Perhitungan Viskositas
Perhitungan Viskositas
Perhitungan Viskositas
2008 40
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
1
2 White & Mungal
1 INTRODUCTION
liquids can produce profound effects on a wide variety of flow phenomena that
tion. This is most evident in turbulent boundary layers, where dissolving parts-
alone (e.g., Virk, 1975). The corresponding effect on the character of the flow
near-wall turbulent structure between Newtonian flow and polymer drag reduced
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
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
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
polymer drag reduction can be determined, it is conceivable that the effect can
additives, among other strategies (Gad-el Hak, 2000, Choi & Karniadakis, 2003).
Success on this front would have tremendous impact on the economics of energy
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
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,
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
The activity that occurred after Toms’ discovery soon produced sufficient exper-
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
is that, for a fixed pipe diameter, the Reynolds number, Re, at which drag reduc-
pend fundamentally on Re (Monin & Yaglom, 1975) and polymer dynamics de-
pend fundamentally on the number of monomers (Bird et al., 1987, Doi, 1996).
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
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
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
Indeed, the data of Nadolink (1987) using monodisperse polymers (i.e., a single
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-
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-
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
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
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
for formulating eddy viscosity models for computational fluid dynamics (CFD)
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
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
& White, 2000). This is best observed and explained from model systems of a
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
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
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.
pretations of the viscous and elastic theories of drag reduction. Two generalized
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
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
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
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
regime that lies in the transitional flow region between laminar and turbulent
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
bulence. These important findings and their implications are discussed in more
The most notable application of polymer drag reducing additives is in the Trans-
Alaska Pipeline System (TAPS). Polymer injection (and the subsequent drag
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
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-
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,
sion during flow (Barnard & Sellin, 1972, Zakin & Houston, 1980, Den Toonder
N with the effect of a decrease in the drag reducing effectiveness of the polymer.
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
3 RECENT ADVANCEMENTS
The equation for the conservation of momentum for the isothermal flow of an
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
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
of τijp and its divergence should provide a necessary thread needed to begin to
day diagnostics τijp cannot be measured directly in a turbulent flow and must be
measurements in which the polymer shear stress can be “inferred” from the mea-
surements, and numerical simulations where the polymer dynamics are modeled
3.1 Experiments
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
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
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
∂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
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
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
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
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
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
Using the FENE-P model, the numerical simulation involves solving the conti-
nuity equation, the equation for the conservation of momentum (in carrying out
Beris & Edwards, 1994) and the equation for the evolution of the configuration
tensor given by
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
Here we provide brief details and examples of our earlier statement that the
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
The statistics measured or computed most often are the mean streamwise velocity
The statistical data from laboratory flows and from numerical simulations of
is extensive (see, Nieuwstadt & Den Toonder, 2001), and much is known about
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
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
stress (ρúv́). The kinetic energy budgets at low DR follow similar trends as the
the turbulent kinetic energy budget suggests that the effects of polymers on the
the mean streamwise component of the pressure-strain term, responsible for the
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
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.
The literature on turbulent boundary layers with injection (i.e., external flows
boundary layer flow with polymer injection (or ejection) is expected to bear some
similarities to fully developed channel flows, but also some significant differences
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
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.,
turbulent mixing and the resulting drop in the near-wall concentration (Fontaine
rized on the DR vs log (K) plot shown in Figure 4, where the “K” parameter
is the volumetric injection rate per unit span, Ci is the polymer concentration,
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,
ferences due to polymer types exist, onset of drag reduction occurs for K > 10−9
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
closely tied to the behavior of the DR. It follows that the slow decrease in the
fluctuations during drag reduction. In brief, polymers are not passive scalars and
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
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-
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
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
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
are performed at speeds up to 18 m/s using a range of polymer types and con-
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.
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.
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
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-
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.
wise velocity streaks (see Figure 1). In the context of the wall-model, these
cycle (Dubief et al., 2004). However, despite the capabilities of the numerical
simulations, complete details of how polymers disrupt the cycle remains open to
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-
of the near-wall structures (i.e, they are non-space filling) presents additional
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
For the most part, numerical simulations directly investigating the interactions
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
The polymer stress gradients near the vortices can arise either from polymers
(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
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
5 SUMMARY
tal measurement techniques, the last decade has produced significant progress in
sponsible for drag reduction. Here we summarize these advancements and provide
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
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
polymer drag reduction (at least for low to moderate Re), for predictive purposes
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
5.3 Outlook
reduced turbulence models for the Large Eddy Simulation (LES) and Reynolds-
niques will allow for the modeling and optimization of applications at high Re
mer additives in flows at high Re, with pressure gradients, and with separation
ACKNOWLEDGMENTS
We thank Y. Dubief for his contributions to this paper. We also thank Costas
Eric Shaqfeh and Vijay Somandepalli for their valuable insights into the problem
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
References
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
Bird RB, Curtiss CF, Armstrong RC, Hassager O. 1987. Dynamics of Polymeric
Brown GL, Thomas ASW. 1977. Large structure in a turbulent boundary layer.
