Vous êtes sur la page 1sur 4

Manifestations of Drag Reduction by Polymer Additives

in Decaying, Homogeneous, Isotropic Turbulence

Prasad Perlekar,1, ∗ Dhrubaditya Mitra,2, † and Rahul Pandit1, ‡


1
Centre for Condensed Matter Theory, Department of Physics,
Indian Institute of Science, Bangalore 560012, India.
2
Observatoire de la Côte d‘Azur, BP 4229, 06304 Nice Cedex 4, France.
The existence of drag reduction by polymer additives, well established for wall-bounded turbulent
flows, is controversial in homogeneous, isotropic turbulence. To settle this controversy we carry out
a high-resolution direct numerical simulation (DNS) of decaying, homogeneous, isotropic turbulence
with polymer additives. Our study reveals clear manifestations of drag-reduction-type phenomena:
arXiv:nlin/0609066v1 [nlin.CD] 27 Sep 2006

On the addition of polymers to the turbulent fluid we obtain a reduction in the energy dissipation
rate, a significant modification of the fluid energy spectrum especially in the deep-dissipation range,
a suppression of small-scale intermittency, and a decrease in small-scale vorticity filaments.

PACS numbers: 47.27.Gs, 47.27.Ak

The dramatic reduction of drag by the addition of iments [18, 21]: In particular, DR increases with c (upto
small concentrations of polymers to a turbulent fluid con- 25% in one of our simulations). For small values of c
tinues to engage the attention of engineers and physicists. the energy spectrum of the fluid is modified appreciably
Significant advances have been made in understanding only in the dissipation range; however, this suffices to
drag reduction both experimentally [1, 2, 3] and theo- yield significant drag reduction. We show that vorticity
retically [4, 5, 6, 7] in channel flows or the Kolmogorov filaments and intermittency are reduced at small spatial
flow [8]. However, the existence of drag-reduction-type scales and that the extension of the polymers decreases
phenomena in turbulent flows that are homogeneous and as c increases.
isotropic [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19] remains The NS and FENE-P (henceforth NSP) equations are
controversial. Some experimental [15, 16, 17, 18], numer- µ
ical [11, 12, 13, 14], and theoretical [9, 10] studies have Dt u = ν∇2 u + ∇.[f (rP )C] − ∇p; (1)
τP
suggested that drag reduction should occur even in ho-
f (rP )C − I
mogeneous, isotropic turbulence; but other studies have Dt C = C.(∇u) + (∇u)T .C − . (2)
refuted this claim [19]. τP
To settle this controversy we have initiated an exten- Here u(x, t) is the fluid velocity at point x and time t,
sive direct numerical simulation (DNS) of decaying, ho- incompressibility is enforced by ∇.u = 0, Dt = ∂t + u.∇,
mogeneous, isotropic turbulence in the presence of poly- ν is the kinematic viscosity of the fluid, µ the viscos-
mer additives. We monitor the decay of turbulence from ity parameter for the solute (FENE-P), τP the polymer
initial states in which the kinetic energy of the fluid is relaxation time, ρ the solvent density (set to 1), p the
concentrated at small wave vectors; this energy then cas- pressure, (∇u)T the transpose of (∇u), Cαβ ≡ hRα Rβ i
cades down to large wave vectors where it is dissipated the elements of the polymer-conformation tensor C (an-
by viscous effects; the energy-dissipation rate ǫ attains gular brackets indicate an average over polymer con-
a maximum at tm , roughly the time at which the cas- figurations), I the identity tensor with elements δαβ ,
