39
2 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS Jacques L. Zakin Department of Chemical and Biomolecular Engineering, Ohio State University, Columbus, OH, USA Wu Ge Department of Chemical and Biomolecular Engineering, Ohio State University, Columbus, OH, USA; presently at: Research, Development and Quality, Kraft Foods, Glenview, IL, USA 2.1 Background and introduction 2.2 General aspects of polymer and surfactant drag-reducing systems 2.2.1 Rheology of polymer and surfactant solutions 2.2.2 Maximum drag reduction asymptotes 2.2.3 Type A and B drag reduction 2.2.4 Mean velocity profiles and limiting asymptotes 2.2.5 Turbulence intensities and stress balance 2.3 Polymer drag reduction 2.3.1 Effect of polymer molecular structure 2.3.2 Effect of molecular weight and its distribution 2.3.3 Effect of concentration 2.3.4 Solvent quality 2.3.5 Mechanical degradation 2.4 Surfactant drag reduction 2.4.1 Surfactant micelles, self-assembly, and nanostructures 2.4.2 Types of effective surfactant drag-reducing additives 2.4.3 Solvent effects 2.5 Proposed polymer and surfactant drag reduction mechanisms 2.6 Summary and future development Polymer Physics: From Suspensions to Nanocomposites and Beyond, Edited by Leszek A. Utracki and Alexander M. Jamieson Copyright © 2010 John Wiley & Sons, Inc. 89

Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

Embed Size (px)

Citation preview

Page 1: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

2POLYMER AND SURFACTANT DRAGREDUCTION IN TURBULENT FLOWS

Jacques L. ZakinDepartment of Chemical and Biomolecular Engineering, Ohio State University,Columbus, OH, USA

Wu GeDepartment of Chemical and Biomolecular Engineering, Ohio State University,Columbus, OH, USA; presently at: Research, Development and Quality, Kraft Foods,Glenview, IL, USA

2.1 Background and introduction

2.2 General aspects of polymer and surfactant drag-reducing systems2.2.1 Rheology of polymer and surfactant solutions2.2.2 Maximum drag reduction asymptotes2.2.3 Type A and B drag reduction2.2.4 Mean velocity profiles and limiting asymptotes2.2.5 Turbulence intensities and stress balance

2.3 Polymer drag reduction2.3.1 Effect of polymer molecular structure2.3.2 Effect of molecular weight and its distribution2.3.3 Effect of concentration2.3.4 Solvent quality2.3.5 Mechanical degradation

2.4 Surfactant drag reduction2.4.1 Surfactant micelles, self-assembly, and nanostructures2.4.2 Types of effective surfactant drag-reducing additives2.4.3 Solvent effects

2.5 Proposed polymer and surfactant drag reduction mechanisms

2.6 Summary and future development

Polymer Physics: From Suspensions to Nanocomposites and Beyond, Edited by Leszek A. Utracki andAlexander M. JamiesonCopyright © 2010 John Wiley & Sons, Inc.

89

Page 2: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

90 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

2.1 BACKGROUND AND INTRODUCTION

Corresponding-states studies of the viscosity–concentration behavior of dilute andsemidilute polymer solutions with Robert Simha led one of the authors (JLZ) totheir applications in turbulent flows, a phenomenon that is generally called dragreduction. About six decades ago, Mysels [Mysels, 1949; Agoston et al., 1954]and Toms [1949] discovered that small amounts of aluminum soaps and highpolymers added to a fluid in turbulent flow could significantly reduce pressurelosses.

Dilute solutions of effective polymer, soap, and surfactant drag-reducing addi-tives (DRAs) can dramatically affect the turbulent structures of flowing fluid. Theseadditives can decrease system energy consumption up to 90%, increase flow rateor system length, and permit reduced pipe and pump size. In addition, they alsodecrease heat transfer [Gasljevic et al., 1998]. The greatest use of DRAs has been incrude oil and finished-product pipelines, first demonstrated in 1979 in the 800-miletrans-Alaska (TAPS or Alyeska) pipeline. Crude throughput in the 1.2 m diameterpipeline was increased by up to 30% by injecting a concentrated solution of ahigh-molecular-weight polyolefin additive at bottleneck sections of the line, givinghomogeneous concentrations as low as 1 ppm downstream of the pumps [Motier et al.,1996]. Polymer DRAs have also been used successfully for energy savings in manyother important crude pipelines, such as the Iraq–Turkey, Bass Strait in Australia,Mumbai Offshore [Nijs, 1995], and North Sea Offshore [Dujmovich and Gallegos,2005] and also in finished hydrocarbon product lines [Motier and Carreir, 1989]. Thepolymer composition needs to be tailored to match the composition of the crude oilor hydrocarbon product being transported.

Here we consider only homogeneous solution additives [Manfield et al., 1999],specifically polymers, sodium soaps, aluminum di-soaps, and surfactants, all of whichare effective, require low capital costs, and are convenient to use. In addition to trans-port of crude oil or other hydrocarbons in pipelines, their use has been proposed inoil field operations [Hellsten and Oskarsson, 2004; Sullivan et al., 2007], slurry orhydraulic capsule pipeline transportation [Golda, 1986; Wul et al., 1998], suppres-sion of atherosclerosis [Mostardi et al., 1978; Unthank et al., 1992], prevention oflethality from hemorrhagic shock [Kameneva et al., 2004], increased water flow andwater jet focusing in firefighting equipment [Fabula, 1971; Figueredo and Sabadini,2003], prevention of water overflow in sewage systems after heavy rains [Sellin, 1978;Dembek and Bewersdorff, 1981], and increase in the volumetric flow rate of water inhydropower and irrigation systems [Singh et al., 1985].

Because high-molecular-weight polymers are susceptible to mechanical degra-dation in regions of high shear, such as in pumps, they are not suitable for use inrecirculating systems. Surfactant systems that have long threadlike network micellemicrostructures are also effective drag-reducing additives. These “living polymers”have the advantage of self-assembly after mechanical degradation and so are suitablefor recirculating flows such as in district heating or cooling (DHC) systems [Chouet al., 1989a; Chou, 1991; Steiff, 1991; Pollert et al., 1994; Harwigsson and Hellsten,1996; Gasljevic et al., 1998].

Page 3: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

BACKGROUND AND INTRODUCTION 91

An inherent characteristic of drag-reducing (DR) solutions is that the reducedfriction factors are accompanied by reduced heat transfer ability, as expected fromthe analogous relationship between heat transport and momentum transport [Aguilaret al., 1998, 1999]. This could be beneficial or detrimental to the practical use of DRsolutions in some applications, such as in DHC systems. On the beneficial side, thecost of heat insulation could be greatly reduced in pipelines carrying water in DHCsystems. However, for heat exchangers in systems designed to supply or remove heat,methods for enhancing the heat transfer capability of surfactant DR solutions have tobe developed.

In turbulent internal flows, drag reduction, often expressed as a percentage, iscalculated as follows:

%DR = fs − f

fs

× 100 (2.1)

where f is the measured Fanning friction factor (a dimensionless number defined astwice the wall shear stress divided by the product of fluid density and velocity squared:f = 2σ12/ρv2 of the solution with additives, and fs is that of the solvent at the sameflow rate, or solvent Reynolds number, where the Reynolds number is a dimensionlessratio of inertial forces to viscous forces.

Heat transfer reduction, also expressed as a percentage, is generally obtained bycomparing the Nusselt number (a dimensionless number defined as the ratio of con-vective to conductive heat transfer across a boundary) of a DR surfactant solution(Nu) with that of the solvent (Nus) at the same solvent Reynolds number:

%HTR = Nus − Nu

Nus

× 100 (2.2)

Over the past 60 years, a great deal of applied and theoretical research has been carriedout on both polymer and surfactant DRAs because of their potential useful applica-tions and the influence of the additives on both turbulent structure and rheology.Important results include the identification of maximum drag reduction asymptotes(MDRAs): both Virk’s MDRA for polymer solutions [Virk et al., 1970; see Eq. (2.5)]and Zakin et al.’s MDRA for surfactant solutions [Zakin et al., 1996; see Eq. (2.6)],relating solution nanostructures and rheological properties to macroscopic DR phe-nomena; hypotheses on the influence of DRAs on turbulent structures, mechanismsfor turbulent drag reduction, developing heat transfer enhancement techniques, andso on.

Techniques such as flow and pressure measurements, laser-Doppler velocimetry,high-speed cameras, and particle image velocimetry (PIV) have been used to revealDR flow patterns. Nanostructures of DR solutions are believed to be the controllingfactor in drag reduction and can be seen directly in the quiescent state in cryogenictransmission electron microscopy (cryo-TEM) images. It is generally accepted thatthe presence of long, flexible threadlike nanostructures such as polymer chains or giantsurfactant micelles are required for effective drag reduction [Zakin et al., 2007]. Underreal flow conditions, the nanostructures change with shear, which is a challenging area

Page 4: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

92 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

of considerable theoretical, experimental, and practical interest. Many techniqueshave been developed or are being developed to identify these changes with shear,including, but not limited to, measurements of rheological properties, small-angleneutron scattering intensity, and rheooptical techniques such as flow birefringenceand various light-scattering techniques.

With the recent advance in computational technology, efforts have also beendevoted to numerical simulations of drag reduction, such as for viscoelastic poly-mers via constitutive equations and finite element methods [Dimitropoulos et al.,1998; Fullerton and McComb, 1999; Mitsoulis, 1999; Beris et al., 2000; Yu andKawaguchi, 2004] and for DR flow with surfactant additives via second-order finite-difference direct numerical simulation (DNS) studies [Yu and Kawaguchi, 2003,2006].

In this chapter we focus on how molecular structures of polymer and surfactantadditives affect DR behavior, the rheology and nanostructures of their dilute solutions,and possible mechanisms of polymer and surfactant drag reduction.

2.2 GENERAL ASPECTS OF POLYMER AND SURFACTANTDRAG-REDUCING SYSTEMS

Since the first report by Toms [1949], polymer drag reduction has been studiedextensively in both aqueous and hydrocarbon media. Many linear polymers hav-ing molecular weight of approximately 106 or higher, synthetic or natural, have beenfound to be good DRAs, and a number of excellent reviews on the subject are avail-able [Lumley, 1969; Patterson et al., 1969; Virk, 1975; White and Hemmings, 1976;Shenoy, 1984; Hoyt, 1986; Kulicke et al., 1989; Gyr and Bewersdorff, 1995; Nadolinkand Haigh, 1995; Manfield et al., 1999; Graham, 2004]. Synthetic water-soluble poly-mers such as poly(ethylene oxide) (PEO) and polyacrylamide (PAM) are extremelyeffective drag reducers, PEO being more efficient but less shear stable than PAM.Highly effective natural polymers include Xanthamonas exudates and guar; PEOis also effective in hydrocarbon solvents [Liaw, 1969]. Synthetic oil-soluble poly-mers such as poly(isobutylene) (PIB), polystyrene (PS), and poly(dimethylsiloxane)(PDMS) have been found to be good DRAs in organic solvents. Commercial polymerDRAs are available, such as the Polyox WSR series from Union Carbide (now DowChemical) or the Separan series from Dow Chemical, along with the hydrocarbon-soluble FLO series from Baker Hughes and the LP series from ConocoPhillips usedin hydrocarbon pipeline transport.

Serious interest in drag reduction research began in the 1960s, and by the 1980sconsiderable progress was achieved [Manfield et al., 1999], mainly in studies ofpolymers in both aqueous and hydrocarbon systems. This led to their applications inonce-through systems as high-molecular-weight polymers degrade when subjected tohigh mechanical shear and elongation stresses such as are encountered in most pumps.The polymer chains are broken and the smaller segments formed are not effectiveas drag reducers. This degradation process of the polymers is somewhat enhancedby temperature. Polymer degradation is irreversible, as there is little chance of the

Page 5: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

GENERAL ASPECTS OF POLYMER AND SURFACTANT DRAG-REDUCING SYSTEMS 93

segments recombining. Thus, for recirculation systems where the fluid is continuallyrecycled through a pump, high polymers in either aqueous or hydrocarbon systemsare not suitable as DRAs.

There are important applications where recirculation of water is desired, suchas in DHC systems. Here hot or chilled water whose temperature is adjusted ata central station is circulated to heat or cool buildings in a district. The water isreturned to the central station for reheating or recooling. Pumping energy costsare an important cost factor in such systems, and the use of DRAs to reducethese costs may be attractive. Thus, there has been considerable interest in recentyears in finding nondegrading or rapidly repairable additives which would be effec-tive for long periods of time in such recirculation systems. Surfactant moleculesform threadlike nanostructures, which are necessary for drag reduction, can repairthemselves (self-assemble) after passing through a region of high shear or highextension rate, and thus have potential applications in recirculating flow systemsor multipump station pipeline flow systems. The recovery time is affected by sur-factant concentration, counterion concentration, and temperature, usually falling inthe range from a fraction of a second to several minutes [Gasljevic et al., 2007].This has led to increased interest in aqueous surfactant drag reduction and a numberof publications, including several good reviews and dissertations in the literature[e.g., Bewersdorff, 1990; Gasljevic, 1995; Lu, 1997; Zakin et al., 1998a, 2007;Manfield et al., 1999; Qi, 2002; Zhang, 2005; Drappier et al., 2006; Broniarz-Presset al., 2007].

Among surfactant DRAs, cationics with organic counterions such as salicylateshave received the most attention, primarily because of their lower costs. Typi-cal cationic surfactants studied for drag reduction are quaternary ammonium salts(QASs) with one long alkyl chain (carbon number from 14 to 22) and methylor hydroxyethyl groups in the other positions. In recent years, however, there hasbeen increasing interest in mixed surfactant solutions such as cationic and anionic[Salkar et al., 1998; Koehler et al., 2000; Qi and Zakin, 2002; Schubert et al., 2003],zwitterionic and anionic [Harwigsson and Hellsten, 1996; Hellsten and Harwigs-son, 1996; Li et al., 1998], and cationic surfactant solutions with different alkylchain lengths [Lin et al., 2000]. Nonionic surfactant DRAs, such as polyoxyethy-lene alkyl ethers, are also effective DRAs. Zwitterionic surfactant DRAs includealkyldimethylamine oxides and alkyl betaines, which can be blended with anionicsurfactants such as sodium dodecyl sulfate (SDS) and sodium dodecylbenzylsulfonate(SDBS).

