Vous êtes sur la page 1sur 19

Colloid & Polymer Science

Colloid & Polymer Sci. 262, 319-337 (1984)

A review on drag reduction with special reference to micellar systemr


A.V. Shenoy
Polymer Scienceand EngineeringGroup, ChemicalEngineeringDivision,National ChemicalLaboratory, Pune (India)
Abstract:A state-of-the-artreviewon drag reductionhas beenpresentedin order to bring

out someimportant aspectsof the dragreductionphenomenonand its potentialfor practical use. The reviewis biasedtowards micellarsystemsand discussesin detailthe morphologicaldifferencesbetweendrag reducingpolymericand micellarsystems.Workrelatingto polymericsystemshas not beendealtin detailas it has beenthe subjectof earlier
reviews. Studiesrelatingto biologicaladditivesas well as suspensionshave been briefly
mentionedwithout detaileddiscussionas their potentialfor practical use presently appears to be limited.
Key words:Drag reduction review,micellarsystem, polymer additive,biologicaladdi-

tive, suspension,theory of drag reduction


1. Introduction

In a number of practical fluid flow situations, turbulence occurs near solid surfaces and the main resistance to the flow of the fluid is associated with this turbulence. The energy losses due to turbulent friction are
of very high magnitude and this provides enough motiv~ition for unabated research to find ways for its reduction. The 1930's witnessed one of the most exciting
discoveries of fluid mechanics, namely, the observation that turbulent skin friction can be reduced by the
presence of certain additives in the flowing fluid. Various types of additives in the flowing fluid have been
found to be effective:
a) macromolecules like those of polymers or surfactants, and
b) simple solids like fine grains or fibers.
The discovery provided a physical means of reducing the tremendous energy losses due to turbulent skin
friction in the flow of fluids through conduits or, conversely, the movement of solid bodies in fluids. The research that followed this discovery focussed attention
on the nature of turbulent skin friction generated at a
solid surface by relative flow of a fluid past it.
Most of the researchers in this field have worked
with fluids at room temperature because of the practical importance of using drag reducers in ship-building
industries, for fire-fighting operations, oil-well fractur* NCL-CommunicationNo. 3384
K 732

ing processes, etc. where high temperatures are not involved. Recently the idea of using drag reducing additives in central heating systems has been evolved and
there has been some study of the effectiveness of drag
reducers at high temperatures. An epitome of the results of the tremendous work done on drag reduction
can be obtained in a number of reviews and reports
[1-17].
The present review deals with the phenomenon of
drag reduction in micellar systems, which has been
consistently neglected in earlier reviews despite the
voluminous literature available on the subject. A brief
recapitulation of the various facets of drag reduction
right from the historical beginning to its present status
has been undertaken with a view of developing a background for understanding and appreciating the phenomenon of drag reduction in general. Specific attention
has been allotted to the morphological differences between micellar and polymeric drag reducing systems.
Most of the potential uses of drag reduction have been
briefly mentioned in order to stimulate an interest for
using micellar systems in applications other than those
in which they are presently used.
2. Review of past work
2.1 Historical

The earliest works that recorded a decrease in pressure drop during turbulent flow were undertaken in

Colloid and Polymer Science, Vol. 262 9No. 4 (1984)

320

the thirties [18-20] and concerned the transportation


of paper pulp. This was, however, not explicitly referred to as a drag reduction phenomenon. Toms [21]
was the first to recognise the tremendous reduction in
wall shear stress caused by the addition of small
amounts of linear macromolecules to a flowing fluid.
He noticed that under certain conditions of turbulent
pipe flow, dilute solutions of polymers required less
specific energy expenditure than that required for the
pure solvent. This effect of specific energy reduction is
occasionally termed the 'Toms effect'. Other names
that this phenomenon has received include the 'nonNewtonian effect' and the 'visco-elastic effect'. Quite
often it is known as the, 'Texas effect' in view of the
contributions of Texas investigators. Recently, it has
come to be known as the 'Texas-Toms effect', but even
now the least objectionable and the most suited name
for it is the 'Drag-Reducing effect'.
Toms attributed this phenomenon to a 'wail effect'
or a region near the wall where the polymer molecules
would be absent due to their bulky size. At that time
Oldroyd [22] attempted to explain the above phenomenon by suggesting that the external constraint imposed by the tube wall may introduce a locally preferred direction in a normally isotropic material.
Hence, there was a possibility of the existence of an abnormally mobile laminar sub-layer of a thickness
comparable with molecular dimensions.
During the Second World War, an effect similar to
the above was observed during the flow studies of gasoline thickened with an anionic surfactant, namely,
Aluminium Soaps. The findings of this work were,
however, first published much later [23].
After about ten years of dormancy since the first
works in drag reduction, the year 1959 witnessed the
dawn of a new era in the field of drag reduction. The
thesis work of D. W. Dodge under Prof. A. B. Metzner
[24, 25] at Delaware was then published at the same
time as the work of R. G. Shaver under Prof. E.W.
Merrill at M. I.T. [26]. Both noticed unusually low
friction factors for certain non-Newtonian solutions
like those of sodium carboxymethylcellulosein water.
At around the same time, industrial researchers
[27, 28] made similar observations with certain gums,
the most prominent being guar gum, which is a polysaccharide derived from a plant. The gums were used
to suspend sand in the sand-water mixtures utilised in
oil-well fracturing operations. The other industrial researchers, who dominated the field of drag reduction,
were Savins and Crawford. The works of Savins [29]
and Crawford [30] revived an interest among the U. S.

Navy workers to explore the potential of the drag-reducing additives for military applications. Among the
Navy researchers, it was Hoyt and Fabula [31, 32]
who made the first significant contributions. Fabula et
al. [33] were the first to identify the still most effective
drag reducer, namely, poly(ethylene oxide).
In the years that followed the above initial efforts, a
number of papers and reports appeared in various
journals for rheologists, chemical engineers, polymer
chemists, hydrodynamicists and oil technologists,
wherein the flow characteristics of drag-reducing solutions were studied for different internal and external
flow situations. In the study of internal flows, almost
every pipe-flow geometry has been considered. Apart
from straight, smooth, rough or corrugated pipes, studies have been made in curved or helicallycoiled tubes
[34-37] in orifices [38], in tube enlargements [39], in
annuli [40, 41] and in open channels [41, 42]. A great
number of theoretical and experimental studies have
been carried out for a number of external flow situations like around rotating disks [31-33, 43-63], rotating cylinders [64-74], flat plates [75-87] and spheres
[88-99].

2.2 Definition of drag reduction


A general and widely accepted definitionof drag reduction was provided by Savins [29,100]. He defined
drag reduction as the increase in pumpability of a fluid,
which is achieved on the addition of small amounts of
certain materials to the fluid when it is flowing under
turbulent conditions. In symbols, drag reduction for
pipe is given by the following expression (4):
drag reduction =

APs APa
[4P[
-

(at constant flow rate)

where APs = pressure loss due to friction in unit length of the pipe
for solvent alone,
APa = pressure loss due to friction in unit length of the pipe
for the solution containing the additive.

Savins [29] also introduced the term 'Drag Ratio' as


a means of measure of the drag reducing ability of an
additive. It was defined by him as the ratio of the pressure drop of a solution containing the additive to the
pressure drop of the pure solvent at the same flow rate:
DR-

APsolution
APsolvent

(at constant flow rate).

A solution is categorised as drag reducing if it gives a


drag ratio which is less than unity.

Shenoy, A review on drag reductionwith ~ecial referenceto micellarsystems


2.3 Types of drag reducing additives
There have been a number of types of additives
which have shown themselves to be successful drag reducers. Basically, they can be broadly classified under
the following four headings:
1 Polymers,
2 Solid-particle Suspensions,
3 Biological Additives, and
4 Surfactants.
It is generally the practice to give 'soaps' as one of
the classifications of the types of drag reducing addifives. The author prefers the term 'surfactants' as it
covers a much broader spectrum and includes not only all soaps but also many other surface-active agents
that have received recent attention as drag-reducers.

2.3.1 Drag reduction with polymer solutions


Most of the studies on drag reduction have been undertaken with solutions containing small quantities, a
few parts per million, of soluble high-molecularweight polymers in both aqueous and organic solvents, and they have been found to be quite effective.
Table 1, though not absolutely complete, serves as a
ready guide to the extensive literature available on drag
reducing polymer systems.
Table 1. Drag reducingpolymer solutions
Solute

Solvent

Guar Gum
Water
Guar Gum
SeaWater
Polyethyleneoxide Water

References

[48, 66, 89,101-106]


[107]
[49, 60, 73, 91,103,
108-123]
Polyethyleneoxide SeaWater
[32]
Polyethyleneoxide Watercontaining [I17,124,125]
an electrolyte
Polyethyleneoxide Hydrocarbons, [126,127]
e. g. benzene
trichloroethylene
Polyethyleneoxide Bloodtransfusion [128]
fluids
Polyacrylamide W a t e r
[103,108,109,119,120,
125,129-134]
Polyacrylamide Brine
[135]
Sodiumcarboxy- Water
[24-26,103,136-139]
methylcellulose
Hydroxyethyl Water
[140]
cellulose
Polyisobutylene Toluene,benzene, [26,141-144]
cyclohexane
Polyisobutylene Lightmineraloil [145]
Polyisobutylene Decalin
[146]
Polyisobutylene Crude oil,kerosene [107,147]

Solute

321
Solvent

Poly(isodecyl
Kerosene
methacrylate)
Polymethyl
Toluene
methaerylate
Polymethyl
Monochloromethacrylate
benzene
Polyvinylalcohol Water
VinylI
Water
SodiumpolystyreneWater
sulfonate
Tri-n-butylHexane
stannylfluoride
Polystyrene
Toluene
Polystyrene
Polystyrene
Dowell-APE
Potassium
polyphosphate
Flaxmeal
Certainrubbers

References
[148]
[141,142,149]
[21]
[108,137]
[29]
[57]
[150]

[118,127,141,142,145,
151-153]
Benzene,
Methyl [127]
ethylketone
Cyclohexane
[118]
Kerosene
[154]
Watercontaining [155]
sodium
pyrophosphate
SeaWater
[156]
Organicsolvent [126]

A study of the drag reducing ability of the above polymers has led t o the conclusion that any macromolecular substance with a linear structure and high molecular weight, i. e. above 500,000, is a good drag reducer. The advantage of polymers is that they are effective
at very low concentrations; but, as they are very prone
to mechanical degradation in dilute solutions [157161] they have limited practical applicability.

2.3.2 Drag reduction with solid particle suspensions


The drag reduction caused by the addition of suspended solid matter has been the subject of study
since the beginning of this century. The study was initiated by the fact that turbid streams of water were
found to flow faster than clear ones. One of the first
comprehensive experimental works was undertaken
by Vanoni [162] in the year 1946. In spite of this early
start, it was more than ten years before investigators
began systematic study of solid suspension drag reduction.
Broadly speaking, suspensions may be considered
to be of two different types, namely, a) granular or
nearly spherical particles and b) fibres. The experimental work on drag reduction with both these types
of suspensions has been reviewed by Zandi [163]. As a
quick aid to some of the available literature, the following table 2 has been provided.