cade is completed. A recent shell-model study [13] has f (rP ) ≡ (L2 − 3)/(L2 − rP2 ) the FENE-P
p potential that
suggested that this peak in ǫ can be used to quantify ensures finite extensibility, rP ≡ T r(C) and L the
drag reduction by polymer additives. Since shell mod- length and the maximum possible extension, respectively,
els are far too simple to capture the complexities of real of the polymers, and c ≡ µ/(ν + µ) a dimensionless mea-
flows, we have studied decaying turbulence in the Navier- sure of the polymer concentration [22]. c = 0.1 corre-
Stokes (NS) equation coupled to the Finitely Extensible sponds, roughly, to 100ppm for polyethylene oxide [1].
Nonlinear Elastic Peterlin (FENE-P) model [20] for poly- We consider homogeneous, isotropic, turbulence, so we
mers. Our study, designed specifically to uncover drag- use periodic boundary conditions and solve Eq. (1) by
reduction-type phenomena, shows that the position of using a massively parallel pseudospectral code [23] with
the maximum in ǫ depends only mildly on the polymer N 3 collocation points in a cubic domain (side L = 2π).
concentration c; however, the value of ǫ at this maxi- We eliminate aliasing errors [23] by the 2/3 rule, to
mum falls as c increases. We use this decrease of ǫ to de- obtain reliable data at small length scales, and use a
fine the percentage drag (or dissipation) reduction DR in second-order, slaved Adams-Bashforth scheme for time
decaying homogeneous, isotropic turbulence; we also ex- marching. For Eq. (2) we use an explicit sixth-order
plore other accompanying physical effects and show that central-finite-difference scheme in space and a second-
they are in qualitative accord with drag-reduction exper- order Adams-Bashforth method for temporal evolution.
2
P
The numerical error in rP must be controlled by choosing total kinetic energy E(t)P≡ k E(k, t), the energy-
a small time step δt, otherwise rP can become larger than dissipation-rate ǫ(t) ≡ ν k k 2 E(k, t), the cumulative
L, which leads to a numerical instability; this time step probability distribution of scaled polymer extensions
is much smaller than what is necessary for a pseudospec- P C (rP2 /L2 ), and the hyperflatness F6 (r) ≡ S6 (r)/S23 (r),
tral DNS of the NS equation alone. Table I lists the where Sp (r) ≡ h{[u(x + r) − u(x)] · r/r}p i is the order-p
parameters we use. We preserve the symmetric-positive- longitudinal velocity structure function and the angular
definite (SPD) nature of C at all times by using[22] the brackets denote an average over our simulation domain
following Cholesky-decomposition scheme: If we define at tm . For notational convenience, we do not display the
J ≡ f (rP )C, Eq. (2) becomes dependence on c explicitly.
Figure (1a) shows that ǫ first increases with time,
Dt J = J .(∇u) + (∇u)T .J − s(J − I) + qJ , (3)
reaches a peak, and then decreases; for c = 0 this peak
where s = (L2 − 3 + j 2 )/(τP L2 ), q = [d/(L2 − 3) − (L2 − occurs at t = tm . The position of this peak changes
3 + j 2 )(j 2 − 3)/(τP L2 (L2 − 3))], j 2 ≡ T r(J ), and d = mildly with c but its height goes down significantly as
T r[J .(∇u) + (∇u)T .J ]. Since C and hence J are SPD c increases. This suggests the following natural defini-
matrices, we can write J = LLT , where L is a lower- tion [13] of the percentage drag or dissipation reduction
triangular matrix with elements ℓij , such that ℓij = 0 for for decaying homogeneous, isotropic turbulence:
j > i. Thus Eq.(3) now yields (1 ≤ i ≤ 3 and Γij = ∂i uj )  f,m
ǫ − ǫp,m