2.2.1 Rheology of Polymer and Surfactant Solutions

The mechanism for turbulent drag reduction remains uncertain (even though manyresearchers have studied this subject) due to the complex nature of the interaction ofDRA nanostructure with turbulence. The influence of extensional viscosity and vis-coelasticity on drag reduction has been examined and keenly debated in theoreticaland experimental investigations. In light of this influence, drag reduction is sometimestermed rheological drag reduction [Graham, 2004]. Since it is widely agreed that the

Page 6: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

94 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

long chains of polymers or threadlike micelles (TLMs) of surfactants are responsiblefor drag reduction, our discussion of surfactant DRAs will be restricted to concentra-tions above their critical micelle concentration (CMC) and those containing TLMs toconcentrations above the second critical micelle concentration, CMCII.

When polymers are dissolved in a solvent, the viscosity of the solution increaseswith polymer concentration and with polymer molecular weight. Similarly, the vis-cosity of a micellized surfactant solution increases with surfactant concentration andthe growth of TLMs. Most polymer or surfactant solutions exhibit shear-thinning(shear viscosity decreasing with increasing shear rate) in a steady simple shearflow, which can be described by empirical constitutive equations. Other impor-tant rheological properties, often seen in polymer or surfactant solutions, includehigh extensional viscosity, high viscoelasticity, and in surfactant solutions, shearthickening.

High extensional viscosity has been suggested as the cause of drag reduction indilute polymer solutions [Landahl, 1972; Hoyt, 1986; Bewersdorff, 1993] since it canbe quite large compared to that of the Newtonian solvent. This leads to additionalresistance against vortex stretching and turbulent eddy growth, and results in reduc-tion of energy dissipation or drag reduction. For surfactant solutions, Savins [1967]was the first to recognize that drag-reducing surfactant solutions show viscoelasticflow behavior. Later, studies on surfactant drag reduction led to a general belief thatdrag reduction of surfactant solutions is related to their rheological properties. A risein shear viscosity at a critical shear rate caused by a shear-induced structure (SIS),viscoelasticity (nonzero first normal stress difference, quick recoil, and stress over-shoot), and high extensional viscosity/shear viscosity ratios (>100) are rheologicalproperties present in many DR surfactant solutions [Bewersdorff, 1996; Qi and Zakin,2002]. It has been proposed that smaller ordered clusters damp the turbulent eddiesin the near-wall region, while in the core region, the SIS damps the large-scale turbu-lence [Ohlendorf et al., 1986; Rehage et al., 1986; Bewersdorff and Ohlendorf, 1988;Chou, 1991]. However, Lu et al. [1997] reported that a highly drag-reducing cationicsurfactant solution of Arquad S-50 (5 mM) with sodium salicylate (12.5 mM), whichis highly birefringent, has high extensional viscosity, and shows a network of TLMsin cryo-TEM images, is not viscoelastic. After reviewing the rheological behaviorof many DR surfactant solutions, Qi and Zakin [2002] concluded that SIS andviscoelasticity are not always observed in DR surfactant solutions, while high exten-sional/shear viscosity ratios may be a requirement for surfactant solutions to be dragreducing.

2.2.2 Maximum Drag Reduction Asymptotes

There is a maximum drag reduction limit for polymer drag reduction [Castro andSquire, 1967; Giles and Pettit, 1967; Virk et al., 1970; Virk, 1975]. The maximumdrag reduction asymptote (MDRA) proposed by Virk is generally accepted for high-polymer drag reduction [Virk, 1975]. Poiseuille’s law for laminar flow [Eq. (2.3)],the von Karman law for Newtonian turbulent flow [Eq. (2.4)], and Virk’s maximum

Page 7: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

GENERAL ASPECTS OF POLYMER AND SURFACTANT DRAG-REDUCING SYSTEMS 95

FIGURE 2.1 Type A and type B drag reduction. (From Liaw [1969].)

drag reduction asymptote (MDRA) for polymer solutions [Eq. (2.5)] are

f−1/2 =

0.0625 Ref 1/2

4.0 log10 (Ref 1/2) − 0.4

19.0 log10 (Re f 1/2) − 32.4

(2.3)

(2.4)

(2.5)

They form the three lines in the plots of Fanning friction factor f versus Reynoldsnumber shown in Figure 2.1, denoted as laminar, von Karman line, and Virk’s MDRA.Drag-reducing flow data for polymer solutions usually lie between the von Karmanline and Virk’s MDRA. The data points shown in Figure 2.1 are for a 200-ppm PEO–benzene solution in a 0.833-mm-ID pipe system and in a 2.32-mm-ID pipe system[Liaw, 1969; Liaw et al., 1971].

For surfactant drag reducers, however, the friction factor can be lower than thatpredicted by Virk’s asymptote [McMillan, 1970; McMillan et al., 1971; Hersheyet al., 1975; Chara et al., 1993]. Based on data for cationic surfactants in water andaluminum disoaps in hydrocarbons, Zakin et al. [1996] proposed a new asymptotefor nonpolymeric drag reducers:

f = 0.32 Re−0.55 (2.6)

This MDRA predicts friction factors which are significantly lower than those predictedby Virk’s MDRA for high polymers (Figure 2.1).

2.2.3 Type A and B Drag Reduction

In 1967, Hershey and Zakin pointed out that for polymer DRAs, there exist two typesof drag reduction, characterized, respectively, by the presence of an onset shear stress

Page 8: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

96 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

above which drag reduction starts and by an extension of the laminar flow line intothe turbulent region [Hershey and Zakin, 1967a]. These two were later labeled typeA and type B drag reduction by Virk and Wagger [1989]. As illustrated in Figure 2.1,for a solution of 200 ppm PEO in a 0.833-mm-ID pipe, there is no transition region,and turbulent region data lie on a gradual extension of the laminar line, approachingVirk’s MDRA. For the same solution in the 2.32-mm-ID pipe, the friction factordata show a normal laminar–turbulent transition and lie on the von Karman lineuntil an onset wall shear stress is exceeded. At an Re value of about 3000, the onsetpoint is reached, drag reduction begins, and the friction factor drops toward Virk’sMDRA.

The data of Liaw et al. show that a given polymer DR solution exhibited what waslater called type A drag reduction in larger pipes with lower wall shear stress, whileshowing type B behavior in smaller pipes with higher wall shear stress at the sameRe value, as demonstrated in Figure 2.1 [Liaw, 1969; Liaw et al., 1971]. Gasljevicet al. [2001] also noted that type A and B drag reduction might occur in two differentflow regions for the same DR solution rather than being two distinct types of DRphenomena.

For type A polymer drag reduction, which is usually observed in randomly coiledpolymer solutions, drag reduction occurs only when a certain wall shear stress isreached. This may be due to uncoiling of the polymer chains in the extensionalflow field or to their entanglements reaching the size of turbulent eddies [Virk andMerrill, 1969; Berman, 1978; Virk and Wagger, 1989]. This onset stress is a func-tion of polymer parameters such as type and molecular-weight distribution or thegoodness of the solvent. It decreases when the molecular weight or radius of gyra-tion increases [Virk, 1975]. It is independent of pipe diameter, which results inthe onset Reynolds number being proportional to the 8

7 power of the pipe diame-ter [Virk and Baher, 1970; Virk, 1975]. While Virk [Virk and Baher, 1970; Virk,1975] and Gold et al. [1973] suggested that the onset wall shear stress was inde-pendent of polymer concentration, a decrease in the shear stress with increasingconcentration has been observed by other researchers [Whitsitt et al., 1969; Patersonand Abernathy, 1970; Hansen and Little, 1971]. Three different models have beenproposed to predict the onset phenomenon: length scale [Virk and Merrill, 1969],time scale [Hershey, 1965; Fabula et al., 1966; Hershey and Zakin, 1967; Walsh,1967], and strain energy storage models [Walsh, 1967]. Details are discussed inSection 2.5.

Type A drag reduction is typically observed and well documented for polymerDR systems [Virk and Wagger, 1989; Gasljevic et al., 2001] but is not often seenclearly in surfactant DR systems, as most DR data reported for surfactant DR systemsshow type B drag reduction behavior. Considering the similarities of DR behaviorof polymer and surfactant solutions, surfactant drag reduction should also follow thetwo types of drag reduction, and onset with surfactant solutions has been observed[Bewersdorff, 1990]. Recently, Ge reported that a zwitterionic surfactant solution of6 mM oleyl aminimide morpholine with 6 mM NaNO2 in 30% glycerol–water showedclear drag reduction onset at every DR effective temperature tested. For that solution,the pipe flow becomes highly unstable for about 1 or 2 minutes when the flow rate is

Page 9: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

GENERAL ASPECTS OF POLYMER AND SURFACTANT DRAG-REDUCING SYSTEMS 97

increased to just above the onset level; then the pressure loss drops significantly anddrag reduction is observed [Ge, 2008].

2.2.4 Mean Velocity Profiles and Limiting Asymptotes

Turbulent flow velocity profiles for Newtonian fluids are divided into three regions:the viscous sublayer, the buffer layer, and the turbulent core. The velocity profile inthe viscous sublayer of DR solutions is similar to that for Newtonian fluids. Figure 2.2shows turbulent core velocities for Newtonian fluids in pipe flow, represented by

u+ = 2.5 ln y+ + 5.5 (2.7)

while an additional term, ∆B, is needed for a parallel profile in the turbulent core forfluids with moderate drag reduction:

u+ = 2.5 ln y+ + 5.5 + ∆B (2.8)

u+ and y+ are the nondimensionalized mean velocity and distance from the wall:

u+ = u

u∗ (2.9)

50

40

30

20

10

0100 101 102

y+

u+

u+ = y+

Eq (2.7)

Newtonian flow

∆B

Eq (2.8)

Drag reducing flow

Eq (2.2)

Virk’s UVP

Viscous

sublayer

Buffer

layer

Turbulent

core

103

FIGURE 2.2 Turbulent sublayers and mean velocity profiles. (From Ge [2008].)

Page 10: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

98 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

FIGURE 2.3 Comparison of limiting turbulent mean velocity profiles. (From Qi [2002].)

y+ = yu∗ ρ

η(2.10)

u∗ =√

τw

ρ(2.11)

where u is the local mean velocity, which varies with y, the distance from the wall,η is the shear viscosity, ρ the density of the fluid at the wall, ∆B the velocity profilecorrection term, τw the wall shear stress, and u* is the friction velocity.

Based on the polymer MDRA [Eq. (2.5)], Virk proposed the following equationfor the mean velocity asymptote in the turbulent core [Virk, 1971]:

u+ = 11.7 ln y+ − 17.0 (2.12)

Equation (2.12), often called Virk’s ultimate velocity profile (UVP), predicts thesteepest profile observed for high-polymer systems. For surfactant solutions at theirMDRA, data showed that the limiting mean velocity profile asymptote in the turbulentcore is [Zakin et al., 1996]

u+ = 23.4 ln y+ − 65 (2.13)

which is almost twice as steep as that for high polymers, as shown in Figure 2.3.

2.2.5 Turbulence Intensities and Stress Balance

Peak root-mean-square time-averaged axial turbulent intensities, u′, of a DR surfactantsolution were shown to be smaller than solvent intensities at low Re and greater

Page 11: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

GENERAL ASPECTS OF POLYMER AND SURFACTANT DRAG-REDUCING SYSTEMS 99

FIGURE 2.4 Turbulent stress profiles for solution with zero Reynolds stress [Eq. (2.14)].(From Qi [2002].)

at high Re [Warholic et al., 1999]. However, the magnitude of v′, the root-mean-square time-averaged turbulent intensities normal to the flow direction for both high-polymer and surfactant DR systems, is always considerably lower than those forNewtonian solvents [Chara et al., 1993; Gampert and Rensch, 1996; Kawaguchiet al., 1996; Myska et al., 1996; Park et al., 1996; Warholic, 1997]. Whereas forNewtonian solvents v′ data peak at about y+ ≈ 100 [Zakin et al., 1998], v′ valuesfor high-polymer DR solutions peak at y+ values >100 and surfactant data peak aty+ < 100. This difference, along with their different MDRAs and different limitingvelocity profile slopes, suggest that polymer and surfactant DR mechanisms maydiffer.

Reynolds stresses (−ρu′v′) are defined as time-averaged products of u′ and v′ andare the major component of total stress away from the wall in conduit flow. Severalauthors have reported that Reynolds stresses for DR solutions are lower than requiredfor total stress balances and attribute this to low v′ and differences in phase between u′and v′. Recently, zero Reynolds stress profiles for highly DR surfactant solutions havebeen observed, clearly illustrating a major stress deficit in these systems [Kawaguchiet al., 1996; Myska and Chara, 1999; Warholic et al., 1999a]. An additional dragreduction stress, τDR, is needed to satisfy the total stress balance. Evaluating andinterpreting this stress is a major research challenge. Figure 2.4 is a schematic illus-trating the magnitude of the additional term, τDR, needed to satisfy the stress balanceat any point for systems with zero Reynolds stresses:

τtotal = τviscous + τReynolds + τDR (2.14)

τviscous = ηdu

dy(2.15)

τReynolds = −ρu′v′ (2.16)

Page 12: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

100 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

2.3 POLYMER DRAG REDUCTION

Decades of research in polymer drag reduction have shown that the effects of variablessuch as polymer molecular structure, flexibility, molecular weight, concentration, andsolvency are important. They are described below along with polymer mechanicaldegradation.

2.3.1 Effect of Polymer Molecular Structure

Molecular structure, chain flexibility, and molecular weight determine the DR effec-tiveness of high polymers. Polymer backbone structure and side groups as well assolvent–polymer interactions determine the flexibility of the molecule [Patterson et al.,1969] as well as the polymer electrostatic charge in aqueous systems.

Straight-Chain Polymers Liaw et al. [1971] showed that in nonpolar systems, flexi-ble polymers were the most effective DRAs. For a given pipe diameter, they measuredthe critical concentration (Cc) at which polymer solution drag reduction changed fromtype A to type B behavior and compared polymer pervaded volume fractions, Cc[η],where [η] is the intrinsic viscosity. This measure of pervaded volume was utilized bySimha and his co-workers in their solution viscosity–concentration studies [Weiss-berg et al., 1951; Simha and Zakin, 1960, 1962]. Values of Cc[η] are sensitive to theflexibility of the polymer molecules, β, the ratio of the root-mean-square end-to-endlength of a polymer molecule in a theta solvent to the length for a freely rotating molec-ular model. Values of Cc[η] increase (becoming less DR effective) with increasing β.PDMS and PEO, which are linear and flexible, with β values of 1.30 and 1.38, arethe most effective polymer DRAs [Liaw et al., 1971; Hoyt, 1986]. Polyacrylic acid(PAA), PAM [Kim et al., 1973; Little et al., 1975], and polysaccharides, which arelinear in nature, are also effective DRAs, whereas highly branched polymers such asgum arabic and dextran are not very effective [Hoyt, 1985]. DRAs used commerciallyin hydrocarbon transport pipelines are linear copolymers of α-olefins and 1-buteneof very high molecular weight. They are effective at very low concentrations (1 ppmor less) [Motier and Carreir, 1989]. It has been reported that although large numbersof short side chains decrease the polymer DR efficiency, fewer but long branchescan actually increase that efficiency. Cross-linking can enhance or destroy the DRcapability of a linear molecule, depending on the degree of cross-linking [Wade,1975].

