for the low speed data, are essentially
contained in a 4-pcrcent data band as shown on figure 4. One has to be aware that
the rib let performances ha\'e been obtained with different techniques for estimating
the drag ,"ariations (changes in momentum thickness, direct drag measurements, ... ).
Let us add that all the considered riblet models had the worldwide tested symmetric
V-shape, with an aspect ratio equaJ to 1, explaining why any rib parameter (height,
h, or spacillg, s) can be used for plotting riblet performances. Let us recall that
experiments performed by Coustols, [9] or COllstols et aL [121, pointed out that the
129

a
6~d I (%)
t
10
o
f
o

o 0

O~--------~--------~~------~----L---~

o
o
o
o
.
o

00')


-10
s=h=0.023 mm -1{"w

10 15 20 25
Figure .5: Effect of free-stream flow angle

zero-drag reduction cross-over point was smaller when the aspect ratio (s/h) increases
from 1 to 3. The effect of varying the aspect ratio seems to affect the le\"el of skin-
friction reduction; one has to be aware that the greater this ratio is, the smaller the
increase of \\"etted area is.

2.1.4 Inft uenee of free-stream ftow angle


The effect of misalignment between grooves and the external flow direction has been
investigated by several authors for two-dimensional boundary layers developing at
low speeds, [9], [12], [2.5], at transonic speeds, [22], or even in flight conditions. [19].
In this present experiment, the model axis is maintained aligned \\"it h the direc-
tion of the ex(ernal free-stream flow; on the other hand, the streamwise groO\"es are
at a given angle, 'Y, to the cylinder symmetry line. \Ve assume there is no induced
three-dimensional or helicoidal effect. Data are plotted on figure .j for a single rib let
model (s=h=O.023mm) and two angles of yaw: 10 and 20. It can be observed that
the groO\"es keep a beneficial effect, even at 20 of angle of yaw. However. the recorded
drag reductions are less important than at zero angle of yaw, and the higher y is, the
smaller the zero drag reduction crossover point is. Thus, this effect is quite compa-
rable to that evidenced at low speeds, for flat plate-type experiments, [9], [12], [2:j].
AbO\'e 20 of angle of yaw, one could guess that drag reduction would quickly vanish,
as clemonstra.led by Walsh et aI, [2.5]. During their flight test experiments, IvIcLean et
ai, [19], found that a deliberate yaw angle of 1.5 decreased the riblet drag reduction
of about .50%. This is a stronger effect of lTlisalignment than that observed either in
this set of experiments or in others; that could be explained by the additive effect of
pressure gradient and angle of yaw, or by the difficulty to achie\"e a good measure-
ment accuracy in flight, using changes in momentum thickness for evaluating riblet
performances.
130

2.2 CAST 7 Aerofoil


The study of the efficiency of grooved surfaces for reducing skin-friction drag has
been pursued in a context closer to aircraft applications, by considering the flow
developing on both sides of an aerofoil in transonic conditions. \Vith a view of a
possible application of this type of passive devices, not only to the fuselage of a
transport aircraft but also to the wings, fin and horizontal taiL it is necessary to
analyse the behaviour of riblet models under adverse and/or favourable pressure
grad ients.

2.2.1 Experimental apparatus - Drag measurements


The considered aerofoil is a CAST 7 one, the chord length of which is 200mm : it is
steadily lwld between the two lateral walls of the test section, at an angle of attack
of 0, and an angle of s\\'eep of 0. The boundary layer is tripped at .570 chord length
from the leading edge. on the suction and pressure sides, by using a carborundum
banet the average height of which is close to 4.5f.lm.
Four symmetric V-groove riblets, manufactured by the 31\1 Company in an
adhesi\'e backed film, with an aspect ratio (s/h) of one, have been tested: h = 0.017.
0.02:3,0.03:3 and 0.051mm. The grooved surfaces have been applied bet\\'een l.jo/c and
10070 chorcllength ; in front of the leading edge of the ribs, a smooth vinyl sheet has
been stuck, in order to avoid the facing step effect. The thicknesses of the adhesiw
bcd-plate of the rihlet film (~ 90pm) and of the smooth vinyl sheet (~ 100pm) are
almost identical; then, as it was the case for the cylinder, whatever groove depth is
considered. the ribs trough plane is approximately mounted flush with the adjacent
upstream smooth part, [20]. The reference configuration, i.e the "smooth" aerofoiL
will consist of the upper and lower sides covered completely with the smooth vinyl
sheet. Of course, the carborundum band will lay upon the smooth surface.
The tests ha\'e been performed at ambiant stagnation temperature. for a stagna-
tion pressure close to 1.6 bar, and for a free-stream Mach number range, Mso : 0.65 -
0.76. \Vhate\'er the value of 1\100 is, the lower and upper walls of the test section
have been adapted. For a stagnation pressure, Pi, of 1.6 bar and a free-stream l\Iach
number of 0.76, the chord Reynolds number, R c ) is close to 3.3 10 6 . For these specific
tests, no variation of Pi has been taken into account ; the slight variation on Rc is
induced by the l\Tach number effect.
The performances of these four rib let models, under different pressure gradients,
have been determined through drag variations estimated from Pitot tube surveys in
the aerofoil wake; that streamwise location corresponds to x/ c= 1..5. The displace-
ment has been adjusted according to the total wake height : 31mm (resp. 71mm)
when :-10::' : 0.65 - 0.72 (resp. 0.74 - 0.76) ; this wake increase corresponds to the ap-
pearance of a shock on the suction side of the aerofoil. Special care has been taken in
order to lay hold of measurement accuracy: number of data points within the wake,
wake survey duration, same value of free-stream Mach number, precise 'Nail adapta-
tion, ... , [20]. For 1100 ::::: 0.74, the data scatter on the wake momentum thickness
or on the total drag coefficient is l:,Cd/Cd ~ 0.7% ; for greater values of the Mach
number, the scatter is greater mainly because the experiments are carried out very
close to the divergence Mach number, from which the total drag coefficient rapidly
lIlcreases.
131

M~ ~ 0,76
/

a=O
Hc=3.5 10(,

-..X/c
Ol-------
o
Pigure 6: Local Mach number distributions (CAST 7 Aerofoil)

2.2.2 Local Mach number distributions

The local 11ach number distributions, 1l" are plotted on figure 6 for three values of
the free-stream 11ach number, Moo = 0.65, 0.70 and 0.76. As it can be observed,
for an angle of attack of 0, the chosen increase in 1\100 allows a significant variation
of pressure gradients. On the pressure side, whatever the value of Moo is, the flow
is decelerating from xl c=35% to 90%, then accelerating towards the trailing edge.
On the other hand, on the suction side, all along the distance subject to future
manipulation (0.15 5::x/c5:: 1.00), the data exhibit a flat plate-type evolution for
11::0 5:: 0.7-1 ; above that value, a shock appears at about 40% chord length.
For the reference configuration, when the pressure and suction sides of the
aerofoil are covered with a smooth vinyl sheet, some of the 103 pressure taps have been
uncorked in order to check eventual modifications of the pressure distributions, [20].
\Vithin the experimental uncertainty, there was no obvious variation; furthermore,
the total drag coefficient estimated through wake surveys was almost identical to the
one measured with an uncovered aerofoil. Though grooved surfaces ha\'e a significant
effect on fricLion drag, none experiment clearly establishes they might also modify
the pressure field. Thus, one could believe that replacing part of the smooth film by
riblct film \\'ould not have any influence on the pressure drag.
KnO\\'ing the local Mach number distributions, the same boundary layer code,
as the one mentioned in the preceding sections, has been run in order to compute
the local skin-friction coefficient and to integrate it all along the aerofoil, so that one
can get an estimate of the friction contribution to the total measured drag. That
contribution slightly decreases as the free-stream Mach llumber increases: 58% at
l\Lxo=0.65 dowll to 48% at 1100=0.76.
From the boundary layer code, it is also possible to compute the streamwise
evolution of the non-dimensionalisecl rib height parameter, h~, on either the pressure
or suction side of the CAST 7 aerofoil. Because of the different pressure gradients, h~
132

1,3
I>
0 smooth
100 Cd
h=O.023mm
I>
1,2 >--- + h=O.033mm
r I> h=O.OSlmm
1>1>
I>
+

+
1,1
0
1/--
I> 0

+
;I- 0
[D
+ ~
1,0 -'1]
+



0,9
I ) 1=
0,63 0,67 0,71 0,75
Figure 7: Synthesis of drag measurements

does not behaxe as smoothly as, for instance, on the cylinder. Nevertheless, one can
end up with an average value, integrated all along the rib let model between 15% and
100% chord length. This value is not identical on both sides of the aerofoil : indeed,
for a free-stream I\Iach number of 0.72, on the pressure (resp. suction) side, ht=0.7
(resp. 0.6) for a given groove height h=lflm. As these average values are rather close,
a mean value of 0.6.5 per pm would characterize the riblet model, root of passive
turbulent manipulation on both sides of the aerofoil. That mean value h~=0.65 is
slightly modified by the variations of the free-stream Mach number.

2.2.3 Performances of grooved surfaces


For the reference configuration, the total drag coefficient of the aerofoil - Cd - covered
with smooth ,iml sheet. has been determined as a function of the free-stream Mach
number at a str~amwise abscissa located half-chord behind the trailing edge : Cd
increases smoothly with 1\1 00 , Data concerning three grooved surfaces are plotted
on figure 7 ; for each model, 7 surveys have been, at least, performed in order to
scrutillize the whole free-stream Mach nurnber range. For h=2:3pm, drag reductions
(:::::: -:3.:3%) are recorded; moreover, the level of reduction seems to be independant of
1'11'::0' On the other hand, for higher rib sizes drag increases have been obtained:::::: +1 %
(resp. :::::: +10%) for h=33pm (resp. 51flm). 'vVithin the experimental ullcertainty,
results given by the smaller rib let (17 pm) are almost identical to those provided with
the model h=23/1m. For the model h=23pm, wake surveys have been carried out for
an intermeelia le configuration: grooved surfaces on the pressure side between 15%
anel 100% chord length and smooth film everywhere else. A decrease in the total
drag coefficient was recorded:::::: -l.;,)%. That reduction is approximately half the one
obtained with ribs on both sides of the aerofoil, which means that, besides the fact
that the intensities of pressure gradients are different, the grooves behan> in a similar
manner. In fact, this test was performed at the beginning of the campaign undertaken
133

[1
25

20
- I
/:::,. Cd
- (%) c CAST 7
~
Cd 15 -- - Cylinder
friction
- Cylinder V
over L
~
I
10
V ------
/ V/
Jr V
5

/ /
0

-5
- _h:
-B-
L-------'" -
-
-10
o 10 20 30 40 SO
Figure 8: Synthesis of drag data (with and without pressure gradient)

in the T2-wind tunneL As the level of recorded skin-friction reductions was relatively
low, it was decided to cover not only the pressure side but also the suction one in
order to increase the percentage of manipulated wetted area, and with that the drag
variations_ From these preliminary results, it was tried to estimate the variations of
the friction drag coefficient over the manipulated aerofoil surface, the percentage of
which. towards the total wetted area, is close to 85%. Let us recall that boundary
layer computations showed up that the contribution of the friction part in the total
drag balance represented between 58% and 48% depending upon the ,-alue of the
free-st ream :-1ach number. The variations of the friction drag coefficient. over L, are
plot t eel on figure S versus h1;;, mean value of both integrated h-:: parameters e,-aluated
along the manipulated surface on pressure and suction sides. Data are compared to
those obtained on the cylinder; because of the scatter in this laUer experiment the
upper and 10\\'er curves of the data domain are represented. R('sults obtained by
lTlanipulating the turbulent boundary layer which develops on the CAST 7 aerofoil
are plotted in the shape of rectangular boxes: the horizontal box side expresses the
variation of h~ with 1\L;;v, when the vertical one denotes the slight dependency of drag
varia tions \\'i tll 1\ I 0::' and the experimental scatter. The skin-friction drag variat ions
brought about b.y the two smallest riblet models (2:3 and 33{Lm) are in agreement \\"ith
those recorded on the cylinder without any pressure gradient. Howewr, the lewl of
cl rag increase obtained from the deeper rib (51pm) is higher. This difference could be
explained probably by a bad quality of the model surface finition or a cross-section
non uniforlllity, since the other two models do not seem to show up a noticeable effect
of pressure gradients on riblet performances.
Thus, when grooved surfaces cover about 8-5% of the aerofoil wetted area, fric-
tion drag reclLlctiollS have been recorded [or ht :::; 20. 1\Iaximum total drag decrease
of about 3.5% was obtained for the smaller rib hei~, which corresponds to max-
imum skin-friction reduction of almost 7.5 - 8% at h1;; close to 12 - 16. This result
is obtained by assuming thai grooves do not modify the pressure drag but act only
on the frictioll drag. For free-stream rvIach numbers less than the di\-ergence iV1ach
134

number, the average pressure gradient parameter estimated along the manipulated
area, {J - Clauser parameter - for instance, is weak on the suction side since the local
l\Jach number is almost constant over 35 - 40% chord length; on the pressure side,
this parameter is stronger but still moderate.
The performances of grooved surfaces under pressure gradient conditions have
been investigated by some researchers in incompressible flows by Choi et ai, [8], Cous-
tols, [9], Truong et ai, [23], in transonic flows by Squire et ai, [22] and also in flight
by ?I1cLean et ai, [19]. The data collected from all these experiments seem to be
consistent, though the drag variations were estimated through different measurement
techniques. In summary, significant riblet data are available, for flows subject to ad-
verse and/ or favourable pressure gradients, to suggest their drag reduction efficiencies
as the pressure gradient is moderate, which is the case for most of the wing surface,
where grooved surfaces might be applied.

