Perhitungan Viskositas

Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Ann. Rev. Fluid Mech.

2008 40

Mechanics and Prediction of Turbulent Drag

Reduction with Polymer Additives

Christopher M. White1 and M. Godfrey Mungal2

1 Department of Mechanical Engineering, University of New Hampshire, Durham, New

Hampshire 03824; email: [email protected]

2
Mechanical Engineering Department, Stanford University, Stanford, California 94305;

email: [email protected]

Key Words polymer, drag reduction, turbulence, boundary layer, flow control

Abstract

This article provides a review of recent progress in understanding and predicting polymer drag

reduction in turbulent wall-bounded shear flows. The reduction in turbulent friction losses by

the dilute addition of high molecular weight polymers to flowing liquids has been extensively

studied since the phenomenon was first observed over sixty years ago. While it has long been

reasoned that the dynamical interactions between polymers and turbulence are responsible for

the drag reduction, it is not until recently that progress has been made to begin to elucidate

these interactions in detail. These advancements come largely from numerical simulations of

viscoelastic turbulent flows and detailed turbulence measurements in flows of dilute polymer

solutions using laser-based optical techniques. A selective overview of the current state of the

numerics and experimental techniques and their impact on understanding the mechanics and

prediction of polymer drag reduction is discussed. The review includes a discussion of areas in

which our understanding is incomplete, warranting further study.

1
2 White & Mungal

1 INTRODUCTION

The addition of small quantities of high molecular weight polymers to flowing

liquids can produce profound effects on a wide variety of flow phenomena that

appear incommensurate with the small concentration of polymers added to solu-

tion. This is most evident in turbulent boundary layers, where dissolving parts-

per-million quantities of long-chain flexible polymers into solution can reduce

turbulent friction losses by as much as 80% compared to that of the solvent

alone (e.g., Virk, 1975). The corresponding effect on the character of the flow

is equally impressive as illustrated in Figure 1. The acute differences in the

near-wall turbulent structure between Newtonian flow and polymer drag reduced

flow, as observed in the figure, can be described phenomenologically as follows:

As a consequence of the reduced wall friction, the mean velocity profile is mod-

ified and the shear in the boundary layer is redistributed. This effect alters the

nature and strength of the vortices formed, resulting in a significant modifica-

tion of the near-wall structure of the turbulent boundary layer. The unknown in

this simple description is the lack of coupling between the near-wall turbulence

and the skin friction, such that one cannot be merely a consequence of the other

(Robinson, 1991, Choi et al., 1994, Fukagata et al., 2002). It is this complexity of

the near-wall turbulence dynamics, further coupled with the dynamics of dilute

polymers in solution, that has made determination of a detailed mechanism of

polymer drag reduction an enigma for nearly sixty years.

Polymer drag reduction is studied for both practical and fundamental pur-

poses. Practical applications are pipe flows (or other internal flow geometries)

and marine vehicles, though the former has had much more success with polymers

than the latter (see Section 2.3). One reason for the fundamental interest is the
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 3

idea that studying the effects of polymers on turbulence provides valuable insight

into the physics of fluid turbulence, particularly, the self-sustaining mechanisms

of wall turbulence. Moreover, if a detailed understanding of the mechanics of

polymer drag reduction can be determined, it is conceivable that the effect can

be reproduced by other means, such as surface modification, sensor actuation, or

additives, among other strategies (Gad-el Hak, 2000, Choi & Karniadakis, 2003).

Success on this front would have tremendous impact on the economics of energy

propulsion and pollutant emission reduction from vehicles. Ideally, research on

polymer drag reduction should lead to improved and expanded practical applica-

tions and advance our understanding of fluid turbulence and turbulence control.

Since the discovery of the phenomenon by Toms (1948), there has been ex-

tensive, and continuing, research on the subject. The focus of this review is to

highlight recent progress in understanding and predicting polymer drag reduction

resulting from recent advances in computational simulations and experimental di-

agnostics. While it is not possible to be inclusive of all subject matter in the space

provided here, the present study provides a representative overview of the current

state of the research. Although we have strived to make the article self-contained,

our succinct recapitulation of several topics, most notably the fundamentals of

polymer drag reduction, do not adequately represent their importance. The re-

views by Lumley (1969), Liaw et al. (1971), Hoyt (1972), Landahl (1973) and

Virk (1975) provide greater breadth on various topics. The later reviews of Mc-

Comb (1990), Gyr & Bewersdorff (1995), and Nieuwstadt & Den Toonder (2001)

expand upon the early reviews in the context of more recent data.
4 White & Mungal

2 PREDICTION

2.1 Onset of Drag Reduction

The activity that occurred after Toms’ discovery soon produced sufficient exper-

imental evidence to conclude that the underlying physical mechanisms of drag

reduction involve dynamical interactions between polymers and turbulence. Two

results provide the basic evidence. The first is that laminar pipe flow of dilute

polymer solutions show no significant differences in the skin friction (or other flow

characteristics) compared to laminar pipe flow of Newtonian fluids. The second

is that, for a fixed pipe diameter, the Reynolds number, Re, at which drag reduc-

tion is first observed depends on the number of monomers1 in the macromolecule.

This implies an incipient interaction since, in general, turbulence dynamics de-

pend fundamentally on Re (Monin & Yaglom, 1975) and polymer dynamics de-

pend fundamentally on the number of monomers (Bird et al., 1987, Doi, 1996).

Figure 2 provides some illustration of the basic characteristics of a polymer

molecule and polymer stretch (and relaxation) in a shear flow.

Scaling arguments and experimental data led to the so-called time criterion for

drag reduction (Hershey & Zakin, 1967, Lumley, 1969). The criterion requires

that for drag reduction to occur the polymer relaxation time must be longer than
µs
a representative time-scale of the near wall turbulence, i.e. Tz > ρu2τ
, where

Tz is the average time it takes for a stretched polymer to return to a coiled

configuration (Zimm, 1956), µs the viscosity of the solution, ρ the density of the
p
solution, uτ = τw /ρ is the wall friction velocity, and τw is the shear stress at

the wall. For flexible linear polymers in solution, the relaxation time can be
1
A monomer is the repeat unit from which a polymer is built. For polyethylene oxide (PEO),

for example, the repeat unit is CH2 –CH2 –O (see Figure 2a).
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 5
µs (N 3/5 a)3
approximated by Tz ≈ κT , where N is the number of repeating monomers

in the molecule, a is length of a single monomer, κ is the Boltzmann constant,

and T is the solution temperature (Flory, 1971). According to the time criterion,

onset of drag reduction occurs when the ratio of the polymer time scale to the

flow time scale of the near-wall turbulence, defined as the wall-shear Weissenberg
Tz ρu2τ
number (W eτ = µs ), is order unity. While experimental data has generally

shown that at the onset of drag reduction W eτ ≈ 1, the omission of polymer

concentration in the time criterion is somewhat limiting for predictive purposes.

Indeed, the data of Nadolink (1987) using monodisperse polymers (i.e., a single

well-defined Tz as opposed to a distribution of Tz found for polydisperse polymers)

clearly demonstrates that the onset of drag reduction depends systematically on

polymer concentration. It follows that for predictive purposes more detailed

explanations are required.

Detailed explanations for the onset of drag reduction can generally be divided

into two classes based on the proposed effects of polymer stretching on the flow.

The first focuses on viscous effects (Lumley, 1969, Ryskin, 1987, L’vov et al., 2004)

and the second focuses on elastic effects (Tabor & de Gennes, 1986, Joseph, 1990).