322

Colloid and Polymer Science, VoL 262 9No. 4 (1984)

Table 2. Drag reducing solid suspension solutions


Suspension

FlowingF l u i d

References

Sand
Coal,FlyAsh,
Clay,Activated
Charcoal
Woodand
WoodPulp
Fibrous
WoodPulp
Emery
Thoria
NylonFibres
NylonFibres
NylonFibres

Water
Water

[164,165]
[163]

Water

[166-170]

Solutionof
GuarGum
Water
Water
Water
Polymer
Solution
Polymer
Solution
+ AerosolOT
Water
Water
PolymerSolution
AerosolOT
Solution
Theirmotherliquor

[171]

RayonFibres
AsbestosFibres
AsbestosFibres
AsbestosFibres
YellowDye
Crystals

[172]
[173,174]
[175-181]
[179]
[176]
[180,182]
[157,182,183-188]
[179]
[185]

be more effective in alkaline media. They are also


found to show long-term stability and are thus superior to poly(ethylene oxide), which was used initially
in towing tanks [198].
Sea-water slime and other fresh-water biological
growths have also been found to be effective drag reducers. Some of the effective biological additives are
given in table 3.
Table 3. Drag reducing biologicaladditives
BiologicalAdditive

References

PorphyridiumAerugineum
ProphyridiumCruentum
XanthomonasCampestris
ChaetocerosDidymus
ProtocentrumMicans
Fish Slimes

[31]
[193]
[193-195,198]
[199]
[199]
[131,199-204]

[189]

To date the effects of solids suspened in flowing liquids has not been investigated as thoroughly as that of
polymer additives, but, from the industrial point of
view, suspensions of solids have an edge over that of
polymers for the following two reasons:
1. the solid matter can be added and removed from
the flowing liquid with great ease, and
2. they suffer no mechanical degradation in most
cases except in a few as noted by some workers
[185,190,191].

2.3.3 Drag reduction with biological additives


The drag reducing ability of biological additives
came to light as a serendipity. When models of prospective new ships were tested in towing tanks to determine their drag and propulsion characteristics,
Hoyt [3!, 192] was the first to notice the drastic variation in the turbulent flow properties of water in those
tanks. He explained that the long-chain polysaccharides produced by living organisms during growth
were the cause of this drag reduction.
The drag reduction measurements made by Hoyt
and Soli [193] initiated the study of the polysaccharides of several fresh water and marine algae [194,195].
Kenis studied the effect of pH [196] and temperature
[197] on the bacterial production of drag reducing polysaccharides. He concluded that the bacteria could be
cold-lovingor heat-loving but they are often found to

2.3.4 Drag reduction with su~actant solutions


'Surfactant' is a very convenient contraction of the
term 'Surface-active agent'. It connotes an organic
molecule or an unformulated compound having surface-active properties. Three classes of surfactants, namely, anionic, cationic and nonionic, have been investigated for drag reduction.

2.3.4a Anionic su~factants as drag reducers


(i) Aqueous systems
An extensive and pioneering work on anionic surfactants as drag reducers in aqueous solutions was carried out by Savins [205-207]. He made use of alkali
metal and ammonium soaps; and obtained [205] a
drag reduction of 30 % for 0- 2 % sodium oleate in water at only 0.3 m/sec.
The addition of an electrolyte helped to increase the
drag reduction. He observed that the drag reduction
could be raised from 45 to 82 percent at a fixed wall
shear stress (= 150 dynes/cm2) by increasing the concentration of the electrolyte, namely, KC1 from 3.5 to
10 percent. He explained that the presence of KC1
helped in the enhancement of the association of the
soap molecules and that the soap micelles, which were
initially spherical in the aqueous solution [208], were
rearranged under the influence of the electrolyte into
cylindrical shapes which in turn formed a network of
interlaced rod-like elements [209]. The amount of
electrolyte normally ranged from 2 to 14 percent in order to produce stable association colloids for drag

Shenoy, A review on drag reduction with special reference to micellar systems

reduction. The soap concentrations involved were of


the order of 0-1%, which are considerably higher than
the typical polymer concentrations.
Savins [205] observed an interesting stressed controlled drag reduction effect in the soap solutions,
which is contrary to what is observed in the case of polymer solutions. He found that the percentage reduction in drag increased with an increase in (8 V/D) up to
a critical shear stress (T~,)c.However, forr~ > (~)c the
percentage reduction in drag began to decrease and at
some higher values of (8 V/D), the soap solution became virtually indistinguishable from the soap-free
electrolyte solution. According to Savins, this occurred because of a temporary disentanglement induced by turbulent vortices and eddies in fully developed turbulent flow. He noted another interesting
fact that the maximum drag reduction always occurred at the critical shear stress and the critical shear
stress essentially depended on the amount of electrolyte present. He also observed the amazing reversibility of the stress controlled drag reduction phenomenon. If the shear stress was reduced from above (Tw)~
to below (r~)~ then the bonds reformed and the drag
reducing ability of the solution was restored. This is in
contrast to polymer molecules, which, when broken,
cause the fluid to lose its drag reducing ability permenantly. The highest critical shear stress up to which effectiveness was obtained by Savins was 1000 dynes/
cm2 for a concentration of 2000 ppm of his soaps. He
also showed that the soaps were resistant to mechanical degradation. He described a test on a 2500 ppm
soap solution which initially showed 77.1 ~ drag reduction as compared to 78.2 ~ after 88 continuous
hours of circulation of the fluid by a centrifugal pump
at a wall shear stress of about 200 dynes/cm2.
Savins used the jet thrust technique and detected no
normal stress differences in the shear rate range of
2000 to 125000 sec-1, in an aqueous solution containing 0.2 0/0sodium oleate, 10 0/0KC1 and 0.6 % KOH.
He also did not observe any critical shear stress for 'onset' in the soap systems that he studied. The critical
shear stress that he makes allusion to is actually the
'threshold' shear stress beyond which the drag reduction activity of the soap solutions decreases steadily
and virtually becomes nil at some higher shear stress.
Savins found that temperature did affect the drag reduction ability of the soaps he investigated [206]. For a
solution of 1 part by weight of sodium oleate, 3 parts
by weight of sodium chloride, 0.1 part by weight of
sodium hydroxide and 95.9 parts by weight of water,
the percentage reduction in pressure drop at 30 ~ was

323

more than that at 65 ~ for lower flow rates and vice


versa for higher flow rates. Thus, Savins concluded
that it is profitable to use a lower concentration of soap
for greater efficiency if the environmenttemperature is
lower.
Undoubtedly, the utility of these soaps in industry
is where long-term stability of the drag reducing effect is of importance. These conventional soaps are relatively inexpensive and mechanically stable; yet, they
have limited applicability as the calcium and other insoluble soaps are precipitated out by the interaction
with calcium and other ions that are generally present
in tap and sea water.

(ii) Nonaqueoussystems
The earliest work on anionic surfactants in nonaqueous systems dates back to the Second World War
when, in the course of designing and testing flame
throwers, it was felt that there was a need for information on the flow characteristics of gasoline thickened
with aluminium soaps which were the constituents of
napalm. K.J. Mysels and his associates investigatedthe
various aspects ofthese thickened solutions and found
that the pressure loss per unit length of the pipe when
thickened gasoline fowed under turbulent conditions
was much lower than that of the pure solvent. The findings were published much later [23] and the effect
was attributed to a variable viscosity with shear rate of
the thickened fluid. However, it was Agoston et al.
[210], who later suggested, that there was a different
mode of flow of the thickenedjelly-like substances and
the gasoline. They observed drag reduction in a
0.3 cm diameter pipe at about 15 m/sec. They also
showed that in more concentrated solutions there was
a pronounced viscoelastic effect with long relaxation
times and that the pressure gradients peresisted for
several minutes after the pumping had been terminated. Apart from [23] and [210], an account of Mysels'
experiments [211] appeared many years later in 1971.
In 1960, Osterhout and Hall [28] showed that diesel
oil gelled with an unnamed soap at an unspecified concentration yielded a drag reduction of 40 to 50 % at
about 10 m/sec in 5.1 and 6.3 cm diameter pipes.
Many ye~trs after the initial demonstration of drag
reduction in nonaqueous anionic surfactant solutions,
a systematic study of these systems was initiated by
Radin et al. [212]. Among the first systems studied
were aluminium dioleate and aluminium palmitate
soap in toluene. Unlike the aqueous anionic surfactants, drag reduction did not occur until a concentration as high as 0-75 0/0 was reached. The viscosity at

324

this concentration was found to be two or three times


that of the solvent toluene. The viscosity was found to
drop slightly when reproducibility of data was being
checked, thus indicating a slight mechanical degradation. Measurements were made for solutions aged for
12 days and it was found [212] that the reduced viscosity decreased greatly at all concentrations for every
initial concentration. They also found that the aged solutions were not as sensitive to mechanical degradation as fresh solutions.
Contrary to the results of Savins [205], significant
values of the first normal stress differences were observed by Radin et al. [212] in the 1% solution of aluminum dioleate in toluene for the shear rate range
2000 to 30000 sec-1. Furthermore, they did not find
the critical shear stress observed by Savins. In fact,
they observed no loss in drag reduction even up to a
wall shear stress of 5000 dynes/cm2. They suggested
that this was due to the forces holding the micelles together in hydrocarbon solvents being stronger than in
aqueous solvents.
McMillan [213] and Radin [214] observed no time
or shear degradation effects in the concentrated solutions of aluminum disoaps that they worked with. McMiUan suggested an equilibrium model on the basis of
which he explained his experimental results. He observed that a permanent drag reduction in aluminum
distearate cannot be obtained unless a certain minimum concentration is reached. Above this concentration, he suggested that a highly structured association
colloid existed which is stable with respect to time and
is recovered completelyafter a shear breakdown. Below this critical concentration, a metastable structure
exists in the solution, which may readily be transformed into its dispersed equilibrium state by aging or
by high shear or by both. Lee [215,216] also observed
the same phenomenon, thus confirming the above findings.
McMillan et al. [217] noticed that the method of
preparation of the disoap solutions strongly affected
their turbulent flow characteristics. Solutions prepared by diluting more concentrated batches or prepared at temperatures of 60 ~ to 80 ~ showed higher
viscosities and were less effective drag reducers than
those prepared so as to be completely mixed. Moreover, the properties of such solutions prepared by dilution were time dependent, but the differences between such nonequilibrium states were reduced with
increased aging and/mixing.
McMillan et al. [217] also found that aluminum
dioctoate is the best drag reducer among soaps or po-

Colloid and Polymer Science, VoL 262. No. 4 (1984)

lymers for organic solvents. Aluminum dioctoate in


toluene yielded equivalent amounts of drag reduction
at concentrations that were an order of magnitude
lower than those for aluminum distearate in toluene
[217] or aluminum dioleate in toluene [212].
A marked difference between the anionic aqueous
surfactant solutions investigated by Savins and the
nonaqueous ones was that the temperature dependency of drag reduction showed opposite trends.
Larger percentages of drag reduction were found at
15 ~ than at 25 ~ or 35 ~
McMillan's light scattering studies [213] showed
that large agglomerates or micelles caused appreciable
drag reduction by affecting both the viscosity and the
viscoelasticity of the solution. The effect of aging is to
increase the number and size of the micelles. Zakin
[218] noticed that aging of aluminum dioctoate solution in toluene helped to increase both the apparent
viscosity and the drag reducing ability by causing a
more complete dispersion, and a possible increase in
the agglomerate size. Rodriguez [219] suggested that
the stability to aging of aluminum di-2-ethylhexanoate
was superior to that of other aluminum soaps investigated. A sample which had been stored as a 2 % gel in
toluene for over five years gave nearly as good a drag
reduction on dilution as the initial preparation. Sheffer
investigated [220] the effect of temperature on the rate
of decrease of viscosity of a 0.1 ~ aluminum dilaurate
solution in benzene. He found the rate of decrease to
be lowest near room temperature but there was a rapid
fall in viscosity for either lower or higher storage temperatures.
The presence of light, moisture and traces of other
substances have a significant effect on the performance
of disoap-hydrocarbon systems [221]. Sheffer [220]
observed that deterioration of the soap systems was
accelerated by trace amounts of water and acid. On
the contrary, the observations by Zakin [218] showed
that aging was actually retarded by the presence of
water, probably because water molecules slowed
down the break-up of the agglomerates or micelles.
Other hydrogen bonding compounds like crotyl chloride were also found to enhance the drag reducing ability of soap solutions, by speeding up the initial soap
dispersion in the solvent [216]. The presence of such
hydrogen bonding molecules may cause either rapid
disperion, rapid aging or stabilization of the aluminum
disoap micelles depending on the kind of interactions
of the hydrogen bonding molecules with the soap.