2.2.4 Miscellaneous
For these experiments, the grooved surfaces were applied onto the trailing edge of
the CAST 7 aerofoil. Then, depending upon the considered vinyl sheet, smooth
ftlm or rib let one, the thickness of the aerofoil base val'ies from 0.2111111 (reference
case) up to 0.282mm (h=0.0.51mm). Compared to the aerofoil covered with smooth
film, changing the rib size from 2:3llm to 51pm would raise the base drag of about
25%. For the reference configuration, at 1\100=0.7, an estimate pointed out that
the base drag accounts for a little bit more than 1 %. As in all these experiments,
one is looking for small variations of the total drag coefficient, it would be better
to keep constant the thickness of the trailing edge, so that, when manipulating the
turbulent boundary layer, the only recorded changes ,\'attld be devoted to skin-friction
reductions or increases. Of course, this remark is useful for small models in wind
tunnel. but "'oldd be useless for aircraft applications.

3 Experhnents in the Sl-wind tunnel of Modane


Following the successful tests, as regards the efficiency of ribbed surfaces, either on
an axisymmetric body for zero-pressure gradient flows or on an aerofoil, in transonic
regime, further experiments were carried out at the ONERA/Modane Sl-wind tunnel
in collaboration with Aerospatiale, [10], [l i l]. The area of the test section, the length
of which is Um, is close to 40 m 2 .

3.1 Experimental apparatus


The model is a 1/11 th scale Airbus A:320 model, mounted on a straight sting, which
is maintained through a tripod set-up. The considered model configuration has no fill
and no horizontal tail. The fuselage length, I, is 3.416m and the mean aerodynamic
chord length is 0.:381m.
The experiments have been performed at ambient stagnation temperature (c::::
300 K) at a stagnation pressure equal to the local atmospheric pressure (0.9 bar). The
unit Reynolds number range is : .5 ..5 - 1l.8 10 6 . The free-stream ~\lach number varied
from 0.:3 up io 0.82 ; this latter value corresponds to the configuration "fuselage
135

alone", for which the maximum Reynolds number based on the fuselage length is
a pproximaiely equal to 40 10 6 . The angle of attack of the model could vary between
_2 and +3.
Boundary layers were tripped on the fuselage and on the wings, using carborun-
dum bands. Their locations had been defined by Aerospatiale during other sets of
experiments performed with this specific Airbus model.
Only one symmetric V-groove riblet, manufactured by the 31\11 Company in an
adhesive backed film, with an aspect ratio of one, has been tested on the fuselage and
wings. The knowledge of the flow field around the fuselage, obtained from the wing-
body configuration, allowed us to perform boundary layer calculations, based upon
the resolution of integral equations ; thus, for instance, at ]\1==0.7 and 0=3.7,
one could get the evolution of h~ along the upper and lower symmetric lines, the
lateral mid-lines and so forth. It appeared that this quantity was not varying too
much with either the streamwise abscissa or the peripheral co-ordinate; so, a mean
\"alue - though it might change with the angle of attack and the free-stream 11ach
number - would be representative of the rib let scale. Thus, in the case of ~roove
depth of 0.02:3mm, for the rib let material set on the fuselage, an optimized h;;, value
close to 8 - 0 was found at ~-Ioo =0. 7.
1Ieasuremenis have been performed for fuselage and \\"ing-body type arrange-
ments. Several configurations have been considered:
- Fuselage alone without riblets ;
- Fuselage coyered with riblets ;
- Wing-body configuration without riblets ;
- Wing-body configuration with riblets set only on the fuselage;
- Vving- body configuration covered with riblets.
Along 1he wings, the grooves were approximately aligned with the external
free-stream flow direction. No grooved surface was set in areas subject either to high
geometric curvature (nose or tail cone) or to important streamline curvature (more
than 1:5 from the ribs direction), i.e. wing-fuselage intersection, upper surface of the
wing near the trailing edge, ... The corresponding percentage of wetted areas co\"ered
with longitudinal grooves is approximately:
- Fuselage alone: 7:3% ;
- Wing- body configuration ,,ith riblets set only upon the fuselage: 47o/c :
- Wings and fairings: .57% ;
- Wing-body configuration (fuselage, wings and fairings covered with riblets) : 66%.

3.2 Drag measurements and repeatability


The total forces, especially drag and moments, were measured through an internal
six-component balance; the total drag coefficient, Cd, was obtained by taking into
account the model reference surface (1.012m2). As the expected differences in the
drag forces would be very small. the measurements consistancy was a t first checked
by making a fel\" continuous sweeps in incidence for any cOllfigu~'ation and at different
free-stream ?llach numbers. As shown on figure 9, an excellent repeatability on Cd
measurements has been achieved, the scatter being:
- for the fuselage-configuration, ~Cd < 0 ..5 10- 4 in 09% cases and < 0.:3 10- 4 in
0 1% cases ;
c

- for the wiug-body configuration, ~Cd < 1.0 10- 4 in all cases and < 0.5 10- 4 in
136

CL
test Mao
test Mo>
47-0.800 0,5
53 --- 0.800 104 - 0.699
2
63 -0.800 113 --- 0.700
73 0.800 114-0.699
1
( ) 0,3
6C"x10'

-2 0 2 0,2
b)
-1
a) 0,1
-2 6C Dx10'
~

-2 0 2

Figure 0: Repeatability of drag measurements a) Fuselage configllration b) (cing-body


configu ration

93% cases.
As tbis repeatability is of excellent quality, it was witb confidence that the
efficiency of grooved surfaces has been checked under very good conditions. Further-
more, let us point out that for both configurations, a rather good repeatability on
a.ngle of attack measurements was recorded since c.a < 0.010 in 00% cases.
A dozen of pressure taps was inserted within the fuselage (sting cavity) allowing
to check the homogeneity of these pressure measurements as well as their constancy
during all tIle drag measurements.

3.3 Performances of grooved surfaces


For the fuselage configuration, the variations of the total drag coefficient are plotted.
on figure 10, \'ersus the model angle of attack for three values of the free-stream ]Vlach
number: O..j, 0.7 and O.S. Thus, a very positive effect in terms of drag reduction
has been e\'idcnced, whatever the value of the angle of attack is, especially for large
\'alues of :-1.::0' Indeed, at 1'1'1==0.7 and O.S, the total drag coefficient decrease is
almost constant when the angle of attack, a, varies.
For C\ corresponding to the Cl cruise value (Cl=0.5), nett drag reductions of
1.9%, :2.:3% and :2.'1% were recorded respectively for 1\100=0.5, 0.7 and O.S. The angle
of attack corresponding to CI = 0.5 varies, of course, with the free-stream Mach
number.
\Vhen considering the wing-body configuration with riblets set only on the fuse-
lage, nett drag reductions were also obtained at 1\L)()=0.5 and 0.7, for different Cl
values lying lJctween 0.1 and 0.6. Furthermore, adding riblets on the wings and fair-
ings allowed to get 10\\'er Cd values, than the ones corresponding to the preceding
configuration. The nett reduction in Cd was 1.6% at 1\'1,:0=0.7 and at cruise level,
137

( 6 Cd I 1.9 % ( 6 Cd I 2.3 %
Cd )U:CI cruise Cd )0( CI cruise

'" = 0,7 '" = 0,8

2 2

o o
t.Cd=Cd ribleC Cd smooth
-1 -1 -1

-2 -2 -2

figure 10: Synthesis of drag data (Fuselage configuration)

(~ Cd)
Cd Clcrulse
= 1.2 % ( ~CCdd) = 1.6 %
Clcrulse

Cl

'" = 0,5 '" = 0,7


0,5 0,5

0,4 0,4

0,3 0,3

0,2 0,2

0,1 0,1

t.Cd=Cd ribleC Cd smooth

Figure 11: Synthesis of drag data (Wing-body configllration)


138

with negligible changes in llCd over the Cl range: 0.1 - 0.6 (figure 11).
Thus, at these explored free-stream Mach numbers, the benefit of covering wings
with riblets was observed, though the percentage of grooved surface compared to the
wing wet ted area was rather small (~ 57%). Moreover, though the grooves were
aligned with the external free-stream velocity, the effect of behaviour of riblets in
three-dimensional flows might be different from the influence of misalignment in hYo-
dimensional flo,,-s. Indeed, the main difference comes from the feature that the yeloc-
ity vector varies very quickly in the wall vicinity; so, non negligible angle deflections
exist over distances within the turbulent boundary layer as large as groo,-e depths.
Let us mention that the performances of riblets are under investigation at CERT in
three-dimensional flows: riblets are covering parts of the pressure and suction sides
of an aerofoil, set at an angle of sweep close to that of an Airbus-type wing.
\Vhate,-er configuration is concerned, it is rather difficult to guess a correct
estima te of the skin friction reduction mainly because the sting caused pressure field
modification and, consequently, drag interferences. Anyway, at i\L;,,=0.7 and at cruise
level, though the model has no fin and no horizontal tail, if one assumes that friction
represents about 50% of the total drag, the average nett skin-friction drag reduction,
estimated m-er ille 66% manipulated wetted area, is close to 4.8.5%.
This benefit is not certainly the highest one, because:
- the percentage of wetted areas covered with ribbed surfaces was rather smalL espe-
cially on the wings due to boundary layer tripping;
- on the fuselage, close to wing-body junction, the grooves had not been re-aligned
because of streamline curvature. Compared to the fuselage configuration, tested at
first. several strips of riblet film had only been taken off ;
- the grom-es size had not been optimized on the wings; the same depth and geom-
etry as the ones used on the fuselage were considered, mainly because of material
availa bili ty.
In spite of these observations, substantial drag gains have been recorded on the
wing-body configuration. Furthermore, these results roughly agree with the friction
drag gains measured under laboratory conditions, for instance with data obtained at
CERT on either the cylinder or the CAST 7 aerofoil in the T2-wind tunnel.

4 Conclusions
All the experiments carried out under transonic conditions have alloweu to verify the
efficiency of internal manipulators, riblets or grooved surfaces, for the purpose of
reducing turbulent skin-friction drag. Thus, from different research groups, significant
I'iblet data are now available to firmly establish the potential of such devices for drag
reduction performances. It appears that maximum drag reduction is found \\' hen
groow geometries, such as the height or the spanwise spacing is t}-pically of the order
of 10 - 1;3 v JUT.
Furthermore, experiments performed on a complete A320 model, in the Sl-wind
tunneL showed up that important total drag coefficient reductions could be achieved,
at cruise conditions. Since the fuselage Reynolds number reaches around 40 10 6 , it
is expected that these results could be rather easily applied to practica.l flight test
conditions where the Reynolds number is only increased by a factor :3.
Although the mechanisms involved in such a drag reducing process have not,
139

yet, been understood, a couple of flight tests have already been performed with fuse-
lage, wings, fin, horizontal tail and nacelles equipped with riblets. Some information
concerning the flight tests, carried out by Airbus Industrie and its partners, as a
consequence of the promising Sl-wind tunnel results, on the Airbus A320 N1, could
be found in [21].
Conscquently to these useful results obtained in transonic conditions, the next
step would consist to look at the behaviour of grooved surfaces in supersonic regime.
Indeed, for a Concorde-type aircraft, the friction drag represents about 30 - 35% of
the total drag, which implies that improvement of aerodynamic performances of such
supersonic planes could be achieved by friction drag reduction. To our knowledge,
the only supersonic drag measurements of a riblet film were performed by Gaudet
at a Mach number of 1.25, [18]. \Valsh et aI, [27], reported that the supersonic data
falls within the transonic data band. Let us mention that experiments will be carried
out at Ol'\ERA, under zero-pressure gradient conditions, for higher values of Mach
numbers up to 2..5.

Acknowledgements
Financial support was provided by Airbus Industrie and the "Service Technique des
Programmcs Aeronautiques". Special thanks are due to F. l\'larentic from 3?11-USA
and A. Dcladwnal from 31\I-France, for providing us \yith all of the riblet material.

References
[1] Archambaud J.P., Blanchard A., Seraudie A. : 11th Congress on Instrumentation
in Aerospace Simulation Facilities - Stanford, (1985)
[2] Arnal D. : V.K.I. Lecture Series - AGARD Report W 709 (1984)
[3] Arnal D., Jelliti 1\-1. : CERT Internal Technical Report (1985)
[4] Bacher E.V., Smith C.R. : AIAA Journal, Vol. 24, W8, pp. 1:382-138.5, (1986)
[.5] Blackwelder R.F. : AIAA Paper 89-1009 (1989)
[6] Bushnell D.l\1., IVlcGinley C.B. : Ann. Rev. Fluid 1\1ech., Vol. 21, pp. 1-20, (1989)
[7] Choi K.S. : J. Fluid Mech., Vol. 208, pp. 417-458, (1989)
[8] Choi K.S., Pearcey H.H., Savill A.M. : Int. Conf. on Turbulent Drag Reduction
by Passive Means, London (1987)
[9] Coustols E. : AIAA Paper 89-0963 (1989)
[10] Coustols E. : 4th Int. Conf. on Drag Reduction, Davos (1989)
[11] Coustols E., Gleyzes C., Schmitt V., Berrue P. : 24ieme Colloque AAAF Poit-iers-
France (1987)
[12] Coustols E., Cousteix J. : 16th ICAS Congress, Jerusalem (1988)
[n] Coustob E., Seraudie A., 1\Iignosi A, Breil J.F. : CERT Internal Technical Report
(1988)
[11] Coustols E., Savill. A.1\1. : Applied Scientific Research, Vol. 46, pp. 183-196,
( 1989)
[1.5] Coustols E., Cousteix J. : 2nd IUTAM Symp., Zurich (1989)
[16] Coustols E., Cousteix J. : 7th Symp. on Turb. Shear Flows, Stanford (1989)
[17] Djenicli L., Liandrat J., Anselmet F., Fulachier 1. : 2nd European Turb. Conf.
(1988)
140

[18] Gaudet L. : Applied Scientific Research, Vol. 46, pp. 245-25'1, (1989)
[19] }\1cLean J.D., George-Falvy D.N., Sullivan P.P. : Int. Conf. on Turbulent Drag
Reduction by Passive Means, London (1987)
[20] Prudhomme S., Coustols E., Mignosi A., Dor J.B., Plazanet M. : CERT Internal
Technical Report (1989)
[21] Robert J.P. : to be published in that Proceedings Volume (1990)
[22] Squire L.e., Savill A.~1.: Int. ConL on Turbulent Drag Reduction by Passi\"e
!\leans, London (1987)
[2:3] Truong T.Y., Pulvin P. : Applied Scientific Research, Yolo 46, pp. 217-227, (1989)
[24] Walsh 1\1.3. : AIAA Paper 82-0169 (1982)
[25] \\"alsh :\1.3., Lindeman A.IV!. : AIAA Paper 84-0:347, (1984)
[26] Walsh :\1.3., Sellers III \V.L., McGinley C.B. : AlA A Paper 88-2.,)54, (1988)
[27] 'Walsh 11.J., Anders J1'. J.B. : Applied Scientific Research, Yol. c16, pp. 2.5,')-262.
( 1989)
Effect of Riblets on either Fully Developed
Boundary Layers or Internal Flows in Larninar
Reghne
J. LIANDRAT 1 , E. COUSTOLS 2 , L. DJEI'\IDIl,
F. ANSELMET 1 , X. de SAINT-VICTOR2 , F. FIOC 2 &
L. FULACHIER 1
1 Illstitut de Mecanique Statistique de la Turbulence, Unite Mixte [inivasite.
CNRS I\'" 38033, Marseille, France;
2 ONERA-CERT, Aerothernwdynamics Department. Toulouse. France.