The basic premise of the viscous explanation is that the effect of polymer stretch-

ing in a turbulent flow produces an increase in the effective viscosity. In wall-

bounded turbulent shear flows, polymers are primarily believed to be stretched

just outside the viscous sublayer in the so-called buffer layer. The strain rate

and vorticity fields associated with the buffer layer are assumed suitable to cause

full extension of the polymers (i.e., the coil-stretch transition) leading to a corre-

sponding large increase in the elongational viscosity (Metzner & Metzner, 1970,

Hinch, 1977). It is argued that a large increase in the effective viscosity just out-
6 White & Mungal

side the viscous sublayer will suppress turbulent fluctuations, increase the buffer

layer thickness, and reduce the wall friction (Lumley, 1973). Using scaling argu-

ments and a model polymer, Ryskin (1987) derived an expression for the effective

viscosity increase due to polymer stretch in a turbulent flow. In the derived ex-

pression, the effective viscosity is a function of polymer concentration and the

maximum extensibility of a given polymer-solvent pair. The effect of turbulence

is included through a numerical prefactor that modifies the maximum extensibil-

ity. A limited comparison with experimental data suggests that the Ryskin model

has merit and may be useful for predicting drag reduction, though more detailed

comparisons are required to make a definitive conclusion. In more recent work

L’vov et al. (2004) (see also Benzi et al., 2006) proposed that polymer stretching

produces a space-dependent effective viscosity that grows linearly from the wall.

Using scaling arguments and the assumption that polymer viscosity dominates

the momentum transfer (i.e., neglecting the Reynolds stress contribution), an

expression for the mean streamwise velocity profile is derived that is consistent

with the ultimate profile given by Virk et al. (1967). While the assumption that

the Reynolds stress momentum flux can be neglected does not strictly hold, the

simplicity of the model makes it intriguing for predictive purposes, in particular,

for formulating eddy viscosity models for computational fluid dynamics (CFD)

simulations of polymer drag reduced flows. Numerical simulations (De Ange-

lis et al., 2004) have shown that a space dependent viscosity model that varies

linearly from the wall produces drag reduction and captures the correct physics.

Tabor & de Gennes (1986) (see also de Gennes, 1990) argued that the viscous

theory cannot hold in a wall-bounded turbulent shear flow since the strain rates

near the wall, though high, fluctuate both in time and space and can produce
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 7

only partial stretching of the polymers (Ryskin’s polymer model addresses this

criticism). The elastic theory proposed by Tabor & de Gennes (1986) postulates

that the elastic energy stored by the partially stretched polymers is the important

variable for drag reduction and that the increase in the effective viscosity is small

and inconsequential. The elastic theory predicts that the onset of drag reduc-

tion occurs when the cumulative elastic energy stored by the partially stretched

polymers becomes comparable to the kinetic energy in the buffer layer at some

turbulent length scale larger than the Kolmogorov scale. The usual Kolmogorov-

type energy cascade is then terminated prematurely, and scales below this cutoff

scale are believed to behave elastically (Joseph, 1990). It is argued that these

effects yield a thickened buffer layer and subsequent drag reduction. The polymer

concentration is included in the onset criterion since the cumulative elastic energy

of the polymers is a function of the concentration. Experiments (Sreenivasan &

White, 2000) and numerical simulations (Min et al., 2003, 2004) have shown the

elastic theory to have merit and may be useful for predicting drag reduction.

The two classes of explanations given for the onset of polymer drag reduc-

tion appear fundamentally different, yet we have stated that both appear to have

merit when compared to experimental data. On the surface this implies an incon-

clusive finding. However, the issue is more complicated, since an elastic effect can

formally be interpreted in terms of a corresponding viscous effect (Sreenivasan

& White, 2000). This is best observed and explained from model systems of a

polymer, such as the FENE-P (Finite Elastic Non-Linear Extensibility-Peterlin)

model discussed in Section 3.2. We therefore delay an explanation of this dual

interpretation of polymer stretch until after the FENE-P model is introduced.


8 White & Mungal

2.2 Maximum Drag Reduction

If the friction drag for pipe flows is plotted in Prandtl-Kármán (P-K) coordinates

(see Figure. 3), onset of drag reduction is determined as the point of departure

from the P-K law. Following the onset of drag reduction, for a given Re, drag

reduction initially increases with polymer concentration but saturates beyond a

certain value. The bound on the drag reduction is the so-called maximum drag

reduction (MDR) or Virk asymptote (Virk et al., 1967). Similarly, for a given

polymer concentration, with increasing Re, the drag reduction initially increases

along a unique trajectory that depends on concentration (the so-called slope

increment, Virk, 1975) but abruptly changes trajectory at a certain Re. The

change in trajectory indicates where the drag reduction curve merges with the

MDR asymptote.

Despite recent progress, explanations for the unique bounding mechanisms of

MDR remain largely phenomenological or empirical. These include various inter-

pretations of the viscous and elastic theories of drag reduction. Two generalized

phenomenological explanations chosen as representative are briefly described: (i)

MDR occurs when the effects of polymers are felt over all flow scales causing

the buffer layer thickness to extend across the entire boundary layer (Virk, 1975,

Sreenivasan & White, 2000) and (ii) MDR occurs when the Reynolds stresses are

strongly diminished and the mechanisms that sustain turbulence are primarily

driven by the fluctuating polymer stresses (Warholic et al., 1999, Ptasinski et al.,

2001, Min et al., 2004). Surveying the literature one will find several gaps in these

explanations as well as contradictory results. Here we describe two contradictory

findings. The first is that polymer injection experiments have shown large DR

(near MDR) with polymers primarily within the near-wall region of the boundary
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 9

layer (Fontaine et al., 1992, see also McComb & Rabie (1979)). Furthermore, the

effect of polymers on the turbulence in the outer portion of the boundary layer

(where there is little or no polymers) is relatively weak compared to the effect on

the near-wall region. The second is that polymer stress work has been found in

some studies to dissipate (and not produce) turbulent kinetic energy (Ptasinski

et al., 2003). The empirical models (L’vov et al., 2004, Roy & Larson, 2005), for

the most part, are based on the assumption that the mean velocity profile follows

the MDR asymptote and crosses over to a Newtonian “plug flow” of constant

slope (in this respect these models are based on explanation (i)). However, there

is sufficient experimental data that demonstrates that that this assumption does

not strictly hold. Furthermore, these models do not work for injection experi-

ments since they do not account for inhomogeneities in the polymer concentration

field (Dimitropoulos et al., 2006).

Based on comparisons to experimental data, predicting MDR has had less suc-

cess than predicting the onset of drag reduction. The underlying difficulty is

that for high drag reduction the flow does not relaminarize but remains largely

turbulent in character, suggesting the existence of a unique self-sustaining flow

regime that lies in the transitional flow region between laminar and turbulent

flow. The uniqueness suggests that the self-sustaining mechanisms of turbulence

are different from that of a Newtonian fluid. This view is corroborated by recent

experimental and numerical results indicating that the polymer stresses play a

prominent role in the self-sustaining mechanisms of polymer drag reduced tur-

bulence. These important findings and their implications are discussed in more

detail in Sections 4.1.1 and 4.3.


10 White & Mungal

2.3 Practical Considerations

The most notable application of polymer drag reducing additives is in the Trans-

Alaska Pipeline System (TAPS). Polymer injection (and the subsequent drag

reduction) is used to increase design throughput or to maintain throughput with

off-line pumping stations (Burger et al., 1980). Applications in external flows are

less common although full-scale testing on a U.S. Navy submarine in the 1970’s

found that with polymer ejection hull drag was reduced and speed increased by

10-15% (NRC, 1997). The limitations with marine vehicle applications are the

costs and logistics associated with carrying polymers onboard.