Shenoy, A review on drag reduction with special reference to micellar systems

2.3.4b Cationic surfactants as drag reducers


The cationic suffactant that has been investigated
for drag reduction in most detail is cetyltrimethylammonium bromide (CTAB). Nash [222,223] was the
first to show that mixtures of CTAB and naphthalene
derivatives in water produced viscoelastic gels. He
made use of the swirl decay times to find the conditions under which dilute solutions of CTAB and naphthol became viscoelastic [224].
Nash showed that the optimum mixture for viscoelasticity was not far from equimolaar and that there
was a general lowering of the concentration limits in
the presence of salt. He realised that there was a pronounced effect on varying the order of addition of the
components to the mixture. He found that adding
naphthol to the CTAB solutions produced stronger
gels than viceversa, even though the final concentrations of the reagents were identical.
It was Gadd [225] who suggested the possibility of
using the CTAB-naphthol mixture to reduce turbulent
friction, because the mixture showed shear-thinning
characteristics. White [226] used the findings of Nash
and Gadd to carry out turbulent drag reduction measurements in circular pipes using an aqueous equimolar solution of CTAB and 1-naphthol at a total concentration of 508 wppm. The solution was prepared by
dissolving CTAB in water and then adding the naphthol, which had been previously dissolved in alcohol,
drop by drop. The viscosity of this resulting complex
soap solution was found to be only slightly greater
than that of water. White noticed no threshold stress
effect of the kind that is noticed in many dilute high polymer solutions. He found that the drag reduction terminated at some upper Reynolds number depending
on the pipe diameter. The breaks in the Reynolds
number curves for different pipe diameters corresponded to a constant value of shear stress, rw ---48 dynes/cmL He explained this as the critical shear stress at
which there was a scission of the micelles. The drag reduction ability was, however, restored below the upper Reynolds number because of the reformation of
the disrupted micelles. He also did not note any permanent shear degradation over a period of several
days.
Recently, Zakin et al. [227] also investigated the
drag reducing ability of CTAB-naphthol mixtures.
They carried out tests to determine which weight ratio
of CTAB to naphthol gave the highest critical shear
stress. They found that the maximum shear stability
occurred for the 2.1 to I solution.

325

They also investigated the effect of varying the total


concentration of the complex soap solution and found
.that the critical shear stress increased with concentranon. But as the more concentrated solutions became
more viscous, they were less drag reducing at lower
velocities. However, at about 0.18 % total soap concentration, maximum drag reduction was observed
over the whole range of flow rates.
They also observed that as the temperature was increased, the critical shear stress decreased gradually for
a 0. 2150/0 concentrated solution, in contrast to the
sharp change observed by White [228] in the range of
35 ~ to 40 ~ for a 0.05 % concentration of his complex soap solution. The solutions were tested by Zakin
et al. for temperatures between 26 ~ and 48 ~ It was
found that the drag reducing behaviour was completely lost after the 48 ~ test. This loss was observed to be
permanent as this solution showed no drag reduction
characteristics even when cooled to 26 ~
White [228] has reported some]oss of drag reduction for his complex soap solutions after 5 days and
found that the solutions were rendered completely
ineffective after 12 days. Contrary to this, Zakin et al.
[227] observed a drastic drop in the critical shear stress
from over 500 dynes/cm2 to less than 200 dynes/cm2
after 3 days accompanied by a tremendous loss in the
drag reduction capacity. Seven days later there was
further loss observed but the rate of change of critical
shear stress was found to be slower.
From White's study and that of Zakin et al., it is
clear that CTAB-naphthol mixtures behave like concentrated drag reducing solutions as they show no
transition region but only a gradual deviation from the
extension of the laminar curve in the friction factor
versus Reynolds number plots. The minimum concentration for this type of drag reducing behaviour
was, of course, found to increase with increasing tube
diameter.
Zakin et al. [227] also studied the effect of adding
the CTAB-naphthol mixtures to moderately concentrated sand-water suspensions. They found that the
pressure drops and critical velocities in the soap suspensions flows were considerably lower than in water
suspensions flows, thus giving major reductions in the
energy requirements per pound of solid transported.
They also noticed that there was an interaction, between the surface of the solid particles and the soap, resulting in a reduction of the effective soap concentration in the suspending fluid. As before, they noticed
that the solutions were quite stable to mechanical degradation.

326

One marked advantage of cationic surfactants over


the anionic ones discussed, is that these complex soaps
do not precipitate in the presence of calcium ions. But
they are expensive and degrade chemically in aqueous
solutions in a matter of a few days. Though mechanically stable, they are not thermally stable and hence
cannot be put to a variety of uses.

2.3.4c Nonionic su~factantsas drag reducers


The studies on nonionic surfactants as effective drag
reducing additives have been reported only by Zakin
and Chang [229,230]. They studied the effect of temperature, electrolyte concentration and type, surfacrant concentration and composition, and the effect of
mechanical shear on three nonionic surfactants formed from straight-chain alcohols and ethylene oxide
moieties of proper sizes.
Nonionic surfactants have an upper and a lower
temperature limit for solubility in water. They noted
that the nonionics studied were effective drag reducers
near their upper critical solubility temperature or
cloud point. This cloud point is the point at which a
nonionic dissolved in water becomes turbid as the
temperature is raised. With the approach of the cloud
point, the micelles become larger and larger till finally
a surfactant rich phase separates out. This is possibly
the result of the dehydration of the hydrophobic ether
linkages in the chain, thereby causing an increase in the
hydrophobic nature of the chain [231]. The cloud
point is more or less a constant over a wide range of
concentrations of the surfactants; but it is affected by
other additives. Hectrolytes lower the cloud point; alcohols and other polar solvents may raise it; soap and
other anionic surfactants may displace it.
Zakin and Chang found that the effect of increase of
temperature was to change the drag reducing behaviour from 'dilute' to 'concentrated'. 'Dilute' drag reducing behaviour is one in which there exists a transition period even after fully developedturbulent flow is
reached before the solution shows drag reduction.
When no such transition points are present, the behaviour is described as 'concentrated' drag reducing behaviour. The terms 'dilute' and 'concentrated' as defined above should not be confused with the usual
meanings of these words used to refer to the amount of
additive present.
They found that on increasing the concentration of
the electrolyte K 3 P O 4 from 0.3 N to 0-4 N, they
could obtain 'concentrated' drag reduction from what
was initially no drag reduction. On investigations of
various types of electrolytes, they concluded that mul-

Colloid and Polymer Science, Vol. 262. No. 4 (1984)

tivalent anions were more effective than monovalent


ones, and that the cations generally had very little effect.
Zakin and Chang found that 1% solutions of the
commercial surfactants like Alfonic 1214 were more
effective than their 0.5 0/0solutions. The critical shear
stress for mechanical degradation in the case of the former was greater than 1800 dynes/cm2 compared to
500 dynes/cm2 in the case of the latter. The critical
shear stress for mechanical degradation in the case of
nonionics is dependent on the surfactant concentration, electrolyte type and its concentration, and on the
temperature of operation.
The particular chemical structure of the surfactant
has an important effect on its micelle size and shape
[232] and thus in turn has a pronounced effect on the
drag reduction. Zakin and Chang suggest that there is
need for studies of additional structures in order to find
the optimum for drag reduction.
Undoubtedly, nonionic surfactants have an advantage over all the drag reducing additives studied so far.
They are both mechanically and chemically stable.
They do not precipitate out in the presence of calcium
ions and hence can be used in all impure waters, sea
water, brackish water or concentrated brine solutions.
Detailed studies were carried out by Shenoy [233] in
order to exploit the capacity of nonionic surfactants to
their fullest extent and it was shown that these types of
additives have excellent potential for drag reduction at
high temperatures.

2.4 The proposed theories of drag reduction


Since drag reduction is characterised by large
changes in the flow caused by the presence of minute
quantities of additives, the objective of drag reduction
studies is to seek an equally sensitive mechanism
which can predict such a large effect at the dilutions involved. A number of theoretical explanations have
been offered but none of them have been satisfactory
and exhaustive.
In 1948, Oldroyd [22] was the first to suggest that
polymer molecules, during flow, affected the region
near the wall most strongly. His slip-at-the-wall theory has remained a mere engineering correlation method, rather than a fundamental answer to the mechanism of drag reduction, and will remain so until an
improved method for the measurement of velocity
profiles very near the wall is found. Based on Oldroyd's theories, Toms [21] proposed the idea of a
shear-thinning wall layer with an extremely low viscosity which resulted in lower friction coefficients for

Shenoy, A review on drag reduction with special reference to micellar systems

the drag reducing solutions than for pure solvents. But,


later, looking at the rheograms of drag reducing polymer solutions [34,196] it was evident that they were
not shear-thinning but, in fact, Newtonian by conventional viscometry method. However, it was the work
of Walsh [234] which completely shattered the shearthinning theory of Toms. He showed that solutions of
polymethacrylic acid, which is essentially a shear-thickening substance, gave considerable drag reduction.
Further, it was Lumley [235], who showed that, as viscosity is dominated by inertial forces, turbulence
could not be very sensitive to shear-induced viscosity
changes. Though the shear-thinning wall effect theory
crumpled, El'perin et al. [236] proposed that there was
another aspect of the wall effect which could possibly
explain the phenomenon of drag reduction. They suggested that there could exist an absorbed layer of polymer molecules at the pipe wall during flow and this
could lower the viscosity, create a slippage, dampen
turbulence pulsations, and prevent any initiation of
vorticity at the wall. Experiments by Davies and Ponter [237] and later by Little [238] confirmed the presence of an apparent absorption layer. However, it was
later realised that this absorption effect was a mere experimental artifact [239,240] and though polymer
molecules do adhere to clean surfaces in thin films,
they have no projections into the bulk of the solution
which could alter the flow properties. Thus, it is evident that there is no possible relationship between polymer absorption on surfaces and the drag reducing effect.
The existence of a nonisotropic viscosity was
thought of as another possible way of explaining the
phenomenon of drag reduction. It was quite probable
that the viscosity was low in the flow direction and
high enough in all other direction to cause significant
damping in the turbulent fluctuations. The theoretical
work of Elata and Poreh [240] showed that the differences in 'normal stresses' could be a mechanism for
drag reduction. But Gadd [225] found that for concentrations of equal drag reducing ability, poly(ethylene
oxide) alone showed differences in the normal stress
components, while no such difference was observed in
the case of guar gum solutions. However, Patterson
and Zakin [241] developed a viscoelastic model for
calculating reductions in turbulent bursts. They evaluated the normal stress differences by measuring the
thrust from a jet of the solutions, which of course were
extremely concentrated by drag reducing standards
and hence obviously viscoelastic. Their calculations,
though compatible with their results of drag reduction