Copolymers Singh and his co-workers developed copolymers with both good DReffectiveness and good shear stability by grafting flexible chains of highly DR effectivePAM onto less flexible, less DR effective but highly shear-stable polymer backbonessuch as polysaccharides as well as onto a number of other backbones, such as guargum, xanthan gum, carboxymethyl cellulose, low-molecular-weight starch, and otherpolymers, such as poly(vinyl alcohol) [Deshmukh and Singh, 1987; Ungeheuer et al.,1989; Deshmukh et al., 1991; Singh et al., 1991]. They found that copolymers withlonger and fewer PAM branches are more DR effective and more shear stable than

Page 13: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

POLYMER DRAG REDUCTION 101

those with shorter and more PAM branches. The biodegradation resistance of thesecopolymers is enhanced because the altered molecular structures inhibit enzymeattack [Singh, 1995]. It has also been found that these copolymer DRAs are notaffected by the salts present in seawater and they have synergetic flocculation charac-teristics [Singh, 1990]. McCormick et al. also reported that copolymers of PAM withpolyelectrolytes showed better DR efficiency than that of PAM alone [McCormicket al., 1990; Mumick and McCormick, 1992].

Polymer Aggregates and Mixed Polymers Polymers with molecular weights below105 are generally not very effective drag-reducing additives (DRAs) [Hoyt, 1986].Low-molecular-weight ionic polymers capable of forming aggregates, however, havebeen found to reduce drag [Kowalik et al., 1987; Kowalik, 1988; Malik et al., 1993].The aggregation may be induced by electrostatic, hydrogen bonding, or hydropho-bic interactions. It has been reported that aqueous solutions of partially saponifiedpoly(vinyl acetate) (molecular weight ≈ 19,000 g/mol) with various degrees ofhydrolysis showed drag reduction close to Virk’s MDRA [Lodes and Macho, 1989].It has been found that 0.4 to 0.6 wt% tylose, a methylhydroxyl cellulose (molecularweight 6000, g/mol) from Hoechst showed a drag reduction about half that of itshigh-molecular-weight (300,000) counterpart, sodium carboxymethylcellulose at thesame concentration. It also has good shear stability [SaPereira and Pinho, 1994], pos-sibly because low-molecular-weight aggregates, which are broken up by shear, canreassemble. Cross-linking was found to enhance drag reduction for hydroxypropylguar and guar gum solutions at concentrations below the gel formation level, whereastheir shear stability was not affected [Bello et al., 1996].

Reddy and Singh [1985] found that polymer–polymer mixtures may increase ordecrease the DR efficiency, depending on the composition, flow rate, and polymerspecies. Singh concluded that “random coil size, rigidity and solvation of polymermolecules appear to be responsible for synergism observed in drag reduction causedby (polymer–polymer) mixtures” [Singh, 1995] and that maximum drag reductionis usually obtained at a particular ratio of the polymers [Singh, 1990]. The %DRof the mixture can be predicted from those of the individual polymers by using thesimple mixing rule with either a positive or a negative interaction parameter [Singh,1995]. To have a DR-effective polymer mixture, at least one of the polymers needsto be rigid. For example, Dingilian and Ruckenstein [1974] found that mixtures suchas PEO–carboxymethylcellulose or PAM–carboxymethylcellulose gave better dragreduction than did either component, while a mixture of PEO–PAM gave poorer dragreduction than did either component.

2.3.2 Effect of Molecular Weight and Its Distribution

Figures 2.5 and 2.6 compare DR results for three narrow-molecular-weight (MW)-distribution polystyrenes in toluene, a good solvent, and in a theta solvent, respectively.In each case, at fixed polymer concentration, increased MW leads to greater dragreduction. Also, at fixed MW, increased polymer concentration leads to increasedDR. Moreover, at fixed MW and polymer concentration, the magnitude of DR appears

Page 14: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

102 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

5

10

20

30

Dra

g R

educt

ion (

%)

40

50

60

70

10 20 50 100Polystyrene Concentration in Toluene (ppm)

200 500

MW=7,000,000

Tube diameter 0.158 cm

Reynolds number = 16,000

MW=4,000,000MW=2,000,000

1000

FIGURE 2.5 Drag-reducing effectiveness versus polystyrene concentration in toluene, agood solvent. (From Zakin and Hunston [1977].)

significantly higher in the good solvent versus the theta solvent [Zakin and Hunston,1977]. Based on the effect of MW on drag reduction, Hoyt developed a method thatis “extremely rapid and convenient to use” to utilize drag reduction as an estimatorof MW for DR linear high polymers [Hoyt, 1966]. However, Paterson and Abernathyconcluded from the lack of correlation between intrinsic viscosity and drag reductionthat the average MW may not be a good measure of DR efficiency [Paterson andAbernathy, 1970]. Instead, it was pointed out by Hunston and Reischman that theDR efficiency depends strongly on the highest-MW components of a given MWdistribution. They showed that the addition of 25 ppm of a narrow-MW-distribution2 × 106 polystyrene into a benzene solution of 2.5 ppm of a narrow-MW-distribution7 × 106 polystyrene had little effect on drag reduction (Figure 2.7), demonstrating thatthe highest-MW species present determines the level of drag reduction in a mixture[Hunston and Reischman, 1975].

Liaw et al. [1971] examined a reduced-MW or entanglement capacity, definedas the ratio of the polymer molecular weight to its critical entanglement molecularweight (the minimum molecular weight for molecular entanglements to occur inpolymer melts). They found that polymer samples with entanglement capacity aboveabout 60 are relatively effective DRAs, whereas those with entanglement capacitylower than 15 are poor regardless of their flexibility.

2.3.3 Effect of Concentration

Figures 2.5 and 2.6 show that DR effectiveness increases with concentration, presum-ably due to the increasing number of molecules present, which causes the decay ofmore turbulent eddies. At high concentrations %DR levels off [Hoyt, 1986], due to

Page 15: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

POLYMER DRAG REDUCTION 103

70 Tube diameter 0.158 cm

Reynolds number = 17,000

MW = 7,000,000MW = 4,000,000MW = 2,000,000

60

50

40

30

20

10

010 50 100 200

POLYSTYRENE CONCENTRATION IN POOR SOLVENT, ppm

% D

RA

G R

ED

UC

TIO

N

500 1000 2000

FIGURE 2.6 Drag-reducing effectiveness versus polystyrene concentration in 58%toluene/42% isooctane, a poor solvent. (From Zakin and Hunston [1977].)

the increase in solution viscosity [Zakin and Hunston, 1977]. Virk and Baher foundthat polymer DR solutions obey a linear relationship on semilogarithmic Prandtl–vonKarman coordinates (f -0.5 versus Re f 0.5) with a slope that is proportional to thesquare root of the concentration [Virk and Baher, 1970]. Liaw et al. [1971] foundthat increased concentration and decreased tube diameter reduced the onset Reynoldsnumber (type A drag reduction), leading to the disappearance of the laminar–turbulent

40 I.D. = 0.158 cm

MW=2,000,000 (25ppm)

MW=7,000,000 (2.5ppm)

Mixture of two 2M (25ppm) – 7M(2.5ppm)30

20

10

0 2 4 6 8 10 12 14 16 18 20 22

WALL SHEAR STRESS

VISCOSITY

PE

RC

EN

T D

RA

G R

ED

UC

TIO

N

X 10-4 (SEC-1)

FIGURE 2.7 Molecular weight effects on drag reduction. (From Huntson Reischman, withpermission. Copyright © 1975, American Institute of Physics.)

Page 16: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

104 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

transition (type B drag reduction) followed by a gradual departure from the laminarline.

2.3.4 Solvent Quality

Polymer molecules in good hydrocarbon solvents are expanded and stretched morefreely, leading to higher drag-reducing efficiencies. For nonpolar linear polymers thatform coils, such as PIB, Hershey and Zakin [1967b] found that the DR efficiencydecreased by 40% in benzene, a poor solvent, compared with cyclohexane, a goodsolvent. Figures 2.5 and 2.6 illustrate the differences in drag reduction of polystyrenein a good solvent and in a poor solvent, showing that drag reduction is greater in theformer. The onset of drag reduction is observed at a lower Re value for good solvents[Hunston and Zakin, 1980a].

In aqueous systems, solvent goodness can be changed by adjusting the electrolyteconcentration or the pH value. Sellin and Loeffler [1977] observed that MgCl2 causeda sharp drop in the drag reduction of PAA. As salt is added, the polymer chains shiftfrom rodlike shapes stiffened by repulsive charges on the chain to coiled shapes,which gave a lower drag reduction efficiency. The effect of pH on drag reduction foran aqueous PAA solution was reported by Hand and Williams [1969]. While ionizedPAA showed a small drag reduction at a pH of 9.0, 5.0, and 3.0, reducing pH to 2.1,drag reduction increased abruptly and remained high at a pH of 1.0. They suggestthat at pH values of 3.0 or higher, the PAA molecules are fully ionized and extended(rod-shaped). At lower pH values, the polymer chains undergo a second-order phasetransition from an extended rod to a more compact helical structure, which is veryDR effective. At a little lower pH (0.7), only a slight drag reduction was observed.The authors offer no explanation for this observation, but clearly the PAA molecularconformation is sensitive to high hydrogen ion concentration in the low-pH range.

Drag reduction of xanthan gum solutions is greatly reduced at high salt concen-trations [Rochefort and Middleman, 1987]. From their rheological measurements,Rochefort and Middleman concluded that the xanthan molecules underwent a tran-sition from a highly extended semiflexible coil with disordered backbone struc-ture, to a collapsed, semirigid rod with an ordered backbone structure at high saltconcentrations.

2.3.5 Mechanical Degradation

Mechanical degradation of polymers, which refers to a process “in which the activationenergy for scission of the polymer is provided, at least in part, by mechanical actionon the polymer” [Hunston and Zakin, 1980b], imposes severe limitations on the use ofpolymers as DRAs and also exerts detrimental effects on other practical applications,such as in polymer preparation and in polymer use as viscosity index improversfor motor oils. Therefore, it has received considerable attention over the last severaldecades, and a large number of papers have been published, among which there areseveral review articles [Pollert and Sellin, 1989; Singh, 1990; Moussa and Tiu, 1994;Choi et al., 2000b].

Page 17: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

POLYMER DRAG REDUCTION 105

A number of variables affect shear degradation in high polymers. They includethe stress level, polymer concentration, molecular structure, molecular weight and itsdistribution, solvent quality, and temperature. Many degradation mechanisms havebeen proposed, and most of them can be roughly categorized into two groups. The firstgroup, which is the most commonly proposed, focuses on the shear and/or elonga-tional stresses acting on individual polymer molecules, which cause the molecules tofracture, leading to changes in the molecular weight distribution by breakdown of thehighest-molecular-weight species (the most effective drag reducers) and decrease inthe average molecular weight. Studies have suggested a direct relationship betweendepletion of the high-molecular-weight species and losses in the effectiveness ofpolymer drag reduction [Patterson et al., 1966; Paterson and Abernathy, 1970;Hunston, 1976; Yu et al., 1979; Hunston and Zakin, 1980; Chang and Darby, 1983;Brostow et al., 1990; Choi et al., 2000a,b, 2002]. The second group, which hasreceived less attention, associates the degradation with (a decrease of) aggregation ofpolymer molecules [Dunlop and Cox, 1977; Malik et al., 1993; Kalashnikov, 2002;Vlachogiannis et al., 2003; Liberatore et al., 2004].

The rate of mechanical degradation on DRAs increases with increasing MW andwall shear stress and decreases with concentration [Zakin and Hunston, 1977; Yu et al.,1979; Hunston and Zakin, 1980]. Choi et al. [2002] reported that the double-strandedλ-DNA, which is a good drag reducer, loses its drag-reducing ability when denaturedinto two single-stranded molecules and the DNA is always cut in half in turbulentflow when degradation occurs. Yu et al. [1979] studied mechanical degradation of PIBand polystyrene in dilute organic solvents and confirmed the molecular breakage byobserving decreased average MW and changes in MW distribution. They found thatthe breakage does not necessarily always occur at the midpoint of the polymer chain,although the midpoint has the highest probability of being the location of the scissionpoint. In their study of DR solutions of PIB in organic solvents, Patterson et al. [1966],found that the highest MW fractions are most susceptible to shear degradation. In theirstudy of the degradation of PEO in a high-speed stirring device, Minoura et al. [1967],showed that degradation could be represented by a rate constant for molecular scissionand a critical (minimum) MW below which no scission occurs at a given shear stress.Similarly, in their study of mechanical degradation of DNA through capillary tubing,Levinthal and Davison [1961] found that for a given polymer chain length and strengthof bond, there is a critical value of wall shear stress below which shear degradation isnegligible. Hunston [1976] noticed that a high-MW narrow-distribution polystyrene inbenzene solution had greater resistance to degradation than a broader distribution PEOaqueous solution with a high-MW tail at equal weight concentration, each initially ofcomparable DR ability. He pointed out that mechanical degradation depends stronglyon MW distribution and proposed a mechanism for molecular scission, which showsthat at a given shear stress, MW distribution determines the fraction of DR effectivemolecules subject to scission. He took into account for each chemical structure theobservations of Minoura et al. [1967] and Levinthal and Davison [1961] that there isa critical lower MW below which no degradation occurs.

At a fixed stress, mechanical degradation rates are lower in a good solvent than ina poor one [Zakin and Hunston, 1978; Hunston and Zakin, 1980a]. Although this may

Page 18: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

106 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

seem counterintuitive since polymer molecules are more compact in a poor solvent andone might expect them to be less susceptible to mechanical degradation than the moreextended molecules in a good solvent. However, polymer molecules in a poor solventare more likely to associate with other polymer molecules than in a good solvent andthese larger entangled masses may have higher susceptibility to mechanical degrada-tion as they encompass larger ranges of shear gradients. Brostow and his colleagues[1983, Brostow et al., 2007] proposed a polymer conformation model, assuming thatin each polymer chain there exist: “two kinds of sequences, good (oriented along theflow direction or close to it, and strongly solvated) and poor (oriented approximatelyperpendicularly to the flow and poorly solvated).” Based on this model, the authorsclaimed that “solvation numbers go symbatically (a similar tendency) with DR effi-cacy,” which implies that not only the goodness of the solvent, but also the goodnessof the polymer chain sequences, play a significant role in their mechanical stability.In addition, they proposed an equation “that describes quantitatively the degradation,thus predicting drag reduction as a function of time and of concentration.”