SUMMARY - The present paper summarizes the status of the numerical research carried
out both at OKERA/CERT and IMST, as regards the efficiency of internal manipulators,
more commollly named by many researchers, riblets, in external as well as internal flows,
under laminar conditions. First results obtained at CERT during preliminary studies will
briefly be recalled. From a numerical point of view, many difficulties arose; most of them
stemmed from the singularity at the riblet edge. First of all, results issued from different
grid refinements will be presented. Then, emphasis will be put on analogies and differences
between manipulated fully developed internal and external flows in the presence of either
\'- or L-shaped riblets ; the latter corresponds in fact to half of a U-shaped groove. These
theoretical studies showed up that the wetted area increase did not induce extra skin friction
in laminar boundary layers over V- or L-shaped grooves.

1 Introduction
It is clear from various experiments, performed up to now in turbulent boundary
layers, that l'iblets of suitable dimensions can reduce wall friction by about 6 - S%
(see. for instance, Bushnell [:3], Coustols [.5] or Walsh et al [19]). In the case of
internal turbulent flows there are much fe\ver experimental results but it seems from
Liu et al (sec [:20]), Lowson et al [Ue] and Rohr et al [16], that similar reduction could
exist. This reduced skin-friction seems indeed to be related to an increased stability
of near-wall structures (i.e. Bacher and Smith [1], Choi [4]). But, what is the physic~l
mechanism responsible for this stabilisation? By considering the most usual situation
of Y-groO\'es with s=h, (s denotes the spacing between two adjacent grooves and h the
rib depth), the wetted a.rea increase can be greater than 100%. If one considers that
this specific increase tends to enhance skin friction, as pointed out experimentally
by Pulles et al [1.5] and numerically by Launder et al [13], it is difficult to accept
that turbulence Oll its own can compensate for and even overshoot this negative
effect. Thus, it is reasonable to think that a purely viscons mechanism is involved
and either entirely or partly counteracts this areil increase. This is the reason why
an experimenlal study on laminar boundary layers o\,er triangular l'iblets has bee;!
performed in the I.lVI.S.T water tunnel [11]. Results have shown that, over grooves

141
142

with s=2.3h, [10], [7], (32% wetted area increase) and riblets with s=l.2h (94 % wetted
area increase, unpublished results so far), the overall skin friction is practically the
same as that oyer the smooth wall: there is, thus, no straight relationship between
skin friction a11(1 wetted area.
In order to specify whether skin friction is slightly increased or eventually re-
ducecL a numerical study has been undertaken at Il\lST, in the same conditions as
in experiments, based on a code initially developed at CERT [6]. First results, ob-
tained at IMST, [8], showed a small drag reduction of about 3%. although the grid
mesh had to be refined in particular in the crest vicinity [9]. These calculations were
ca.rried out over V-shaped riblets and the rectangular grid was not orthogonal to
walls: this is, mainly, one of the reasons why computations over L-shaped grooyes
have been pursued at CERT and performed using different grid refinements. \' ever-
theless, these first Il\JST results were in agreement with Kahn's ones, [12] : indeed,
Kahn computed the boundary layer development over a V-groove riblet, s=2h, and
reported a 17c reduction in laminar regime.
As far as laminar internal flows are concerned, there seems to be only a single
experiment<d result relative to a fully developed pipe flow with triangular grom-es
with an aspect ratio (s/h) of 1, [16] : within the measurement accuracy, there was
no effect of riblets on skin friction. Preliminary calculations carried out at CERT in
a rectangular channel with V-shaped grooves [6], and computations run at Ul\IIST
in channel with L-shaped grooves [13], have shown a drag increase, especially in
Reference [13], where the skin friction variation is significant, about a 307c increase
for s=h and h/II=0.05 (II denotes the channel half-width).
It should be noted that a large amount of theoretical work has also been per-
formed by 13echert et ai, [2] ; indeed, he and his colleagues looked at several groove
cross-sections by considering transformal mapping and solving Laplace's equation.
Their results were analysed in terms of "protusion height", a measure of the influ-
ence of the riblets upon the boundary layer, rather than in terms of ability of such
surfaces to decrease shear stress.
In order to try to clarify some specific features, attention is focused, within this
paper, upon:
- analogies and/or differences between fully developed internal flows and boundary
layer flows in the presence of either V- or L-shaped riblets
- existence or lack of skin friction increase.

2 Equations and N unl.erical Methods for Bound-


ary Layer Flows
\Vhatever the groove cross-section is (for instance, V-, U- or L-shaped riblets), the
equations to be solved are identical. Those will be defined in a (a;, y, ::) Cartesian
co-ordinate system; :r is the streamwise abscissa, whereas the independant y and z co-
ordillates define the plane at right angles to the mean flow. For laminar steady three-
dirnensional flows, the P arabolized N avier Stokes equations have been considered.
This set of equations can be soh-ed with either a semi-elliptic or parabolic approach,
according as the streamwise pressure gradient is either unknown or gi\'en. Both
approaches \\-ere considered by de Saint-Victor in his preliminary 5t udies concerning
the manipulation of laminar flows over triangular cross-section groon's, [6].
143

In the afore-mentioned co-ordinate system, the P arabolized N avier Stokes equa-


tions can be read:
DUi 1 ap a 2Ui a 2Ui
- = - - - + v ( - +-) (1 )
Dt P aXi ay2 0:;2
where Ui = (U, 11, TY) and Xi = (x, y, z).
The continuity equation is added to that system:

(2)

Though the diffusion terms, taken along the streamwise direction, are neglected, the
elliptic feature holds according to the streamwise pressure gradient. Thus, the preced-
ing system corresponds to what has been called "semi-elliptic case". If the pressure
is given, the clliptic behaviour disappears and the system to be solved provides us
with a "parabolic case".
This set of equations is solved using an x-marching method, where it finite
\'olume schemc is considered for discretising any deduced equation. The numerical
code is based on a Semi-Implicit Method for Pressure Linked Equations, [IS]. During
iterations, en'ry equation (momentum and pressure correction equations) arc solved
by either the Modified Strongly Implicit algorithm, [17], or t\\o sweeps in y- and
:;-clirections. The efficiency of one or another method depends strongly upon the
mesh size within the groove, the number of points within the computational domain,
the groove cross-section shape, and so forth.
Initial conditions arc given on the velocity components at the initial (Y. z) plane,
whereas the pressure or its derivative with respect to :r, is fixed at the last streamwise
(Y. z) plane. No-slip boundary conditions apply at all rigid surfaces and symmetry
conditions along the lateral boundaries within the fluid (aU/az = 0, aF/az = O.
TV = 0) : at the upper limit of the domain, which corresponds to approximately 20
boundar}' layer thicknesses, the first derivative with respect to y of the streamwise
velocity component is equal to zero.
First calculations were performed on a V-shaped groove such as s=211, 11/8 :::
0.125. [6]. whcre 8 denotes the thickness of the laminar boundary layer at the leading
edge of 1he riLlet model. Computations started on a smooth plate. with a Blasius t~'IW
profile. After 0.25m downstream distance, the laminar boundary layer dewlopcd o\'er
the riblet model. the length of which was 0.2.5m. Then. the wall geometry went back
to a smooth plate. A single configuration was only studied because of computational
economy due to the great number of nodes: 7.5 planes in the streamwise direction wit h
all * 19 mesh in the y * z directions \\'ith 10 point.s coyering the groo\'e depth. Doth
approaches, semi-elliptic and parabolic ones, prm'idec! local skin-friction reductions
oyer the riLlet modeL compared to a smooth surface. Hmveyer. some differences
su bsisted. mainh' in the vicinity of the geometric singularity. i.e junet ion of the riblet
configuration alld the downstream smooth onc, due to the downward step effect. for
installce. a lilrge skin-friction increase \\'a5 recorded upstream of this singularity, in
the semi-elliptic approach. III facL that same behayiour was also obsen'ecl when
computing the development of laminar flow nearby a downward step.
Besides these considerations, one of the most significant results has been that
the transyersc velocity component, TV, is negligible in all the computa.tional domain,
solving the system of equations with either a semi-elliptic or parabolic approach. This
144

has also been evidenced through experiments performed at IMST, [11], [7]. Neglecting
this transverse velocity component is, moreover, consistent with Kahn's work, [12].
Indeed, those last computations revealed, by solving the parabolized streamwise mo-
mentum and vorticity equations, the absence of secondary flows, which do not form
a vortex within the groove, in laminar flows. However, additional computations
performed by the same author but, in turbulent regime, indicated the presence of
secondary flows which "are found helping the viscous forces in the drag reduction
process" .
It should be noted that such an assumption is very useful since it allows to
simplify tremendously the system of equations and, consequently, to diminish the
computational time. Thus, the system of parabolic equations can be written as
follows:
o(UU) o(UF) 1 dP 02U 02U
-ox- + -oy- = ---
p dx
+ v( - +-)
oy2 OZ2

aU + oV = 0 (-)
ox oy
That system of equations is solved using an x-marching method; the strealmvise ve-
locity component, U, is obtained from equation (3), taking equation (4) into accounL
until convergence. Then, from the continuity equation the normal to the wall velocity
component, F, is deduced. The initial and boundary conditions are identical to those
used for soh'ing the complete three-dimensional case.
This "semi three-dimensional approach with Hi = 0", has been developed both
at IMST and CERT for looking at laminar boundary layer manipulation respectively
with V- and L-shaped grooves. Particular attention has been devoted to the estima-
tion of the mean local skin friction coefficient, since small variations of this quantity
are seeked [01'. Depending upon the groove cross-section, that estimation will be
more or less easy. The choice of L-shaped riblets made at CERT differs from that of
either earlier studies [6] or Il\IST work [7], [9], [10]. ;\Ievertheless, these specific ribs
gave similar drag reductions to ones of triangular section, as demonstrated experi-
mentally by Wilkinson et al [21] for turbulent boundary layer flows; it was expected
that it would be the same case in laminar external flows. That choice was made
in the interest of computational "easiness" since the rectangular grid mesh could be
easily defined. ]\loreover, the skin friction coefficient can be obtained directly from
streamwise velocity component derivative with respect to y or z co-ordinate. The
main resulls o[ these studies will be discussed in !1.

3 Larninar Internal Flows


For the last decade, a lot of experiments has been perforrned by several investiga-
tors, in order to confirm that drag reduction was possible through streamwise micro-
grooves geometry in developing boundary layer flows. Indeed, one of the first possible
approach of such a passive method was devoted to aircraft applications. IIO\\,ever,
there has been much less work investigating the efficiency of riblets in fully devel-
oped internal flows, though the technique for estimating skin friction reduction seems
easier since it is deri\'ed from the pressure drop.
145

The efficiency of riblets in laminar internal flows had briefly been considered
at CEnT, [G]. To our knowledge, the only other numerical work concerning the ef-
fect of ribleted surfaces in internal flows has been performed, later on, at Ul\IIST
by Launder and Li [13]. Nevertheless, based on the results obtained from developing
turbulent boundary layers, these authors made the assumption that "the contribution
of the conVE'cti"e terms to the flow pattern in the vicinity of the riblets is negligi-
ble", since the groove depth does not exceed a few percent of the boundary layer
thickness. Thus, they performed a complete examination, from a numerical point of
view, of mani j)ulation of fully developed flow through a plane channel under laminar
conditions, [1:3] ; it should be noted that they considered knife-edge riblets in internal
turbulent flows.
The computations carried out at CERT were performed using equations (1)
and (2), soh'eel under the parabolic form. The main difference with the developing
boundary flow is that the mass flow will be kept constant ; that constraint allows
one to obtain the mean streamwise pressure gradient at each computed plane in the
streamwise direction. The numerical code was run for as many streamwise abscissa
as necessary, until convergence was reached. Since the geometry was symmetrical.
the flow \\'as studied for 0 ::; y ::; H, where H is the channel haH-width, i.e half
the distance bet\\'een the two facing grooves valleys. The number of grid points
was ,11 * 19 in the y * : ; directions, with 10 points covering the rib height. The
boundary conditions were similar to those used for external flows calculations. 1'\0-
slip boundary conditions applied at all rigid surfaces and symmetry conditions along
the lateral boundaries within the fluid (aU / az = 0, D1T / az = O. TV = 0) ; at the
upper limit of the domain (y=II), the first derivati"e with respect to y of all the
velocity components was equal to zero.
Three configurations were examined:
- Configuration A : h/5=0.:300 , h/1I=0.290 ;
- Configuration 13 : h/s=0.:360 , h/H=0.OS.5 ;
- Configuration C : h/5=0.0:36 . h/H=0.009 .
Once tlie convergence was reached, the computed streamwise pressure gradient
\\'as compared to the one corresponding to a smooth \\'all for the same flow rate and
the same chanllel cross-section. Thus. for configuratioll A. an increase of friction
drag close to 2G% was recorded. As the ratio h/lI was rather important. it \\'as
bclicH'cl thai numerical uncertainties could exist due to the small amount of grid
points used lo define the groO\e. It \\'as then decided to decrease the yaluc of this
height: channel width ratio, as performed in configurations 13 and C. Ne"ertheless,
drag increases were still obtained: 4% (resp. c::::: 0%) for h/1I=0.OS.5 (resp. 0.009).
Then, when the rib depth \\'as adapted to the half-channel width, let say h/E
less than 10 (!ci, skin friction drag increases were recorded for laminar internal flO\\s.
Though the mesh grids were not carefully defined (especially around the crest vicin-
ity). these results might be subject to caution. Nevertheless, when running the same
three-dimensional parabolic code on developing boundary layers, wilh the same type
of mesh size, one ended up with almost 4-5% skin friction drag reduction o\'er a
similar grooye geometry h/ 8=0.10 and h/s=0 ..50. Thus, because of non grid refine-
ment, the afore-mentioned interpretations have to be considered as a qualitative even
compa.rative (internal - external flow behaviour) approach and analysecl in such a
way.
This skin friction drag increase, for internal laminar flows, is consistent with
146

results obtained later by Launder and Li [13]. Their reported results concerned
manipulation of fully developed flows through plane channels with knife-edge riblets.
Different parameters were varied; for instance, h/s from 0.5 to 3.0 and h/I-I up to
0.08. The equation, they solved, is the Poisson equation which could be read in the
same co-ordillate system:

1 dP cPU fJ2U
0= - - -
p dx
+ v( - + -)
f} y 2 f} z 2
(5)

This equation could be in fact derived from the semi three-dimensional approach with
HT =0 by neglecting the convective terms in equation (3). Information concerning
computational details could be found in [13]. An interesting feature was that be-
cause of very small dimensions of the riblet, these authors applied a one-dimensional
treatment at heights greater than 2h above the rib crest. Then, this computational
economy allo"'ed Launder and Li to consider several geometry configurations. A
typical grid 111esh was 28 * 46 for the two-dimensional region, with an extra 22 lines
y=constant for the one-dimensional domain. In fact, that grid size \\"as quite compa-
ra ble to the one we used for triangular grom-es. It is unfortunate that they did not
appear to have investigated the effect of grid refinement since they themseh-es have
stated that "it is not clear that any of the studies has adopted a sufficiently fine grid
to provide a convincing demonstration that drag reduction is achieved, or. indeed is
achieva ble in laminar flow" .
Anyway, whatever geometry configuration was considered, the results did not
indicate a skin friction drag reduction compared with the case of a smooth channel.
for a constant aspect ratio his, the drag augmentation factor increases with hilI, (see
figure 10). nesults obtained at CEnT qualitatively agree with those from Launder
and Li, though the considered system of equations and numerical method were rather
different.
So, in conclusion, this preliminary numerical study, performed over micro-
grooves geometry with triangular cross-section, clearly showed up that the efficiency
of riblet appears to be different between internal and external flows. nesults, deduced
from the same numerical code with the same type of grid mesh for almost identical
aspect ratio, indicate a drag decrease (resp. increase) for developing boundary layer
flows (resp. fully developed plane channel flows).

4 Lall1inar Boundary Layer Flo,vs


U sing preliminary results obtained from three-dimensional elliptic and paTabolic
codes and assuming that the local transverse velocity, TV, is equal to zero in all
the computational domain, the system of equations to be soh-ed is derived from
equations (:3) and (il). This numerical approach has been considered in detail for
evaluating rilJlet performances in developing laminar boundary layers under uniform
conditions, "'ith V-shaped grooves at I1\IST and L-shaped ones at CEHT. further-
more, the pressure gradient is set equal to zero in the momentum equation for the
streamwise velocity component.
All through this section, f will refer to the friction force relative to either the
ribld surface (F H ) or the smooth surface (fE) : in this last case the width is s. i.e the
147

I!
III I II III
I I II II
I I I I
II III ,I I
II I I i Iii II II I I i if
y
II I III I j II III Ii
II I III II II III II
II
II I
I III
I I!
I II
III
I III
I i 1111 Ii
Iii t
I, I I I I III I I1111 I I I

11111111 I I! I , l l II I III
Iii' I I III, I I' 1I11
I! I i I I !III I II I 1111/ I
I I~ I I I I: I I I 1'1 I :.11 I
Ii I r II 11111111111111111 I I I
II I I 1'-' I I' II 1111111 I I I '
1111 t " ,II Illl! 1 ' 1 1 1 1'1
Ii 1111 I II' III I II 1111 I

,
I I II II 11111 111111

I::!: ~ : :
I,' 11111 III 111111'
II!
111111 : : I I , : : : I ~ I:
111111 I I, ;111/)11 II III i
IIII! I 1I11 I III III" 1 I I
11I11 II 11II III '1111 I I

Figure 1: Schematic repartition of grid points within the V-g7'00 ve

groove spacing. The mean skin friction coefficient is, then, obtained when dividing
that force by the considered wetted area. 1\'1oreovcr, b is the thickness of the laminar
boundary layer taken at the riblet leading edge.

4.1 V-shaped grooves


The calculat ion domain was subdivided in two parts along the :r-flow direction; the
upstream length corresponded to the smooth wall and included 25 constant steps,
and the downstream one simulated a V-groove, the crest of which was in the same
plane as the smooth wall. At each :I,-station, the grid mesh was rectangular; the steps
in the v-direction, i.e along the line perpendicular to the crest plane, and in the :;-
direction. i.e parallel to this plane, could be varied: the number of grid points was in
the range from ,11 * 19 up to 115*9.5. It should be mentioned that the most commonly
used one was 81 * 39. Along the v-direction, the step \\"as taken constant within the
riblet : it followed a geometrical progression from the plane of the crests to the upper
limit of the domain, located at 20 times the boundary layer thickness at the inlet
station where a I3lasius profile was applied. HO\ve\'er, the v-number of points within
the V. lly: was equal to (n o +1)/2, where n z was the z-number of points. The :;-step
,,'as constant (figure 1). Since the main objective was skin friction compu tatioll. the
crucial problem lied in the estimate of the normal to the wall velocity gradient at
the crest. (DU /a!))c : in fact, the latter strongly decreases when the distance to the
\\',dl. !::"Yc increases. \Ve shall define !::"Yc as the mesh thickness just above the groove
crest. ;\ systemiltic study of this quantity showed (figure 2, ",here (aC/DY)B is the
veloci ty gradient over the smooth plate in the same conditions) that it was constant
whell !::"Yc was less than 1O- 7 m ~ 7.5 10- 6 b ; b was about 13.3 1O- 3 m for an outer
velocity U c =6.7cms- 1 . It should be noted that these calculations were performed
with flow characteristics corresponding to experiments carried out in the Il\lST water
tUl1lwl, [7]. It was then decided to keep 6Yc=7.5 1O- 6 b. In these conditions, the
148

('Jlj)
\JY
C

(~~)
II

o ~
LI __________________________________________ ~ ______ ~

10- 11 10- 9
6 Y (In)
c

Figure 2: InJlllencc of the size of the first step ovcr the V-grool'c crEst (h=lmm. 0
s/II=0.:3 . s/h=L3. Ue =6.7cms- 1 . grid mesh: 81*39)

number of grid points (figure 3) did not influence the computed overall skin friction
provided that it \\'as larger enough (more than 41 * 19).
The overall wall friction was inferred from the spanwise distribution of velocity
gradient along the line connecting two adjacent crests. Indeed, it was shown., ac-
cording to references [7] and [10], that friction along that line was equal to friction
along the \'-\\'all. since momentum fluxes corresponding to the ach'ectin' terms in the
elementary volmne delimited by this line and the wall were negligiblE'. ;"Ioreo\,e1', the
velocity gradient distribution along the V-wall exhibited much larger spanwise \'aria-
tions than that related to velocity gradients along the crest line: estimating the area
beneath the curve related to the second distribution is easier for the determination
of the overi,ll wall friction, especially in the crest vicinity. Three methods were used
to perform this integration. In the first one, the trapezia method was applied to the
gradient velocity values obtained at the grid mesh points: friction was overestimated
even if the number of grid points was increased and did not depend upon this number
(figure 3). Indeed, when the grid was refined, velocity gradients per unit surface were
getting larger since they were related to smaller surfaces. This integration technique
gave drag enhancements of about 0.3% with respect to tbe smooth wall. The second
method dealt with eye's smoothing of the data (i.e. drawing a best-fit curve by eye)
and integration with a large enough number of trapezia in order to produce an in-
tegration error as much negligible as possible. About 1..5% drag reduction was then
obtained: this was corroborated using the standard Newton-Cotes method.
For the parametric study of the riblet geometry [7], the grid mesh was 81 * :39
with ~YL=7.;j 10- 6 b. Keeping either s or h constant allowed one to let varying the
aspect ratio s/h. Then, a 1% up to 2% drag reduction was achieved \\'hen 0.0:3 <h/b <
OAO for s= 1111111 , (figure il).
149

L1FF ('!)
.,


a
0 0 0
0

IOU 1000 10000


ny* nz

Figure 3: II/fluence oj lhe number oj grid points on jriction jorce 1'arialion (h=l mm.
s=1.5mm. Uc =G.7 cms- 1 , a integmtion jl'OTn grid points only. _ integratioll llsing c:rtm
points j/'Um eye's smoothing jitting)

- 4

o
o
o
"T
o o
o

sill
o ~---------------L________________L -______________~______~
o
Figure cl: Influence oj aspect ratio on jridion jorce variatian in V-gl'Oa['( ajtEl' eye's
smaathingfiiting (a h=lml7l," s=lmm, Ue =6.7cms- 1 , grid mesh,' 81*39)
150

4.2 U- or L-shaped grooves


This cross-section shape has been mainly considered at CERT because experiments,
concerning the performances of such riblet models, will be performed in low speed
conditions for 5e\'era1 aspect ratio s/h. Results recorded by Wilkinson et aI, [21],
showed up not only the efficiency of such thin-element grooves but also some striking
features as regards alterations of the sireamwise turbulence intensity profiles. So,
these data basis would help in testing different numerical codes or, more exactly, the
validity of different turbulence models. But, before undertaking such an approach. it
is necessary to adapt the code for laminaT boundaTY layer flows.
Since the ribs are symmetric, the equations were solved only for half of the U-
shaped rib in order to a\'oid unnecessary computations. The x-momentum equation
allo\\"ed one to obtain the streamwise velocity component U. through a double sweep
met hod until conyergence was reached for U. The convergence criterion \\'as such
that maxlLT.t1 - Ui~;I < 10- 6 , where n corresponds to the iteration number. Then.
F was deduced from the continuity equation.
The riblets were prescribed of zero thickness and the calculations could start
at 0.1-1m upstream of the riblet model, the length of which was O..SOm. It should be
noted that the bottom of the grooye was mounted flush with the upstream smooth
plate. At a gi \"en streamwise location, the computation domain was bounded on the
left by the wrtical line passing through the left branch of the grOO\'e (z = 0). on
the right by the symmetry line half-way between the left and right branch of the
U-shaped groo\'e (:::=s/2) and on the lower side by the smooth wall or grooye trough
(V = 0) : the upper side of the domain corresponded to a distance which is an order
of magnitucle higher than the boundary layer thickness.
In all cases presented in this paper, flat plate velocity profiles were applied at
the illitial ~talioll located at O.l1m dO\\'llstream of the origin of the larninar bound-
ary la~"er. These calculations \\'ould simulate experiments which will he performed at
CERT in subsonic wind tunnel. Thus, the external free-stream yelocity was 10 111S- 1
wi tit a boundary layer thickness 15 close to 6 1O- 3 m. Tile houndary conditions along
the computational domain were: no-slip boundary conditions applied at all rigid sur-
faces. symmelry conditions along the lateral boundaries within the fluid (aU/D.: = o.
DF/a::: = 0) and (aU/aV = o. aV/Dv = 0) at the upper limit of the domain.
As poillted out by de Saint-Victor [6], Djenidi [7] and Djenidi et al [9], the
value of the mean local skin-friction coefficient is strongly dependent upon the grid
refinement. especially close to the crest vicinity. If the estimation of the skin friction
codficient is not wry easy for V-shaped groO\"es, it is t rue that it is less difficult for
L-shaped riblcts. since mainly the Cartesian grid is orthogonal to the walls. Then, tl]('
sheilr stress (DLr / 071)"'011 is eyaluated in the middle of each grid cell which constitutes
t he "solid" wall of the calculation domain.
First pint of the work consisted in looking for a sufficiently fine grid to proyide
consistellt skin friction coefficient evolutions. For this reason, a single groO\'e was
cOllsidered 5=11=0.5111m (h/15 = 0.OS:3). Three different kinds of grid, involving the
same number of points: 10 * 20 in the z * 11 directions, were used. The differences
occurecl especially within the groove; above the crest, the outward !'::"y is kept constant
O\Tr four gricllines (!'::"y=!::,.yc!, then !::"V increases according to an expansion factor.
Typical grid lCiyouts are giwn on figure 5 where:
- grid I : the spacing between aeljacent griellines was constant in the V and::: directions,
151

5,Oe-4 5,Oe-4

5,Oe-" """'~~=~~,r
y(m)

2,Se--<1

1- - ----
t 2,Se-4 2,5e-..:j

--- ---

-.-Z(m)

2,5e-4 2,Se-4 2 ,Se - ~l

II III

Figure :s: Typical grid layouts within the groove: h=s=0,5I7lln

D.y=constallt and D.z=constant ;