The expression given for the polymer relaxation time (see, Section 2.1) shows

that the number of monomers in the molecule, N , is the important factor in deter-

mining the drag reducing effectiveness of a polymer (i.e., a long relaxation time).

Other properties relevant to drag reduction are a linear molecular structure, flexi-

bility (or extensibility), solubility, and viscoelastic characteristics (Lumley, 1969).

Common polymers used in the laboratory and for application are polyethylene

oxide (PEO), polyacrylamide (PAM), and polyisobutylene (PIB). With a few ex-

ceptions (e.g., Nadolink, 1987), the distribution of molecular weights are generally

broad, and typically the weight averaged N > 105 and Tz > 10−3 s. In practice,

DR is not sustained indefinitely but eventually decreases due to polymer scis-

sion during flow (Barnard & Sellin, 1972, Zakin & Houston, 1980, Den Toonder

et al., 1995). The degradation of the polymers corresponds to reduced values of

N with the effect of a decrease in the drag reducing effectiveness of the polymer.

Degradation correlations developed by Vanapalli et al. (2005) indicate that the

critical wall shear rate for chain-scission in PEO solutions follows the relation-

ship γ̇w ∼ M −n , where γ̇w is the critical wall shear rate, M is the weight average
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 11

molar mass (g/mol), and n ≈ −2.2.

3 RECENT ADVANCEMENTS

The equation for the conservation of momentum for the isothermal flow of an

incompressible polymer solution is given by

ρ∂t ui + ρuj ∂j ui = −∂i p + µs ∂j ∂j ui + ∂j τijp . (1)

where ρ is the fluid density, t is time, ui are the components of the velocity

vector, p is the hydrostatic pressure, µs is the solution viscosity, and τijp is the

polymer stress tensor arising from the entropic restoring force that tries to return

a stretched polymer molecule to its equilibrium configuration (i.e., a coiled chain).

In general, the shear viscosity of a dilute polymer solution differs little from the

solvent viscosity, and the effect of polymers on the flow is felt entirely through

the divergence of τijp . It follows that a detailed understanding of the generation

of τijp and its divergence should provide a necessary thread needed to begin to

unravel a detailed mechanism of polymer drag reduction. However, with present-

day diagnostics τijp cannot be measured directly in a turbulent flow and must be

estimated by other methods. These other methods include detailed turbulence

measurements in which the polymer shear stress can be “inferred” from the mea-

surements, and numerical simulations where the polymer dynamics are modeled

and the modeled polymer stress tensor is directly computed.

3.1 Experiments

Laser velocimetry measurements acquired in flows of drag-reducing polymer so-

lutions are well-established (see, Nieuwstadt & Den Toonder, 2001). As the capa-

bilities of these measurement systems have improved over the last decades, so has
12 White & Mungal

the detail and complexity of the measurements. For example, multi-component

single point velocity measurements using Laser Doppler Velocimetry (LDV) (e.g.,

Willmarth et al., 1987, Tiederman, 1990, Den Toonder et al., 1997, Warholic et al.,

1999, Ptasinski et al., 2001) and instantaneous planar maps of the velocity field

using Particle Image Velocimetry (PIV) (e.g., Warholic et al., 2001, White et al.,

2004, Hou et al., 2007) are common. The accuracy and precision with which

these measurements can be made, provide a means to obtain estimates of the


p
mean polymer shear stress profile (i.e., τxy (y), where the overbar denotes a mean

quantity) using a stress deficit approach. In this approach, the mean total shear

stress is decomposed into three terms consisting of the viscous, Reynolds, and

polymer shear stress contributions:

∂u p
τ =µ − ρúv́ + τxy , (2)
∂y

where τ is the total shear stress and the components of velocity are decomposed

into a mean and fluctuating quantity, i.e., ui = ūi + úi . The polymer shear stress

profile is then constructed from the difference between the total shear stress profile

and the measured Reynolds and viscous shear stress profiles (i.e., a stress deficit).

The stress deficit approach has been applied to channel flows (Tiederman, 1990,

Warholic et al., 1999, Ptasinski et al., 2001) and external boundary layer flows

(Koskie & Tiederman, 1991, Hou et al., 2006, White et al., 2007). The data

from these studies, discussed in Section 4, has provided both valuable insights

into the physical mechanisms of polymer drag reduction and benchmark data for

validation of numerical models.


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 13

3.2 Numerical simulations

Although the capability to numerically simulate flow of viscoelastic fluids is rel-

atively new (Sureshkumar et al., 1997, Den Toonder et al., 1997), the technique

has since been widely applied (e.g., Dimitropoulos et al., 1998, Sibilla & Baron,

2002, Ptasinski et al., 2003, De Angelis et al., 2003, Dubief et al., 2004, Min et al.,

2003). For the most part, the numerical simulations use constitutive equations

derived by modeling the polymer molecule as two beads connected by an elastic

spring (i.e., a single dumbbell, see Bird et al., 1987, Beris & Edwards, 1994). The

polymer dynamics are then entirely described by the evolution of the end-to-end

vector connecting the two beads, represented as the phase averaged configuration

tensor defined as cij = hqi qj i, where the q’s are the components of the end-to-end

vector. In a flow field, the evolution of cij is governed by the stretching and

restoring forces acting on the dumbbell. The restoring force is identically the

polymer stress tensor, such that, evaluation of the configuration tensor provides

a direct measure of the modeled polymer stress tensor.

The polymer model most often implemented for the study of drag reduction

is the FENE-P model (Bird et al., 1987). Although there are competing models

(e.g. FENE, Oldroyd-B), the FENE-P model is preferred since it accounts for

the finite extensibility of the molecule and uses a simple second-order closure

model in the equation for the polymer stress tensor. The former characteris-

tic, aside from being physically consistent with real polymers, reduces numerical

instabilities, while the latter reduces computational costs. Although the single

dumbbell FENE-P model is capable of capturing the basic rheological properties

of a polymer solution in many types of flows, there are clear circumstances where

the model does not capture the correct physics (Wedgewood et al., 1991, van Heel
14 White & Mungal

et al., 1998, Somasi et al., 2002, Zhou & Akhavan, 2003, Vincenzi et al., 2007).

The limitations of the model are primarily a consequence of the closure approxi-

mation (higher order moments are not accounted for) and the fact that a polymer

molecule consisting of typically N ≈ 105 monomers has been reduced to a single

dumbbell. In addition, polymer-polymer interactions, important even for dilute

polymer solutions in which polymers have been found to organize into super-

molecular structures (Kalashnikov, 1994), are not incorporated into the model.

The reader is referred to the review by Suen et al. (2002) for further details on

the computational modeling of viscoelastic fluids.

Using the FENE-P model, the numerical simulation involves solving the conti-

nuity equation, the equation for the conservation of momentum (in carrying out

the simulations Eq. 1 is modified slightly to incorporate concentration effects, see

Beris & Edwards, 1994) and the equation for the evolution of the configuration

tensor given by

∂t cij + uk ∂k cij = ckj ∂k ui + cik ∂k uj − τijp , (3)

1 cij
τijp = ( ). (4)
W eτ 1 − cLkk2 − δij

W eτ is the wall shear Weissenberg number (see, Section 2.1), ckk is the trace of

the confirmation tensor, and δij is the Kronecker delta .