In general, initial mechanical degradation rates are relatively rapid and the degrad-ing process can usually be characterized by an exponential time-decay expression[Choi et al., 2000a; Sung et al., 2004]. Choi et al. [2000a] found that a single expo-nential decay model may not be universally suitable for all polymeric drag reducers,but it can effectively represent shear-resistant DRAs, including the polysaccharidesthey studied, and can also be applicable to describing short-time degradation behavior.Patterson et al. [1984] reviewed reports on temperature effects on mechanical degra-dation, and concluded that the effects are not significant. Slower degradation wasreported as the temperature rose, which was attributed to greater molecular flexibility.

2.4 SURFACTANT DRAG REDUCTION

Surfactant solutions containing TLMs that resemble polymer chains may reducefriction energy loss in turbulent flows by up to 90% at relatively low surfactantconcentrations under appropriate flow and temperature conditions.

2.4.1 Surfactant Micelles, Self-Assembly, and Nanostructures

In aqueous solutions of surfactants at concentrations above the critical micelle con-centration (CMC), the molecules self-assemble to form micelles, vesicles, or othercolloidal aggregates. These may vary in size and shape depending on solution con-ditions. In addition to surfactant molecular structure, the effects of concentration,pH, other additives, cosolvents, temperature, and shear affect the nanostructure of themicelles. The presence of TLMs or cylindrical, rodlike, or wormlike micelles at con-centrations > CMCII are generally believed to be necessary for surfactant solutionsto be drag reducing [Zakin et al., 2007].

This nanostructure is similar to that of long, flexible polymer chains. It differs,however, because the chains are held together by secondary forces. They can self-assemble or break up, whereas polymers, which are held together by covalent bonds,

Page 19: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

SURFACTANT DRAG REDUCTION 107

FIGURE 2.8 Schematic phase diagram for cationic surfactant solutions. (From Chou [1991].)

do not re-form after chain scission. Thus, surfactant DRAs can be used in recirculationsystems that have to be subjected to high stresses passing through pumps, as theycan reassemble after shear degradation, whereas polymer DRAs which are degradedirreversibly cannot.

In turbulent conduit flows of surfactant DR solutions, the level of drag reductiongenerally increases with flow rate to a maximum. Above that critical wall shear stress,the availability of TLMs near the wall is reduced and drag reduction diminishes athigher flow rates. However, if the stress level is reduced by decreasing the flow rate,maximum drag reduction can be restored when the rate of micelle recovery nearthe conduit wall exceeds the rate of shear degradation. The critical wall shear stressreaches a maximum at some temperature for each particular surfactant solution anddiminishes at higher temperatures.

The phase diagram in Figure 2.8 shows the effects of surfactant concentration andtemperature on the structural organization of a typical cationic surfactant in aqueoussolution. As the concentration rises above the CMC, spherical micelles form untilat higher concentrations, the micelles transform into TLMs. This is referred to asCMCII. Higher CMCII concentrations are required at higher temperatures. If theelectrostatic repulsions between the positively charged surfactant head-groups arereduced by increasing the ionic strength of the surrounding media following additionof organic counterions or other salts to the solution, the head-group area is reduced,promoting formation of TLMs. Additions of organic counterions such as sodiumsalicylate lower the concentration requirements to form TLMs and are used to enhancedrag reduction. The micelles form entangled networks, and at higher counterion/cationratios may form branched structures which are more resistant to shear [Chellamuthuand Rothstein, 2008]. The cross-sectional diameters of the spheres or of the TLMs

Page 20: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

108 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

are close to twice the length of the alkyl chains of the cationic surfactant (typically5 nm), while the length of the DR effective TLM chains may be a few micrometers.

Small-angle neutron scattering (SANS) [Cates and Candau, 1990; Rehage andHoffmann, 1991] and cryo-TEM studies [Hoffmann et al., 1985] have shown thatsurfactant solutions with vesicle structures in the quiescent state can transform toTLMs under shear, accounting for the rare and unexpected observations that solutionswith vesicle structures can be DR in turbulent flow fields [Zheng et al., 2000].

2.4.2 Types of Effective Surfactant Drag-Reducing Additives

Surfactant molecules are classified by their charges. Cationics are positively charged,anionics are negatively charged, nonionics are not charged, and zwitterionics haveboth positive and negative charges, with a zero net charge. All types have been shownto be effective drag reducers, including some mixtures.

Cationic Surfactants Quaternary ammonium cationic surfactants with alkyl groupsof C16 to C22 with organic counterions have been studied extensively [Chou, 1989b]and have been shown to be effective at concentrations as low as 50 ppm [Kawaguchiet al., 2003]. This effectiveness is enhanced by increasing the surfactant concentrationand by increasing the counterion/cation ratio, ξ.

Chemical Structure of a Cationic Surfactant Quaternary ammonium cationics havean ammonium head-group and are positively charged in solution. Most drag reductionstudies have been with one saturated or unsaturated alkyl group and combinationsof three methyl or hydroxyethyl groups attached to the ammonium. Their anion iscommonly chloride or bromide, although there are reports on acetates. In general,they are not effective DRAs, as the anion resides outside the micelle surface andlong TLMs are rarely formed in solutions. Addition of organic counterions, whichpenetrate the micelle surface at depths of as much as four CH2 groups of the alkylchain [Ge et al., 2008], strongly enhances formation of DR effective TLMs.

The temperature range over which cationic systems are effective drag reducers isimportant for practical applications. Longer chains increase the upper temperaturelimit. They also increase the critical wall shear stress above which drag reductiondecreases. An increase in the upper temperature limit for drag reduction of about8.5◦C for each additional CH2 group in the alkyl chain was reported by Ohlandorfet al. [1986]. As longer chains also reduce low-temperature solubility, they increasethe lower limiting temperature for drag reduction. Unsaturated alkyl chains haveincreased polarity and improved low-temperature solubility and DR while having littleeffect on the high-temperature limit [Rose and Foster, 1989; Chou, 1991; Laughlin,1996].

The effect of cis versus trans configuration of oleyl (cis) and elaidyl (trans)trimethyl ammonium salicylate systems was studied by Qi and Zakin [2002]. Littledifference in the temperature ranges for effective DR were found at ξ = 1.0 and 1.5.However, the kinked chain of the cis (oleic) structure with its larger volume resultsin a larger packing parameter (p = ve/a0lc, where ve is the effective volume of the

Page 21: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

SURFACTANT DRAG REDUCTION 109

hydrocarbon chain, a0 the optimal headgroup area, and lc is the critical chain length,or the maximum effective length that the chains can assume [Israelachvili, 1991])with concomitant growth of TLMs. The resulting cis nanostructure gives a highercritical wall shear stress for drag reduction. At ξ = 2.5 the cis (oleyl) surfactant wasnot soluble below 50◦C, while the trans (elaidyl) was an effective DRA from 4 to80◦C.

Odd versus even chain lengths were studied by Lin et al. [2002]. For alkyltrimethyl ammonium chloride/3-chlorobenzoate systems, lower Krafft temperaturesand decreased lower drag-reducing temperatures were observed for odd-length chainsthan would be predicted from measurements with even chains. They attributed theseeffects to differences in nanostructure caused by differences in packing orientationsin the micelle cores.

Substitution of hydroxyl ethyl groups for up to three of the conventional methylgroups in quaternary ammonium surfactants increases their low-temperature solu-bility and hence their drag reduction at lower temperatures [Rose and Foster, 1989;Chou, 1991]. The increased size of the substituent on the ammonium head-group givessmaller packing parameters, which tends to reduce micelle size. Zhang et al. [2005a]noted that replacing one methyl by an ethyl on a trimethyl ammonium headgroupreduces the packing parameter and the upper temperature limit for drag reduction.Nevertheless, in a series of experiments with a saturated C18 surfactant and sodiumsalicylate (ξ = 2.5) and 0, 2, and 3 hydroxyethyls on the headgroup, Chou reportedthat their DR temperature ranges were the same, while the critical wall shear stress forloss of drag reduction at low temperature increased with the substitution of hydrox-yethyls for methyls [Chou, 1991]. He hypothesized that this resulted from the greaterhydrogen-bonding ability of the hydroxyethyls to the salicylate counterions and towater, giving stabilized and strengthened micelles due to formation of a thick hydra-tion layer at the micelle surface. However, Horiuchi et al. [2001a,b] found that theorder of DR effectiveness with different hydroxyethyl substitutions for methyl onoleyl cationic surfactants varied with ξ values and with surfactant concentration.

Counterion Structure and Ratio As noted above, organic counterions that bondstrongly to the cationic surfactant micelles are essential for effective drag reduc-tion [Chou, 1991]. Important counterion properties affecting the binding are thestrength of van der Waals forces and hydration. Salicylate, tosylate, halobenzoates,and hydroxy-naphthoates are examples of effective counterions for promoting dragreduction [Zakin et al., 1998a]. Their organic portion penetrates the micelles’ hydro-carbon cores, and their aromatic rings’ π-electron cloud interacts with the positivecharge of the ammonium head-group, neutralizing the effects of their positive chargerepulsions, thus promoting the growth of long chains or TLMs. Due to their resonancestructure and delocalized negative charges, these counterions form intramolecularhydrogen bonds [Shapley et al., 1998; Rakitin and Pack, 2005] and their polar por-tions form bonds with water molecules [Underwood and Anacker, 1984; Bijma et al.,1998]. Salicylate, which has been used by many drag reduction researchers, gener-ally gives a large drag reduction with a high DR upper temperature limit. This hasbeen attributed to hydrogen bonds between the adjacent carboxylate and hydroxyl

Page 22: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

110 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

FIGURE 2.9 Orientation of 2-OH-benzoate, 3-OH-benzoate, and 4-OH-benzoate counteri-ons in micelles. (From Smith [1992].)

groups binding to the cationic head-group, further strengthening the micelle networkstructure.

Jamieson and his co-workers found that the ratio of the counterion to cationicsurfactant affected the critical shear rate for shear thickening (SIS). At 1 mMcetyltrimethylammonium bromide, the minimum occurred at 1 mM sodium salicylate[Wang et al., 1994; Yu et al., 1994]. These authors also noted that there was a shearrate below which SIS was not observed [Hu et al., 1993a] and observed strong bire-fringence at shear rates above the critical shear rate for SIS [Hu et al., 1993b]. Flowbirefringence relaxation time also reached a maximum at the 1 : 1 ratio and decreasedat higher ratios [Hu et al., 1994].

The relative position of substituent groups on the aromatic ring has a majoreffect on the DR ability of a cationic/counterion system and its rheological behav-ior because of their influence on the micelle network structure [Gravsholt, 1976;Smith et al., 1994; Kreke et al., 1996; Lu et al., 1998a]. For example, Lu et al.compared the nanostructure, rheological, and DR behaviors of three chlorobenzoatesin cetyltrimethylammonium chloride solutions all at ξ = 2.5. The 4-chlorobenzoate(chloro is strongly hydrophobic) formed well-defined TLMs and was a very effectiveDRA over a broad temperature range, with strong viscoelasticity and high extensionalviscosity. In contrast, the 2-chlorobenzoate system showed only 5-nm-diameter spher-ical micelles: no viscoelasticity and no drag reduction. The 3-chlorobenzoate systemwas intermediate with a DR effective temperature range of only 30 to 50◦C. The threemethylbenzoates gave similar results, but since the methyl group is less hydropho-bic than chloro, the 4-methylbenzoate was less effective than the 4-chlorobenzoate[Smith, 1992; Smith et al., 1994]. We can surmise that the hydrophobic chloro groupin the 4 position is strongly embedded in the hydrocarbon micelle core, giving verystable TLMs, while in the 2 position it may reside in the aqueous phase, where itprovides no contribution to enhancing micelle chain growth. In support of this, Smithet al. suggested the counterion binding orientations shown in Figure 2.9 based on NMRstudies [Smith, 1992; Smith et al., 1994]. The dominant carboxylate in the 1 position isalways in the aqueous phase, forcing the chlorine in the 2 position to reside there also.

Micelle growth is favored by an increase in ξ, as is critical wall shear stress for dragreduction [Rose and Foster, 1989; Chou, 1991; Lu, 1997; Lu et al., 1998b; Lin et al.,2001]. However, at high values of ξ, some systems become insoluble [Qi, 2002]. Theeffect of ξ on TLM formation is complex, depending on both molecular structures

Page 23: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

SURFACTANT DRAG REDUCTION 111

of the cationic surfactant and the counterion. At low values of ξ (<1) the addition ofcounterion promotes micelle growth. At ξ = 1 not all counterions are bound and themicelles surface remains positively charged. Ge et al. [2008] studied the effects ofchemical structures of four 4-halobenzoate counterions (F, Cl, Br, I) with different ξ

values in cetyltrimethylammonium chloride (CTAC) solutions. Using zeta-potentialmeasurements, they reported that for 4-fluorobenzoate, 100% counterion bindingcould not be achieved even when ξ = 8. 4-Chlorobenzoate gave 100% counterionbinding at ξ = 8, 4-bromobenzoate at ξ = 4, and 4-iodobenzoate at ξ = 2.5, becauseof the increasing hydrophobicity of these counterion molecules. Their drag-reducingability increases in the order F, Cl, Br, and I, the order of decreasing electronegativityof these halide ions. Branched micelles which have high critical wall shear stress forloss of drag reduction have been observed at high ξ values [Zhang, 2005c].

Nonionic Surfactants In the temperature region near their coacervation tem-peratures or cloud points, solutions of nonionic surfactants such as polyethyleneglycol monoalkyl ether and ethoxylated alkyl ethers, ethoxylated fatty acidmonoethanolamides, and their mixtures can be DR [Shenoy, 1984; Chang and Zakin,1985; Harwigsson and Hellsten, 1996]. Usually, DR effectiveness extends to 5 to10◦C below and above the cloud point [DeRoussel, 1993]. The polyethylene oxidehead-groups are dehydrated as temperature rises. This loss of hydrophilicity leads toaggregation and micelle growth until, just above the cloud point, phase separationoccurs [Shenoy, 1984; Chang and Zakin, 1985].