327

experiments, could not be relied upon as the solutions


used during these experiments were much more dilute
than those used for the thrust experiments. Contrary
to the simple Maxwell model used by Patterson and
Zakin, Kinnier and Reister [242] and Prather [243]
suggested that a Maxwell distribution of the elements
would be necessary to account for the drag reducing
property of poly(ethylene oxide). It was Boggs and
Thompson [244] who found that viscoelasticitygave a
destabilizing effect and concluded that transition
should occur earlier though experiments showed contrary results. They, however, suggested that in the turbulent zone, frictional drag was a function of one-third
the power of the Weissenberg number, which is the ratio of the elastic forces to the viscous forces. Later, Lockett [245] made calculations to find that viscoelasticity did give destabilization but he inferred that a vorticity component was introduced in the flow direction.
Further, Black [246] developed a new theoretical model of turbulent shear flow which indicated greater
sublayer stability, but showed an increase in the longitudinal velocity fluctuations when polymers are present during the flow. Ruckenstein [247] suggested that
drag reduction could be due to two effects of viscoelasticity:
1. the intantaneous shear stress at the wall being
smaller for viscoelastic fluid than for the corresponding Newtonian fluid, and
2. the renewal of the elements of liquid along the
wall taking place more slowly in the case of viscoelastic
fluid than in the case of a Newtonian fluid.
A detailed discussion on the relationship between
drag reduction and the random surface renewal in turbulent pipe flow is available in Fortuin and Klijn [248].
Another possible explanation for drag reduction
could involve the idea of elongational flow. It is quite
likely that the addition of small quantities of polymer
to a solvent leads to a substantial increase in the resistance to elongational flow thereby resulting in less turbulent bursts and thus lowering the turbulent drag.
Shin [73] suggested the possibilky of anisotropic viscosity effects due to coil extension of polymer molecules. Tulin [249] was, however, among the first to visualize that polymer molecules undergo an extension
during shear, whereby the strain energy from the turbulent eddies is absorbed and transmitted as elastic
shear waves which, of course, eventually die off due to
viscosity. Tulin verified from the experimental data of
Hoyt and Fabula [32] that the turbulent dissipation in
the polymer solution is proportional to the product of
concentration and square root of molecular weight.

Colloid and Polymer Science, VoL 262 9No. 4 (1984)

328

He further postulated that the increased turbulent dissipation, which could be as high as three times that of a
normal fluid, was proportional to the increase in the
sublayer thickness of the boundary layer and thus resuked in drag reduction. Lumley [250] concurred
with Tulin's theory that the growth of the laminar sublayer could be explained by molecular extension, but
he found by calculation that the elongation itselfis very
slight. Cottrell et al. [251-253] also performed a number of experiments and found that there was only a very slight macromolecular elongation during shear
flow. However, Lumley postulated that it was the molecular entanglements, which extended cross-stream
and resisted the.formation of streamwise vortices, that
caused an ,ncrease in the sublayer thickness. But Paterson [157] found pipe drag reduction with poly(ethylene oxide) solutions as dilute as 0.03 ppm, and, at the
molecular separation distances and volume fractions
involved, it is quite unlikely that molecular entanglement could play any part in drag reduction. Nevertheless, Ellis et al. [254] showed that polymer solutions
stored for several weeks showed lower drag reduction
than freshly mixed ones, the implication being that
there has been a molecular disentanglement during
storage. But as polymer solutions may degrade in
more ways than one, their observations must be taken
with a pinch of salt. Thus the situation not being very
clear, one can disregard the 'molecular entanglement'
hypothesis only with some reservations. Little [225]
found that poly(ethylene oxide) solutions of different
molecular weights gave the same drag reduction when
their concentration was proportional to the critical
concentration at each molecular weight (i. e. the computed concentration for the polymer coils to touch
each other). Kinnier [116] made similar computations
but used the concept of 'equivalent concentration' to
correlate his data. He found that, to have equal drag reduction for different molecular weight polymer solutions, one has to have equal volumes of polymer based
upon the hydrodynamic sphere considerations.
Pfenninger [256] postulated that the molecular
stretching and the possible breaking-up of the polymer
molecules interfered with and weakened the vorticity
of the disturbances or amplified boundary-layer oscillations.
The molecular stretching itself is induced by the
mutual interactions of convoluted vortex filaments
starting as non-linear laminar oscillations close to the
wall. The drag reducing additives help to delay the
breakdown of these oscillations into turbulence. The
net effect is a partial absorption of the kinetic energy of
9

the vortices, thereby allowing them to grow in size and


penetrate further away from the wall before being unstable, thus resulting in an increased sublayer thickness
and a drag reduction. Peterlin [257] pointed out that
there was an extremely high energy dissipation of a
vortex when it hit a macromolecule, the latter absorbing the energy by molecular stretching. The polymer
molecules help to stabilize the surface layer by mechanical interference which suppresses the turbulence and
reduces the growth of the vortex. Further, using the
Corino and Brodkey [258] picture of the turbulent
bursting process which is initiated in the sublayer and
mainly responsible for the turbulent energy production, Gordon [259, 260] showed that polymer molecules suppressed the turbulent bursts and thereby increased the resistance to stretching. He suggested that,
if the ejection frequency of turbulent bursts is unchanged in polymer solutions, the resistance to
stretching inhibits the magnitude of the bursts and
hence the flow becomes less turbulent. Further, the
measurements of Latto and Shen [261] have shown
that the presence of polymers during turbulent solvent
flow reduce the size range of the turbulent eddies.
Quite recently, Gyr [262] talked of a reversal in the relative occurrence of the streamwise and spanwise vortices in the presence of polymers. A decrease in the velocity fluctuations in the spanwise direction was also
observed earlier by Rudd [263,264]. Gyr's results
have also been in agreement with those of Brennen
[265] and,. Fortuna and Hanratty [266]. Kim et al.
[267] have suggested that the presence of ionic sidegroups in molecules like polyacrylamide cause an extension due to charge repulsion, thus resulting in a
drag reduction effectiveness proportional to the ionic
content of the solution. Banijamali et al. [268] have later shown their concurrence with the conclusions of
Kim et al. Lumley [7, 8,269], ofcourse, has postulated
a very comprehensive and physically plausible drag
reduction mechanism in which macromolecular elongation initiates a sequence of changes in mean and turbulent flow structures, and cause a reduction in friction.
It was Astarita [270] who suggested that turbulence
in viscoelastic liquids is perhaps not suppressed, but is
less dissipative than in viscous liquids. But Gadd
[271,272] was among the first to suggest that drag reduction occurred not due to reduced turbulence dissipation, but rather due to a decreased production of
turbulence. Johnson and Barchi [273] performed the
first few experiments to show that there was a decreased production of small eddies in a developing

Shenoy, A review on drag reduction with special reference to micellar systems

boundary layer containing a polymer. Walsh [234,


274] proposed a comprehensive theory of drag reduction that large scale disturbances which produce Reynolds stresses some distance downstream were, previously, small disturbances at the edge of the viscous
sublayer some distance upstream. The polymer molecules help to alter the energy balance of the turbulent
fluctuations close to the wall, thus allowingthe viscous
dissipation to destroy the disturbances. This decrease
in the number of disturbances moving out from the
viscous sublayer eventually alters the structure of tur=
bulence in the outer part of the boundary layer, thus
resulting in lower Reynolds stresses and hence drag reduction. Walsh's theory does not predict turbulence
damping in a free flow and is based upon the assumption that the phenomenon is essentially due to the existence of a wall-boundary-layer flow. This concept was
well borne out by the experiments of Wakers and
Wells [275] who exuded polymers from a porouswalled pipe section in flowing water and observed
drag reduction occurring in the downstream sections
of the pipe where the polymer was in contact with the
walls and pure water was flowing through the interior
of the pipe. Kilian [276] has also shown that in drag reducing polymer flows the turbulence energy is considerably reduced at high frequencies and is essentially
due to the decreased production of turbulence.
One more possible mechanism by which drag reduction can occur is the developmentof a resistance to
vortex stretching due to the presence of the additives.
Gadd [277] has pointed out that the larger polymer
molecules, with larger relaxation times, would probably affect the large eddies with less intense stretching
rates and cause their more rapid decay. Further support of vortex stretching inhibition, on the basis of
large elongational viscosity, has been presented by
Kuo and Tanner [278] in their grid turbulence analysis. Gyr [279], on the other hand, pointed out that the
process of vortex stretching inhibition was effective
only in the case of the smallest eddies near the wall.
Metzner and Metzner [280], too, have shown the remarkable resistance to elongational stress of dilute polymer solutions, the implication being the resistance to
vortex stretching. Vortex stretching inhibition has also
been considered by Lacey [281] who postulated that
there was a net macromolecular movement towards
the centre of the vortices, followed by a resistance to
stretching due to an interaction in the long overlap regions of aligned molecules. The phenomenon of vortex inhibition as observed by Gordon and Balakrishnan [282,283] too, could be explained by a resistance

329

to vortex stretching caused by filament formation in


drag reducing polymer solution.
Virk [284,285] has introduced the physical notion
of an elastic sublayer, which according to him is characteristic of the drag reduction phenomenon. He has
suggested that the stimulation of polymer molecules
by a turbulent shear flow could create an elastic sublayer at the onset of drag reduction, which could grow
in size with increasing drag reduction and eventually
occupy the entire pipe cross-section at maximum drag
reduction. He [121,286], thus, proposed the idea of a
maximum drag reduction asymptote beyond which
no further viscous drag reduction could be found experimentally. He expressed this asymptote as
f = 0-42 (Re -~
as compard to
[=

16 (Re- 1.o)

Poiseuille Law

and
f = 0. 079 (Re -~

Blasius Law

Thus, it is evident that for turbulent flow drag reduction, the position of the friction factor versus Reynolds number plot is always confined to the portion
between the Blasius line and Virk's maximum drag reduction asymptote.
In this section, an attempt has been made to illustrate the various approaches to a rational explanation
of the drag reduction phenomenon by the different
schools of thought. Though none of the above are
complete in themselves, it is quite clear that each embodies an element of truth on account of which it cannot simply be disregarded.