DR ≡ × 100; (5)
X 1h sℓi1 i ǫf,m
Dt ℓi1 = Γki ℓk1 + (q − s)ℓi1 + (−1)(i mod 1) 2
2 ℓ11 here (and henceforth) the superscripts f and p stand, re-
k
ℓi2 X spectively, for the fluid without and with polymers and
+ (δi3 + δi2 ) Γm1 ℓm2 the superscript m indicates the time tm . Figure (1b)
ℓ11 m>1
shows plots of p DR versus c, for the Weissenberg num-
ℓ233 ber W e ≡ τP ǫf,m /ν ≃ 0.35, and versus W e, for
+ δi3 Γi1 , for i ≥ 1;
ℓ11 c = 1/11 ≃ 0.1. DR increases with c in qualitative accord
X ℓi1 X with experiments on channel flows (where DR is defined
Dt ℓi2 = Γmi ℓm2 − Γm1 ℓm2
ℓ11 via a normalized pressure difference); but it drops gently
m>2 m>2
as W e increases, in contrast to the behavior seen in chan-
1h (i+2) ℓi2
 ℓ221 i nel flows (in which τP is varied by changing the polymer).
+ (q − s)ℓi2 + (−1) s 2 1+ 2
2 ℓ22 ℓ11 In decaying turbulence, the total kinetic energy E(t)
h ℓ2  ℓ21  ℓ21 ℓ31 i of the fluid falls as t increases; the rate at which it falls
+ δi3 33 Γ32 − Γ31 +s 2 , for i ≥ 2; increases with c [Fig. (1c)], which suggests that the ad-
ℓ22 ℓ11 ℓ11 ℓ22
hX Γ ℓ i Γ ℓ ℓ ℓ dition of polymers increases the effective viscosity of the
3m 3m 31 32 21 33
Dt ℓ33 = Γ33 ℓ33 − ℓ33 + solution. This is not at odds with the decrease of ǫ with
ℓmm ℓ11 ℓ22
m<3 increasing c since the effective viscosity because of poly-
ℓ21 ℓ31 ℓ32 1h mers turns out to be scale-dependent. We confirm this
−s 2 + (q − s)ℓ33
ℓ11 ℓ22 ℓ33 2 by obtaining the kinetic-energy spectrum E p,m (k) for the
s  X ℓ2  sℓ221 ℓ232 i fluid in the presence of polymers at t = tm . For small
3m
+ 1+ + . (4) concentrations (c ≃ 0.1) the spectra with and without
ℓ33 m<3 mm
ℓ2 ℓ211 ℓ222 ℓ33
polymers differ substantially only in the deep dissipa-
The SPD nature of C is preserved by Eq.(4) if ℓii > 0, tion range, where E f,m (k) ≪ E p,m (k). As c increases,
which we enforce explicitly [22] by considering the evolu- to say c ≃ 0.4, E p,m (k) is reduced relative to E f,m (k)
tion of ln(ℓii ) instead of ℓii . at intermediate values of k [Fig. (2a)]; however, deep in
We use the following initial conditions (superscript the dissipation range E f,m (k) ≪ E p,m (k). We now de-
0): Cmn 0
(x) = δmn for all x; and u0m (k) = fine [12] the effective scale-dependent
P viscosity νe (k) ≡
0
Pmn (k)vn (k) exp(iθn (k)), with m, n = x, y, z, Pmn = ν + ∆ν(k), with ∆ν(k) ≡ −µ k−1/2<k′ ≤k+1/2 uk′ · (∇ ·
2 ′ 2 p,m ′
(δmn − km kn /k ) the transverse projection operator, k J )−k′ /[τP k E (k )], where (∇ · J )k is the Fourier
the wave-vector with components km = (−N/2, −N/2 + transform of ∇ · J . The inset of Fig. (2a) shows that
1, . . . , N/2), k = |k|, θn (k) random numbers distributed ∆ν(k) > 0 for k < 15, but ∆ν(k) < 0 around k =
uniformly between 0 and 2π, and vn0 (k) chosen such 20. This explains why E p,m (k) is suppressed relative to
that the initial kinetic-energy spectra are either of type E f,m (k) at small k, rises above it in the deep-dissipation
I 2 4
I, with E (k) = k exp(−2k ), or of type II, with range, and crosses over from its small-k to large-k be-
E II (k) = k 4 exp(−2k 2 ). haviors around the value of k where ∆ν(k) goes through
In addition to u(x, t), its Fourier transform uk (t), and zero.
C(x, t) we monitor the vorticity ω ≡ ∇ × u, the kinetic- Given the resolution of our DNS, inertial-range inter-
energy spectrum E(k, t) ≡ k−1/2<k′ ≤k+1/2 |u2k′ (t)|, the
P
mittency can be studied only by using extended self sim-
3