3.2.1 Elastic and Viscous Interpretation of Polymer Stretch

Here we provide brief details and examples of our earlier statement that the

effects of polymer stretch on the flow can be formally interpreted as either an

elastic effect or a viscous effect. In the framework of the FENE-P model (or

similar polymer model) the elastic energy stored by a stretched polymer is pro-
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 15

portional to the trace of the confirmation tensor, ckk = cxx + cyy + czz (Min et al.,

2003). If the problem is approached from the perspective of the elastic theory of

drag reduction, the transport equation for the elastic energy can be studied to

understand the energy transfer between the polymers and the flow (Min et al.,

2004, 2003). Alternatively, and within the framework of the FENE-P model,

Benzi et al. (2006) approached the problem from the perspective of the viscous

theory and found that the important component of the confirmation tensor is

cyy , which appears in the momentum and kinetic energy equations as an effective

viscosity.

4 MECHANICS

4.1 Velocity Statistics

The statistics measured or computed most often are the mean streamwise velocity

profile, root-mean-square (r.m.s.) fluctuating velocity profiles, Reynolds stress

profiles, turbulent energy spectra, and kinetic energy budgets.

4.1.1 Internal Flows with Homogeneous Polymer Concentration

The statistical data from laboratory flows and from numerical simulations of

internal flows with homogenous polymer concentration (i.e., a polymer ocean)

is extensive (see, Nieuwstadt & Den Toonder, 2001), and much is known about

the turbulent velocity field. Of particular interest is the experimental work of

Warholic et al. (1999) in which distinct differences in the statistical trends of the

turbulence velocity field between flows at low drag reduction (LDR) and high

drag reduction (HDR) were clearly identified. Similar trends were also found

by Ptasinski et al. (2001), Min et al. (2004) and Dubief et al. (2004), though

important differences in the data between these studies compared to Warholic


16 White & Mungal

et al. were reported. In general, drag reductions greater than approximately

40% are considered HDR flows.

The velocity statistics of LDR flows are consistent with the phenomenologi-

cal description that, with increasing DR, the buffer layer increases and displaces

the Newtonian “plug flow” away from the wall (Virk, 1975). Overwhelmingly,

experimental and numerical data show that the logarithmic region of the mean

streamwise velocity profile shifts upwards with increasing DR but remains paral-

lel to that of the Newtonian flow. Furthermore, the profiles of low-order turbulent

statistics are similar in shape to Newtonian profiles except that there is a shift

away from the wall that increases with increasing DR. Detailed changes in the

statistics with increasing DR include an increase in the peak of the mean stream-

wise velocity fluctuation (ú), and a decrease in the wall-normal (v́) and spanwise

(ẃ) velocity fluctuation, as well as a monotonic decrease in the Reynolds shear

stress (ρúv́). The kinetic energy budgets at low DR follow similar trends as the

turbulence statistics (Dimitropoulos et al., 1998). A term-by-term analysis of

the turbulent kinetic energy budget suggests that the effects of polymers on the

low-order turbulent velocity statistics are a consequence of the large decrease in

the mean streamwise component of the pressure-strain term, responsible for the

transfer of streamwise turbulent kinetic energy to the spanwise and wall-normal

components (Dimitropoulos et al., 1998, Ptasinski et al., 2001).

Drag reductions greater than approximately 40% exhibit different statistical

trends in the velocity field than observed for LDR flows. Notable changes are

an increase in the slope of the log-region of the mean streamwise velocity profile,

a decrease in the peak of the mean streamwise velocity fluctuation, and an in-

crease in the role the polymer shear stress (as determined from the stress deficit
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 17

approach) plays in sustaining turbulence. The latter is perhaps the most inter-

esting and controversial characteristic of HDR flows. Warholic et al. (1999) for

DR’s near MDR find an almost complete depletion of the Reynolds shear stress,

and conclude that for MDR the self-sustaining mechanisms of Newtonian wall-

turbulence can not exist, and that turbulence is sustained entirely by the polymer

stresses. However, other studies (Ptasinski et al., 2001, Min et al., 2004, Dubief

et al., 2004) find that although the Reynolds shear stress is strongly attenuated

for HDR flows it remains finite. Consideration of these unresolved differences

leads to the general conclusion that for HDR flows polymer stresses play a dom-

inant role and Reynolds stresses play a lesser role in the near-wall dynamics of

the flow.

4.1.2 External Flows with Inhomogeneous Polymer Concentration

The literature on turbulent boundary layers with injection (i.e., external flows

with inhomogeneous concentrations) is less extensive, probably since it has never

been shown to be cost effective in marine vehicle applications. The external

boundary layer flow with polymer injection (or ejection) is expected to bear some

similarities to fully developed channel flows, but also some significant differences

due to the continuous decrease of the polymer concentration with downstream

distance.

The development of the drag reduction in external flows with polymer injec-

tion will ultimately depend on the evolution of the polymer concentration field.

Experimental observations have shown that three regions of drag reduction are

produced with increasing downstream distance: (1) a development region just

downstream of the injector where DR first increases, (2) a steady-state region

where DR is nominally constant, and (3) a depletion region where the DR reduces
18 White & Mungal

asymptotically towards zero (Poreh & Hsu, 1972, Winkel et al., 2006, Hou et al.,

2007). The decrease in drag-reduction with downstream distance results from

turbulent mixing and the resulting drop in the near-wall concentration (Fontaine

et al., 1992, Somandepalli, 2006). The downstream evolution of DR is summa-

rized on the DR vs log (K) plot shown in Figure 4, where the “K” parameter

developed by Vdovin & Smol’yakov (1978) is defined by K ≡ Qi Ci /ρXs U∞ , Qi

is the volumetric injection rate per unit span, Ci is the polymer concentration,

ρ is the solvent density, Xs the downstream distance, and U∞ the free-stream

velocity. Results from Fontaine et al. (1992), Petrie & Fontaine (1996), Winkel

et al. (2006), Hou et al. (2007) suggest a similar plot for a range of polymers,

injection rates, polymer concentrations, and free-stream velocities. Although dif-

ferences due to polymer types exist, onset of drag reduction occurs for K > 10−9

and MDR is generally achieved for K > 10−6 .

The downstream transport of polymers away from the wall is found to be

significantly slower than the transport of a passive scalar (Dimitropoulos et al.,

2006). The resulting effect is that the downstream evolution of the drag reduction

with a single axial injection location is fairly robust. Using simultaneous planar

laser induced fluorescence (PLIF) and PIV, Somandepalli (2006) found that the

wall normal flux, Ć v́ where C is the polymer concentration, is seen to be most

closely tied to the behavior of the DR. It follows that the slow decrease in the

dispersion of polymers is indicative of the reduction in the wall-normal velocity

fluctuations during drag reduction. In brief, polymers are not passive scalars and

their ability to attenuate the near-wall turbulence enables a fairly concentrated

polymer solution to persist for long distances downstream.