Mixtures of nonionics with composition C18H35(OCH2CH2)xOH have cloudpoints between those of their individual components [Chang and Zakin, 1985]. Sincethey are effective as DRAs near their cloud points, effective DR temperature canbe targeted by adjusting the ratio of the components. Addition of multivalent saltssuch as sodium sulfate or of cosolvents such as phenol, which promote dehydration,will lower the cloud point, increase relative viscosity, and enhance drag reduction.Composition changes that enhance hydrophilicity, such as shorter alkyl chain length,unsaturated alkyl chains, or more ethylene oxide groups will raise the cloud pointand the temperature for effective drag reduction. Therefore, the cloud temperatureand DR effective temperature range can be tuned by selection of alkyl chain lengthand structure, number of ethylene oxide groups, salt addition, and pairing of differentnonionic surfactants [Chang and Zakin, 1985; DeRoussel, 1993].

Combinations of ethoxylated fatty acid monoethanolamide and alkyl ether non-ionics with 6 mM MgSO4 salt are effective DRAs at low temperatures [Harwigssonand Hellsten, 1996]. At moderate concentration, the addition of monovalent sodiumnitrite in place of the bivalent salt did not give drag reduction. However, for ethoxy-lated alkyl ether surfactants in seawater, high concentrations of sodium nitrite didenhance drag reduction [Harwigsson and Hellsten, 1996].

Zwitterionic Surfactants There are environmental concerns regarding the use ofcationics as DRAs, as they are not easily biodegradable. Zwitterionics are biodegrad-able and are not very sensitive to the presence of salts. Despite their higher costs, theirmore environmentally benign character has led to an interest in zwitterionics and their

Page 24: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

112 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

mixtures with anionics for commercial use. Carboxylate betaines, amine oxides, andaminimides have all been shown to be effective DRAs. The latter have been shownto have good DR properties from 5 to 30◦C [Zhang, 2005], which is in the range ofapplication for district cooling systems, typically 5 to 15◦C.

Anionic Surfactants Early studies by Savins [1967] showed that sodium oleatesoaps with potassium hydroxide and potassium chloride in aqueous solution hadgood DR effectiveness. Increasing the concentration of KCl from 5% to 10% gavebetter drag reduction results. Unfortunately, the soaps precipitate and are ineffectiveas DRAs in the presence of calcium ions, which are present in most aqueous systems.Solutions of anionic surfactants such as SDS and SDBS are not drag reducing. Littleresearch on anionics as DRAs has been carried out.

Mixed-Surfactant DR Systems

Mixtures of Cationic Surfactants with Different Chain Lengths A C22 cationic whichis DR to about 100◦C becomes less soluble and loses its DR ability at about 50◦C.Addition of small amounts of a C12 cationic reduced the lower temperature limitfor drag reduction of the mixture to 2◦C with only minor reduction of the upper-temperature DR limit [Chou, 1991; Lin, 2000; Lin et al., 2000]. This was attributedto the less-ordered arrangement of long and short alkyl chains in the micelle core,which lowered the Krafft temperature [Chou, 1991]. Thus, by changing the relativeconcentrations of surfactants of different chain lengths, the effective DR temperaturerange can be varied. Of course, the C12–C22 mixture lost DR ability at high C12concentration [Lin et al., 2000]. Tomasic et al. [1999] noted that the spacing betweenadjacent chains increases in proportion to the difference in their chain lengths becausea mixture of alkyl chains cannot pack in the hydrocarbon core as efficiently as uniformchains, which increases the packing parameter, promoting micelle growth. Since boththe long- and short-chain surfactant molecules in the mixed micelle must positiontheir polar groups at the micelle surface, cavities exist in the core due to bending andtwisting of the long alkyl chains. This less-ordered packing gives the reduced lowerdrag reduction temperature of the C12–C22 mixture. It also causes some reduction inthe upper temperature limit due to the weaker van der Waals forces [Chou, 1991; Lin,2000; Lin et al., 2000].

Cationic/Anionic Mixtures Qi [2002] studied the drag reduction, rheological behav-ior, and nanostructure of 10 mM solutions of cetyltrimethylammonium tosylate(CTAT) mixed with SDBS. The mixtures had higher solution viscosity, greater vis-coelasticity, and higher upper-temperature limits for drag reduction than did CTAT.Similar results were observed with added sodium dodecyl sulfate as well as withaddition of excess tosylate, but addition of sodium chloride had much smaller effects.Branching points in the TLM networks were shown by cryo-TEM for the CTAT/SDBS(96/4) system with excess tosylate. The presence of branches generally is accompa-nied by higher resistance of the micelle networks to shear degradation [Chellamuthuand Rothstein, 2008] which was observed in this system [Zhang et al., 2005b].

Page 25: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

SURFACTANT DRAG REDUCTION 113

Zwitterionic/Anionic Mixtures Harwigsson and Hellsten [1996] studied the DRbehavior of C16 and C18 sarcosinate zwitterionic surfactants and observed that ata concentration of only 500 ppm, they were effective DRAs from 66 to 98◦C. How-ever, their usefulness is limited, as they tend to crystallize at lower temperatures andpH affects the head-group net charge. They also studied C18 betaines with addedSDBS (at a molar ratio of 6.5 : 1), resulting in a mixture that was pH insensitivefrom 7 to 10.5 and which was DR effective from 50 to 85◦C at concentrations of250 to 2000 ppm. A mixture of 2.5 mM alkylbetaine/0.5 mM SDBS was DR effec-tive from 6 to 60◦C for C16 and from 20 to 100◦C for C18 [Hellsten et al., 1996].Akzo Nobel developed a product for commercial drag reduction use that containeda mix of C16 and C18 betaines and sodium salt of C11–C13 alkyl benzene sulfonate(molar ratio 4 : 1). For a tetradecyltrimethylamino oxide and SDBS mixture, Hu et al.[1993a] found that the induction time for the start of SIS was inversely proportionalto shear rate.

Although there have been theoretical and experimental studies of the interactionsof zwitterionic and anionic surfactants, their mechanisms are still uncertain. Gen-erally, mixtures containing up to 20% anionic surfactant are the most DR effectivecompositions. Denser packing of the mixed-surfactant molecules compared with theindividual components was proposed by Hoffmann et al. [1994] as the reason for thesynergetic effects observed.

Amine oxide zwitterionics are not strongly DR effective. At low temperatures oleyldihydroxyethyl amine oxide (OHAO) shows weak drag reduction from 5 to 15◦C.At high temperatures, behenyl dihydroxyethylamine oxide (BMAO) is weakly DRfrom 60 to 80◦C. However, when 2-lauryl carboxymethyl hydroxyethyl imidazoliumbetaine was added to OHAO at a ratio of 1 to 1.5, at 15◦C DR was increased from55% to 70%. The surfactant also increased drag reduction from 10% to 88% at 80◦Cwhen added to BMAO [Nobuchika et al., 2000].

2.4.3 Solvent Effects

If the freezing point of circulating liquid could be lowered to −5◦C, the useful rangeof recirculating coolant in district cooling systems could be increased compared withthat for water: that is, 5 to 15◦C (∆T = 10◦C) to −5 to 15◦C (∆T = 20◦C). Thismight be accomplished by use of cosolvents such as ethylene glycol and would reduceby about half the mass flow requirements and hence the pumping energy requirementsfor heat removal. The requirements could be further reduced by at least another halfby use of suitable DRAs. However, CMC would be increased and aggregation numberlowered by the addition of a polar nonaqueous solvent such as ethylene glycol becauseof its effects on micelle self-assembly. This would affect DR effectiveness.

Zhang et al. [2005b] studied the effects of various percents of ethylene glycol inwater (15, 20, and 28%) on cationic surfactant solution properties. Using commercialcationics from Akzo Nobel (Ethoquad O12 and O13), they observed that the cosolventreduced the upper temperature limit for drag reduction, maximum percent drag reduc-tion, maximum critical wall shear stress, and relative shear viscosity. The formationof TLMs was hampered, but the addition of excess sodium salicylate promoted TLM

Page 26: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

114 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

formation at lower temperatures, which reduced deleterious effects compared withwater. TLMs at 0◦C were better developed than at 25◦C, and drag reduction improved,although the surfactants were still not as effective in the ethylene glycol/water solventas in water.

The Akzo Nobel alkyl ammonium carboxylate (zwitterionic)/SDBS (molar ratio4 : 1) surfactant mixture is DR in both water and ethylene glycol/water systems.However, better DR results were obtained at lower temperatures by a new alkyl amin-imide zwitterionic in both water and ethylene glycol/water solutions [Zhang et al.,2005b]. This was attributed, in part, to the presence of the oleyl chain, which enhanceslow-temperature solubility.

2.5 PROPOSED POLYMER AND SURFACTANT DRAG REDUCTIONMECHANISMS

Due to the complexity arising from turbulence and non-Newtonian fluid mechanics,the origin and the mechanism of drag reduction are not well understood, despiteextensive efforts. In the late 1960s, three theories were proposed to predict the onsetof polymer drag reduction: length scale [Virk and Merrill, 1969], time scale [Hershey,1965; Fabula et al., 1966; Hershey and Zakin, 1967b; Walsh, 1967], and strain energystorage models [Walsh, 1967]. In the length-scale model it was suggested that onsetoccurs when the ratio of the polymer length scale to the turbulent length scale reacheda certain value. The time-scale model states that drag reduction begins when therelaxation time of the polymer molecule in solution equals a characteristic flow time orsome multiple of it. This is equivalent to a constant Deborah number (a dimensionlessnumber defined as the ratio of a relaxation time, characterizing the intrinsic fluidity of amaterial, and the characteristic time scale of an experiment probing the response of thematerial). It was observed that the time scale onset prediction gave consistent resultsin good solvents but not in poor solvents [Hunston and Zakin, 1980a]. Predictionsby the length-scale model are opposite to certain experimental results [Berman andGeorge, 1974]. The energy storage model suggests that the polymer molecules storeexcess energy of small-scale disturbances near the wall so that these small-scaleinstabilities will not grow. There exists the possibility of converting energy from thehighly strained wall region to the unstrained fluid in the core of the pipe flow. Thismodel predicts that drag reduction begins when strain energy convection becomescomparable to kinetic energy diffusion. Berner and Scrivener measured the velocityfluctuations in polymer DR flow and found that the buffer layer thickness is increasedand the number of small-scale energy dissipative structures is decreased (reductionin high-frequency components of the energy spectra), suggesting that polymer dragreduction is due to changes in energy exchanges mainly in the buffer layer [Bernerand Scrivener, 1980].

Wells and Spangler [1967] and Smith and Tiederman [1991] provided furtherevidence that drag reduction does not occur unless the polymer is present near the wall(y+ < 50). It has also been found that onset for polymer drag reduction is unaffectedby pipe roughness [Virk, 1971; Bewersdorff, 1993]. Based on these observations

Page 27: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

SUMMARY AND FUTURE DEVELOPMENT 115

and some velocity profile measurements, Virk concluded that the region of interestfor drag reduction extends to y+ of about 50 [Virk, 1975]. Virk also suggested thatthe polymer–turbulence interaction responsible for drag reduction takes place in thebuffer zone near the position of peak turbulent energy production, y+ = 15. He thenconcluded that drag reduction occurs because polymer molecules interfere with theturbulent bursting process [Virk, 1975]. However, Bewersdorff [1984] showed that theintroduction of nonhomogeneous polymer threads at the pipe center gave significantdrag reduction even though it was unlikely that any DRA reached the pipe wall. Thisapparent contradiction has not been resolved.

Polymer drag reduction is also associated with the following turbulence observa-tions: The frequency of turbulent eddy bursts decreases as drag reduction increases;the spacing between nondimensionalized fluid streaks increases as drag reductionincreases; high-frequency components of turbulent velocity fluctuations are attenu-ated while low-frequency ones are enhanced; axial turbulence intensity maxima shiftaway from the wall; radial and tangential turbulence intensities are reduced, turbulentshear stresses (Reynolds stresses) are reduced; and so on.

Theoretical studies of the dynamics of self-assemblies of wormlike surfactantmicelles have been reported by a number of investigators, such as Cates and co-workers [Turner and Cates, 1991; Marques et al., 1994]. Since they are subject toreversible scission and recombination, they are called “living polymers.” The con-tinuous breaking and repair of the micellar chains provides more complex solutionbehavior than do reptating polymer chains; that is, their stress relaxation mechanismsare a combination of reptation and breaking followed by reassembly. At low fre-quencies, linear viscoelastic (Maxwell) behavior is predicted and observed for somesurfactant systems. However, non-Maxwell behavior was observed in Cole–Cole plotsof a number of cationic surfactant systems [Lu, 1997; Lin, 2000].

Other proposed drag reduction mechanisms for both types of additives includehigh extensional viscosity, providing additional resistance to vortex stretching andturbulent eddy growth, and viscoelasticity, which dampens small turbulent eddiesand stores turbulent fluctuation energy that would otherwise be dissipated. Recently,a small number of researchers have suggested that chain entanglements or association/aggregation of polymer structures can be important or even predominant in dilute orsemidilute polymer solution’s drag-reducing ability [Dunlop and Cox, 1977; Kowaliket al., 1987; Kowalik, 1988; Malik et al., 1993; Liberatore et al., 2003; Vlachogiannisand Hanratty, 2004].

2.6 SUMMARY AND FUTURE DEVELOPMENT

Turbulent drag reduction in homogeneous flows by polymer or surfactant additivesis a striking phenomenon with both theoretical and practical implications. On thetheoretical side, it remains a challenge to fully understand the drag reduction mech-anism and the interaction details between DRAs and the turbulent flow field. Newmethods, especially computational ones, have been developed to solve this problem,such as direct numerical simulations and stochastic simulations. On the application

Page 28: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

116 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

side, polymer DRAs have been used routinely in crude oil and hydrocarbon producttransport and oil production and operations. Surfactant DRAs have also been usedsuccessfully in small-scale DHC systems. Methods for improving polymer DRAs’shear stability, such as by grafting or mixing different polymer species, are beingdeveloped. Also, novel surfactant DRAs are being synthesized and tested to satisfythe needs of targeted application areas.

REFERENCES

Agoston, G. A., Harte, W. H., Hottel, H. C., Klemm, W. A., Mysels, K. J., Pomeroy, H. H., andThompson, J. M., Flow of gasoline thickened by napalm, J. Ind. Eng. Chem., 46, 1017–1019(1954).

Aguilar, G., Gasljevic, K., and Matthys, E. F., On the relationship between heat transfer anddrag reduction for polymer and surfactant solutions, HTD (Am. Soc. Mech. Eng.), 361(1),139–146 (1998).