2.5 Morphology of micellar and polymenc systems


When one compares the data for CTAB-naphthol
solutions [226] on one hand with that for poly(ethylene oxide) solutions [287] on the other, it becomes
obvious that the drag reduction behavioural patterns
in these two cases are different. The soap solution exhibits drag reduction at low wall shear stress values
and beyond a certain critical value the solution is found
to rapidly revert to normal turbulent behaviour.
On the other hand, poly(ethylene oxide) solutions
exhibit relatively small drag reduction at low Reynolds
numbers and increasingly large reduction in drag at
high Reynolds numbers. These two types of beha-

330

viour are a consequence of the significant morphological difference between a micellar-type drag reducer
and a polymeric-type drag reducer.
In the case of micellar systems, there are a number
of different types of structures suggested. All surfacrant molecules or ions, at concentrations above a minimum value characteristic of each solvent-solute system, associate to form particles called micelles. The
most common and widely supported structures for
these micelles are the spherical and the lamellar ones.
The spherical micelle is generally conceived as a small
ball-like particle of colloidal dimensions and fairly
constant in size for a given species [288]. These spherical micelles are said to exist only in relatively dilute solutions.
In concentrated solutions, however, the lamellar
micelle is favoured, the existence of which for an ionic
micelle is compatible with the thermodynamic data
and the double layer theory of charged colloids [289].
A double layer micelle, of course, is one in which the
micelle formed by 50-100 molecular size particles is
lined up in two flat layers.
Debye and Anacker [290] were the first to find that
in the presence of salt, C14 and C16 quaternary compounds had rod-shaped micelles, wherein the ions are
disposed radically about a cylindrical axis with tails
pointing inward and heads pointing outward. Successive planes of these ions are disposed parallel to one
another and perpendicular to the length of the rod.
This model fits well with the theoretical calculations
on the energy of formation of the micelles and does not
preclude the existence of double layer micelles of the
same ions under different environments [291]. It has
been postulated [208] that in a surfactant system, initially spherical micelles rearrange into cylindrical or
rod-like micelles under the influence of an electrolyte.
The current views on micelle structure are shared
by all the leading investigators in the field. They all
agree that several different structures are possible and,
in fact, exist. It is possible for transition from one type
to another to occur. Which of these structures exists in
any particular system depends entirely on the system
parameters such as the chemical nature of the surfactant, the solvent, the presence of other components
like salts, solubilized materials, temperature and the
concentration of each component.
Since drag reducing solutions show viscoelasticity
with increased concentration of the additives, a study
of the same was essential in order to develop an insight
into the morphology of the systems. Pilpel [292] has
suggested that, in soap solutions, viscoelasticity arises

Colloid and Polymer Science, Vol. 262 9No. 4 (1984)

from the formation of large interlinkingsecondary micelles. He reasoned out that if this viscoelasticity is merely due to physical entanglementof the micellar units,
then it was quite immaterial whether the units are lamellar, cylindrical or bead-like strings of spherical micelles. An idea of the morphological parameters for
micellar systems, for the conditions of maximum drag
reduction activity, can be obtained from Savins [207]
who estimated the same using the information of Pilpel
[293]. The shape of the micelles of a complex soap
causing the Toms effect has been discussed in detail very recently by Myska and Simeckova [294] who confirm the existence of fibrillous structure.
In the case of polymeric systems, it is generally
agreed that the morphology of a polymer in solution
resembles a loose three-dimensionalnetwork consisting of variously extended macromolecule segments
[205]. A number of studies have been made to determine the important molecular parameters of polymer
molecules for drag reduction to occur, the most recent
being those of Berman [295,296], Ting [297] and Zakin and Hunston [298]. It has been found that the molecules should be flexible [299], loosely coiled and
long-chained [118,125,300-302] with relatively small
side groups [303,304]. It has also been found that the
drag reduction effectiveness of a homologous series of
polymers increases with increasing chain length
[134,157] and that as the highest molecular weight
components disappear, the polymer solution becomes
less effective.
Thus, the morphological difference between micellar and polymeric systems explains the contrasting
drag reducing behaviour of the two systems. These
two kinds of behaviour can be reconciled if one can
realise that
(a) the flexible polymer molecule needs to be elongated be a large velocity gradient before its full drag reducing ability is developed, and
(b) the surfactant particles are oriented much more
easily at lower velocity gradients, but are broken down
at the high shear stresses associated with large velocity
gradients. Further, in terms of equivalent molecular
weight, micelles are known to have larger values [305]
than polymers and hence they would naturally shift
the onset of drag reduction to a lower shear stress value.

3. The potential of drag reduction


Even after about thirty-five years of research since
Toms discovered the phenomenon of drag reduction,

Shenoy, A review on drag reduction with special reference to micellar systems

it seems to be too early to expect this effect to be put


into wide practical use, though a few cases of pragmatic value have been recorded over the years. However,
it would be beneficial to outline the areas where the
first steps towards putting the drag reduction phenomenon into effect have been taken, to look into most
of the possible applications and thus to obtain an idea
of its potential.
It is a common practice, nowadays, to use a guar
gum in the hydraulic fracturing of oil wells in order to
obtain a reduction in the turbulent friction. It has been
found by Pruitt et al. [104] that by using an optimum
concentration of guar gum, determined by a nomograph as a function of well depth, pumping rate, and
total volume of the fluid to be pumped, the horsepower required during the operation could be reduced by
a factor of over five as compared with the use of water
alone. Melton and Malone [306] have also provided
extensive data on the estimated friction resistance for
oil-well pipelines with different additive concentrations.
One of the most promising applications of friction
reduction has been found to be in fire-fighting operations [307-310]. It has been ascertained that addition
of small quantities of drag reducers causes a considerable reduction in pressure drop through the hose, increases the volume of flow and furthers the throw of
the water stream [311-316]. An emergency increase in
the system capacity can also be achieved in times of
high demand like a major fire by the addition of drag
reducers to the entire city water supply system [317].
Irrigation systems offer an equally promising area of
application, as it has been found that an irrigation
pump can cover twice as much area with the presence
of drag reducing additives in the supply system [311,
318,319]. The drag reduction studies in open channels
by Sellin and Barnard [320,321] can be efficiently applied to hydraulic problems like coping up with peak
discharges during flooding of waterways. The practical use of polymeric drag reducers has been seen in
Dallas, USA and Bristol, England [322,323] where
they are injected into storm sewers at times of peak
loads during heavy storms to handle the increased runoff.
Ram, Finkelstein, and Elata [107] have showed that
addition of high molecular weight polyisobutylene
greatly reduced the turbulent friction in the transportation of crude oil and kerosene. The recent successful
tests of polymer injection into the Alaska pipeline to
increase the throughput of Alaska crude [324, 325] demonstrate the pragmatic use of the drag reduction phe-

331

nomenon. Poreh et al. [326] have shown that the presence of drag reducers considerably reduces the power
levels in the transportation of suspensions of sand and
slag particles. Field tests in Czechoslovakia using drag
reducers for hydrotransport of fly ash from power
plants burning high ash coals [327] is also another
example of the practical use of drag reducing agents.
The only factor limiting the widespread use of polymers as drag reducers in increasing the capacity of pipelines, is the problem of degradation [328]. But, in the
case of oil pipelines which happen to be in cold areas,
the drag reducers not only decrease the pumping costs
but also retard the drop in oil temperature by lowering
the liquid-pipe wall heat transfer coefficient.
In the ship-building industry, drag reducers find
good use in the model testing of prospective new ships
[75,192,329,330]. Their presence in towing tanks has
a two-fold advantage - first, by being able to simultaneously scale the Froude and Reynglds numbers, the
operating conditions are improved, and secondly, by
lowering the drag on the model they help to bring the
wave-making resistance coefficient of the model to the
same as that of the full-scale ship.
The concept of drag reduction in external flow situations can be effectively used for reducing the drag
on full-scale ships and other submerged vessels [105,
331-335]. Since large quantities of additives are required [32, 336] the economics depend on finding very cheap drag reducers. Nevertheless, Kowalski [337]
has suggested pulse ejection of the drag reducers in order to reduce the additive requirements. Wu [338] has,
however, pointed out that ejecting dilute drag reducing solutions in the sublayer around a travelling ship
could produce extremely economical results. Canham
et al. [339] hive reported full scale trials on the 140 feet
long ship, the 'I-IMS Highburton', and obtained very
encouraging results. The tremendous research in this
area will certainly lead to the eventual commercial use
of drag reduction for ships.
In recent years, there has been an increasing interest
in using drag reducing additives as possible circulation
aids for diseases such as arteriosclerosis and coronary
thrombosis during which the blood flow is turbulent
[340-342]. Hoyt and White [128] have suggested that
poly(ethylene oxide) with plasma and dextran can be
used to reduce the pressure drop in cardiovascular systems, provided, of course, they have no adverse physiological effects. Hoyt, in particular, has patented the
use of drag reducing additives with blood transfusion
fluids [343]. Recently, it has been found that okra
gum, apart from being a good drag reducer, is a suc-

332

cessful blood plasma replacement like dextran and has


a considerable resistance to shear degradation [344]. A
number of other in vitro experiments have been undertaken by Greene et al. [345]. These have been followed by a number of toxicity studies by Nokes et al.
[346], which obviously are the first steps to in vivo ex-"
perimentations.
Killen [347] and Brady [289] have performed certain sophisticated experiments with typical drag reducers and concluded that their presence causes a
marked reduction in flow noise radiated from the
boundary layer in the high frequency range. A surfacepressure fluctuations decrease was also found to occur
due to the presence of these additives [69,348]. The
drag reducers cause a suppression of the turbulent
bursts in the boundary layer and hence they can be
used effectively to decrease the noise levels.
Current applied research has shown that the presence of drag reducing additives improves the efficiency
and head characteristics of turbomachines [349]. For
example in the case of large single-suction, centrifugal
pumps, the backing plate drag can be reduced by as
much as 50 % by the use of drag reducers. Very recently, Hampsan and Naylor [350] have shown that high
molecular weight polymers could be used for reducing
friction in hydrodynamic journal bearings.
In current studies, hydrofoils are widely used [351]
among others like keels, rudders, propeller blades, stabilizing fins, or lifting surfaces for hydrofoil boats, and
one of the problems is to increase the lift-to-drag-ratio.
However, it has been found by Lehman and Suessmann [352] that the lift problems in hydrofoils can be
solved by the use of drag reducers, but special hydrofoil designs and special ejection processes are essential,
lest there should be large reductions of lift probably
due to a contracted wake induced by viscoelasticity as
observed by Wolff and Cahn [353], Sarpkaya [354]
and Kawada and Tagori [355]. Sanders et al. [60] have
also tested propellers at low Reynolds numbers and reported results comparable to those of Sarpkaya.
It has been found that typical drag reducing additives inhibit the tendency of water to form a vortex in
freedraining flows [282,283]. This pheonomenon of
vortex inhibition has been shown to correlate very
well with the drag reducing ability of the additives and
hence could be used to predict the effectiveness of an
additive as a drag reducer [282]. Gordon and Balak~
rishnan [283] have also shown that this effect could be
used in molecular-weight distribution studies.
The phenomenon of friction reduction could be
useful in studying molecular properties like molecular

Colloid and Polymer Science, VoL 262 9No. 4 (1984)

configurations [356] and determination of molecular


weight [357]. Kenis [194] has suggested that it could
also be used for following polymerisation - depolymerisation reactions in biological systems. The findings of
Greskovich and Shries [358] and Beattie [359], that
polymers provide a reduction in friction even in twophase systems like a gas-liquid one, could be put to very good use.
Drag reducers could be put into excellent use in central heating systems, where the main resistance to heat
flow is not on the liquid side and hence the presence of
the additives have practically no inhibiting effect upon
the heat transfer. Fitzgerald [360] studied the use of
polymer drag reduction additives in domestic central
heating systems but concluded that they are not suitable mainly because of the problem of polymer degradation. The problem of degradation is quite acute in
central heating systems as the additi.ves are liable to
both mechanical and thermal degradation. The possibility of using micellar systems as dragreducers for this
application was considered by Shenoy [233], but the
study is far too limited for immediate practical use.
The recent work of Ahrnbom and Hagstrand [361]
shows that the interest in this application is still current.
A number of other applications do exist though
most of the major ones have been cited above, and
these can be found in the comprehensive work of Sellin et al. [17]. With the massive efforts that have gone
into the area of drag reduction and the continuing research that is going on presently, it is reasonable to expect the potential of drag reduction phenomenon to be
exploited to the fullest extent with the elapse of time.
Acknowledgement

The author would like to express his appreciation to Dr. H.-W.