The velocity statistics measured (Koskie & Tiederman, 1991, Fontaine et al.,
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 19

1992, White et al., 2004) and computed (Dimitropoulos et al., 2006) in external

flows with polymer injection at LDR show similar trends to data obtained in

internal flows with homogeneous polymer concentration. With increasing DR,

however, differences in the statistics between the two flow types become increas-

ingly apparent. There are three main observations compared to internal flows

with a homogeneous polymer concentration at HDR: (i) the slope of the logarith-

mic region of the mean velocity profile is lower, (ii) the overall decrease in the

Reynolds shear stress is not as large, and (iii) in the near-wall region the statisti-

cal profiles are similar, but the differences increase with increasing distance from

the wall. Comparative studies (Petrie & Fontaine, 1996, Dimitropoulos et al.,

2006) between boundary layer flow with homogeneous (“ocean”) and inhomoge-

neous (injection) polymer concentration suggests that a lack of polymer in the

outer part of the boundary layer for the injection case can explain the observed

differences in the statistics. Given that DR’s near MDR are reached in the injec-

tion experiments, these results suggest that the effect of polymers on the flow in

the outer portion of the boundary layer may not play a central role in the skin

friction reduction. While the importance of the near-wall region of the flow to the

turbulent skin friction is expected, Fukagata et al. (2002) derived a quantitative

relationship between flow statistics and friction drag in turbulent channel, pipe,

and plane boundary layer flow of a Newtonian fluid that clearly illustrates the

importance of the near-wall region of the flow. The analysis was extended by

White et al. (2007) (see also Hou et al., 2006) to include the effects of polymer

additives, in which the importance of the near-wall region of the flow to the drag

reduction was demonstrated.


20 White & Mungal

4.2 Drag Reductions at High Reynolds Number

The majority of laboratory flows investigating polymer drag reduction have been

carried out at fairly moderate Re. High Re has been addressed by the HiPlate

experiment completed recently (Winkel et al., 2006). The HiPlate experiments

are performed at speeds up to 18 m/s using a range of polymer types and con-

centrations. The measurements performed at various locations extending over 13

m downstream include direct wall shear stress measurements, near wall velocity

measurements, and polymer concentrations. The HiPlate results suggest that in-

creasing molecular weight and increasing flux of polymer leads to increasing DR.

Following the development region, near-MDR is achieved, then the continued

process of mixing causes a steady decrease of DR. The extent of the development

region varied with the free-stream speed, while the decay regions showed much

less speed dependence. It was also found that the higher molecular weight poly-

mers appeared to undergo chain scission degradation which limited their drag

reducing effectiveness.

4.3 Effect on the Near-Wall Structure of Turbulence

Prevalent structures of wall-bounded turbulent shear flows are streamwise veloc-

ity streaks and quasi-streamwise vortices. A generally accepted view regarding

first-order effects in the self-sustaining cycle of wall turbulence is that quasi-

streamwise vortices extract energy from the mean flow, create streamwise velocity

streaks, and in turn, an instability of the streaks gives rise to the quasi-streamwise

vortices (Kim et al., 1971, Swearingen & Blackwelder, 1987). Direct evidence of

this cycle at low-Reynolds numbers is provided by the numerical simulations of

Jiménez & Pinelli (1999), though at high Reynolds numbers the interactions with
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 21

the outer layer are likely important and it can be expected that the regenera-

tion cycle is more complex than given by the wall-model (Brown & Thomas,

1977, Hunt & Morrison, 2000, Hunt & Carlotti, 2001). Despite the shortcom-

ings in our understanding, a simple yet fundamental principle of any turbulent

drag reduction strategy is to disrupt the turbulence regeneration cycle (Choi &

Karniadakis, 2003). In this context and based on the level of drag reduction pos-

sible, polymers (as well as surfactants, see Myska et al., 1995, Li & Kawaguchib,

2004) have proven to be the most effective strategy for disrupting the turbulence

regeneration cycle.

The effect of polymer additives on the near-wall structure of turbulence results

in a reduction in the strength and numbers of the quasi-streamwise vortices and a

stabilization (i.e., a reduced spanwise meandering) and thickening of the stream-

wise velocity streaks (see Figure 1). In the context of the wall-model, these

characteristics are consistent with the disruption of the turbulence regeneration

cycle (Dubief et al., 2004). However, despite the capabilities of the numerical

simulations, complete details of how polymers disrupt the cycle remains open to

interpretation. The variations in the interpretations of the results can include

viscous vs. elastic effects or different views regarding the regeneration cycle of

wall turbulence (e.g., compare Ptasinski et al., 2003, Dubief et al., 2004). The

difficulty is due to the fact that the interactions between polymers and turbulence

are complex, such that, the data mining required to fully understand the inter-

actions from the numerical simulations is nontrivial. The intermittent character

of the near-wall structures (i.e, they are non-space filling) presents additional

difficulties (Dubief et al., 2005).

In general, a detailed understanding of the polymer interactions with the near-


22 White & Mungal

wall turbulent structures requires an understanding of the formation of (τijp ) and

its coupling with the flow field. The latter is accomplished by studying the

effects of the polymer body force term (∂j τijp ) in the momentum equation and

the polymer stress work term (ui ∂j τijp ) in the balance equation for the turbulent

kinetic energy. Figure 5 provides a three-dimensional snapshot of the near-wall

vortices and contours of polymer stretch obtained in a DNS of a channel flow

with polymer drag reduction.

For the most part, numerical simulations directly investigating the interactions

between polymers and the near-wall structure of turbulence generally agree on

two main points: (1) polymers stretch primarily in the near-wall region of the

flow and (2) polymers directly interact with and dampen the quasi-streamwise

vortices. The latter point is discussed first. The dampening mechanisms of the

near-wall vortices results from spatial gradients in the polymer stress surrounding

the vortices, producing two effects that lead to a weakening of the vortices. First,

the polymer body force opposes the motion of the vortices (Dubief et al., 2005),

and in turn, the polymer stress work transfers energy from the vortices to the

polymers (Ptasinski et al., 2003). The latter effect can be observed in the energy

budget of the mean turbulent kinetic energy since the polymer stress work is

found to be a negative term (Dimitropoulos et al., 1998, Ptasinski et al., 2003).

As a consequence of this vortex suppression, the self-sustaining cycle of the wall

turbulence is disrupted and the turbulent skin friction drag is reduced.

The polymer stress gradients near the vortices can arise either from polymers

stretching or relaxing. At present, the numerical evidence is somewhat con-

flicting on these details. Using Brownian dynamic simulations, Terrapon et al.

(2004) found that polymers stretch primarily in straining flows next to the quasi-
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 23

streamwise vortices (see also Massah, 1993, Sibilla & Baron, 2002, Dubief et al.,

2004). Intermittent bursts of strong biaxial flows cause the polymer chains to un-

ravel at low W e and fully extend (i.e., coil-stretch transition) at high W e. These

strong biaxial flows are found to be strongly correlated with the flow next to the

quasi-streamwise vortices. Other studies (Stone et al., 2004, Davoudi & Schu-

macher, 2006) suggest that polymers stretch primarily in the streak regions and

relax as they move from the streaks into the vortices. We reiterate the fact that

regardless if the polymers are stretching or contracting around the vortices, in

general, the vortex suppression mechanisms remains valid. For historical purposes

we note that the attenuation of the near-wall vortices observed with polymer ad-

ditives has long been speculated as a possible cause of the drag reduction (Seyer

& Metzner, 1969, Paterson & Abernathy, 1970, Virk, 1975).

5 SUMMARY

Aided by the technological advancements in computing resources and experimen-

tal measurement techniques, the last decade has produced significant progress in

formulating a detailed understanding of the polymer turbulence interactions re-

sponsible for drag reduction. Here we summarize these advancements and provide

general comments regarding some of the remaining open questions.

5.1 Mechanisms

In the context of the FENE-P model and at low Re, the numerical simulations

provide direct evidence that polymers disrupt the near-wall turbulence regener-

ation cycle and reduce the turbulent friction drag by directly interacting with

and dampening the quasi-streamwise vortices. The vortex suppression is a result


24 White & Mungal

of spatial gradients in the polymer stress surrounding the vortices that leads to

the transfer of energy from the vortices to the polymers. This proposed mech-

anism is consistent with most experimental and numerical data and has long

been speculated as a possible cause of the drag reduction with polymer additives.