Aguilar, G., Gasljevic, K., and Matthys, E. F., Coupling between heat and momentum transfermechanisms for drag-reducing polymer and surfactant solutions, J. Heat Transfer, 121, 796–802 (1999).

Bello, J. B., Mueller, A. J., and Saez, A. E., Effect of intermolecular crosslinks on drag reductionby polymer solutions, Polym. Bull. (Berlin), 36, 111–118 (1996).

Beris, A. N., Dimitropoulos, C. D., Sureshkumar, R., and Handler, R. D., Direct numericalsimulations of polymer-induced drag reduction in viscoelastic turbulent channel flows, inProceedings of the 13th International Congress on Rheology, Cambridge, UK, Aug. 20–25,2000, Vol. 2, pp. 190–192.

Berman, N. S., Drag reduction by polymers, Annu. Rev. Fluid Mech., 10, 47–64 (1978).

Berman, N. S., and George, W. K., Jr., Onset of drag reduction in dilute polymer solutions,Phys. Fluids, 17, 250–251 (1974).

Berner, C., and Scrivener, O., Drag reduction and structure of turbulence in dilute polymersolutions. in viscous flow drag reduction, in Progress in Astronautics and Aeronautics, Vol.72, Hough, G. R., Ed., American Institute of Aeronautics and Astronautics, New York, 1980,pp. 290–299.

Bewersdorff, H.-W., Heterogeneous drag reduction in turbulent pipe flows, Rheol. Acta, 23,522–543 (1984).

Bewersdorff, H.-W., Drag reduction in surfactant solutions, in Structure of Turbulence andDrag Reduction, Gyr, A., Ed., Springer-Verlag, Berlin, 1990, p. 293.

Bewersdorff, H.-W., Rheology of drag reducing surfactant solutions, Proc. ASME Fluids Eng.Div., 237, 25–29 (1996).

Bewersdorff, H.-W., and Ohlendorf, D., The behavior of drag-reducing cationic surfactantsolutions, Colloid Polym. Sci., 266, 941–953 (1988).

Bewersdorff, H.-W., Gyr, A., Hoyer, K., and Tsinober, A., An investigation of possible mecha-nisms of heterogeneous drag reduction in pipe and channel flows, Rheol. Acta, 32, 140–149(1993).

Bijma, K., Rank, E., and Engberts, J. B. F. N., Effect of counterion structure on micellar growthof alkylpyridinium surfactants in aqueous solution, J. Colloid Interface Sci., 205, 245–256(1998).

Page 29: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 117

Broniarz-Press, L., Rozanski, J., and Rozanska, S., Drag reduction effect in pipe systems andliquid falling film flow, Rev. Chem. Eng., 23, 149–245 (2007).

Brostow, W., Drag reduction and mechanical degradation in polymer solutions in flow, Polymer,24, 631–638 (1983).

Brostow, W., Ertepinar, H., and Singh, R. P., Flow of dilute polymer solutions: chain confor-mations and degradation of drag reducers, Macromolecules, 23, 5109–5118 (1990).

Brostow, W., Lobland, H. E. H., Reddy, T., Singh, R. P., and White, L., Lowering mechanicaldegradation of drag reducers in turbulent flow, J. Mater. Res., 22, 56–60 (2007).

Castro, W., and Squire, W., Effect of polymer additives on transition in pipe flow, Appl. Sci.Res., 18, 81–96 (1967).

Cates, M. E., and Candau, S. J., Statics and dynamics of worm-like surfactant micelles, J. Phys.Condens. Matter, 2, 6869–6892 (1990).

Chang, H. F. D., and Darby, R., Effect of shear degradation on the rheological properties ofdilute drag-reducing polymer solutions, J. Rheol., 27, 77–88 (1983).

Chang, R. C., and Zakin, J. L., Drag reduction of non-ionic surfactant mixtures, in The Influenceof Polymer Additives on Velocity and Temperature Fields. Symposium Universitat-GH, Essen,Germany, Gampert, B., Ed., Springer-Verlag, Berlin, 1985, pp. 61–70.

Chara, Z., Zakin, J. L., Severa, M., and Myska, J., Turbulence measurements of drag reducingsurfactant systems, Exp. Fluids, 16, 36–41 (1993).

Chellamuthu, M., and Rothstein, J. P., Distinguishing between linear and branched worm-like micelle solutions using extensional rheology measurements, J. Rheol., 52, 865–884(2008).

Choi, H. J., Kim, C. A., Sohn, J.-I., and Jhon, M. S., An exponential decay function forpolymer degradation in turbulent drag reduction, Polym. Degrad. Stabil., 69, 341–346(2000a).

Choi, H. J., Kim, C. A., Vinay, S. J., and Jhon, M. S., Effect of degradation on polymer-inducedturbulent drag reduction, Book of Abstracts, 219th ACS National Meeting, San Francisco,Mar. 26–30, 2000, POLY-184, 2000b.

Choi, H. J., Lim, S. T., Lai, P.-Y., and Chan, C. K., Turbulent drag reduction and degradationof DNA, Phys. Rev. Lett., 89, 088302/1–088302/4 (2002).

Chou, L. C., Drag-reducing cationic surfactant solutions for district heating and cooling sys-tems, Ph.D. dissertation, Ohio State University, Columbus, 1991.

Chou, L. C., Christensen, R. N., and Zakin, J. L., Drag reducing additives for district cool-ing systems, in Proceedings of the International District Heating and Cooling Conference,Virginia Beach, VA, 1989, pp. 352–367.

Chou, L. C., Christensen, R. N., and Zakin, J. L., The influence of chemical composition ofquaternary ammonium salt cationic surfactants on their drag reducing effectiveness, in DragReduction in Fluid Flows, Sellin, R. H. J., and Moses, J. T., Eds., Ellis Horwood, Chichester,UK, 1989b, pp. 141–148.

Dembek, G., and Bewersdorff, H.-W., Short-time increase of sewer capacity by addition ofwater-soluble polymers, Das Wasser./Abwasser., 122, 392–395 (1981).

DeRoussel, P., Nonionic surfactant drag reduction, B.S. honors thesis, Department of ChemicalEngineering, Ohio State University, Columbus, 1993.

Deshmukh, S. R., and Singh, R. P., Drag reduction effectiveness, shear stability and biodegra-dation resistance of guar gum-based graft copolymers, J. Appl. Polym. Sci., 33, 1963–1975(1987).

Page 30: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

118 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

Deshmukh, S. R., Sudhakar, K., and Singh, R. P., Drag-reduction efficiency, shear stabil-ity, and biodegradation resistance of carboxymethyl cellulose-based and starch-based graftcopolymers, J. Appl. Polym. Sci., 43, 1091–1101 (1991).

Dimitropoulos, C. D., Sureshkumar, R., and Beris, A. N., Direct numerical simulation ofviscoelastic turbulent channel flow exhibiting drag reduction: effect of the variation ofrheological parameters, J. Non-Newtonian Fluid Mech., 79, 433–468 (1998).

Dingilian, G., and Ruckenstein, E., Positive and negative deviations from additivity in dragreduction of binary dilute polymer solutions, AIChE J., 20, 1222–1224 (1974).

Drappier, J., Divoux, T., Amarouchene, Y., Bertrand, F., Rodts, S., Cadot, O., Meunier, J., andBonn, D., Turbulent drag reduction by surfactants, Europhys. Lett., 74, 362–368 (2006).

Dujmovich, T., and Gallegos, A., Drag reducers improve throughput, cut costs, Offshore, 65,1–4 (2005).

Dunlop, E. H., and Cox, L. R., Influence of molecular aggregates on drag reduction, Phys.Fluids, 20, S203–S213 (1977).

Fabula, A. G., Fire-fighting benefits of polymeric friction reduction, J. Basic Engi., 93, 453–455(1971).

Fabula, A. G., Lumley, J. L., and Taylor, W. D., Some interpretations of the Toms effect, inModern Developments in the Mechanics Continua, Eskanazi, S., Ed., Academic Press, NewYork, 1966, p. 145.

Figueredo, R. C. R., and Sabadini, E., Firefighting foam stability: the effect of the drag reducerpoly(ethylene) oxide, Colloids Surf. A, 215, 77–86 (2003).

Fullerton, G., and McComb, D., Investigation of isotropic non-Newtonian fluid turbulence usingdirect numerical simulation, in Turbulence and Shear Flow Phenomena 1, First InternationalSymposium, Santa Barbara, CA, Sept. 12–15, 1999, pp. 633–638.

Gampert, B., and Rensch, A., Polymer concentration and near wall turbulence structure ofchannel flow of polymer solutions, Proc. ASME Fluids Eng. Div., 237, 129–136 (1996).

Gasljevic, K., An experimental investigation of drag reduction by surfactant solutions andof its implementation in hydronic systems (turbulent flow, heat transfer), Ph.D. disserta-tion, Department of Mechanical Engineering. University of California at Santa Barbara,1995.

Gasljevic, K., Hoyer, K., and Matthys, E. F., Field test of a drag-reducing surfactant additivein the hydronic cooling system of a building: Phase 2. Heat transfer control, HTD (Am. Soc.Mech. Eng.), 361(3), 255–263 (1998).

Gasljevic, K., Aguilar, G., and Matthys, E. F., On two distinct types of drag-reducing fluids,diameter scaling, and turbulent profiles, J. Non-Newtonian Fluid Mech., 96, 405–425 (2001).

Gasljevic, K., Hoyer, K., and Matthys, E. F., Temporary degradation and recovery of drag-reducing surfactant solutions, J. Rheol., 51, 645–667 (2007).

Ge, W., Studies on the nanostructure, rheology and drag reduction characteristics of dragreducing cationic surfactant solutions, Ph.D. dissertation, Ohio State University, Columbus,2008.

Ge, W., Kesselman, E., Talmon, Y., Hart, D. J., and Zakin, J. L., Correlation of micellarnanostructure, drag reduction and rheological behavior of dilute CTAC–halogen benzoatecounterion systems with different halogen substitutions and counterion to CATC ratio, J.Non-Newtonian Fluid Mech., 154, 1–12 (2008).

Giles, W. B., and Pettit, W. T., Stability of dilute viscoelastic flows of polymers, Nature(London), 216, 470–472 (1967).

Page 31: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 119

Gold, P. I., Amar, P. K., and Swaidan, B. E., Friction reduction degradation in dilutepoly(ethylene oxide) solutions, J. Appl. Polym. Sci., 17, 333–350 (1973).

Golda, J., Hydraulic transport of coal in pipes with drag reducing additives, Chem. Eng.Commun., 43, 53–67 (1986).

Graham, M. D., Drag reduction in turbulent flow of polymer solutions, in Rheology Reviews2004. British Society of Rheology, Binding, D. M., and Walters, K., Eds., 143–170 (2004).

Gravsholt, S., Viscoelasticity in highly dilute aqueous solutions of pure cationic detergents, J.Colloid Interface Sci., 57, 575–577 (1976).

Gyr, A., and Bewersdorff, H.-W., Drag Reduction of Turbulent Flows by Additives, KluwerAcademic, Dordrecht, The Netherlands, 1995.

Hand, J. H., and Williams, M. C., Effect of secondary polymer structure on the drag-reducingphenomenon, J. Appl. Polym. Sci., 13, 2499–2503 (1969).

Hansen, R. J., and Little, R. C., Pipe diameter, molecular weight, and concentration effects onthe onset of drag reduction, Chem. Eng. Prog. Symp. Ser., 67, 93–97 (1971).

Harwigsson, I., and Hellsten, M., Environmentally acceptable drag-reducing surfactants fordistrict heating and cooling, J. Am. Oil Chem. Soc., 73, 921–928 (1996).

Hellsten, M., and Harwigsson, I., Use of a betaine surfactant together with an anionic surfactantas a drag-reducing agent, WO patent application, 96-EP950, 9628527, 1996.

Hellsten, M., and Oskarsson, H., A drag-reducing agent for use in injection water at oil recovery,WO patent application, 2003-SE1015, 2004007630, 2004.

Hellsten, M., Harwigsson, I., Blais, C., and Wollerstrand, J., Drag reduction by N-alkylbetaines:a type of zwitterionic surfactants, Proc. ASME Fluids Eng. Div., 237, 37–46 (1996).

Hershey, H. C., Drag reduction in Newtonian polymer solutions, Ph.D. dissertation, Universityof Missouri–Rolla, 1965.

Hershey, H. C., and Zakin, J. L., The existence of two types of drag reduction in pipe flow ofdilute polymer solutions, Ind. Eng. Chem. Fundam., 6, 381–387 (1967a).

Hershey, H. C., and Zakin, J. L., Molecular approach to predicting the onset of drag reduc-tion in the turbulent flow of dilute polymer solutions, Chem. Eng. Sci., 22, 1847–1857(1967b).

Hershey, H. C., Kuo, J. T., and McMillan, M. L., Drag reduction of straight and branched chainaluminum disoaps, Ind. Eng. Chem. Prod. Res. Dev., 14, 192–199 (1975).

Hoffmann, H., Schwandner, B., Ulbricht, W., and Zana, R., The transition from isotropic micel-lar solutions to lamellar phases, in Proceedings of the International School of Physics EnricoFermi, Varenna on Lake Como, Degiorgio, V., and Corti, M., Eds., Elsevier, Amsterdam,1985, pp. 261–280.

Hoffmann, H., Hofmann, S., and Illner, J. C., Phase behavior and properties of micellar solutionsof mixed zwitterionic and ionic surfactants, Prog. Colloid Polym. Sci., 97, 103–109 (1994).

Horiuchi, T., Majima, T., Yoshii, T., and Tamura, T., Effect of chemical structure of quaternaryammonium salt-type cationic surfactants and temperature on aggregate size, Nippon KagakuKaishi, 7, 423–427 (2001a).

Horiuchi, T., Yoshii, T., Majima, T., Tamura, T., and Sugawara, H., Effect of alkyl chain lengthand number of 2-hydroxyethyl groups on drag reduction behaviors of quaternary ammoniumsalt-type cationic surfactant solutions, Nippon Kagaku Kaishi, 415–421 (2001b).

Hoyt, J. W., Friction reduction as an estimator of molecular weight, J. Polym. Sci. B, 4, 713–716(1966).

Page 32: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

120 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

Hoyt, J. W., Turbulent-flow interactions and drag reduction, AIP Conf. Proc., 137, 95–115(1985).

Hoyt, J. W., Drag reduction. in Encyclopedia of Polymer Science and Engineering, Vol. 5,Kroschwitz, J., Mark, H. F., Bikales, N. M., Overberger, C. G., and Menges, G., Eds., Wiley,New York, 1986, 129–151.