- grid II : D.y=constant within the rib height and the spacing in the z direction
enlarged as one proceeded away from the groove left branch (expansion factor close
to L5) :
- grid III : in the z direction the grid \\"as identical to that of grid II : however. in
the y direction, D.y was tightened close to the bottom wall (y = 0) and to the groow
peak (y=h) to a minimum value D.Yc'
Computations were run for these three grids, Results showed that the skin
friction inside the groove was lower than the corresponding flat plate Yalue and that
the skin friction at the top of the rib was higher than the corresponding flat plate
value, Then, the evolutions of the average friction coefficient (based upon the wetted
area) were plotted versus the streamwise abscissa on figure 6 and compared to the
Blasius \'ahles, One could notice important discrepancies which are explained by the
fact that the shear stress increases rather quickly when moving towards the crest, The
i Ill'ee C1'( z )-curves. along the groove trough, did not vary too mnch witII the chosen
grid. which means that there was not a great dependency of the results to the grid
refinernenl close 10 y = O. However, a ratio of about ,5 could be deduced on the Cf
\'alue at the rib crest, according to the D.y and D.z values close 10 (y. z)=(h.O) : that
remark justifies the observed variation on the mean skin friction coefficient evolution
\'ersus the streamwise abscissa. These results clearly showed up the necessity to ()Pal
with grids snch as grid III, though it \Vas not maybe necessary to tighten the wrtical
spacing close to !J = O.
\\"orking with such grids, the influence of the number of nodes within the
groove - typically n z * ny - ha.s then been studied (figure 7) ; this llumber is di-
rectly related to the minimum values of f::.,y and f::.,z. It should be noted that there
are about SO extra grid lines y=constallt, above the rib peak when, for instance,
(ny*nz)=(l0 * 20) . Anyway, one could notice that an asymptotic state seemed to
be reached since there was a slight noticeable variation on the Cf( a; )-curve by in-
creasing ny from 60 to 80. In that case, a small skin friction reduction has been
152

2-
10 3 Cf

O;,-_'-----'~ _ _ I_ _ _ I_ _ I. ___ L ____ 1 _____ 1__1 ____ 1_ _ _ 1. ___ _


o
1_ _ 1 _____

.2 .4 .6
Figure 6: Influence of the grid shape on the streamwise evolution of the mean Cf
(h=s=O.Smm)

2-
10 3 Cf

t Alnslus
<10' no

1-

OL1
o
__ I_ _ I_ _ I _ _ _ I_ _ _ I_ _ _ I_ _ _ I _ _ _ I ____ I. _______ I. _ _ I __

.2 .4 .6
7 _~_:J1~J
Figure 7:fllfl uence of the number of node points on the st7'eamwise evolution of the
mean Cf (h=s=O.Smm, grid type III)
153

le-7 .............................................[............................................................................................ ]

[s=h =O,5m m ~~ (m) :


...............................................................................................................................r - - ' - - - - - ,
8e-8
+ 2.8 e-S
t;. 1.3 e-S
6e-8 o .. 2.8 e-6
+ o 1.3 e-6

8.6e-7
o 6.4 e-7

:: o --+---+~y c (m)
o D ..
Oe+O+--~~~~~~r--~~~~~~--~~~~~~
10- 7 10. 6 10- 5 10- 4

Figure 8: Influence of the v-step at the groove peak

recorded; moreover, after about 50 boundaTY layer thicknesses downstream the ri-
blet leading edge, this drag gain seemed to be independent of Reynolds number.
This observation justifies the fact that the value of Cf taken at the last computed
streamwise abscissa (x = 0.75m) will be retained for evaluating riblet performances.
\Vhen plotting the evolution of the average skin friction coefficient versus momen-
tum thickness Reynolds number, R() with () estimated from integration all along the
groove width, one ended up with the same amount of overall drag reduction at this
last computed streamwise abscissa. In fact, in order to judge of riblet performances,
one would have to compute the overall skin friction drag by integrating the local skin
friction over the streamwise rib surface. It should be noted that the evolution of
the skin friction coefficient just downstream of the leading edge of the riblet model
was strongly dependent upon the 6.x step; the latter was chosen very small at this
specific location and then became larger according to an expansion factor.
The smaller 6.Yc is, i.e. the smaller the mesh thickness just above the peak is,
the larger, at the groove peak, the velocity gradient per unit surface is. In fact, when
plotted the local shear force versus 6.Yc at a given streamwise location (:1' = 0.6m,
figure 8), one could notice that this force did not depend anymore upon !:::.Yc as long
as this latter "'as small enough. This minimum 6.Yc may vary with either the groove
height or the aspect ratio (s/h) ; for the considered groove, s/h=l and h=0.5111m,
(6.Yc)min ~ 1.28 10- 3 h ~ 1.10 10- 4 8, which is an order of magnitude higher than
the minimun1 value 6.Yc considered for V-shaped grooves. It should be noted that,
although both Il\IST and CERT calculations were performed with essentially the same
code, the estimations of the average skin friction coefficient were different. In these
preliminary calculations of laminar flows over L-grooves, the main purpose was in
fact to deal with grid independence rather than grid refinement. vVhen dealing with
parametric study, for h/s=0.5, h=0.25mm and ny = 80, then (6.Yc)m.;n ~ 1.76 10- 3 h.
154

-2 ~~~,.-.~~~~
. ~
Q
I
l1F
I i i
r
F(%)


0 I

0 I Q !I

i
0 I

T 0
I
-4

r
-~~n

hi 0= 0.042

! ...
hi 0= 0.083
,,
hi 0=0.170 i
i I
l
-6 I

s / h
I i~
-8
o 1 2 3 4
Figure 9: Parametric study for L-shaped grooves

\Vhen comparing the mean skin friction coefficient at the trailing edge of the riblet
model to the one obtained at the same location over a smooth plate, one could plot the
percentage of friction drag variations, 6.F /F, versus the aspect ratio s/h for different
values of b//5 (figure 9). For some values of this aspect ratio, a couple of symbols are
present if variations on the friction force have been recorded by increasing ny from 60
to 80 : COl1\'crgence might not be completely achieved but is not very far from being ...
For all the cxamined configurations, small friction drag decreases were obtained. It
seems that:
- for constant s/h. the level of reduction decreases as h/ /5 increases:
- for constant h//5. t.he level of reduction decreases as s/h increases.
Thus, the striking feature is that. weak drag reductions, or more exactly, no
drag increasc has been evidenced when manipulating a laminar boundary layer with
L-shaped grooves. though the increase of wetted area could be t.remendous in some
cases : '100S{ (res]). 200%) when s/h=0 ..5 (resp. l.0). This tendancy is completely
opposite to the one obtained, from a numerical point. of \'iew 1)" Launder and Li [1:3].
in int.ernal channel flows with the same device geometry.

5 Concluding Renlarks
Thus, according to those laminar boundary layer calculations over v- or L-shaped
groO\'es, the skin friction drag is not increased despite the important wetted area
increase; the latter can even be very large especially for L-shaped ribs. Tbe obtained
results hayc been recorded \\'ith various grids which \\'('re refined in regions of large
velocity gradients. Since it is rather difficult to compute these strong gradients with
accuracy, it is more reasonable to conclude only that there is. at least. no drag
155

1,4 ,--------rV-r-/~:' --;-1---....,-------,------,


.' i internal flows
/i,:'- - - . . , _ __
1,3 -+-----/+1+- L-groove h/s=1 1131
1 L-groove h/s=3 1131
Cf

--/.:?'I'----- ]
external flows
Cf V-groove h/s=O.50 1121
o 1,2 L-groove h/s=1.00

i .,,' j
L-groove h/s=O.50
V -groove h/s=O.67
V-groove h/s=O.40
1,1

: I
I
1,0 ..f.....~-n;::.:""'""'~=6+1~~.~~=.-..+~~==~.y.....=D........=-+==!::,.r==.........,

-+ h/~ = h/o
0,9+-~~~-+1~~~~+-~~~-T~~~~+-~~~--j
0,00 0,05 0,10 0,15 0,20 0,25

figure 10: 1r(l1'iations of laminar skin friction coefficient

reduction. i.e. that riblets have practically no effect on friction, in the laminar regime.
It is beyond doubt that friction enhancements, similar to those computed in a laminar
channel flow, [13], could not be obtained. This is in agreement with experimental
results on V-groo\'es, [7]. The results collected on figure 10 for either L- or V-shaped
riblcts in both illternal and external flows support these conclusions. It should be
ITlentionecl that first results reported by de Saint- Victor, [6], are not plotted on that
graph. Although those indicated a trend similar to the U1IIST and Il\IST ones, the
results might be subject to caution because the mesh grids were not carefully defined
in the regions of high \'elocity gradients, i.e. in the vicinity of the l'iblet edge.
Results reported from 1:\ISl' and CERT computations have shown the numerical
difficult ies stemmed from the singularity at this groo\'e crest, whate\'er riblet-shape si
concerned. It should be noted that additional work is under consideration at Ul\IIST
and Camhriclge as regards greater refinement tests (orthogonal mesh, conformal map
ping ... ,) for various groove cross sections.
In addition, we belie\'e that the analogy, propounded by Launder and Li. [13].
cannot. in fact. demonstrate there is an obvious drag increase to support their con
tent ion that the drag must be increased in fully developed internal flows. This anal
ogy considers "the problem of heat conduction in an infinite plate of uniform thermal
condudi vity "'ithin which a uniform heat generation rate is occuring. The surface
of the plate is maintained at a uniform known temperature. In tbe steady state, ... ,
two situations have to be compared: in one, the plate surface is completely smooth
while, in the other, miniature slits are cut in the surface which are also kept at the
156

uniform cold surface temperature. Thus, for h/ s=l, the effective surface area is three
times that of the plate without slits. Suppose the same internal heat.-generation rate
applies in each case. In which of them will the average temperature be higher? "
Launder and Li, [13], thus argue that it will obviously be the case of the smooth
plate, since in the slit case, the cooling surface is increased. However, for h/s=3
(resp. 1), the cooling surface is larger by 600% (resp. 200%), whereas calculations
reveal 32% (resp. 27%) heat flux increase, when, for instance, h/H=0.05 (figure 10).
Thus. the heat flux increase is much weaker than the wetted area increase and. there
is no straight relationship between cooling and wetted area. It is, a priori, rather
difficult to know temperature or velocity gradient distributions and, even if it is not
easy to imagine there is maybe a small decrease in wall heat transfer or friction, it
is not obvious - and, certainly, not demonstrated - that there is an increase. In fact,
there is a large region where gradients are negligible, [10], and their distribution is
such that enlargement in the crest vicinity is practically balanced by strong damping
in the valley. III addition to that, experiments performed by Rohr et al [16] in a fully
de\'eloped pipe flow suggested that riblets (V-shaped grooves, s/h=l, h/H=O.Oll and
0.02-1 where II denotes the pipe radius) "while greatly increasing the wetted surface
area, redistribute wall shear stress to provide both: no drag increase in laminar flow
and drag reduction in certain regimes of turbulent flow".
Thus, because of the velocity gradient redistribution within V- or L-shapecl
riblets (which may be also true for other geometries), the wetted area increase does
not induce extra skin friction in laminar boundary layer; in fully developed internal
laminar flows, measurements and computations give conflicting results. In turbulent,
as well as laminar boundary layers, the increase of wetted area does not have to
he considered. However, one should not forget the basic difference between fully
developed internal flows and boundary layer flows : there is no acl\'ection in the
former ones. This feature could be essential in laminar regime since the skin friction
coefficient represents an integral effect and not a local one. Looking at optimal
geometries one could, then, imagine that the wall viscosity effect is to replace the solid
wall by a fluicl wall located in the crest plane which would induce the same friction
but would smooth turbulent perturbations; there would then be a rearrangement of
wall struct ures which would tend to lower turbulence production and, consequently,
overall friction.

Acknowledgements
These numerical studies performed, both, at Il\IST and ONERA/CERT have been
supported by the "Service Technique des Programmes Aeronautiques".

References
[1] Bacher E.V., Smith C.R. : AIAA Journal, Vol. 24, W8, 1986, pp. 1382-1385.
[2] Bechert D.W., Bartenwerfer 1\1. : J. Fluid Mech., Vol. 206, 1989, pp. 105-130.
[:3] Bushnell D.M. : AGARD Rep. 723, 1985, pp. 5.1-5.26.
[4] Choi 1(.S. : J. Fluid Mech., Vol. 208, 1989, pp. 417-458.
[5] Coustols E. : AIAA Paper 89-0963 (1989).
[6] de SainL-Victor X. : CERT Internal Technical Report (1987).
[7] Djenidi L. : Ph.D. Dissertation (1989).
157

[S] Djenidi 1., Liandrat J., Anselmet F., Fulachier L. : 2nd European Turb. Conf.
Berlin, (19SS).
[9] Djenidi 1., Liandrat J., Anselmet F., Fulachier L. : Applied Scientific Research,
Vol. 46, 19S9, pp. 263-270.
[10] Djenidi L., Liandrat J., Anselmet F., Fulachier L. : Drag Reduction in Fluid
Flows. 4th lnt. ConL on Turbulent Drag Reduction, Davos, 19S9, pp. 3.5-4l.
[11] Fulachier L., Djenidi L., Anselmet F. : LM.S.T Report (19S7).
[12] Khan M.M.S. : AlAA Paper S6-1127 (19S6).
[13] Launder B.E., Li S. : Applied Scientific Research, Vol. 46, 19S9, pp. 271-2S0.
[14] Lowson M.V., Bates J.H.T., Jewson A.R. : Drag Reduction in Fluid Flows, 4th
Int. Conf. on Turbulent Drag Reduction, Davos, 19S9, pp.77-S.3.
[15] Pulles C.J.A., Prasad K.K.,Nieuwstadt F.T.M. : Applied Scientific Research,
Vol. 46, 19S9, pp. 197-20S.
[16] Rohr J., Anderson G.W., Reidy 1.W. : Drag Reduction in Fluid Flows. 4th Int.
ConL on Turbulent Drag Reduction, Davos, 19S9, pp.263-270.
[17] Schneider G.E., Zedan }.II. : Numerical Heat Transfer, Vol. 4, 19S1.
[IS] Van DOOl'maal J.P., Raithby G.D. : Numerical Heat Transfer, Vol. 7, 19S4.
[19] Walsh l\U., Lindeman A.M. : AIAA Paper S4-0347, (19S4).
[20] Walsh "\l.J., Anders Jr. J.B. : Applied Scientific Research, Vol. 46, 19S9, pp.
255-262.
[21] Wilkinson S.P., Anders J.B., Lazos B.S., Bushnell D.M. : Int. Conf. on Turbulent
Drag l1cduction by passive means, London, 19S7.
TURBULENT BOUNDARY LAYER OVER A RIBLETED
SURFACE WITH TANDEM MANIPULATORS
USING SURF ACE DRAG BALANCES

V.D. Nguyen, J. Dickinson, Y. Jean


Y. Chalifour, A. SmaiIi, A. Page, F. Paquet

Department of Mechanical Engineering


Laval University
Quebec - G lK 7P4

ABSTRACT

Low speed wind tunnel tests have been carried out to assess the combined drag
reduction effects of longitudinally grooved or "ribleted" surfaces and upstream tandem
turbulence manipulators or LEBUs (Large Eddy Break-Up devices). Skin friction and
surface drag measurements are presented covering the regions upstream, over and
downstream of the ribleted plate, both with and without LEBUs. Surface drag surveys
using servo-controlled, floating element balances indicated non-linear drag reduction
effects when the two devices are used together. On their own the asymmetrically shaped
triangular riblets produced an average drag reduction of 5% over the ribleted section
based on local direct balance measurements.