However, limitations in the FENE-P model and the low Re of the simulations still

leaves some measure of uncertainty that will require further work to fully resolve.

However, there is little doubt that, with continued experimental and numerical

efforts, the gaps in our understanding can be reduced and eventually eliminated.

5.2 Predictions

Despite a somewhat detailed understanding of the underlying mechanisms of

polymer drag reduction (at least for low to moderate Re), for predictive purposes

a model of the process is still required. Numerical simulations and experiments

are now advanced enough to conduct comprehensive tests of existing predictive

models and develop, in coordination, improved models that better capture the

physics of the problem. Given the importance of the polymer interactions with

the vortices, a reasonable first step for predicting the onset of drag reduction is

to relate the polymer relaxation time to a characteristic time scale of the quasi-

streamwise vortices (Min et al., 2004, Dubief et al., 2005), or similarly to an eddy

turnover time (Ptasinski et al., 2003). Two important yet challenging tasks are

to incorporate inhomogeneities in the polymer concentration field and polymer

degradation into the models.


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 25

5.3 Outlook

Improved and expanded practical applications of polymer drag reduction will

greatly be advanced by the development of robust polymer models and drag

reduced turbulence models for the Large Eddy Simulation (LES) and Reynolds-

Averaged Navier Stokes Numerical Simulations (RANS) techniques. These tech-

niques will allow for the modeling and optimization of applications at high Re

and in complex geometries. In addition to the experiments needed to support

these efforts, detailed experimental investigations to study the effects of poly-

mer additives in flows at high Re, with pressure gradients, and with separation

offer unique opportunities to provide valuable insights on the fundamental fluid

physics in these complex flows.

ACKNOWLEDGMENTS

We thank Y. Dubief for his contributions to this paper. We also thank Costas

Dimitropoulos, Sanjiva Lele, Parvis Moin, John Paschkewitz, Vincent Terrapon,

Eric Shaqfeh and Vijay Somandepalli for their valuable insights into the problem

of polymer drag reduction. We also acknowledge the Defense Advanced Re-

search Projects Agency, Advanced Technology Office, Friction Drag Reduction

Program for their financial support of much of our work reported here. C.M.W.

thanks Joseph Klewicki and Greg Chini for improvements on the manuscript and

Katepalli Sreenivasan for his tutorage.


26 White & Mungal

References

Barnard BJS, Sellin RHJ. 1972. Degradation of dilute solutions of drag-reducing

polymer. Nature 236(62):12–14

Benzi R, De Angelis E, L’vov VS, Procaccia I, Tiberkevich V. 2006. Maximum

drag reduction asymptotes and the cross-over to the newtonian plug. J. Fluid

Mech. 551:185–195

Beris AN, Edwards RJ. 1994. Thermodynamics of Flowing Systems with Internal

Microstructure. Oxford, UK: Oxford University Press

Bird RB, Curtiss CF, Armstrong RC, Hassager O. 1987. Dynamics of Polymeric

Fluids, vol. 2. New York: Wiley

Brown GL, Thomas ASW. 1977. Large structure in a turbulent boundary layer.

Phys. Fluids 20(10):243–252

Burger ED, Chorn LG, Perkins TK. 1980. Studies of drag reduction conducted

over a broad range of pipeline conditions when flowing Prudhoe Bay crude oil.

J. Rheol. 24(5):603–626

Choi H, Moin P, Kim J. 1994. Active turbulence control for drag reduction in

wall-bounded flows. J. Fluid Mech. 262:75–110

Choi KS, Karniadakis GE. 2003. Mechanisms on transverse motions in turbulent

wall flows. Ann. Rev. Fluid Mech. 35:45–62

Davoudi J, Schumacher J. 2006. Stretching of polymers around the kolmogorov

scale in a turbulent shear flow. Phys. Fluids 18(2):25103–11

De Angelis E, Casciola CM, L’vov VS, Piva R, Procaccia I. 2003. Drag reduction

by polymers in turbulent channel flows: energy redistribution between invariant

empirical modes. Phys. Rev. E 67:056312


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 27

De Angelis E, Casciola CM, et al. 2004. Drag reduction by a linear viscosity

profile. Phys. Rev. E 70(5):55301–4

de Gennes PG. 1990. Introduction to Polymer Dynamics. Cambridge, UK: Cam-

bridge University Press.

Den Toonder JMJ, Draad AA, Kuiken GDC, Nieuwstadt FTM. 1995. Degrada-

tion effects of dilute polymer solutions on turbulent drag reduction in pipe

flows. Appl. Sci. Res. 55(1):63–82

Den Toonder JMJ, Hulsen MA, Kuiken GDC, Nieuwstadt FTM. 1997. Drag

reduction by polymer additives in a turbulent pipe flow: numerical and labo-

ratory experiments. J. Fluid Mech. 337:193–231

Dimitropoulos C, Dubief Y, Shaqfeh ESG, Moin P. 2006. Direct numerical sim-

ulation of polymer-induced drag reduction in turbulent boundary layer flow of

inhomogeneous polymer solutions. J. Fluid Mech. 566:153–62

Dimitropoulos CD, Sureshkumar R, Beris AN, Handler RA. 1998. Budget of

Reynolds stress, kinetic energy and streamwise enstrophy in viscoelastic tur-

bulent channel flow. J. Non-Newtonian Fluid Mech. 79:433–468

Doi M. 1996. Introduction to Polymer Physics. Oxford, UK: Oxford University

Press

Dubief Y, Terrapon VE, et al. 2005. New answers on the interaction between

polymers and vortices in turbulent flows. Flow, Turbul. Combust. 74(4):311–29

Dubief Y, White CM, et al. 2004. On the coherent drag-reducing turbulence-

enhanching behaviour of polymers in wall flows. J. Fluid Mech. 514:271–280

Flory PJ. 1971. Principles of Polymer Chemistry. Ithaca, New York: Cornell

University Press
28 White & Mungal

Fontaine AA, Petrie HL, Brungart TA. 1992. Velocity profile statistics in a tur-

bulent boundary layer with slot-injected polymer. J. Fluid Mech. 238:435–66

Fukagata K, Iwamoto K, Kasagi N. 2002. Contribution of Reynolds stress distri-

bution to the skin friction in wall-bounded flows. Phys. Fluids 14:L73–76

Gad-el Hak M. 2000. Flow Control: Passive, Active, and Reactive Flow Manage-

ment. Oxford, UK: Oxford University Press

Gyr A, Bewersdorff HW. 1995. Drag Reduction of Turbulent Flows by Additives.

Netherlands: Kluwer Academic Publishers

Hershey HC, Zakin JL. 1967. A molecular approach to predicting the onset of

drag reduction in the turbulent flow of dilute polymer solutions. Chem. Engng

Sci. 22:1847–1856

Hinch EJ. 1977. Mechanical models of dilute polymer solutions in strong flows.

Phys. Fluids 20(10):22–30

Hou Y, Somandepalli VSR, Mungal MG. 2006. A technique to determine total

shear stress and polymer stress profiles in drag reduced boundary layer flows.

Exp. Fluids 40:589–600

Hou Y, Somandepalli VSR, Mungal MG. 2007. Streamwise development of tur-

bulent boundary layer drag reduction with polymer injection. J. Fluid Mech.

submitted

Hoyt JW. 1972. Effect of additives on fluid friction. Trans. ASME: J. Basic Engng

94(2):258–285

Hunt JCR, Carlotti P. 2001. Statistical structure at the wall of the high reynolds

number turbulent boundary layer. Flow, Turbul. Combust. 66(4):453–475


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 29

Hunt JCR, Morrison JF. 2000. Eddy structure in turbulent boundary layers.