Hu, Y., Wang, S. Q., and Jamieson, A. M., Kinetic studies of a shear thickening micellarsolution, J. Colloid Interface Sci., 156, 31–37 (1993a).

Hu, Y., Wang, S. Q., and Jamieson, A. M., Rheological and flow birefringence studies of ashear-thickening complex fluid: a surfactant model system, J. Rheol., 37, 531–546 (1993b).

Hu, Y., Rajaram, C. V., Wang, S. Q., and Jamieson, A. M., Shear thickening behavior of arheopectic micellar solution: salt effects, Langmuir, 10, 80–85 (1994).

Hunston, D. L., Effects of molecular weight distribution in drag reduction and shear degradation,J. Polym. Sci. Polym. Chem. Ed., 14, 713–727 (1976).

Hunston, D. L., and Reischman, M. M., The role of polydispersity in the mechanism of dragreduction, Phys. Fluids, 18, 1626–1629 (1975).

Hunston, D. L., and Zakin, J. L., Effect of molecular parameters on the flow rate dependence ofdrag reduction and similar phenomena, in Progress in Astronautics and Aeronautics, Vol. 72,Hough, S. R., Ed., American Institute of Aeronautics and Astronautics, New York, 1980a,pp. 373–385.

Hunston, D. L., and Zakin, J. L., Flow-assisted degradation in dilute polystyrene solutions,Polym. Eng. Sci., 20, 517–523 (1980b).

Israelachvili, J. N., Intermolecular and Surface Forces, Academic Press, San Diego, CA, 1991.

Kalashnikov, V. N., Degradation accompanying turbulent drag reduction by polymer additives,J. Non-Newtonian Fluid Mech., 103, 105–121 (2002).

Kameneva, M. V., Wu, Z. J., Uraysh, A., Repko, B., Litwak, K. N., Billiar, T. R., Fink, M. P.,Simmons, R. L., Griffith, B. P., and Borovetz, H. S., Blood soluble drag-reducing polymersprevent lethality from hemorrhagic shock in acute animal experiments, Biorheology, 41,53–64 (2004).

Kawaguchi, Y., Tawaraya, Y., Yabe, A., Hishida, K., and Maeda, M., Active control of turbulentdrag reduction in surfactant solutions by wall heating, Proc. ASME Fluids Eng. Div., 237,47–52 (1996).

Kawaguchi, Y., Yu, B., Wei, J., and Feng, Z., Rheological characterization of drag-reducingcationic surfactant solution: shear and elongational viscosities of dilute solutions, in Pro-ceedings of the 4th ASME/JSME Joint Fluids Engineering Conference, Honolulu, HI, 2003,pp. 721–728.

Kim, O. K., Little, R. C., and Ting, R. Y., Polymer structural effects in turbulent drag reduction,AIChE Symp. Ser., 69, 39–44 (1973).

Koehler, R. D., Raghavan, S. R., and Kaler, E. W., Microstructure and dynamics of wormlikemicellar solutions formed by mixing cationic and anionic surfactants, J. Phys. Chem. B, 104,1, 1035–11044 (2000).

Kowalik, R. M., Drag reduction with interpolymer complexes, Polym. Prepr. (Am. Chem. Soc.,Div. Polym. Chem.), 29, 447–448 (1988).

Kowalik, R. M., Duvdevani, I., Peiffer, D. G., Lundberg, R. D., Kitano, K., and Schulz, D. N.,Enhanced drag reduction via interpolymer associations, J. Non-Newtonian Fluid Mech., 24,1–10 (1987).

Page 33: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 121

Kreke, P. J., Magid, L. J., and Gee, J. C., 1H and 13C NMR studies of mixed counterion,cetyltrimethylammonium bromide/cetyltrimethylammonium dichlorobenzoate, surfactantsolutions: the intercalation of aromatic counterions, Langmuir, 12, 699–705 (1996).

Kulicke, W. M., Koetter, M., and Graeger, H., Drag reduction phenomenon with special empha-sis on homogeneous polymer solutions, Adv. Polym. Sci., 89, 1–68 (1989).

Landahl, M. T., Drag reduction by polymer addition, in Proceedings of the 13th Interna-tional Congress on Theoretical and Applied Mechanics (ICTAM), Moscow (USSR), 21–26August 1972, pp. 177–199, Becker, E., and Mikhailov, G. K., Eds., Springer-Verlag, Berlin(Germany), 1973.

Laughlin, R. G., The Aqueous Phase Behavior of Surfactants, Academic Press, Harcourt Brace& Company, London, 1996.

Levinthal, C., and Davison, P. F., Degradation of deoxyribonucleic acid (DNA) under hydro-dynamic shearing forces, J. Mol. Biol., 3, 674–683 (1961).

Li, F., Li, G.-Z., and Chen, J.-B., Synergism in mixed zwitterionic-anionic surfactant solu-tions and the aggregation numbers of the mixed micelles, Colloids Surf. A, 145, 167–1741998.

Liaw, G. C., Effect of polymer structure on drag reduction in nonpolar solvents, Ph.D. disser-tation, University of Missouri–Rolla, 1969.

Liaw, G.-C., Zakin, J. L., and Patterson, G. K., Effects of molecular characteristics of polymerson drag reduction, AIChE J., 17, 391–397 (1971).

Liberatore, M. W., Pollauf, E. J., and McHugh, A. J., Shear-induced structure formation insolutions of drag reducing polymers, J. Non-Newtonian Fluid Mech., 113, 193–208 (2003).

Liberatore, M. W., Baik, S., McHugh, A. J., and Hanratty, T. J., Turbulent drag reduction ofpolyacrylamide solutions: effect of degradation on molecular weight distribution, J. Non-Newtonian Fluid Mech., 123, 175–183 (2004).

Lin, Z., The effect of chemical structures of cationic surfactants or counterions on solution dragreduction effectiveness, rheology and micellar microstructures, Ph.D. dissertation, Ohio StateUniversity, Columbus, 2000.

Lin, Z., Chou, L. C., Lu, B., Zheng, Y., Davis, H. T., Scriven, L. E., Talmon, Y., and Zakin, J.L., Experimental studies on drag reduction and rheology of mixed cationic surfactants withdifferent alkyl chain lengths, Rheol. Acta, 39, 354–359 (2000).

Lin, Z., Lu, B., Zakin, J. L., Talmon, Y., Zheng, Y., Davis, H. T., and Scriven, L. E., Influence ofsurfactant concentration and counterion to surfactant ratio on rheology of wormlike micelles,J. Colloid Interface Sci., 239, 543–554 (2001).

Lin, Z., Mateo, A., Zheng, Y., Kesselman, E., Pancallo, E., Hart, D. J., Talmon, Y., Davis, H. T.,Scriven, L. E., and Zakin, J. L., Comparison of drag reduction, rheology, microstructure andstress-induced precipitation of dilute cationic surfactant solutions with odd and even alkylchains, Rheol. Acta, 41, 483–492 (2002).

Little, R. C., Hansen, R. J., Hunston, D. L., Kim, O.-K., Patterson, R. L., and Ting, R.Y., Drag reduction phenomenon: observed characteristics, improved agents, and proposedmechanisms, Ind. Eng. Chem. Fundam., 14, 283–296 (1975).

Lodes, A., and Macho, V., The influence of poly(vinyl acetate) additive in water on turbulentvelocity field and drag reduction, Exp. Fluids, 7, 383–387 (1989).

Lu, B., Characterization of drag reducing surfactant systems by rheology and flow birefringencemeasurements, Ph.D. dissertation, Ohio State University, Columbus, 1997.

Page 34: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

122 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

Lu, B., Li, X., Zakin, J. L., and Talmon, Y., A non-viscoelastic drag reducing cationic surfactantsystem, J. Non-Newtonian Fluid Mech., 71, 59–72 (1997).

Lu, B., Li, X., Scriven, L. E., Davis, H. T., Talmon, Y., and Zakin, J. L., Effect of chemicalstructure on viscoelasticity and extensional viscosity of drag-reducing cationic surfactantsolutions, Langmuir, 14, 8–16 (1998a).

Lu, B., Zheng, Y., Davis, H. T., Scriven, L. E., Talmon, Y., and Zakin, J. L., Effect of variationsin counterion to surfactant ratio on rheology and microstructures of drag reducing cationicsurfactant systems, Rheol. Acta, 37, 528–548 (1998b).

Lumley, J. L., Drag reduction by additives, Annu. Rev. Fluid Mech., 1, 367–384 (1969).

Malik, S., Shintre, S. N., and Mashelkar, R. A., Enhancing the shear stability in drag-reducingpolymers through molecular associations, Macromolecules, 26, 55–59 (1993).

Manfield, P. D., Lawrence, C. J., and Hewitt, G. F., Drag reduction with additives in multiphaseflow: a literature survey, Multiphase Sci. Technol., 11, 197–221 (1999).

Marques, C. M., Turner, M. S., and Cates, M. E., Relaxation mechanisms in worm-like micelles,J. Non-Cryst. Solids, 172–174, 1168–1172 (1994).

McCormick, C. L., Hester, R. D., Morgan, S. E., and Safieddine, A. M., Water-soluble copoly-mers: 31. Effects of molecular parameters, solvation, and polymer associations on dragreduction performance, Macromolecules, 23, 2132–2139 (1990).

McMillan, M. L., Drag reduction and light scattering studies of aluminum disoaps in toluene,Ph.D. dissertation. Ohio State University, Columbus, 1970.

McMillan, M. L., Hershey, H. C., and Baxter, R. A., Effects of aging, concentration, tempera-ture, method of preparation, and other variables on the drag reduction of aluminum disoapsin toluene, Chem. Eng. Prog. Symp. Ser., 67, 27–44 (1971).

Minoura, Y., Kasuya, T., Kawamura, S., and Nakano, A., Degradation of poly(ethylene oxide)by high-speed stirring, J. Polym. Sci., Polym. Phys. Ed., 5, 125–142 (1967).

Mitsoulis, E., Computational rheology with integral constitutive equations, Appl. Rheol., 9,198–203 (1999).

Mostardi, R. A., Thomas, L. C., Greene, H. L., VanEssen, F., and Nokes, R. F., Suppression ofatherosclerosis in rabbits using drag reducing polymers, Biorheology, 15, 1–14 (1978).

Motier, J. F., and Carreir, A. M., Recent studies on polymer drag reduction in commercialpipelines, in Drag Reduction in Fluid Flows: Techniques for Friction Control, Sellin, R., andMoses, R., Eds., Ellis Horwood, Chichester, UK, 1989, pp. 197–204.

Motier, J. F., Chou, L. C., and Kommareddi, N. S., Commercial drag reduction: past, present,and future, in 1996 ASME Fluids Engineering Division Summer Meeting, San Diego, CA,Coleman, H., Ed., ASME, New York, 1996, pp. 229–234.

Moussa, T., and Tiu, C., Factors affecting polymer degradation in turbulent pipe flow, Chem.Eng. Sci., 49, 1681–1692 (1994).

Mumick, P. S., and McCormick, C. L., Hydrophobically associating water soluble copolymersfor the study of drag reduction, ANTEC-1992, Proceedings of the Society of Plastic Engineers50th Annual Technical Conference, Detroit, Michigan, USA, May 4–7, 1992, 50, 2095–2099(1992).

Mysels, K. J., Flow of thickened fluids, U.S. patent 2,492,173, 1949.

Myska, J., and Chara, Z., New view of the surfactant drag reduction ability, in 11th Euro-pean Drag Reduction Working Meeting, Prague, Czech Republic, Chara, Z., Ed., Institute ofHydrodynamics, AS, Czech Republic, 1999.

Page 35: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 123

Myska, J., Zakin, J. L., and Chara, Z., Viscoelasticity of a surfactant and its drag-reducingability, Appl. Sci. Res., 55, 297–310 (1996).

Nadolink, R. H., and Haigh, W. W., Bibliography on skin friction reduction with polymers andother boundary-layer additives, Appl. Mech. Rev., 48, 351–460 (1995).

Nijs, L., New generation drag reducer, in 2nd International Pipeline Technology Conference,Ostend, Belgium, Denys, R., Ed., Elsevier, Amsterdam, 1995, pp. 143–149.

Nobuchika, K., Nakata, T., Sato, K., Yamagishi, F., Tomiyama, S., Inaba, H., Horibe, A., andHaruki, N., Drag reduction and heat transfer enhancement of water flow in a pipe usingnew surfactants for environmentally acceptable drag reducing additives, Proceedings of theSymposium on Energy Engineering in the 21st Century, Hong-Kong, Cheng, P., Ed., 2000,pp. 695–700.

Ohlendorf, D., Interthal, W., and Hoffmann, H., Surfactant systems for drag reduction: physic-ochemical properties and rheological behavior, Rheol. Acta, 25, 468–486 (1986).

Park, S. R., Suh, H. S., Moon, S. H., and Yoon, H. K., Pump and temperature effects on andflow characteristics of drag reducing surfactants, Proc. ASME Fluids Eng. Div., 237, 177–182(1996).

Paterson, R. W., and Abernathy, F. H., Turbulent flow drag reduction and degradation withdilute polymer solutions, J. Fluid Mech., 43, 689–710 (1970).

Patterson, G. K., Hershey, H. C., Green, C. D., and Zakin, J. L., Effect of degradation bypumping on normal stresses in polyisobutylene solution, Trans. Soc. Rheol., 10, 489–500(1966).

Patterson, G. K., Zakin, J. L., and Rodriguez, J. M., Drag reduction: polymer solutions, soapsolutions, and solid particle suspensions in pipe flow, Ind. Eng. Chem., 61, 22–30 (1969).

Patterson, G. K., Shen, A.-M., and Chen, I.-M., Rates of shear degradation of drag reducingpolymer, in 3rd International Conference on Drag Reduction, University of Bristol, Bristol,UK, Sellin, R., and Moses, R., Eds., 1984, pp. E4-1, E4-6.

Pollert, J., and Sellin, R. J., Mechanical degradation of drag reducing polymer and surfactantadditives: a review, in Drag Reduction in Fluid Flows: Techniques for Friction Control,Sellin, R., and Moses, R., Eds., Ellis Horwood, Chichester, UK, 1989, pp. 179–188.

Pollert, J., Zakin, J. L., Myska, J., and Kratochvil, P., Use of friction reducing additives indistrict heating system field test at Kladno-Krocehlavy, Czech Republic, in Proceedings ofthe 85th International District Heating and Cooling Association, Seattle, 1994, pp. 141–156.