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
De Angelis E, Casciola CM, L’vov VS, Piva R, Procaccia I. 2003. Drag reduction
Den Toonder JMJ, Draad AA, Kuiken GDC, Nieuwstadt FTM. 1995. Degrada-
Den Toonder JMJ, Hulsen MA, Kuiken GDC, Nieuwstadt FTM. 1997. Drag
Press
Dubief Y, Terrapon VE, et al. 2005. New answers on the interaction between
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-
Gad-el Hak M. 2000. Flow Control: Passive, Active, and Reactive Flow Manage-
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.
shear stress and polymer stress profiles in drag reduced boundary layer flows.
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
Hunt JCR, Morrison JF. 2000. Eddy structure in turbulent boundary layers.
Joseph DD. 1990. Fluid Dynamics of Viscoelastic Liquids. New York, New York:
Springer Verlag
Kim HT, Kline SJ, Reynolds WC. 1971. The production of turbulence near a
Koskie JE, Tiederman WG. 1991. Polymer drag reduction in a zero pressure
Landahl MT. 1973. Drag reduction by polymer addition. In 13th Intl Congr.
Liaw GC, Zakin JL, Patterson GK. 1971. Effects of molecular characteristics of
Lumley JL. 1969. Drag reduction by additives. Ann. Rev. Fluid Mech. 1:367–384
a turbulent flow field using PIV and FENE bead-spring model. Ph.D. thesis,
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
Metzner AB, Metzner AP. 1970. Stress levels in rapid extensional flows of poly-
Min T, Choi H, Yoo JY. 2003. Maximum drag reduction in a turbulent channel
Min T, Yoo JY, Choi H, Joseph DD. 2004. Drag reduction by polymer additives
Myska J, Zakin JL, Chara Z. 1995. Viscoelasticity of a surfactant and its drag-
Nadolink RH. 1987. Friction reduction in dilute polystyrene solutions. Ph.D. the-
Verlag
NRC. 1997. Submarine platform technology. In Technology for the United States
Paterson RW, Abernathy FH. 1970. Turbulent flow drag reduction and degrada-
Petrie HL, Fontaine AA. 1996. Comparison of turbulent boundary layer modi-
Ptasinski PK, Boersma BJ, et al. 2003. Turbulent channel flow near maximum
490:251–291
Ptasinski PK, Nieuwstadt FTM, van den Brule BHAA, Hulsen MA. 2001. Ex-
Robinson SK. 1991. Coherent motions in the turbulent boundary layer. Ann. Rev.
Roy A, Larson RG. 2005. A mean flow model for polymer and fiber turbulent
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
Somasi M, Khomami B, Woo NJ, Hur JS, Shaqfeh ESG. 2002. Brownian dynamics
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-
Sureshkumar R, Beris AN, Handler RA. 1997. Direct numerical simulation of the
Swearingen JD, Blackwelder RF. 1987. The growth and breakdown of streamwise
Lett. 2(7):519–522
Terrapon VE, Dubief Y, Moin P, Shaqfeh ESG, Lele SK. 2004. Simulated polymer
Tiederman WG. 1990. The effect of dilute polymer solutions on viscous drag
www.AnnualReviews.org•Mech. & Pred. of Turbulent DR with Polymer Additives 33
Toms BA. 1948. Some observations on the flow of linear polymer solutions through
straight tubes at large Reynolds numbers. Proc. 1st Intern. Congr. Rheol.,
van Heel APG, Hulsen MA, van den Brule BHAA. 1998. On the selection of
271
Vanapalli SA, Islam MT, Solomon MJ. 2005. Scission-induced bounds on maxi-
Virk PS, Merril EW, Mickley HS, Smith KA, Mollo-Christensen EL. 1967. The
Mech. 20(10):22–30
Warholic MD, Heist DK, Katcher M, Hanratty TJ. 2001. A study with particle-
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
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-
Zakin JL, Houston DL. 1980. Effect of polymer molecular variables on drag re-
Zhou Q, Akhavan R. 2003. A comparison of fene and fene-p dumbbell and chain
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).
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%
polymer molecule. The PEO molecule is constructed from N CH2 –CH2 –O monomers
by the change in q.
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
boundary layer flows with polymer injection (Hou et al., 2007). DR is plotted
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.
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
(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
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
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
boundary layer flows with polymer injection (Hou et al., 2007). DR is plot-
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
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-