ilarity [24] as we will report elsewhere. However, we N δt L ν τP c


explore dissipation-range statistics further by calculat- NSP-96 96 1.0 × 10 −2
100 10−2
0.1 − 3 0.1, 0.2, 0.3, 0.4
ing the hyperflatness F6 (r) [Fig. (2b)]. The addition of NSP-192 192 1.0 × 10−2 100 10−2 1 0.1, 0.4
polymers slows down the growth of F6 (r), as r → 0, NSP-256A 256 1.0 × 10−2 100 10−2 1 0.1, 0.4
which signals the reduction of small-scale intermittency.
NSP-256 256 4.0 × 10−3 100 10−3 1 0.1, 0.4
This is further supported by the iso-|ω| surfaces shown
in Fig. (3). If no polymers are present, these iso-|ω| sur- TABLE I: The parameters N , δt, L, ν, τP and c for our
faces are filamentary [25] for large |ω|; polymers suppress four runs NSP − 96, NSP − 192, NSP − 256A, and NSP − 256.
a significant fraction of these filaments. NSP − 96, NSP − 192, NSP − 256A use type I initial conditions;
We use a rank-order method [26] to obtain P C (rP2 /L2 ) NSP − 256 uses an initial condition of type II. We also carry
and find that, as c increases [Fig. (2c)], the extension of out DNS studies of the NS equation with the √ same numerical

f,m
resolutions as our NSP runs. Re ≡ 20E / 3νǫf,m
the polymers decreases. We have checked that, in the p
and W e ≡ τP ǫf,m /ν; NSP-96: Re = 47.1 and W e =
passive-polymer version of Eqs.(1) and (2), the extension 0.03, 0.17, 0.24, 0.28, 0.31, 0.41, 0.48, 0.55, 0.62, 0.68, 1.03;
of polymers is much more than in Fig. (2c). NSP-192 and NSP-256A: Re = 47.1 and W e = 0.35; NSP-256:
Our study contrasts clearly drag reduction in homoge- Re = 126.6 and W e = 0.76.
neous, isotropic, turbulence and in wall-bounded flows.
In both these cases the polymers increase the overall vis-
cosity of the solution (see, e.g., Fig. (1c) and Ref.[12]). will stimulate more experimental studies of drag or dis-
In wall-bounded flows the presence of polymers inhibits sipation reduction in homogeneous, isotropic turbulence.
the flow of the stream-wise component of the momentum
into the wall, which, in turn, increases the net through- We thank C. Kalelkar, R. Govindarajan, V. Kumar,
put of the fluid and thus results in drag reduction, a S. Ramaswamy, L. Collins, and A. Celani for discus-
mechanism that can have no analog in homogeneous, sions, CSIR, DST, and UGC(India) for financial support,
isotropic turbulence. However, the decrease of ǫ(t) with and SERC(IISc) for computational facilities. DM is sup-
increasing c [Fig. (1b)] yields a natural definition of DR ported by the Henri Poincaré Postdoctoral Fellowship.
[Eq.(5)] for this case [27]. Thus, if the term drag reduc-
tion must be reserved for wall-bounded flows, then we
suggest the expression dissipation reduction for homo-
geneous, isotropic, turbulence. We have shown that νe

must be scale-dependent; its counterpart in wall-bounded Electronic address: perlekar@physics.iisc.ernet.in

flows is the position-dependent viscosity of Refs. [4, 7]. Electronic address: Dhrubaditya.MITRA@obs-nice.fr