Qi, Y., Investigation of relationships among microstructure, rheology, drag reduction and heattransfer of drag reducing surfactant solutions, Ph.D. dissertation, Ohio State University,Columbus, 2002.

Qi, Y., and Zakin, J. L., Chemical and rheological characterization of drag-reducing cationicsurfactant systems, Ind. Eng. Chem. Res., 41, 6326–6336 (2002).

Rakitin, A. R., and Pack, G. R., Necessity of aromatic carboxylate anions to be planar to inducegrowth of cationic micelles, Langmuir, 21, 837–840 (2005).

Reddy, G. V., and Singh, R. P., Drag reduction effectiveness and shear stability of polymer–polymer and polymer–fiber mixtures in recirculatory turbulent flow of water, Rheol. Acta,24, 296–311 (1985).

Rehage, H., and Hoffmann, H., Viscoelastic surfactant solutions: model systems for rheologicalresearch, Mol. Phys., 74, 933–973 (1991).

Page 36: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

124 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

Rehage, H., Hoffmann, H., and Wunderlich, I., A rheological switch: shear induced phasetransitions in aqueous surfactant solutions, Ber. Bunsenges., 90, 1071–1075 (1986).

Rochefort, W. E., and Middleman, S., Rheology of xanthan gum: salt, temperature, and straineffects in oscillatory and steady shear experiments, J. Rheol., 31, 337–369 (1987).

Rose, G. D., and Foster, K. L., Drag reduction and rheological properties of cationic viscoelasticsurfactant formulations, J. Non-Newtonian Fluid Mech., 31, 59–85 (1989).

Salkar, R. A., Mukesh, D., Samant, S. D., and Manohar, C., Mechanism of micelle to vesicletransition in cationic–anionic surfactant mixtures, Langmuir, 14, 3778–3782 (1998).

SaPereira, A., and Pinho, F. T., Turbulent pipe flow characteristics of low molecular weightpolymer solutions, J. Non-Newtonian Fluid Mech., 55, 321–344 (1994).

Savins, J. G., A Stress controlled drag reduction phenomenon, Rheol. Acta, 6, 323–330(1967).

Schubert, B. A., Kaler, E. W., and Wagner, N. J., The microstructure and rheology of mixedcationic/anionic wormlike micelles, Langmuir, 19, 4079–4089 (2003).

Sellin, R., Drag reduction in sewers: first results from a permanent installation, J. Hydraul.Res., 16, 351–371 (1978).

Sellin, R. H. J., and Loeffler, E. J., Drag reduction measurements with poly(acrylic acid)under different solvent pH and salt conditions, Paper C2, in Proceedings of the 2nd Inter-national Conference on Drag Reduction, Cambridge University, Cambridge, UK, Clarke, J.A., Ed., British Hydromechanics Research Association Fluid Engineering, Cranfield, UK,1977, pp. 11–20.

Shapley, W. A., Bacskay, G. B., and Warr, G. G., Ab initio quantum chemical studies of the pKa’sof hydroxybenzoic acids in aqueous solution with special reference to the hydrophobicityof hydroxybenzoates and their binding to surfactants, J. Phys. Chem. B, 102, 1938–1944(1998).

Shenoy, A. V., A review on drag reduction with special reference to micellar systems, ColloidPolym. Sci., 262, 319–337 (1984).

Simha, R., and Zakin, J. L., Compression of flexible chain molecules in solution, J. Chem.Phys., 33, 1791–1793 (1960).

Simha, R., and Zakin, J. L., Solution viscosities of linear, flexible, high polymers, J. ColloidSci., 17, 270–287 (1962).

Singh, R. P., Drag reduction and shear stability mechanism, in Encyclopedia of Fluid Mechanics,Vol. 9, Gulf Publishing, Houston, TX, 1990, pp. 425–480.

Singh, R. P., Turbulent drag reduction by polymer-based mixtures and graft copolymers, Curr.Sci., 68, 736–739 (1995).

Singh, R. P., Singh, J., Deshmukh, S. R., and Kumar, A., The effect of grafted and ungraftedguar gum on turbulent flow of water and on hydraulic conductivity of soil, in: Proceedings ofthe IUTAM Symposium on the Influence of Polymer Addititives on Velocity and TemperatureFields, Essen, Germany, 1984, Gampert, B., Ed., Springer-Verlag, Berlin, 1985, pp. 131–139.

Singh, R. P., Jain, S. K., and Lan, N., Drag reduction, flocculation and rheological characteristicsof grafted polysaccharides, Polym. Sci., 2, 716–721 (1991).

Smith, B. C., Flow birefringence, nuclear magnetic resonance and corrosion measurements ondrag-reducing cationic surfactant solutions for district heating and cooling systems, Ph.D.dissertation, Ohio State University, Columbus, 1992.

Page 37: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 125

Smith, B. C., Chou, L. C., and Zakin, J. L., Measurement of the orientational binding of coun-terions by nuclear magnetic resonance measurements to predict drag reduction in cationicsurfactant micelle solutions, J. Rheol., 38, 73–83 (1994).

Smith, R. E., and Tiederman, W. G., The mechanism of polymer thread drag reduction, Rheol.Acta, 30, 103–113 (1991).

Steiff, W. A., Groth, S., Helmmig, J., Kleuker, H. H., Wocadlo, T., and Weinspach, P. W.,Progress in the application of drag reducing additives in district heating systems, in Interna-tional Symposium on Fluids for District Heating, Copenhagen, Denmark, Bohm, B., Eds.,1991, pp. 35.

Sullivan, P., Nelson, E. B., Anderson, V., and Hughes, T., Oilfield applications of giant micelles,in Giant Micelles: Properties and Applications, Zana, R., and Kaler, E. W., Eds., CRC Press,Taylor & Francis, New York, 140, 2007, pp. 453–472.

Sung, J. H., Kim, C. A., Choi, H. J., Hur, B. K., Kim, J. G., and Jhon, M. S., Turbulent dragreduction efficiency and mechanical degradation of poly(acrylamide), J. Macromol. Sci.Phys., B43, 507–518 (2004).

Tomasic, V., Stefanic, I., and Filipovic-Vincekovic, N., Adsorption, association, and precipi-tation in hexadecyltrimethylammonium bromide/sodium dodecyl sulfate mixtures, ColloidPolym. Sci., 277, 153–163 (1999).

Toms, B. A., Some observations on the flow of linear polymer solutions through straight tubesat large Reynolds numbers, in Proceedings of the 1st International Rheology Congress,Scheveningen, The Netherlands, North-Holland, Amsterdam, 1949, pp. 135–141, 47–49,51–52.

Turner, M. S., and Cates, M. E., Linear viscoelasticity of living polymers: a quantitative probeof chemical relaxation times, Langmuir, 7, 1590–1594 (1991).

Underwood, A. L., and Anacker, E. W., Organic counterions and micellar parameters: sub-stituent effects in a series of benzoates, J. Phys. Chem., 88, 2390–2393 (1984).

Ungeheuer, S., Bewersdorff, H.-W., and Singh, R. P., Turbulent drag effectiveness and shearstability of xanthan-gum-based graft copolymers, J. Appl. Polym. Sci., 37, 2933–2948 (1989).

Unthank, J. L., Lalka, S. G., Nixon, J. C., and Sawchuk, A. P., Improvement of flow througharterial stenoses by drag reducing agents, J. Surg. Res., 53, 625–630 (1992).

Virk, P. S., Elastic sublayer model for drag reduction by dilute solutions of linear macro-molecules, J. Fluid Mech., 45, 417–440 (1971).

Virk, P. S., Drag reduction fundamentals, AIChE J., 21, 625–656 (1975).

Virk, P. S., and Baher, H., Effect of polymer concentration on drag reduction, Chem. Eng. Sci.,25, 1183–1189 (1970).

Virk, P. S., and Merrill, E. W., Onset of dilute polymer solution phenomena, in Proceedingsof the Symposium on Viscous Drag Reduction, Dallas, TX, Wells, C. S., Ed., Plenum Press,New York, 1969, pp. 107–130.

Virk, P. S. and Wagger, D. L., Aspects of Mechanism in Type B drag reduction, in SecondIUTAM Symposium on Structure of Turbulence and Drag Reduction, Zurich, Gyr, A., Ed.,Springer-Verlag, Berlin, 1989, pp. 201–213.

Virk, P. S., Mickley, H. S., and Smith, K. A., Ultimate asymptote and mean flow structure ofToms’ phenomenon, J. Appl. Mech., 37, 488–493 (1970).

Vlachogiannis, M., and Hanratty, T. J., Influence of wavy structured surfaces and large scalepolymer structures on drag reduction, Exp. Fluids, 36, 685–700 (2004).

Page 38: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

126 POLYMER AND SURFACTANT DRAG REDUCTION IN TURBULENT FLOWS

Vlachogiannis, M., Liberatore, M. W., McHugh, A. J., and Hanratty, T. J., Effectiveness ofa drag reducing polymer: relation to molecular weight distribution and structuring, Phys.Fluids, 15, 3786–3794 (2003).

Wade, R. H., A Study of Molecular Parameters Influencing Polymer Drag Reduction, NavalUndersea Center, San Diego, CA, 1975.

Walsh, M. A.,On the turbulent flow of dilute polymer solutions, Ph.D. dissertation, CaliforniaInstitute of Technology, Pasadena, 1967.

Wang, S.-Q., Hu, Y., and Jamieson, A. M., Formation of nonequilibrium micelles in shear andelongational flow, ACS Symp. Ser., 578, 278–287 (1994).

Warholic, M. D., Modification of turbulent channel flow by passive and additive devices, Ph.D.dissertation, University of Illinois, Urbana, 1997.

Warholic, M. D., Massah, H., and Hanratty, T. J., Influence of drag-reducing polymers onturbulence: effects of Reynolds number, concentration and mixing, Exp. Fluids, 27, 461–472(1999a).

Warholic, M. D., Schmidt, G. M., and Hanratty, T. J., The influence of a drag-reducing surfactanton a turbulent velocity field, J. Fluid Mech., 388, 1–20 (1999b).

Weissberg, S. G., Simha, R., and Rothman, S., Viscosity of dilute and moderately concentratedpolymer solutions, J. Res. Natl. Bur. Stand., 46, 298–314 (1951).

Wells, C. S., Jr., and Spangler, J. G., Injection of a drag-reducing fluid into turbulent pipe flowof a Newtonian fluid, Phys. Fluids, 10, 1890–1896 (1967).

White, A., and Hemmings, J. A. G., Drag Reduction by Additives: Review and Bibliography,BHRA Fluid Engineering, Cranfield, UK, 1976.

Whitsitt, N. F., Harrington, L. J., and Crawford, H. R., Effect of wall shear stress on drag reduc-tion of viscoelastic solutions, in Proceedings of the Symposium on Viscous Drag Reduction,Dallas, TX, Wells, C. S., Ed., Plenum Press, New York, 1969, pp. 265–280.

Wul, G., Xu, J., and Miles, J., Polymer drag reduction in large diameter coal log pipeline,Proceedings of the International Technical Conference on Coal Utilization and Fuel Systems,23, 889–900 (1998).

Yu, B., and Kawaguchi, Y., Effect of Weissenberg number on the flow structure: DNS study ofdrag-reducing flow with surfactant additives, Int. J. Heat Fluid Flow, 24, 491–499 (2003).

Yu, B., and Kawaguchi, Y., Direct numerical simulation of viscoelastic drag-reducing flow: afaithful finite difference method, J. Non-Newtonian Fluid Mech., 116, 431–466 (2004).

Yu, B., and Kawaguchi, Y., Parametric study of surfactant-induced drag-reduction by DNS,Int. J. Heat Fluid Flow, 27, 887–894 (2006).

Yu, J. F. S., Zakin, J. L., and Patterson, G. K., Mechanical degradation of high molecular weightpolymers in dilute solution, J. Appl. Polym. Sci., 23, 2493–2512 (1979).

Yu, Z., Yahsi, U., McGervey, J. D., Jamieson, A. M., and Simha, R., Molecular weight-dependence of free volume in polystyrene studied by positron annihilation measurements,J. Polym. Sci. B, 32, 2637–2644 (1994).

Zakin, J. L., and Hunston, D. L., Mechanical degradation and drag reducing efficiency ofdilute solutions of polystyrene, Paper C5 in 2nd Drag Reduction International Conference,University of Cambridge, Cambridge, UK, Clark, J. A., Ed., BHRA Fluid Engineering,Cranfield, UK, 1977, pp. 41–54.

Zakin, J. L., and Hunston, D. L., Effects of solvent nature on the mechanical degradation ofhigh polymer solutions, J. Appl. Polym. Sci., 22, 1763–1766 (1978).

Page 39: Polymer Physics (From Suspensions to Nanocomposites and Beyond) || Polymer and Surfactant Drag Reduction in Turbulent Flows

REFERENCES 127

Zakin, J. L., Lu, B., and Bewersdorff, H.-W., Surfactant drag reduction, Rev. Chem. Eng., 14,253–320 (1998a).

Zakin, J. L., Myska, J., and Chara, Z., New limiting drag reduction and velocity profileasymptotes for nonpolymeric additives systems, AIChE J., 42, 3544–3546 (1996).

Zakin, J. L., Myska, J., and Lin, Z., Similarities and differences in drag reduction behaviorof high polymer and surfactant solutions, in International Symposium on Seawater DragReduction, Newport, RI, Meng, J. C. S., Ed., Systems Resource Management, 1998b, pp.277–280.

Zakin, J. L., Zhang, Y., and Ge, W., Drag reduction by giant micelles, in Giant Micelles:Properties and Applications, Zana, R., and Kaler, E. W., Eds., CRC Press, Taylor & Francis,New York, 2007, pp. 140, 473–492.

Zhang, Y., Qi, Y., and Zakin, J. L., Headgroup effect on drag reduction and rheological propertiesof micellar solutions of quaternary ammonium surfactants, Rheol. Acta, 45, 42–58 (2005a).

Zhang, Y., Schmidt, J., Talmon, Y., and Zakin, J. L., Co-solvent effects on drag reduction,rheological properties and micelle microstructures of cationic surfactants, J. Colloid InterfaceSci., 286, 696–709 (2005b).

Zhang, Y., Correlations among surfactant drag reduction, additive chemical structures, rheolog-ical properties and microstructures in water and water/co-solvent systems, Ph.D. dissertation,Ohio State University, Columbus, 2005c.

Zheng, Y., Lin, Z., Zakin, J. L., Talmon, Y., Davis, H. T., and Scriven, L. E., Cryo-TEMimaging the flow-induced transition from vesicles to threadlike micelles, J. Phys. Chem. B,104, 5263–5271 (2000).