Euro. J. Mech. B/Fluids 19(5):673–694

Jiménez J, Pinelli A. 1999. The autonomous cycle of near-wall turbulence. J.

Fluid Mech. 389:335–59

Joseph DD. 1990. Fluid Dynamics of Viscoelastic Liquids. New York, New York:

Springer Verlag

Kalashnikov VN. 1994. Shear-rate dependent viscosity of dilute polymer solu-

tions. J. Rheol. 38(5):1385–1403

Kim HT, Kline SJ, Reynolds WC. 1971. The production of turbulence near a

smooth wall in a turbulent boundary layer. J. Fluid Mech. 50:133–60

Koskie JE, Tiederman WG. 1991. Polymer drag reduction in a zero pressure

gradient boundary layer. Phys. Fluids A 3:2471–2473

Landahl MT. 1973. Drag reduction by polymer addition. In 13th Intl Congr.

Theor. Appl. Mech. Moscow, eds. E Becker, GK Mikhalov, 177–179, Springer

Li FC, Kawaguchib Y. 2004. Investigation on the characteristics of turbulence

transport for momentum and heat in a drag-reducing surfactant solution flow.

Phys. Fluids 16:3281–3295

Liaw GC, Zakin JL, Patterson GK. 1971. Effects of molecular characteristics of

polymers on drag reduction. AIChE J. 17(2):391–397

Lumley JL. 1969. Drag reduction by additives. Ann. Rev. Fluid Mech. 1:367–384

Lumley JL. 1973. Drag reduction in turbulent flow by polymer additives. J.

Polym. Sci. Macromol. Rev. 7:263–90

L’vov VS, Pomyalov A, Procaccia I, Tiberkevich V. 2004. Drag reduction by

polymers in wall bounded turbulence. Phys. Rev. Lett. 92(24):244503–4


30 White & Mungal

Massah H. 1993. Studies of the interaction between drag-reducing polymers and

a turbulent flow field using PIV and FENE bead-spring model. Ph.D. thesis,

University of Illinois, Urbana.

McComb W. 1990. The Physics of Fluid Turbulence. Oxford, UK: Oxford Uni-

versity Press

McComb W, Rabie LH. 1979. Development of local turbulent drag reduction due

to nonuniform polymer concentration. Phys. Fluids 22:183–185

Metzner AB, Metzner AP. 1970. Stress levels in rapid extensional flows of poly-

meric fluids. Rheol. Acta. 9:174–181

Min T, Choi H, Yoo JY. 2003. Maximum drag reduction in a turbulent channel

flow by polymer additives. J. Fluid Mech. 492:91–100

Min T, Yoo JY, Choi H, Joseph DD. 2004. Drag reduction by polymer additives

in a turbulent channel flow. J. Fluid Mech. 486:213–238

Monin A, Yaglom A. 1975. Statistical Fluid Mechanics: Mechanics of Turbulence.

Cambridge, Massachusetts: MIT Press

Myska J, Zakin JL, Chara Z. 1995. Viscoelasticity of a surfactant and its drag-

reducing ability. Appl. Sci. Res. 55(4):297–310

Nadolink RH. 1987. Friction reduction in dilute polystyrene solutions. Ph.D. the-

sis, University of California at San Diego

Nieuwstadt F, Den Toonder J. 2001. Drag reduction by additives: a review. In

Turbulence structure and motion, eds. A Soldati, R Monti, 269–316, Springer

Verlag

NRC. 1997. Submarine platform technology. In Technology for the United States

Navy and Marine Corps, 2000-2035: Becoming a 21st-Century Force, vol. 6:


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 31

Platforms, 92, Commission on Physical Sciences, Mathematics, and Applica-

tions, Washington DC: National Academy Press

Paterson RW, Abernathy FH. 1970. Turbulent flow drag reduction and degrada-

tion with dilute polymer solutions. J. Fluid Mech. 43:689–710

Petrie HL, Fontaine AA. 1996. Comparison of turbulent boundary layer modi-

fication with slot-injected and homogeneous drag-reducing polymer solutions.

ASME, Fluids Engng 237:205–210

Poreh M, Hsu KS. 1972. Diffusion of drag reducing polymers in a turbulent

boundary layer. J. Hydronaut. 6(1):27–33

Ptasinski PK, Boersma BJ, et al. 2003. Turbulent channel flow near maximum

drag reduction: Simulations, experiments and mechanisms. J. Fluid Mech.

490:251–291

Ptasinski PK, Nieuwstadt FTM, van den Brule BHAA, Hulsen MA. 2001. Ex-

periments in turbulent pipe flow with polymer additives at maximum drag

reduction. Flow, Turbul. Combust. 66(2):159–182

Robinson SK. 1991. Coherent motions in the turbulent boundary layer. Ann. Rev.

Fluid Mech. 23:601–639

Roy A, Larson RG. 2005. A mean flow model for polymer and fiber turbulent

drag reduction. Applied Rheology 15(6):370–89

Ryskin G. 1987. Turbulent drag reduction by polymers: a quantitative theory.

Phys. Rev. Lett. 59(18):2059–62

Seyer FA, Metzner AB. 1969. Turbulence phenomena in drag reducing systems.

AIChE J. 15(3):426–434
32 White & Mungal

Sibilla S, Baron A. 2002. Polymer stress statistics in the near-wall turbulent flow

of a drag-reducing solution. Phys. Fluids 14(3):1123–1136

Somandepalli VSR. 2006. Combined PIV and PLIF measurements in a polymer

drag reduced turbulent boundary layer. Ph.D. thesis, Stanford University

Somasi M, Khomami B, Woo NJ, Hur JS, Shaqfeh ESG. 2002. Brownian dynamics

simulations of bead-rod and bead-spring chains: numerical algorithms and

coarse-graining issues. J. Non-Newton. Fluid Mech. 108(1-3):227–255

Sreenivasan KR, White CM. 2000. The onset of drag reduction by dilute poly-

mer additives, and the maximum drag reduction asymptote. J. Fluid Mech.

409:149–164

Stone PA, Roy A, Larson RG, Waleffe F, Graham MD. 2004. Polymer drag reduc-

tion in exact coherent structures of plane shear flow. Phys. Fluids 16(9):3470–

3482

Suen JKC, Joo YL, Armstrong RC. 2002. Molecular orientation effects in vis-

coelasticity. Ann. Rev. Fluid Mech. 34:417–444

Sureshkumar R, Beris AN, Handler RA. 1997. Direct numerical simulation of the

turbulent channel flow of a polymer solution. Phys. Fluids 9(3):743–755

Swearingen JD, Blackwelder RF. 1987. The growth and breakdown of streamwise

vortices in the presence of a wall. J. Fluid Mech. 182:255–290

Tabor M, de Gennes PG. 1986. A cascade theory of drag reduction. Europhys.

Lett. 2(7):519–522

Terrapon VE, Dubief Y, Moin P, Shaqfeh ESG, Lele SK. 2004. Simulated polymer

stretch in a turbulent flow using brownian dynamics. J. Fluid Mech. 504:61–71

Tiederman WG. 1990. The effect of dilute polymer solutions on viscous drag
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 33

and turbulence structure. In Structure of Turbulence and Drag Reduction, ed.