Electronic address: rahul@physics.iisc.ernet.in;
Furthermore, as in wall-bounded flows, an increase in c
also at Jawaharlal Nehru Centre For Advanced Scientific
leads to an increase in DR [Fig. (1b)]. In channel flows Research, Jakkur, Bangalore, India.
an increase in W e leads to an increase in DR, but we find [1] P. Virk, AIChE 21, 625 (1975).
that DR falls marginally as W e increases [Fig. (1b)]. [2] P. van Dam, G. Wegdam, and J. van der Elsken, J. Non-
Our DNS of the Navier-Stokes equation with poly- Newtonian Fluid Mech. 53, 215 (1994).
mer additives [Eqs. (1) and (2)] resolves the controversy [3] J. D. Toonder, M. Hulsen, G. Kuiken, and F. Nieuwstadt,
J. Fluid Mech. 337, 193 (1997).
about drag reduction in decaying homogeneous, isotropic
[4] J. Lumley, J. Polym. Sci 7, 263 (1973).
turbulence and shows clearly that Eq. (5) offers a natu- [5] K. Sreenivasan and C. White, J. Fluid Mech. 409, 149
ral definition of DR for this case in a far more realistic (2000).
model than those of Refs. [11, 13]. We also find a non- [6] P. Ptasinski et al., J. Fluid Mech 490, 251 (2003).
trivial modification of the fluid kinetic-energy spectrum [7] V. L’vov, A. Pomyalov, I. Procaccia, and V. Tiberkevich,
especially in the deep-dissipation range [Fig. (2b)] that Phys. Rev. Lett. 92, 244503 (2004).
can be explained in terms of a polymer-induced, scale- [8] G. Boffetta, A. Celani, and A. Mazzino, Phys. Rev. E
dependent viscosity. Experiments [16, 17] do not resolve 71, 036307 (2005).
[9] M. Tabor and P. D. Gennes, Europhys. Lett. 2, 519
the dissipation range as clearly as we do, so the experi- (1986).
mental verification of the deep-dissipation-range behav- [10] J. K. Bhattacharjee and D. Thirumalai, Phys. Rev. Lett.
ior of Fig. (2a) remains a challenge. Earlier theoretical 67, 196 (1991).
studies [10, 11] have also not concentrated on this dis- [11] R. Benzi, E. de Angelis, R. Govindarajan, and I. Procac-
sipation range. The reduction in the small-scale inter- cia, Phys. Rev. E 68, 016308 (2003).
mittency [Fig. (2b)] and in the constant-|ω| isosurfaces [12] R. Benzi, E. Ching, and I. Procaccia, Phys. Rev. E 70,
026304 (2004) consider a scale-dependent viscosity for a
[Fig. (3)] is in qualitative agreement with channel-flow
shell model (but use an artificial diffusivity for polymers
studies [2], where a decrease in the turbulent volume frac- for numerical stability).
tion is seen on the addition of the polymers, and water-jet [13] C. Kalelkar, R. Govindarajan, and R. Pandit, Phys. Rev.
studies [21], where the addition of the polymers leads to E 72, 017301 (2004).
a decrease in small-scale structures. We hope our work [14] E. de Angelis, C. Casicola, R. Benzi, and R. Piva,
4

−4
x 10
4.8
(a) 25 (c)
5 4.4 (b) 0.06

DR
4
20
4 3.6
0.1 0.8 0.05

DR
We

E
ǫ

15
3
10 0.04
2
5 0.03
0.5 1 1.5 0.1 0.2 0.3 0.4 0.2 0.4 0.6 0.8 1 1.2
c t/τ
t/τ

FIG. 1: (Color online) (a) Temporal


p evolution of the energy dissipation rate ǫ (run NSP-256) for concentrations c = 0.1(−−) and
c = 0.4(solid line), with τ ≡ 2E (t = 0)/3L2 ; (b) percentage drag-reduction DR versus c (run NSP-192); the inset shows the
mild variation in DR with W e (runs NSP-96); (c) temporal evolution of the total fluid energy E for concentrations c = 0.1(−−)
and c = 0.4(solid line) (runs NSP-256). In (a) and (c) the plots for c = 0 (o−) are shown for comparison.