Bewersdorff of Abteilung Chemietechnik, Universitiit Dortmund
(E R. G.), for useful discussions and helpful suggestions during the
preparation of the revised version of this article. This was done during the author's three week visit to the Universitiit Dortmund in
December 1983.

References

1. Darby R (1972) NRL Memo Report 2446


2. Fisher MC, Ash RL (1974) NASA TMX - 2894
3. Gadd GE (1971)Encyclopedia of Polym Sci and Techn, Vo115,
John Wiley, New York
4. Hoyt JW (1972) Trans ASME, J1 Basic Engg 94:258
5. Landahl, MT (1972) MIT Cambridge/Mass, Dept Aeronaut
Astronaut, Report No 1 AFOSR-TR-73-1200
6. Little RC, Hansen RJ, Hunston DL, Kim OK, Patterson RL,
Ting RY (1975) Ind Eng Chem Fundamentals 14:283

Shenoy, A review on drag reduction with special reference to miceUar systems

7. Lumley JL (ed) (1969) Sears WR, Annual Reviews of Fluid


Mech, Annual Review Inc, Palo Alto, California 1:367
8. Lumley JL (1973) J Polymer Sci, Macromolecular Reviews
7:263
9. PalyvosJA (1974) Report No 741, Thermodynamics and Transport Phenomena Lab, National Techn Univ, Athens 147,
Greece
10. Patterson GK, Zakin JL, Rodriguez JM (1969) Ind Eng Chem
61:22
11. Virk PS (1975) AIChE J1, Journal Review 21:625
12. White A, Hemmings JA (1976) BHRA fluid Engg, Cranfield
13. Berman NS (1978) Ann Rev Fluid Mech 10:47
14. Giesekus H, BewersdorffHW, Dembek G, Kwade M, Martischius FD, Schaff R (1981) Fortschritte der Veffahrenstechnik
19:3
15. Kirdyashkin AG (1977) Fluid Mechanlcs-Sov Res 6:79
16. Sellin RHJ, Hoyt JW, Scriverner O (1982) J Hydraulic Res
20:29
17. Sellin RHJ, Hoyt JW, Pollert J, Scrivener O (1982) J Hydraulic
Res 20:235
18. Forrest G (1931) Paper Trade J 22:298
19. Brautlecht CA, SethiJR (1933) Ind Eng Chem 25:283
20. Brecht W, Heller H (1939) Das Papier 264
21. Toms BA (1948) Proc 1st Intern Congr on Rheology, Vol II,
p 135, North Holland, Amsterdam
22. OldroydJG (1948) Proc 1st Intern Congr on Rheology, Vol II,
p 180, North Holland, Amsterdam
23. Mysels KJ (27 Dec 1949) US Patent 2, 492:173
24. Dodge DW, Thesis PhD (1958) Univ of Delaware
25. Dodge DW, Metzner AB (1959) AIChE J1 5:189
26. Shaver RG, Merrill EW (1959) AIChE J1 5:181
27. Dever CD, Harbour RJ, Siefert WF (1962) US Patent 3, 023,
760
28. Ousterhout RS, Hall CD Jr (1960) J1 Petro Techn 13:217
29. Savins JG (1964) SPE J1 4:203
30. Pruitt GT, Crawford HR (1963) Westco Research Final Report
on Contract 60530-8250 to Naval Ordnance Test Station
31. Hoyt JW, Fabula AG (1963) Proc 10th Intern Towing Tank
Conf, Teddington
32. Hoyt JW, Fabula AG (1964) Proc 5th Syrup on Naval Hydrodynamics, Bergen, Norway, Office of Naval Research ACR
112:947
33. Fabula AG, Hoyt JW, Crawford HR (1968) Bulletin American
Physical Society, Vol 8
34. Barnes HA, Wakers K (1968) Nature 219:57
35. Barnes HA, Walters K (1969) Proc Roy Soc Lond A 314:85
36. Kelkar JV, Mashelkar RA (1972) J Appl Polym Sci 16:3047
37. Walters K, Barnes HA, Dodson AC (1971) AIChE, Chem Engg
Prog Syrup Series No 111 67:1
38. Tomita Y (1970) Bull ISME 13:926
39. Little RC, Wiegard M (1971) J Appl Polym Sci 15:1515
40. Bilgen E, Boulos R (1972) Trans Can Soc Mech Engrs 1:25
41. Rubin H, Elata C (1971) AIChE J117:990
42. Eckelmann H (1973) Max-Planck-Institut fiir Str6mungsforschung G6ttingen, Bericht Nr 101
43. Bilgen E (30 March-2 April 1971) Fifth Intern Conf on Fluid
Sealing, Univ of Warwick England, Paper H4
44. Bilgen E (1971) Trans ASME, J1 Basic Engg 93:85
45. Bilgen E, Vasseur P (1974) Trans ASME, J1App1 Mech 41:45
46. Gadd GE (1963) National Physical Lab, Ship TM 42
47. Gilbert CG, Ripken JF (ed) (1969) Wells CS, Jr, in: Viscous
Drag Reduction, Plenum Press, NY, 251

333

48. Giles WB (1968) J Hydronautics 2:34


49. Gold IYr, Amar PK, Swaidan BE (1973) J1 Appl Polym Sci
17:333
50. Goldstein S (1965) Proc of the Cambridge Philosophy Soc
31:232
51. Kale DD, Mashelkar RA, Ulbrecht J (1973) Nature Phys Sci
242:29
52. Kato H, Watanabe K, Veda K (1972) Bull JSME 15:1185
53. Kelkar JV, Mashelkar RA, UlbrechtJ (1972) Trans Insm Chem
Engrs 50:343
54. Mashelkar RA (1973) AIChE J119:382
55. Mashelkar RA, Kale DD, UlbrechtJ (1975) Trans Insm Chem
Engrs 53:143
56. Miloh T (1969) MSc Thesis, Technion, Israel Inst of Tech
57. Peyser P (1973) J App1 Polym Sci 17:421
58. Porch M, Miloh T (1971) J Hydronautics 5:61
59. Quraishi AQ, Mashelkar RA, Ulbrecht J (1976) J Non-Newtonian Fluid Mech 1:223
60. Sanders JV, Henderson BH, White RJ (1973) J1 Hydronautics
7:124
61. Smallman JR, Wade JHT (1969) Can Aero-Space Inst Trans
2:37
62. Whitsitt NF, Crawford HR (1967) Western Co Report on Contract No N 60530-12741
63. Whitsitt NF, Moon GM (1967) Western Co Report on Contract No N 60530-12741
64. Belokov VS, Kalashnikov VN (1971) Nature Phys Sci 229:55
65. Denn MM, Roisman JJ (1969) AIChE J115:454
66. Hata C, Tirosh J (1965) Israel J1 Tech 3:1
67. James DF, Gupta OP (1971) AIChE, Chem Engg Prog Syrup
Series No 111 67:62
68. Jones WM, Marshall DE (1969) British J1 Appl Physics 2:809
69. KillenJM, AlmoJ (ed) (1969) Wells CSJr, in: Viscous Drag Reduction, Plenum Press, NY 447
70. Mashelkar RA, Kale DD, Kelkar JV, Ulbrecht J (1972) Chem
Engg Sci'27:973
71. Merrill EW, Mickley HS, Ram A (1962) J Fluid Mech 13:86
72. Rubin H, Elata C (1966) Phys Fluids 9:1929
73. Shin H, Thesis ScD (1965) MIT, Cambridge, Mass
74. Song CS, Tsai FY (1966) St Anthony Falls Hydraulic Lab Project Report Mp - 84
75. Emerson A (1965) Trans NE Coast Inst of Engrs and Shipbuilders 81:201
76. Fruman D, Sulmant P (1969) CR Acad Sci Paris 268:149
77. Fruman DH, Tulin MP (June 1972) Hydronautics, Inc Tech Report 7101-3
78. Kowalski T (Sept 1969) Prof of the Conf on Turbulence Measurements in Liquids, Univ of Missouri, Rolla 102
79. Kumar SM, Sylvester ND (1973) AIChE Chem Engg Prog
Symp Series No 130, 69:1
80. Levy J, Davies S (1967) Intern Shipbuilding Prog 14:166
81. Love RH (1965) Hydronautics Inc, Tech Report 353-2
82. Odai Y, Tagori T (1971) presented at the 49th MeetingJSME,
Preprint No 710 - 15
83. Tagori T, Ashidate I (1969) Proc 12th Intern Towing Tank Conf
Rome
84. White DA (ed) (1969) Wells CS Jr, in: Viscous Drag Reduction,
Plenum Press, NY 35
85. Wu J (ed) (1969) Wells CSJr, in: Viscous Drag Reduction, Plenum Press, NY 331
86. Wu J (1971) Nature Phys Sci 231:150
87. Wu J (1973)J1 Hydronautics 7:129

334
88. Acharya A, Mashelkar RA, Ulbrecht J (1976) Rheol Acta
15:454
89. Acharya A, Mashelkar RA, Ulbrecht J (1976) Rheol Acta
15:471
90. Chhabra RP, Uhlhers PHT, Boger DV (1980)J Non-Newtonian Fluid Mech 6:187
91. Crawford HR, Pruitt GT (1- 5 Dec 1963) presented at the
Symp on Non-Newtonian Fluid Mechanics, 56th Annual
Meeting, AIChE, Houston
92. Lang TG, Patrick HVL (1966) ASME preprint 66-WA/FE-33
93. Ruszczycky MA (1965) Nature 206:614
94. Sanders, JV (1967) Int Shipbuilding Prog 14:140
95. Subbaraman V, Mashelkar RA, Ulbrecht J (1971) Rheol Acta
10:429
96. White A (1966) Nature 211:1390
97. White A (1967) Nature 216:994
98. White DA (1966) Nature 212:277
99. Williams MC (1965) AIChE J1 11:467
100. Savins JG (1961) J1 Inst Petrol 47:329
101. Hoyt JW (1967) US Patent 3, 327, 522
102. Hoyt JW (eds) (1965) Morris AW, Wang JTS, in: Symp on
Rheology, ASME, NY 71
103. Pruitt GT, Crawford HR (April 1965) Report No DTMB-1 of
Western Co under Contract No Nonr 4306 (00)
104. Pruitt GT, Simmons CM, Neill GH, Crawford HR (1964)
SPE Paper No 997
105. Wells CS (ed) (1969) Wells CS, in: Viscous Drag Reduction,
Henum Press, NY 362
106. White A (1966) J Mechanical Engg Sci 8:452
107. Ram A, Finkelstein E, Elata C (1967) Ind Engg Chem Process
Design Dev 6:309
108. Chashehin IP, Shalavin NT, Saenko VA (1975) Int Chem Engg
15:88
109. Ellio~tJH, Stow FS Jr (1971) J Appl Polym Sci 15:2743
110. Fabula AG (ed) (1965) Lee EH, Proc 4th Intern Congr Rheol,
1963, Part 3, Inter Science Publishers, NY 455
111. Felsen IM (1971) PhD Dissertation, Univ of Maryland
112. Felsen IM, Smith TG (May 1972) Paper 13 b, 72nd National
Meeting, AIChE
113. Giles WB, Pettit WT (1967) Nature 217:470
114. Goren Y, Norbury JF (1967) Trans ASME, J1 Basic Engg
89:814
115. Henderson LH (1971) M Sc Thesis, US Naval Postgraduate
School
116. Kinnier JW (1970) Ph D Thesis, Naval Postgraduate School,
Monterey
117. Little RC, Patterson RL (1974)J Appl Polym Sci 18:1529
118. Peyser P, Little RC (1971)J Appl Polym Sci 15:2623
119. Pike EWA (July 1971) Admiralty Res Lab (Teddington) ARL/
C/N9
120. Ripkin JF, Pilch M (1964) St Anthony Falls Hydraulic Lab
Project No 71 to DTMB Nonr 710 (49)
121. Virk PS (1966) Sc D Thesis, MIT, Mass
122. Virk PS, Merrill EW, Mickley HS, Smith KA (1966) in: Modern Developments in tile Mechanics of Continua, Academic
Press, Inc NY
123. White WD (1966) Proc 5th US National Congr ofAppl Mech,
ASME
124. Little RC (1971) J App1 Polym Sci 15:3117
125. Pruitt GT, Rosen B, Crawford HR (1966) Western Co Report
DTMB-2 Nonr 4306 (00)
126. Liaw GC (1968) Ph D Thesis, Univ of Missouri, Rolla