NOMENCLATURE

Cf local wall drag coefficient: 'w/(l!2pUe2)


Cfo local skin friction coefficient over smooth plate
F force registered by the balance
h height of LEBU from the wall, or riblet height
Re Reynolds number: Uee/u
s riblet "pitch" or transverse spacing
S surface area of the floating head of the balance
Ue local free stream velocity
U, friction velocity ('w/p)1/2
x streamwise distance from beginning of test section
x~ streamwise distance downstream from trailing edge of last LEBU blade

159
160

local boundary layer thickness


boundary layer thickness at Xs
= 0: i.e. end of LEBU
normalized distance downstream ofLEBUs: xSJ(\
displacement thickness
momentum thickness
1) kinematic viscosity
p density
average wall "shear" stress: F/S

1. INTRODUCTION

The energy crisis of the early 1970's stimulated a great deal of research and
development work on the aerodynamic drag reduction of both the wings and fuselages of
aircraft. Two particularly innovative techniques were successfully pioneered by NASA
Langley and associated groups; i.e. "riblets", consisting of suitably contoured
longitudinally grooved surfaces, and "Large Eddy Break-Up devices" or LEBUs, formed
by transverse bands or aerofoils mounted parallel to the surface inside the turbulent
boundary layer (cf. [1],[2]). Recent international meetings have high lighted the world-
wide research effort in these two areas ([3],[4],[5],[6]), both from the applied and the
more fundamental phenomenological point of view. There now seems to be some
agreement that drag reductions of up to 6-8% are possible with riblets, at least under
laboratory conditions. Controversy still exists however ([2],[7],[8]) as to whether one
can actually achieve overall drag reductions with LEBUs, taking into account the drag
penalty of the device itself and its support.

Despite extensive efforts to explain the mechanisms by which the riblets and the
LEBU s work, detailed descriptions are still subject to much debate. An excellent analysis
and documentation of the turbulent characteristics of the boundary layer downstream of
the LEBU has been given recently by Lemay in his Ph.D. dissertation [12].

In the present paper we have not attempted to address the detailed aspects of how
the turbulent structure is modified by riblets or LEBUs, but have restricted our efforts to
monitoring local changes in skin friction and surface drag arising from the replacement of
a smooth surface by a riblet section, both with and without the presence of upstream
LEBUs.

The series of tests described here cover the following configurations:

undisturbed flow without the use of any drag reduction device;


tandem LEBUs at three different heights inside the boundary layer;
ribleted plate with an asymmetrical saw-tooth geometry;
combinations of the previous LEBU and riblet devices.

Our particular choice of both LEBU sand rib lets was guided by experimental
work, in both the U.S.A. and Europe, aimed at determining the best configurations to
161

achieve maximum overall drag reductions. Much of this work has been well summarized
in various papers presented very recently at the Fourth International Conference on Drag
Reduction held in Davos [6]. It has been suggested that tandem LEBU devices offer the
best possibility of a net overall drag reduction because of reduced parasitic drag due to the
downstream blade lying in the wake of the upstream manipulator. In low speed flows it
appears that the best position for the LEBU s is in the outer part of the boundary layer (hlo
'" 0.8), whilst in the subsonic, compressible regime (M '" 0.8) they should be placed
nearer the centre (h/o '" 0.5).

In the case of riblets, early work at NASA Langley demonstrated that the key
parameters were the height (h) and spacing (s) Reynolds numbers h+ and s+, based on
the local friction velocity U t (i.e. h+ = U'thlu and s+ = U'ts/u). The general consensus
is that small but significant reductions in smooth wall skin friction are possible with the
appropriate scaling of these two parameters (e.g. h+ < 10 and s+ < 30 : cf. figure 10).
We thus chose a riblet height of 0.2 mm, with a pitch of 0.5 mm, corresponding to
"optimized" h+ and s+ values of around 8 and and 20 for our mid-speed tunnel
conditions.

One of the principal objectives of our investigation was to provide direct local
measurements of both surface drag and skin friction, using a series of self-nulling, servo-
controlled, floating element balances developed over a number of years in the Mechanical
and Electrical Engineering Departments at Laval University. Such measurements would
seem to be particularly appropriate in the case of "rough" or "ribleted" surfaces, where
one might expect that traditional "wall law" techniques for determining surface drag
would be adversely influenced by uncertainties regarding the "effective origin" of the
"equivalent smooth surface".

Originally we had envisaged using ribleted adhesive tape developed by the 3M


Company for NASA. However, rather than attempting to glue the 3M tape to the the
heads of our balances, we decided to machine the test plate and the associated balance
components at the same time, to attain the best control on both geometry and transverse
alignment of the floating element and the surrounding surface.

2. EXPERIMENTAL ARRANGEMENTS

All the tests were conducted in the Laval 61 x 46 x 480 cm low speed wind tunnel
which has been extensively used in a number of previous drag reduction studies
([13],[14],[15]). Figure 1 illustrates this wind tunnel which possesses an IBM-PC
controlled automatic traversing mechanism, data acquisition system and speed control,
described in the previous references. The data presented in the previous paper and also at
the 1st Canadian Symposium on Aerodynamics ([19]), covers a first series of local skin
friction and surface drag measurements, involving both riblets and LEBU s. The
measurements were made in a boundary layer developing along the floor of the tunnel
without the use of any tripping device.
162

2.1 The test section

The tests covered a 2.5 metre section downstream of the trailing edge of the rear
LEBU blade (cf. figure 2), corresponding to approximately one hundred boundary layer
thicknesses (~ = x~/Di = 100: 0i = 21.5 mm: 0i being the boundary layer thickness at x~
= 0). The leading edge of the ribleted plate, 89 cm in length (~ = 41.4), was situated
some 45.5 cm (~ = 21.2) downstream of the tandem LEBU. The riblets were preceeded
and followed by smooth plates, with the riblet peaks mounted flush with the upstream
and downsteam smooth sections. Both the smooth and ribleted plates were equipped
with removeable plugs permitting the insertion of the skin friction and surface drag
balances. Initial measurements over an instrumented smooth plate at the riblet test section
provided the comparative smooth wall data. The main tests were carried out with a free
stream velocity of 16 metres/sec at the entrance of the working section.

2.2 The LEBUs

The LEBUs used in this experiment were a pair of thin steel blades in tandem,
each with a thickness of 0.15 mm (0.0070i), a chord length of 25 mm (1.l60i) and a
spacing of l30mm (6.050i: cf. figure 3). They were placed at the same position in the
tunnel as used by Lemay et al. [14], who employed manipulators with a chord length of
19 mm and a thickness of 0.4 mm. We retained the same spacing and heights of the
manipulators in the boundary layer (h/Oi = 0.32, 0.48 and 0.75) as in Lemay's
experiments (ref. [12]).

2.3 The riblets

Our original intention had been to machine a symmetrical v-shaped riblet pattern
on an aluminium plate. However, a misalignment of the cutting tool resulted in the
asymmetrical saw-tooth wave shape illustrated in figure 4, which we decided to test. The
base angles of the riblets were 60 and 30, with height (h) and spacings (s) respectively
of 0.2 mm and 0.5 mm. As mentioned earlier, the latter dimensions were chosen to
provide a range of spacing and height parameters susceptible to produce the best possible
drag reductions over the ribleted surface based on work by NASA Langley and other
investigators (i.e. s+ = D'Cs/u < 30 and h+ = D'Ch/u < 10: [9]), and compatible with our
range of tunnel speeds (8 < De < 24 rn/s). The plate possessed five inserts for mounting
the floating element drag balance. The ribleted plate, the four removeable plugs, the
floating element and the surrounding surface of the balance casing (cf. figure 6) were all
machined at the same time, to achieve an identical surface geometry and good "mountain-
valley" matching between the various elements.

2.4 The surface drag balances

The standard twin-floating head balance illustrated in figure 5 was used for the
smooth wall measurements, whilst two special single headed versions were built for the
ribleted, "rough" wall tests (cf. figure 6). In the first series of tests, described here, the
163

larger square floating head (25 x 25 mm) balance was employed. The basic principle
behind all the balances is the same. The floating elements (#11 and 19: figure 5) are
mounted on four flexures (#17), which are very stiff in the transverse direction, but free
to move in the longitudinal direction aligned with the skin friction, or surface drag forces
to be measured. Under the action of perturbing forces the floating head starts to move.
This movement is immediately detected by a linear voltage differential transformer
(LVDT:#13), which provides the error signal to drive a small linear motor (#12) forming
part of a closed loop, self-nulling, servo system, with the current in the motor coil
directly proportional to the perturbation force acting on the floating head. Due to the very
small weight of the floating elements (e.g. 1.5 gm for the standard 28.6 mm dia. smooth
head), the balances can be calibrated directly by mounting them in the vertical position,
attaching weights (10, 50, 100, 500 mg etc) and reading the output voltage across a
precision resistor in series with the motor coil. The calibration factor is adjusted to give a
1 millivolt output for 1 mg force. All the balances possess a sensitivity of better than 1
milligram, with a full scale range of over 5 gf. Fuller details can be found in reference
[13] and by referring to the assembly drawings and corresponding nomenclature
illlustrated in figure 5.

3. RESUL TS AND DISCUSSIONS


3.1 Smooth surface alone

To ensure that conditions had not changed since previous experiments,


particularly since we recently replaced the filters at the tunnel entrance after many years'
use, the skin friction over the smooth surface was first measured using both Preston
tubes and a smooth floating element balance. The results presented in figure 7 illustrate
excellent repeatability compared with the previous experiments [4]. This reassured us
that the flow conditions had probably remained unchanged and that the smooth balance
was working well and providing consistent measurements with the Preston tube data.
The flow parameters 8 ,Re and Ue are also indicated in figure 7 at the various measuring
stations (cf. figure 2), for the case of the smooth surface.

3.2 Riblets alone

The distribution of Cf with the riblets alone is shown in figure 8, where the
surface drag and skin friction data has been normalized with respect to Cfo, the skin
friction coefficient over the smooth plate alone (cf. figure 7). A nearly constant reduction
in Cf of around 5% ( 2%) was registered over the ribleted section, with relaxation to the
smooth wall data some 25-30 boundary layer thicknesses downstream.

Net drag reductions were calculated using momentum thicknesses at the


beginning (index 0) and the end (index 1) of the ribleted and smooth plate under the same
flow conditions using the expression:

~Cd = 81 rib let - 81 smooth


Cct 8 1 smooth - 80 smooth
164

employed by a number of workers. The results illustrated in figure 9, corresponding to


different values of the wall units h+, indicated both very large scatter and optimistically
high drag reductions compared with those suggested by both the present balance
measurements (figure 8) and low-velocity drag data for rib lets obtained by other authors
(figure 10). The 5% average drag reduction over our ribleted plate, at a free stream
velocity U e of 16 mis, is slightly larger than the values indicated in figure 10 at the
corresponding value of s+ equal to 20.

3.3 LEBUs alone

A resume of Cf data behind the tandem LEBU s alone is presented in figure 11


with the LEBU s positioned at three different heights in the boundary layer (i.e. h/Oi =
0.32,0.48 and 0.75). Results obtained by Lemay et al. [14] from previous experiments
are also shown. Despite the differences in chord length and thickness of the LEBUs
between the two experiments, under the same nominal flow conditions, good agreement
has been obtained, which again tended to assure us concerning the repeatability of the
flow characteristics and the balance measurements. One point of interest to note is the
location of maximum local drag reduction. This point moves downstream as the height is
increased and as demonstrated in Lemay's thesis [12] corresponds to the region where
the wake behind the LEBU hits the wall. In our present investigation this occurs ahead
of the ribleted plate in the cases of h/Oi = 0.32 and 0.48 and close to the plate for h/Oi =
0.75 (cf. figure 11). The question arises whether it is beneficial or not to have this
phenomenon take place outside the riblet plate?

3.4 Riblets and LEBUs combined

Results of the riblets and LEBU s combined, compared with both the rib lets and
LEBU alone are also illustrated in figure 11 respectively for the three different heights of
the LEBU. The data does not seem to indicate a simple linear additive effect. Ahead of
the ribleted plate the combined effect seems to be worse, whilst downstream the drag
reduction is improved, with an apparently much longer relaxation length. These effects
do seem rather suspicious. Behind the riblet surface, where the effects of the riblets on
their own die out fairly quickly, one might expect the "combined" drag reduction to
picture rather closely that of the LEBU alone. Further tests are underway to investigate
this apparent anomaly. Tests will be repeated over a range of tunnel speeds with the two
drag balances illustrated in figure 6 and their smooth headed counterparts.

4. CONCLUSIONS

-Direct local measurements of surface drag with a servo-controlled floating


element balance indicated a fairly constant, average, drag reduction of approximately 5%
over the complete ribleted surface. This figure is quite consistent with the trend obtained
by other investigators at comparable wall parameters (h+ = 8, s+ = 20: cf. figure 10).
165

-The standard method of determining the overall drag reduction of the plate from
momentum thickness measurements at the leading and trailing edges of the ribleted plate
produced very scattered data, with rather optimistic drag reductions compared with the
balance data. We tend to place more confidence in the direct surface drag measurements,
than the momentum thickness estimates which could be adversely affected by similar
factors to those influencing the Von Karman momentum equation technique.