(a) (b) (c )
E p,m (k) or E f,m (k)

200
−10 −2
10 10

P C (rP2 /L2 )
−4 150
F6 (r)

x 10
8
−4
10
∆ν(k)

−20 5
10 100
2
−1 −6
−30 20 40 60
50 10
10 k
0 1 −3 −2
0.5 1 1.5 2 10
10
k
10 r rP2 /L2 10

FIG. 2: (Color online) (a) Plots of the energy spectra E p,m (k) or E f,m (k) versus k (run NSP-192) for c = 0.1(−−) and
c = 0.4(solid line) [E p,m (k) is unchanged if we use N = 256, with all other parameters the same (run NSP-256A)]; inset:
polymer contribution to the scale-dependent viscosity ∆ν(k) versus k for c = 0.1(−−); ∆ν(k) = 0 (solid line) is also shown
for reference; (b) the hyper-flatness F6 (r) as a function of r (run NSP-256) and concentration c = 0.4(solid line). In (a) and
(b) the corresponding plots with c = 0 (o−) are shown for comparison. (c) The cumulative PDF P C (rP2 /L2 ) versus rP2 /L2 for
c = 0.1(−−) and c = 0.4(solid line) (run NSP-256).

[19] D. Bonn et al., J. Phys. CM 17, S1219 (2005).


[20] A. Peterlin, J. Polym. Sci., Polym. Lett. 4, 287 (1966); H.
Warner, Ind. Eng. Chem. Fundamentals 11, 379 (1972);
R. Armstrong, J. Chem. Phys. 60 724 (1974); E. Hinch,
Phys. Fluids 20, S22 (1977).
[21] J. Hoyt and J. Taylor, Phys. Fluids 20, S253 (1977).
[22] T. Vaithianathan and L. Collins, J. Comput. Phys. 187,
1 (2003). We correct their Eq.(40) and definition of q.
FIG. 3: (Color online) Constant-|ω| isosurfaces for |ω| = [23] A. Vincent and M. Meneguzzi, J. Fluid Mech. 225, 1
h|ω|i + 2σ at tm without (a) and with (b) polymers, (run (1991); C. Canuto, M. Hussaini, A. Quarteroni, and
NSP-256) and c = 0.4; h|ω|i is the mean and σ the standard T. Zang, Spectral Methods in Fluid Dynamics (Spinger-
deviation of |ω|. Verlag, Berlin, 1988).
[24] R. Benzi et al., Phys. Rev. E 48, R29 (1993); S. Dhar, A.
Sain, and R. Pandit, Phys. Rev. Lett. 78, 2964 (1997).
[25] Y. Kaneda et al., Phys. Fluids 15, L21 (2003).
J. Fluid Mech 531, 1 (2005). [26] D. Mitra, J. Bec, R. Pandit, and U. Frisch, Phys. Rev.
[15] E. van Doorn, C. White, and K. Sreenivasan, Phys. Flu- Lett. 94, 194501 (2005).
ids 11, 2387 (1999). [27] In some steady-state simulations [11, 22] DR is associated
[16] W. McComb, J. Allan, and C. Greated, Phys. Fluids 20, with E p (k) > E f (k), for small k. We obtain this for type
873 (1977). II, but not type I, initial conditions; but Eq.(5) yields
[17] C. Friehe and W. Schwarz, J. Fluid Mech 44, 173 (1970). drag reduction for both of these initial conditions.
[18] D. Bonn, Y. Couder, P. van Dam, and S. Douady,
Phys. Rev. E 47, R28 (1993).

Vous aimerez peut-être aussi