A Gyr, 187–200, Springer Verlag

Toms BA. 1948. Some observations on the flow of linear polymer solutions through

straight tubes at large Reynolds numbers. Proc. 1st Intern. Congr. Rheol.,

North Holland II135–141

van Heel APG, Hulsen MA, van den Brule BHAA. 1998. On the selection of

parameters in the FENE-P model. J. Non-Newton. Fluid Mech. 75(2-3):253–

271

Vanapalli SA, Islam MT, Solomon MJ. 2005. Scission-induced bounds on maxi-

mum polymer drag reduction in turbulent flow. Phys. Fluids 17(9):95108–11

Vdovin AV, Smol’yakov AV. 1978. Diffusion of polymer solutions in a turbulent

boundary layer. J. Appl. Mech. Tech. Phys. 19(2):196–201

Vincenzi D, Jin S, Bodenschatz E, Collins LR. 2007. Stretching of polymers in

isotropic turbulence: A statistical closure. Phys. Rev. Lett. 98(2):024503–4

Virk PS. 1975. Drag reduction fundamentals. AICHE J. 21:625–656

Virk PS, Merril EW, Mickley HS, Smith KA, Mollo-Christensen EL. 1967. The

Toms phenomenon—turbulent pipe flow of dilute polymer solutions. J. Fluid

Mech. 20(10):22–30

Warholic MD, Heist DK, Katcher M, Hanratty TJ. 2001. A study with particle-

image velocimetry of the influence of drag-reducing polymers on the structure

of turbulence. Exp. Fluids 31(5):474–483

Warholic MD, Massah H, Hanratty TJ. 1999. Influence of drag-reducing polymers

on turbulence: effects of Reynolds number, concentration and mixing. Exp.

Fluids 27:461–472
34 White & Mungal

Wedgewood LE, Ostrov DN, Byron Bird R. 1991. A finitely extensible bead-

spring chain model for dilute polymer solutions. J. Non-Newton. Fluid Mech.

40(1):119–139

White CM, Somandepalli VSR, Dubief Y, Mungal MG. 2007. The dynamical

contributions to the skin friction in drag reduced polymer flows. Phys. Fluids

submitted

White CM, Somandepalli VSR, Mungal MG. 2004. The turbulence structure of

drag reduced boundary layer flow. Exp. Fluids 36:62–69

Willmarth W, Wei T, Lee CO. 1987. Laser anemometer measurements of

Reynolds stress in a turbulent channel with drag reducing polymer additives.

Phys. Fluids 30:933–935

Winkel ES, Oweis GF, et al. 2006. Friction drag reduction at high reynolds num-

bers using injected polymer solutions. In 26th ONR Symp. Naval Hydrodynam-

ics, Rome, Italy

Zakin JL, Houston DL. 1980. Effect of polymer molecular variables on drag re-

duction. J. Macromol. Sci. B18(4):795–814

Zhou Q, Akhavan R. 2003. A comparison of fene and fene-p dumbbell and chain

models in turbulent flow. J. Non-Newton. Fluid Mech. 109(2-3):115–155

Zimm BH. 1956. Dynamics of polymer molecules in dilute solutions: viscoelas-

ticity, flow birefringence and dielectric loss. J. Chem. Phys. 24:269–279


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 35

FIGURE CAPTIONS

Figure 1: PIV vector plots of the streamwise and spanwise fluctuating ve-

locity field at y + ≈ 20 in (a) water and drag reduction of (b) ≈ 60%. Flow is

from top to bottom. Vector color corresponds to magnitude (White et al., 2004).

Instantaneous visualizations of the near-wall vortex structures (white isosurfaces)

identified using isosurfaces of the positive second invariant of the velocity gradi-

ent tensor, high speed velocity streaks (red isosurfaces), and low speed velocity

streaks (blue isosurfaces) for (c) Newtonian fluid and drag reduction of (d) 60%

(courtesy of Y. Dubief). Flow is from left to right.

Figure 2: (a) Schematic of the basic structure of a polyethylene oxide (PEO)

polymer molecule. The PEO molecule is constructed from N CH2 –CH2 –O monomers

of length a. The end-to-end vector of the molecule is denoted as q. (b) Schematic

of polymer stretch (and relaxation) in shear flow. Polymer stretch is characterized

by the change in q.

Figure 3: A schematic illustrating the onset and different trajectories of poly-

mer drag reduction. The red dashed line represents the case where Re is fixed

(at the value where onset of drag reduction is first observed) and polymer con-

centration, C, is increased. The blue dotted line represents the case where C is

fixed and Re is increased.

Figure 4: Plot representing the downstream development of DR in external

boundary layer flows with polymer injection (Hou et al., 2007). DR is plotted

as a function of K, where K ≡ Qi Ci /ρXs U∞ , Qi is the volumetric injection

rate per unit span, Ci is the polymer concentration, ρ is the solvent density, Xs

the downstream distance, and U∞ the free-stream velocity. The letters denote

data obtained for different “grades” of PEO (Polyox COAG and Polyox 301) and
36 White & Mungal

injection concentrations.

Figure 5: A three-dimensional snapshot of vortices and contours of polymer

stretch obtained in a DNS channel flow with polymer drag reduction. The color

(both in the planes and on the wall) corresponds to polymer stretch and the

white isosurfaces identify the near-wall vortex structures using the positive second

invariant of the velocity gradient tensor (courtesy of Y. Dubief).


www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 37

(a) (b)

(c) (d)
38 White & Mungal

(a)
(CH 2 − CH 2 − O) N

CH 2 − CH 2 − O polymerization

a q

(b)
shear, strain

q
stretch
l

q relax

coiled configuration stretched configuration

Figure 2: (a) Schematic of the basic structure of a polyethylene oxide (PEO) poly-

mer molecule. The PEO molecule is constructed from N CH2 –CH2 –O monomers

of length a. The end-to-end vector of the molecule is denoted as q. (b) Schematic

of polymer stretch (and relaxation) in shear flow. Polymer stretch is characterized

by the change in q.
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 39

40

35 MDR asymptote

30 laminar

25
1
1/√f

20
f
Re

15 onset
C
10
Prandtl-Kármán law
5

0
2 2.5 3 3.5 4
(
log(Re√f)
log Re f )
40 White & Mungal

90
A Coa 2000 ppm
80 B Coa 1000 ppm
A G
A
C Coa 1000 ppm G G A G G
H H B A
D Coa 500 ppm C
H C C B C
70 E Coa 250 ppm I AH B
F Coa 100 ppm H B
I
60 G 301 2000 ppm I C DB D D
H 301 1000 ppm
I 301 500 ppm J
50 J 301 250 ppm E IE
DR

E
K 301 100 ppm K I
D
40 J J
J
E
K D K
F F
30

20 K E
J
F
10 F
K

F
0 -1
10 100 101
8
102
K×10

Figure 4: Plot representing the downstream development of DR in external

boundary layer flows with polymer injection (Hou et al., 2007). DR is plot-

ted as a function of K, where K ≡ Qi Ci /ρXs U∞ , Qi is the volumetric injection

rate per unit span, Ci is the polymer concentration, ρ is the solvent density, Xs

the downstream distance, and U∞ the free-stream velocity. The letters denote

data obtained for different “grades” of PEO (Polyox COAG and Polyox 301) and

injection concentrations.
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 41

Figure 5: A three-dimensional snapshot of vortices and contours of polymer

stretch obtained in a DNS channel flow with polymer drag reduction. The color

(both in the planes and on the wall) corresponds to polymer stretch and the white

isosurfaces identify the near-wall vortex structures using the positive second in-

variant of the velocity gradient tensor (courtesy of Y. Dubief).

You might also like