-The combined effects of the LEBUs and rib lets were roughly additive over the
region of the ribleted plate, but the data indicated a much longer downstream relaxation
length than one might expect from the single manipulator data. Further experiments are
presently underway to determine whether this is actually a true picture. In particular,
besides repeating the skin friction and surface drag measurements, we would like to
monitor carefully variations, or changes in the "two-dimensionality" of the flow over the
length of the tunnel.

REFERENCES
[1] Bushnell, D.M .. "Turbulent drag reduction for external flows", AGARD-Report
No 723 on Aircraft Drag Prediction and Reduction, 1985.

[2] Walsh, M.J., Anders, I.B., "RibletlLEBU research at NASA Langley", Applied
Scientific Research, Vol. 46, No.3, July 1989.

[3] Savill, A.M., Truong, T.V., Ryhming, I.L., "Turbulent drag reduction by
passive means: a review and report on the first European drag reduction meeting",
Journal of Theoretical and Applied Mechanics, Vol. 7, No.4, 1988.

[4] Proceedings of the International Conference on "Turbulent Drag Reduction by


Passive Means", London, RAeS, 1987.

[5] Coustols, E., Savill, A.M., "Resume of important results presented at the Third
Turbulent Drag Reduction Working Party", Applied Scientific Research, Vol. 46, No.3,
July 1989.

[6] Sellin, R.H.J., Moses, R. T. (Editors), "Drag reduction in fluid flows :


techniques for friction control", Ellis Horwood Ltd., Collected papers presented at the
Fourth International Conference on Drag Reduction held at Davos, Switzerland, July
1989.

[7] Anders, J.B., "Large-Eddy breakup devices as low Reynolds number airfoils",
SAE Paper No. 86-1769, presented at the Aerospace Technology Conference and
Exposition, Long Beach (CA), October 1986.
[8] Anders, I.B., "LEBU drag reduction in high Reynolds number boundary layers",
AIAA Paper No. 89-1011, AIAA 2nd Shear Flow Conference, Tempe (AZ), March
1989.
166

[9] Walsh, M.J., Lindemann, A.M., "Optimisation and application of riblets for
turbulent drag reduction", AIAA Paper No. 84-0347, 1984.

[10] Savill, A.M., Koury, E., Sebastian, W., Squire, L.c., "Combined effect of
manipulators and riblets on turbulent boundary layer skin friction", Fourth International
Conference on Drag Reduction, Davos, Switzerland, July 1989.

[11] Coustols, E., Cousteix, J., "Experimental investigation of turbulent boundary


layers manipulated with internal devices: riblets", 2nd IUTAM Symposium, Zi.irich,
July 1989.

[12] Lemay, J., "Etude experimentale du comportement de la turbulence dans une


couche limite incompressible en presence d'un manipulateur externe", Ph.D. Thesis,
Laval University, Quebec (Canada), May 1989.

[13] Nguyen, V.D., Dickinson, J., Jean, Y., Chalifour, Y., Anderson, J., Lemay, J.,
Haeberle, D., Larose, G., "Some experimental observations of the law of the wall behind
large-eddy-breakup devices using servo-controlled skin friction balances", AIAA Paper
No. 84-0346, January 1984.

[14] Lemay, J., Provenr;al, D., Gourdeau, R., Nguyen, V.D., Dickinson, J., "More
detailed measurements behind turbulence manipulators including tandem devices using
servo-controlled balances", AIAA Paper No. 85-0521, March 1985.

[15] Pineau, F., Nguyen, V.D., Dickinson, J., Belanger, J., "Study of flow over a
rough surface with passive boundary layer manipulators and direct wall drag
measurements", AIAA Paper No. 87-0357, 1987.

[16] Beauchamp, C.H., Philips, R.B., "Riblet and polymer drag reduction on a
axisymmetric body", presented at the Symposium on Hydrodynamic Performance
Enhancement for Marine Applications, Newport Rhode Island, Oct. 1988.

[17] Reidy, L.W., Anderson, G.W., "Drag reduction for external and internal
boundary layers using riblets and polymers", AIAA Paper 88-0138, Jan. 1988.

[18] Coustols, E., "Behavior of internal manipulators: "riblets" models in subsonic


and transonic flows", AIAA Paper 89-0963, March 1989.

[19] Nguyen, V.D., Dickinson, J., Jean, Y., Chalifour, Y., Smaili, A., Page, A.,
Paquet, F., "Boundary layer over a ribleted surface with tandem turbulence manipulators
using surface drag balances", Proceedings, 1st Canadian Symposium on Aerodynamics,
Ottawa, December 1989
167

Figure 1: The Laval 2'x 1 1/2' low speed wind tunnel: dimensions in mm.

2270 mm - - - - - - I...

Figure 2: Position of the LEBU s, ribleted plate and various test stations.

0.15 mm (0.007 oil

t
~ -----lh'-
130mm
(6.050i)

\\\\\\\\\\\\\\\\\\\\\\\\\\\~\\\\\

Figure 3: Configuration and dimensions (mm) of the tandem LEBUs:


hloi = 0.32, 0.48 and 0.75

Figure 4: Approximate shape and dimensions of the ribleted surface.


168

\_-----ill

'---------{.y

@-----

-1- External balance housinQ -10- Balance cover


-2- Internal mountinq block -11- Upper floatinq head
-3- Vertical alignment screws -12- Linear motor(s)
-4- Holdinq rinq -13- Position detectors: L VDT's
-5- Connector cover screws -14- ClampinQ screws
-6- Connector cover -15- Screws for cover
-7- Horizontal aliqnment screws -16- Blank holes for removinq
-8- O-rinQ between internal cover
mounting block and external -17- Flexure(s)
housinQ -18- LVDT zeroinq screw
-9- O-ring between top cover -19- Lower floatinq head
and external housinq

Figure 5: Assembly drawings of the standard, twin-headed, differential skin friction


balance for smooth wall measurements: diameter of floating heads 28.6mm (1 1/8 ").
169

Figure 6: Self-nulling, servo controlled drag balances for riblet investigation with their
associated electronic control unit: present data obtained with the larger (25x25mm),
square floating head.
170

3.5.,.-------------------,

lJ. prest. 1
o prest.2
+ bal.
o
o
~ 3.0
)(

o
U

2.5 - t - - - . . - - - - . - - - - r - - - - , , - - - - - . - - - - 1
o 50 100 150
J : x/deltal
.t

Station ~ t/de~ai : J De~a: 0 Rtheta: He Ue


- mm - m/s
0 7.6 22.5 2406 16.6
1 25.1 24.5 3266 16.6
2 33.5 26.5 3470 16.8
3 41.9 28.5 3733 16.8
4 50.2 30.5 4007 16.8
5 58.6 34.5 4156 16.9
6 80.5 38.5 4761 17
7 88.8 42.5 4933 16.8
8 97.2 42.5 5392 17.1
9 105.6 44.5 5617 17.2
10 114 48.5 6029 17.3

Figure 7: Skin friction variation, Cf/Cfo vs x/Oi ' along the completely smooth test
./
section: 0i =21.5 mm, the boundary layer thickness at the end of the last LEBU blade:
L'l. Preston tube: 0 Preston tube (Lemay [4]): + Balance data.
Table: Variation of 0, Re and Ue versus x/oi :.:: J'.
J
171

I. i l
Riblets


Riblets

.
1.00


0.90
,.
I I
,

o 50 100 150

5: ..~deltai
Figure 8: Local drag coefficient Cf along test section compared with smooth wall data
Cfo: balance data.

..
10

f:. exp.1

.
/:::,. exp. 2
- 0

.. . i .. " ~! .
~ /:::,.
><
~ .. /:::,. f:.
U
<1. 10 f:.
/:::,.
f:.
/6 IS ,sT ZO
20 4S'
4 6 10
h+

Figure 9: Overall drag reduction, i1Cd ICd vs h+, on the ribleted plate based on
momentum thickness analysis (section 3.2)

1.10 o Beauchamp et 01 [16]

~~
(> Reidy et 01 [17]
!::, Liu et 01
1.05 Coustols [18]

.2
(.)

B
1.00

0.95
t~t~ Low-speed film dolo
4% bond
0.90 0
10 20 30
s+

Figure 10: Low speed drag data for riblets: M.J. Walsh & J.B. Anders; [2] p. 256
172

1.1
A Riblets

I.,
ALP.32

+
0

I".! .-;1
1.0 .A
Riblets 06~R ALB.32
.~ ALR.32
+
,g 0.9
b 0 AI + +
( a)
u ~ + + +1
t I + I
+
I
0.8

0
I
0.7
0 50
.s :x1deltai
.t
100 150

A
1.1
Riblets
.. Riblets., ALP.48
AI
0
A6A
1.0
OA ,I I'.i~ + ALB.48
ALR.48
,g IA ~!~AI + (b)
g 0.9
+ +
u Ii I ++ +
0.8 If 1+ + + + +1

0.7
0 50 100 150
S' lfIdeltai
~

1.1
A Riblets
0 ALP.75
.. Riblets
i "1 AAA + ALB.75
1.0 .A I I ALR.75

c3 0 I AA AOI ~~8 (c)


t 0.9
+ IA 00 1 1
I : I
+ +
+

I. Gl + +1
0.8 -I-_~=;;;:;;;;;;;;;L---r_ _.--_...--~
o
t: ~deltai
50 100 150

Figure 11: Cf/Cfo vs fBi measurements: A riblets: LEBU alone: + LEBU plus
riblets: 0 LEBU alone (Preston tube) (a) h/oi = 0.32 (b) h/oi = 0.48 (c) h/oi = 0.75
List of referees

F. Anselmel (IMST - Marseille - FRANCE)


D. Arnal (ONERAj CERT - Toulouse - FRANCE)
A. Bertelrud (FFA Aeronautical Reseal'ch Institute, Bromma, SWEDEN -
High Technology Corporation, Hampton, VA, USA)
J.P. Bonnet (CEAT - Poitiers - FRANCE)
D.t.L Bushnell (NASA Langley - Hampton, VA - USA)
KoS. Choi (BMT - Middlesex - UK)
D.G. Clark (Queen Mary College - London - me)
J. Cousteix (ONERAjCERT - Toulouse - FRANCE)
E. Coustols (ONERAj CERT - Toulouse - FRANCE)
L. Djenidi (University of Cambridge - Cambridge - UK)
R. Friedrich (Technische Univel'sitiit Miinchen - Miinchen - RFA)
L. Fulachier (IMST - Marseille - FRANCE)
1. Gaudet (RAE - Bedford - UK)
S. Gavrilakis (EPFL - Lausanne - SWITZERLAND)
H. Klein (Tcchnische Universitiit Miinchen - Miincl~en - RFA)
B.E. Launder (UMIST - Manchester - UK)
J. Lemay (Laval University - Quebec - CANADA)
J. Liandrat (IMST - Marseille - FRANCE)
T. ~Iarkham (BAE Filton - Bristol - UK)
H. Leijdens (Tcchnical University of Delft - Delft - THE NETHERLkVDS)
V.D. Nguyen (Laval University, Laval, QUEBEC - National Research Council,
Ottawa, ONTARIO - CANADA)
K.K. Prasad (Technical University of Eindhoven - Eindhoven - THE
NETHERLANDS)
A.l'.I. Savill (University of Cambridge - Cambridge - UK)
V. Schmitt (ONERAj Chatillon - Paris - FRANCE)
L.C. Squire (University of Cambridge - Cambridge - UK)
C. Tenaud (CEAT - Poitiel's - FRANCE)
T.V. Truong (EPFL - Lausanne - STYITZERLAND)
t.U. Walsh (NASA Langley - Hampton, E4 - USA)

173
Mechanics
From 1990, books on the subject of mechanics will be published under two series.
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R. Moreau
Aims and Scope of the Series
The purpose of this series is to focus on subjects in which fluid mechanics plays a fundamental
role. As well as the more traditional applications of aeronautics, hydraulics, heat and mass transfer
etc., books will be published dealing with topics which are currently in a state of rapid develop-
ment, such as turbulence, suspensions and multiphase fluids, super and hypersonic flows and
numerical modelling techniques, It is a widely held view that it is the interdisciplinary subjects that
will receive intense scientific attention, bringing them to the forefront of technological advance-
ment. Fluids have the ability to transport matter and its properties as well as transmit force,
therefore fluid mechanics is a subject that is particularly open to cross fertilisation with other
sciences and disciplines of engineering, The subject of fluid mechanics will be highly relevant in
domains such as chemical, metallurgical, biological and ecological engineering. This series is
particularly open to such new multidisciplinary domains.

1. M. Lesieur: TlIrblllence ill FllIids. 2nd rev. ed .. 1990 ISBN 0-7923-0645-7


2. :vI. Lesieur and O. Metais (eds.): TlIrblllence alld Coherent S!rIlclllres. 1990
ISBN 0-7923-0646-5
3. R. :-'Ioreau: Magnetohydrodynamics. 1990 ISBN 0-7923-0937-5
4. E. Coustols (ed.): TlIrbulence Control by Passive Means. 1990 ISBN 0-7923-1020-9

SOLID MECHANICS AND ITS APPLICATIONS


Series Editor: G.M.L. Gladwell
Aims and Scope of the Series
The fundamental questions arising in mechanics are: Why?, How?, and How mIlch? The aim of this
series is to provide lucid accounts written by authoritative researchers giving vision and insight in
answering these questions on the subject of mechanics as it relates to solids. The scope of the series
covers the entire spectrum of solid mechanics. Thus it includes the foundation of mechanics;
variational formulations; computational mechanics; statics, kinematics and dynamics of rigid and
elastic bodies; vibrations of solids and structures; dynamical systems and chaos; the theories of
elasticity. plasticity and viscoelasticity; composite materials; rods, beams, shells and membranes;
structural control and stability; soils, rocks and geomechanics; fracture; tribology; experimental
mechanics; biomechanics and machine design.