Colloid and Polymer Science, VoL 262. No. 4 (1984)


127. Ramakrishnan C, Rodriguez F (1973) AIChE Chem Engg
Prog Symp Series No 130, 69:52
128. Hoyt JW, White WD (1966) Proc 19th Annual Conf in Engg
on Medicine and Biology 49
129. Forester RH, Larson RE, HydenJW, WetzelJM (1969)J Hydronautics 3:59
130. Metzner AB, Park MG (1964) J Fluid Mech 20:291
131. Nagarajan R, Davies GS, Venkateswarlu D (1974) Chem
Engg, J1 7:249
132. Pruitt GT, Whitsitt NF, Crawford HR (Sept 1966) Western
Co Report to NASA, Contract NAS 7 - 369
133. Seyer FA, Metzner AB (1967) Can J1 Chem Engg 45:121
134. Whitsitt NF, Harrington JJ, Crawford HR (June 1968) Western Co Report No DTMB-3 Nonr 4306(00)
135. Mel'tser LZ, El'perin IT, Lerenthal LI, Karalenko VS (1972)
Kholod Tekh Tekhno115:36
136. Ernst WD (1966) AIChE J1 12:581
137. Ernst WD (1967) AIAA J1 5:906
138. Kobets GF (1969) Bionika 3:72
139. Ripkin JF, Pilch M (1963) St Anthony Falls Hydraulic Lab
Technical Paper No 42, Series B
140. Meter DM (1964) AIChE J110:881
141. Hershey HC (1965) Ph D Thesis, Univ of Missouri, Rolla
142. Hershey HC, ZakinJL (1967) Ind Eng Chem Fundamentals
6:381
143. Patterson GK (1966) Ph D Thesis, Univ of Missouri, Rolla
144. Rodriquez JM (1966) M S Thesis, Univ of Missouri, Rolla
145. Rodriquez JM (1969) Ph D Thesis, Univ of Missouri, Rolla
146. Ram A, Kadim A (1970) J Appl Polym Sci 14:2145
147. Polishchunk AM, Yu D, Raiskii, Temchin AZ (1972) Neft
Khoz 50:60
148. Holtmeyer MD, ChatterjiJ (1980) Polym Eng Sci 20:473
149. Rodriguez JM, Zakin JL, Patterson GK (1967) SPE J1 7:325
150. Evans AP (1974)J Appl Polym Sci 18:1919
151. Block H, Morgan AM, Walker SM (4- 6 Sept 1974) Intern
Confr on Drag Reduction, Cambridge, Eng
152. Cox LR, North AM, Dunlop EH (4-6 Sept 1974) Intern
Conf on Drag Reduction, Cambridge, Eng
153. Nadolink RH (May 1973) Naval Universea Systems Center
Technical Report 4422
154. Sylvester ND, Smith PS (1979) IndEng ChemFundamentals
18:47
155. Hunston DL, Griffith JR, Little RC (1973) Nature Phys Sci
246:140
156. Lummus JL, Randall BV (1964) Pan American Petrolelum
Corp Research Dept Report F64-P-54 Job No 3918
157. Paterson RW, Abernathy FH (1970) J Fluid Mech 43:689
158. Muller HG, Klein J (1980) Makromol Chem Rapid Communication 1:27
159. Aggarwal SH, Porter RS (1980) J App1 Polym Sci 25:173
160. Chang HFD, Darby R (1983)J Rheol 27:77
161. Brostow W (1983) Polymer 24:631
162. Vanoni VA (1946) Trans ASCE 111:67
163. Zandi I (1967) J1 Am Water Works Assoc 59:213
164. Blatch NS (1906) Trans ASCE 57:400
165. Vanoni VA, Nomicos GN (1960) Trans ASCE 125:1140
166. Bugliarello G, Daily JW (1961) TAPPI 44:881
167. Daily JW, Bugliarello G (1961) TAPPI 44:497
168. Mih W, Parker J (1967) TAPH 50:237
169. Robertson AA, Mason SG (1957) TAPPI 40:326
170. Seely TL (1968) Ph D Thesis, The Inst of Paper Chemistry,
Appleton, Wisconsin

Shenoy, A review on drag reduction with special reference to micellar systems

171. Bilgen E, Boulos R (1973) CanJl Chem Engg 51:405


172. Maude AD, Whitmore RL (1958) Trans Insm Chem Engrs
36:296
173. Eissenberg DM (1964) AIChE J110:403
174. Thomas DG (1962) AIChE J1 8:266
175. Bobkowicz AJ, Gauvin WH (1965) Can J1Chem Engg 43:87
176. Kale DD, Metzner AB (1974) AIChE Jl 20:1218
177. Kale DD, Metzner AB (1981) Ind Eng Chem Fund 20:101
178. Kerekes RJE, Douglas WJM (1972) Can J1 Chem Engg
50:228
179. Lee WK, VaseleskiRC, Metzner AB (1974) AIChE J120:128
180. Radin I, Zakin JL, Patterson GK (1973) Nature Phys Sci
246:11
181. VaseleskiRC, Metzner AB (1974) AIChE J1 20:301
182. Robertson AA, Chang MV (1967) Pulp Paper Mag Can
68:T438
183. Arranaga AB (1970) Nature 225:447
184. Ellis HD (1970) Nature 226:352
185. Hoyt JW (July 1972) Naval Undersea Center Report NUC
TP 299
186. Peyser P (1973) J Appl Polym Sci 17:421
187. Van Driest ER (1970)J1 Hydronautics 4:120
188. Van Driest ER (1971)Proc 9thlntern Symp on SpaceTechand
Science, Tokyo
189. Pirih RJ, Swanson WM (1972) CanJl Chem Engg 50:221
190. Mejean L, Boulos MI (1976) Can J1 Chem Engg 54:382
191. Vanusse R, Coupal B, Boulos MI (1979) Can I1 Chem Engg
57:238
192. Hoyt JW (1966) Proc llth Intern Towring Tank Conf Tokyo
193. Hoyt JW, Soli G (1965) Science 149:1509
194. Kenis PR (1968) Nature 217:940
195. Kenis PR, Hoyt JW (1971) Naval Undersea Center Report
TP 240
196. Kenis PR (1968) AppI Microbiology 16:1253
197. Kenis PR (1969) Int Shipbuilding Progress 16:342
198. Kenis PR (1971)J1 Appl Polym Sci 15:607
199. Hoyt JW (May 1968) Naval Res Rev
200. Hoyt JW (1-12July 1974) presented at the Syrup on Swimming and Flying in Nature, California Inst of Techn
201. Kobets GF, Zar'yalova VS, Komarova ML (1969) Bionika
3:80
202. Merkilov VI, Khotinskaya VD (1969) Bionika 3:96
203. Pyatetskii VE, Yu, Savshenko N (1969) Bionika 3:90
204. Rosen MW, Cornford NE (1971) Nature 234:49
205. SavinsJG (1967) Rheol Acta 6:323
206. SavinsJG (1968) US Patent 3,361, 213
207. SavinsJG (ed) (1969)WellsCSJr, in: Viscous DragReduction,
Plenum Press, NY 183
208. Booij HL (ed) (1949) Kruyt HR, 'Association Colloids' in
Colloid Science II Ch 14:681
209. Pilpel N (1954) Jl Colloid Sci 9:285
210. Agoston GA, Harte HW, Hottel HC, Klemm WA, MyselsKJ,
Pomeroy HH, Thompson JM (1954) Ind Eng Chem 46:1017
211. Mysels KJ (1971) AIChE Chem Engg Prog Symp Series No
111, 67:1017
212. Radin I, ZakinJL, Patterson GK (ed) (1969) Wells CSJr, in:
Viscous Drag Reduction, Plenum Press, NY
213. McMillan ML (1970) Ph D Thesis, Ohio State University, Columbia
214. Radin I (1968) M S Thesis, Univ of Missouri, Rolla
215. Lee KC (1970) M S Thesis, Univ of Missouri, Rolla