1. R.T. Haftka, Z. GUrdel and M.P. Kamal: Elements of StrIlclllral Optimi:ation. 2nd rev.ed.,
1990 ISBN 0-7923-0608-2
2. J.J. Kalker: Threedimel1siol1ai Elastic Bodies. 1990 ISBN 0-7923-0712-7

Kluwer Academic Publishers - Dordrecht / Boston / London


Mechanics
3. E.B. Magrab: Vibrations of Elastic Structural Members. 1979 ISBN 90-286-0207-0
4. R.T. Haftka and M.P. Kamat: Elements of Structural Optimization. 1985
Revised and enlarged edition see under Solid Mechanics and Its Applications. Volume 1
5. J.R. Vinson and R.L. Sierakowski: The Behavior of Structures Composed of Composite
Materials. 1986 ISBN Hb 90-247-3125-9; Pb 90-247-3578-5
6. B.E. Gatewood: Virtual Principles in Aircraft Structures. Volume 1: Analysis. 1989
ISBN 90-247-3754-0
7. B.E. Gatewood: Virtual Principles in Aircraft Structures. Volume 2: Design, Plates,
Finite Elements. 1989 ISBN 90-247-3755-9
Set (Gatewood 1 + 2) ISBN 90-247-3753-2

MECHANICS OF ELASTIC AND INELASTIC SOLIDS


Editors: S. Nemat-Nasser and G.iE. Oravas

1. G.M.L. Gladwell: Contact Problems in the Classical Theory of Elasticity. 1980


ISBN Hb 90-286-0440-5; Pb 90-286-0760-9
2. G. Wempner: Mechanics of Solids with Applications to Thin Bodies. 1981
ISBN 90-286-0880-X
3. T. Mura: Micromechanics of Defects in Solids. 2nd revised edition, 1987
ISBN 90-247-3343-X
4. R.G. Payton: Elastic Wave Propagation in Transversely Isotropic Media. 1983
ISBN 90-247-2843-6
5. S. Nemat-Nasser, H. Abe and S. Hirakawa (eds.): Hydraulic Fracturing and Geother-
mal Energy. 1983 ISBN 90-247-2855-X
6. S. Nemat-Nasser, R.J. Asaro and G.A. Hegemier (eds.): Theoretical Foundation for
Large-scale Computations of Nonlinear Material Behavior. 1984 ISBN 90-247-3092-9
7. N. Cristescu: Rock Rheology. 1988 ISBN 90-247-3660-9
8. G.I.N. Rozvany: Structural Design via Optimality Criteria. The Prager Approach to
Structural Optimization. 1989 ISBN 90-247-3613-7

MECHANICS OF SURFACE STRUCTURES


Editors: W.A. Nash and G.iE. Oravas

1. P. Seide: Small Elastic Deformations of Thin Shells. 1975 ISBN 90-286-0064-7


2. V. Pane: Theories of Elastic Plates. 1975 ISBN 90-286-0104-X
3. J.L. Nowinski: Theory ofThermoelasticity with Applications. 1978
ISBN 90-286-0457-X
4. S. Lukasiewicz: Local Loads in Plates and Shells. 1979 ISBN 90-286-0047-7
5. C. Firt: Statics, Formfinding and Dynamics of Air-supported Membrane Structures.
1983 ISBN 90-247-2672-7
6. Y. Kai-yuan (ed.): Progress in Applied Mechanics. The Chien Wei-zang Anniversary
Volume. 1987 ISBN 90-247-3249-2
7. R. Negruriu: Elastic Analysis of Slab Structures. 1987 ISBN 90-247-3367-7
8. J.R. Vinson: The Behavior of Thin Walled Structures. Beams, Plates, and Shells. 1988
ISBN Hb 90-247-3663-3; Pb 90-247-3664-1
Mechanics
MECHANICS OF FLUIDS AND TRANSPORT PROCESSES
Editors: R.I. Moreau and G.1E. Oravas

I. I. Happel and H. Brenner: Low Reynolds Number Hydrodynamics. With Special


Applications to Particular Media. 1983 ISBN Hb 90-01-37115-9; Pb 90-247-2877-0
2. S. Zahorski: Mechanics of Viscoelastic Fluids. 1982 ISBN 90-247-2687-5
3. I.A. Sparenberg: Elements of Hydrodynamics Propulsion. 1984 ISBN 90-247-2871-1
4. B.K. Shivamoggi: Theoretical Fluid Dynamics. 1984 ISBN 90-247-2999-8
5. R. Timman. A.I. Hermans and G.c. Hsiao: Water Waves and Ship Hydrodynamics. An
Introduction. 1985 ISBN 90-247-3218-2
6. M. Lesieur: Tllrbulence in Fluids. Stochastic and Numerical Modelling. 1987
ISBN 90-247-3470-3
7. L.A. Lliboutry: Very Slow Flows of Solids. Basics of Modeling in Geodynamics and
Glaciology. 1987 ISBN 90-247-3482-7
8. B.K. Shivamoggi: Introduction to Nonlinear Fluid-Plasma Waves. 1988
ISBN 90-247-3662-5
9. V. Bojarevics. Ya. Freibergs, E.I. Shilova and E.V. Shcherbinin: Electrically Indllced
Vortical Flows. 1989 ISBN 90-247-3712-5
10. I. Lielpeteris and R. Moreau (eds.): Liquid Metal Magnetohydrodynamics. 1989
ISBN 0-7923-0344-X

MECHANICS OF ELASTIC STABILITY


Editors: H. Leipholz and G.1E. Oravas

1. H. Leipholz: Theory of Elasticity. 1974 ISBN 90-286-0193-7


2. L. Librescu: Elastostatics and Kinetics of Aniosotropic and Heterogeneous Shell-type
Stmctures. 1975 ISBN 90-286-0035-3
3. c.L. Dym: Stability Theory and Its Applications to Structural Mechanics. 1974
ISBN 90-286-0094-9
4. K. Huseyin: Nonlinear Theory of Elastic Stability. 1975 ISBN 90-286-0344-1
5. H. Leipholz: Direct Variational Methods and Eigenvalue Problems ill Engineering.
1977 ISBN 90-286-0106-6
6. K. Huseyin: Vibrations alld Stability of Militiple Parameter Systems. 1978
ISBN 90-286-0136-8
7. H. Leipholz: Stability of Elastic Systems. 1980 ISBN 90-286-0050-7
8. V.V. Bolotin: Random Vibrations of Elastic Systems. 1984 ISBN 90-247-2981-5
9. D. Bushnell: Computerized Buckling Analysis of Shells. 1985 ISBN 90-247-3099-6
10. L.M. Kachanov: Introduction to Continuum Damage Mechanics. 1986
ISBN 90-247-3319-7
11. H.H.E. Leipholz and M. Abdel-Rohman: Control of Structures. 1986
ISBN 90-247-3321-9
12. H.E. Lindberg and A.L. Florence: Dynamic Pulse Bucklillg. Theory and Experiment.
1987 ISBN 90-247-3566-1
13. A. Gajewski and M. Zyczkowski: Optimal Stl"llctllral Design IInder Stability Con-
straints. 1988 ISBN 90-247-3612-9
Mechanics
MECHANICS: ANALYSIS
Editors: V.J. Mizel and G.fE. Oravas

1. M.A. Krasnoselskii, P.P. Zabreiko, E.L Pustylnik and P.E. Sbolevskii: Integral
Operators in Spaces of Sllmmable Functions. 1976 ISBN 90-286-0294-1
2. V.V. Ivanov: The Theory of Approximate Methods and Their Application to the
Numerical Solution of Singular Integral Equations. 1976 ISBN 90-286-0036-1
3. A. Kufner, O. John and S. Pucfk: Function Spaces. 1977 ISBN 90-286-0015-9
4. S.G. Mikhlin: Approximation on a Rectangular Grid. With Application to Finite
Element Methods and Other Problems. 1979 ISBN 90-286-0008-6
5. D.G.B. Edelen: Isol'ector Methods for Equations of Balance. With Programs for
Computer Assistance in Operator Calculations and an Exposition of Practical Topics of
the Exterior Calculus. 1980 ISBN 90-286-0420-0
6. R.S. Anderssen, F.R. de Hoog and M.A. Lukas (eds.): The Application and Numerical
Soillfion of Integral Equations. 1980 ISBN 90-286-0450-2
7. R.Z. Has'minskil: Stochastic Stability of Differential Equations. 1980
ISBN 90-286-0100-7
8. A.I. Vol'pert and S.L Hudjaev: Analysis in Classes of Discontinuous Functions and
Eqllations of Mathematical Physics. 1985 ISBN 90-247-3109-7
9. A. Georgescu: Hydrodynamic Stability Theory. 1985 ISBN 90-247-3120-8
10. W. Noll: Finite-dimensional Spaces. Algebra, Geometry and Analysis. Volume 1. 1987
ISBN Hb 90-247-3581-5; Pb 90-247-3582-3

MECHANICS: COMPUTATIONAL MECHANICS


Editors: M. Stem and G.fE. Oravas

I. T.A. Cruse: Boundary Element Analysis in Computational Fractllre Mechanics. 1988


ISBN 90-247-3614-5

MECHANICS: GENESIS AND METHOD


Editor: G.fE. Oravas

1. P.-M.-M. Duhem: The Evolution of Mechanics. 1980 ISBN 90-286-0688-2

MECHANICS OF CONTINUA
Editors: W.O. Williams and G.fE. Oravas

1. c.-C. Wang and C. Truesdell: Introduction to Rational Elasticity. 1973


ISBN 90-01-93710-1
2. P.l. Chen: Selected Topics ill Wave Propagation. 1976 ISBN 90-286-0515-0
3. P. Villaggio: Qualitative Methods in Elasticity. 1977 ISBN 90-286-0007-8
Mechanics
MECHANICS OF FRACTURE
Editors: G.c. Sih

1. G.C. Sih (ed.): Methods of Analysis and Solutions of Crack Problems. 1973
ISBN 90-01-79860-8
2. M.K. Kassir and G.C. Sih (eds.): Three-dimensional Crack Problems. A New Solution
of Crack Solutions in Three-dimensional Elasticity. 1975 ISBN 90-286-0414-6
3. G.c. Sih (ed.): Plates and Shells with Cracks. 1977 ISBN 90-286-0146-5
4. G.c. Sih (ed.): Elastodynamic Crack Problems. 1977 ISBN 90-286-0156-2
5. G.c. Sih (ed.): Stress Analysis of Notch Problems. Stress Solutions to a Variety of
Notch Geometries used in Engineering Design. 1978 ISBN 90-286-0166-X
6. G.c. Sih and E.P. Chen (eds.): Cracks in Composite Materials. A Compilation of Stress
Solutions for Composite System with Cracks. 1981 ISBN 90-247-2559-3
7. G.c. Sih (ed.): Experimental Evaluation of Stress Concentration and Intensity Factors.
Useful Methods and Solutions to Experimentalists in Fracture Mechanics. 1981
ISBN 90-247-2558-5

MECHANICS OF PLASTIC SOLIDS


Editors: J. Schroeder and G.1E. Oravas

1. A. Sawczuk (ed.): Foundations of Plasticity. 1973 ISBN 90-01-77570-5


2. A. Sawczuk (ed.): Problems of Plasticity. 1974 ISBN 90-286-0233-X
3. W. Szczepitiski: Introduction to the Mechanics of Plastic Forming of Metals. 1979
ISBN 90-286-0126-0
4. D.A. Gokhfeld and O.F. Chemiavsky: Limit Analysis of Structures at Thermal Cycling.
1980 ISBN 90-286-0455-3
5. N. Cristescu and 1. Suliciu: Viscoplasticity. 1982 ISBN 90-247-2777-4

Kluwer Academic Publishers - Dordrecht / Boston / London


Mechanics
From 1990, books on the subject of mechanics will be published under two series:
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R.J. Moreau
SOLID MECHANICS AND ITS APPLICATIONS
Series Editor: G.M.L. Gladwell
Prior to 1990, the books listed below were published in the respective series indicated below.

MECHANICS: DYNAMICAL SYSTEMS


Editors: L. Meirovitch and G ..tE. Oravas

1. E.H. Dowell: Aeroelasticity of Plates and Shells. 1975 ISBN 90-286-0404-9


2. D.G.B. Edelen: Lagrangian Mechanics of Nonconservative Nonholonomic Systems.
1977 ISBN 90-286-0077-9
3. J.L. Junkins: An Introduction to Optimal Estimation of Dynamical Systems. 1978
ISBN 90-286-0067-1
4. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity. Revised and enlarged edition see under Volume 11
5. L. Meirovitch: Computational Methods in Structural Dynamics. 1980
ISBN 90-286-0580-0
6. B. Skalmierski and A. Tylikowski: Stochastic Processes in Dynamics. Revised and
enlarged translation. 1982 ISBN 90-247-2686-7
7. P.C. Miiller and W.O. Schiehlen: Linear Vibrations. A Theoretical Treatment of Multi-
degree-of-freedom Vibrating Systems. 1985 ISBN 90-247-2983-1
8. Gh. Buzdugan, E. Mihililescu and M. Rade: Vibration Measurement. 1986
ISBN 90-247-3111-9
9. G.M.L. Gladwell: Inverse Problems in Vibration. 1987 ISBN 90-247-3408-8
10. G.!. Schueller and M. Shinozuka: Stochastic Methods in Structural Dynamics. 1987
ISBN 90-247-3611-0
11. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity. Second revised and enlarged edition (of Volume 4). 1989
ISBN Hb 0-7923-0062-9; Pb 0-7923-0185-4
12. W. Szempliriska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume I:
Fundamental Concepts and Methods: Applications to Single-Degree-of-Freedom
Systems. 1990 ISBN 0-7923-0368-7
13. W. Szempliriska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume II:
Advanced Concepts and Applications to Multi-Degree-of-Freedom Systems. 1990
ISBN 0-7923-0369-5
Set ISBN (Vols. 12-13) 0-7923-0370-9

MECHANICS OF STRUCTURAL SYSTEMS


Editors: J.S. Przemieniecki and G ..tE. Oravas

1. L. Fryba: Vibration of Solids and Structures under Moving Loads. 1970


ISBN 90-01-32420-2
2. K. Marguerre and K. Wolfel: Mechanics of Vibration. 1979 ISBN 90-286-0086-8

You might also like