335

216. Lee KC, Zakin JL (1973) AIChE Chem Engg Prog Symp Series No 130 69:45
217. McMillan ML, Hershey HC, Baxter R (1971) AIChE Chem
Engg Prog Symp Series No 111 67:27
218. Zakin JL (1972) Nature Phys Sci 235:97
219. Rodriguez F (1971) Nature Phys Sci 230:152
220. Sheffer H (1948) Can J1 of Research 26B:481
221. Baker HR, Bolster NN, Leach PB, Little RC (1970) Ind Eng
Chem Product Res Develop 9:541
222. Nash T (1956) Nature 177:948
223. Nash T (1958) J Col] Sci 13:134
224. Nash T (1956) J Appl Chem 6:539
225. Gadd GE (1966) Nature 212:1348
226. White A (1967) Nature 214:585
227. Zakin JL, Poreh M, Brosh A, Warsharsky M (1971) AIChE
Chem Engg Prog Symp Series No 111, 67:85
228. White A (1968) Hendson College of Tech, Res Bull No 5,113
229. Zakin JL, Chang JL (1972) Nature Phys Sci 239:26
230. Zakin JL, Chang JL (4- 6 Sept 1974) Proc Intern Conf on
Drag Reduction, Cambridge, Eng
231. Kuriyama K (1962) Kolloid-Z 180:55
232. Becher P (ed) (1967) Schick MI, Dekker M, in: Nonionic Surfactants, NY 496
233. Shenoy AV (1976) Rheol Acta 15:658
234. Walsh M (1967) Ph D Thesis, California Inst of Tech
235. Lumley JL (1964) Physics of Fluids 7:335
236. El'perin IT, Smol'skii BM, Leventhal LI (1967) Intern Chem
Engg 7:276
237. Davies GA, Ponter AB (1966) Nature 212:66
238. Little RC (1967) Naval Res Lab Report 6542
239. Little RC (1969) Ind Eng Chem Fundamentals 8:520
240. Elata C, Poreh M (1966) Rheol Acta 5:148
241. Patterson GK, Zakin JL (1968) AIChE J114:434
242. Kinnier JW, Reister WA (1965) M S Thesis, US Naval Postgraduate School, Monterey
243. Prather RJ (1966) M S Thesis, US Naval Postgraduate School,
Monterey
244. Boggs FW, ThompsonJ (1967) US Rubber Co Research Center Report on Contract Nos Nonr-3120(00) and N00014-66C0322
245. Lockett FJ (1969) Nature 222:937
246. Black TJ (ed) (1969) Wells CS, in: Viscous Drag Reduction,
Plenum Press, NY 383
247. Ruckenstein E (1973) J Appl Polym Sci 17:3239
248. Fortuin, JMH, Klijn pJ (1982) AIChE J1 37:611
249. Tulin MP (1966) Proc 6th Symp on Naval Hydrodynamics,
Washington, ONR, ACR-136, 3
250. Lumley JL (1967) App1 Mech Reviews 20, 1139
251. Cottrell FR (1968) Sc D Thesis, MIT, Mass
252. Cottrell FR, Merrill EW, Smith KA (1969)J Polym Sci Series
A-2 7:1417
253. Cottrell FR, Merrill EW, Smith KA (1970) J Polym Sci A-2
8:287
254. Ellis AT, Ting RY, Nadolink RH (1970) AIAA paper 70-532
255. Little RC (1969) Ind Eng Chem Fundamentals 8:557
256. Pfenninger W (1967) Northrop Corp Norair Division Report
BLC-179
257. Peterlin A (1970) Nature 227:598
258. Corino ER, Brodkey RS (1969) J Fluid Mech 37:1
259. Gordon RJ (1970) J Appl Polym Sci 14:2097
260. Gordon RJ (1970) Nature 227:599
261. Latto B, Shen CH (1970) Can J1 Chem Engg 48:34

336
262. Gyr A (1974) Inst fiir Hydromechanik und Wasserwirtschaft
ETH Ziirich Report R7-74
263. Rudd MJ (197l) AIChE, Chem Engg Prog Syrup Series
No 111, 67:21
264. Rudd MJ (1972)J Fluid Mech 51:673
265. Brennen C (1970) J Fluid Mech 44:51
266. Fortuna G, Hanratty TJ (1972) J Fluid Mech 53:575
267. Kim OK, Little RC, Ting RY (1973) AIChE Chem Engg Prog
Symp Series No 130, 69:39
268. Banijamali SH, Merrill EW, Smith KA, Peebles LHJr (1974)
AIChE J1201:824
269. Lumley JL (1970) Proc Drag Reduction Workshop, ONR
Boston
270. Astarita G (1965) Ind Eng Chem Fundamentals 4:354
271. Gadd GE (1965) Nature 206:463
272. Gadd GE (1966) Nature 212:874
273. Johnson B, Barchi RH (1968) J Hydronautics 2:108
274. Walsh M (1967) Intern Shipbuilding Progress 14:134
275. Walters RR, Wells CS (1971) J Hydronautics 5:65
276. Kilian FP (1970) Mitteiltmgen der Versuchanstalt ffir Wasserbau und Schiffsban, Berlin Heft 51 and 52
277. Gadd GE (1968) Nature 217:1040
278. Kuo Y, Tanner RI (1972) Trans ASME, Jl App1 Mech 39:661
279. Gyr A (1968) Nature 219:928
280. Metzner AB, Metzner AP (1970) Rheol Acta 9:174
281. Lacey PMC (1974) Chem Engg Sci 29:1495
282. Balakrishnan C, Gordon RJ (1971) Nature Phys Sci 231:177
283. Gordon RJ, Balakrishnan C (1972) J Appl Polym Sci 16:1629
284. Virk PS (1971) J Fluid Mech 45:225
285. Virk PS (1971) J Fluid Mech 45:417
286. Virk PS, Mickley HS, Smith KA (1970) Trans ASME, J1Appl
Mech 37:488
287. Gadd GE (1971) Nature Phys Sci 230:29
288. Hartley GS (1949) Nature 163:767
289. Brady AP (1949) J1 Phys Coll Chem 53:947
290. Debye P, Anacker EW (1951) J1 Phys Coil Chem 55:644
291. Halsey GD (1953) J1 Phys Chem 57:87
292. Pilpel N (1966) Trans Faraday Soc 62:2941
293. Pilpel N (1966) Trans Faraday Soc 62:1015
294. Myska J, Simeckova M (1983) Colloid Polym Sci 261:171
295. Berman NS (1975) J AppI Polym Sci 19:908
296. Berman NS (1980) Polym Eng Sci 20:45l
297. Ting RY (1982) Chem Engg Communications 15:331
298. Zakin JL, Hunston DL (1980) J Macromol Sci Phys B18:795
299. Liaw GC, Zakin JL, Patterson GK (1971) AIChE J117:391
300. Hand JH, Williams MC (1969)J Appl Polym Sci 13:2499
301. Hand JH, Williams MC (1971) AIChE Chem Engg Prog
Symp Series No 111 67:6
302. Little RC (1971) J Colloid Interf Sci 37:811
303. Parker CA, Joyce TA (1974)J Polym Sci 18:155
304. Ting RY, Kim OK (ed) (1973) Bikales NM, in: Water-soluble
Polymers, Plenum Press, NY
305. Hersey HC, Kuo JT, McMillan ML (1975) Ind Eng Chem
Prod Res Dev 14:192
306. Melton LL, Malone WT (1974) SPE J1 56
307. Fabula AG (1971) Trans ASME, J Basic Engg 93:453
308. Rubin H (1972)J Sanitary Engg Division, Proc ASCE 98:1
309. Thorne PF (4 - 6 Sept 1974) Proc Intern Confin Drag Reduction, Cambridge, Engg
310. Thorne PF, Theobald CR, Mahendran P (1975) Fire Res Station, Borchamwood, England, Fire Res Notes 1033 and 1043
311. Anon (Aug 1966) Chem Engg, p 36

Colloid and Polymer Science, Vol. 262. No. 4 (1984)


312. Anon (Sept 1969) Fire Engg, p 48
313. Anon (Jan 1971) Fire Engg, p 33
314. Green JH (1971) Naval Undersea R and D Center TN 504
315. Scott D (1969) Science 164:1466
316. Loeb DL (February 1975) Fire Chief
317. Jackson HC, Mayer PG (1970) Georgia Inst of Tech, Water
Resources Center Report WRC 0170
318. Eli~s V, VocelJ (1978) Vodohospod~irsky eas. SAV 26:610
319. Fajzullaev DP et al (1974) Doklady AV USSR 7
320. Sellin RHJ, Barnard BJS (1970) J Hydraulic Res 8:223
321. Sellin RHJ, Barnard BJS (Aug 1971) J1 Inst Municipal Engrs
98:207
322. Sellin RHJ (1978) J Hydraulic Res 16:357
323. Sellin RHJ, Ollis MJ (1980) J Rheol 24:667
324. Burger ED, Charn LG, Perkins TK (1980) J Rheo124:603
325. Burger ED, Munk WR, Wahl HA (21 - 24 Sept 1980) paper
presented at the 55th Annual Fall Conference of Society of Petroleum Engineers of AIME, Dallas, TX
326. Poreh M, Zakin HL, Brosh A, Warsharsky M (1970) Proc of
the ASCE, Jl of the Hydraulics Division 196:903
327. Pollert J (1977) Proceedings Second International Conf on
Drag Reduction, Cambridge, England, BHRA Fluid Engg,
B3 - 37
328. Treiber KL, Sieracki LM (1970) Columbia Res Corp Report
No 101-Z
329. Kowalski T (April 1966) Naval Engrs Jl 293
330. Tothill JT (1967) Marine Tech 4:111
331. Dove HL (1966) Proc 1lth Intern Towing Tank Conf, Tokyo
332. Gollan A, Tulin MP, Rudy SL (1970) Hydronautics, Inc Tech
Report 909-1, Final
333. Lang TG (ed) (1969) Wells CSJr, in: Viscous Drag Reduction,
Plenum Press, NY, 3131
334. Oltmann P (1969) Schiff und Hafen 21:3
335. Vogel VM, Patterson AM (1964) Proc 5th Symp on Naval
Hydrodynamics Berge, ONR-ACR-112, 975
336. Kowalski T (Jan 1968) Paper presented to Eastern Canadian
Section, SNAME
337. Kowalski T (1968) Trans RINA 110:207
338. WuJ (ed) (1969) Wells CSJr, in: Viscous Drag Reduction, Plenum Press, NY, 331
339. Canham HJS, Catchpole JP, Long RF (July 1971) The Naval
Architect, J Rina, No 2, 187
340. Greene HL, Nokes RF, Thomas LC (1970) Medical Res Engg
9:19
341. Greene HL, Nokes RF, Thomas LC (May 1971) presented at
ASME, Symp on Flow, paper 4-4-64, Pittsburgh
342. Greene HL, Thomas LC, Mostordi EA, Nokes RF (4-6 Sept
1974) Proc Intern Conf on Drag Reduction, Cambridge, Eng
343. Hoyt JW (1971) US Patent No 3, 590, 124
344. Castro WE, Neuwirth JG Jr (1971) Chem Tech 1:697
345. Greene HL et a! (1972) Proc 25th ACEMB, Bal Harbour
346. Nokes RF, Greene HL, Thomas LC (1971) Proc 24th Annual
conf on Engg in Medicine and Biology, Las Vegas
347. Killen JM (1972) Project Report No 123, Univ of Minnesota,
St Anthony Falls Hydraulic Lab
348. Barker SJ (1973) Physics Fluids 16:1387
349. Latto B, CzabanJ (4 - 6 Sept 1974) Proc Intern Conf on Drag
Reduction, Cambridge, Eng
350. Hampsan LG, Naylor H (1975) Proc Inst Mech, Engrs
189:279
351. Fruman DH, Sundaram TR, Daugard SJ (4 - 6 Sept 1974) Intern Conf on Drag Reduction, Cambridge, Eng paper E2

Shenoy, A review on drag reduction with special reference to micellar systems


352. LehmanAF, SuessmannRT(Nov 1972) Oceanicslnc, Report
No 72 - 94
353. WolffJH, Cahn RD (May 1971) Naval Ship Res Dev Center
Report
354. Sarpkaya T (1973) Nature 241:114
355. Kawada H, Tagori T (1973) Proc Ann Meeting ISME
356. Hoyt JW (1966) Nature 211:170
357. Hoyt JW (1966) polym Letters 4:613
358. Greskovich EJ, Shries AJ (1971) Ind Eng Chem 10:646
359. Beattie DRH (4 - 6 Sept 1974) Proc Intern Conf on Drag Reduction, Cambridge, England
360. Fitzgerald D (Sept 1967) Paper presented at Brit Soc of Rheol
Symp on Non-Newtonian flow through pipes and passages
(Shrivenham)

337

361. Ahrnbom L, Hagstrand U (1981) Report SVF - 97, Studsvik


Nyktping
Received October 12, 1983;
accepted January 2, 1984

Author's address:
Dr. A. V. Shenoy
Chemical Engineering Division
National Chemical Laboratory
Pune 411008 (India)

23

Vous aimerez peut-être aussi