ArticlePDF Available

Turbulence-induced secondary motion in a buoyancy-driven flow in a circular pipe

Authors:

Abstract

We analyze the results of a direct numerical simulation of the turbulent buoyancy-driven flow that sets in after two miscible fluids of slightly different densities have been initially superimposed in an unstable configuration in an inclined circular pipe closed at both ends. In the central region located midway between the end walls, where the flow is fully developed, the resulting mean flow is found to exhibit nonzero secondary velocity components in the tube cross section. We present a detailed analysis of the generation mechanism of this secondary flow which turns out to be due to the combined effect of the lateral wall and the shear-induced anisotropy between the transverse components of the turbulent velocity fluctuations.
Turbulence-induced secondary motion in a buoyancy-driven flow
in a circular pipe
Yannick Hallez and Jacques Magnaudet
INPT, UPS, IMFT (Institut de Mécanique des Fluides de Toulouse), Université de Toulouse,
Allée Camille Soula, F-31400 Toulouse, France and CNRS, IMFT, F-31400 Toulouse, France
Received 13 June 2009; accepted 17 July 2009; published online 25 August 2009
We analyze the results of a direct numerical simulation of the turbulent buoyancy-driven flow that
sets in after two miscible fluids of slightly different densities have been initially superimposed in an
unstable configuration in an inclined circular pipe closed at both ends. In the central region located
midway between the end walls, where the flow is fully developed, the resulting mean flow is found
to exhibit nonzero secondary velocity components in the tube cross section. We present a detailed
analysis of the generation mechanism of this secondary flow which turns out to be due to the
combined effect of the lateral wall and the shear-induced anisotropy between the transverse
components of the turbulent velocity fluctuations. © 2009 American Institute of Physics.
DOI: 10.1063/1.3213246
Since Prandtl’s group pioneering investigations,1high-
Reynolds-number incompressible flows in curved channels
or in straight noncircular ducts are known to involve second-
ary mean motions. Centrifugal or other nonconservative
body forces perpendicular to the main motion tend to skew
the primary spanwise vorticity in developing flows, leading
to transverse motions whose magnitude can be up to 20%–
30% of the streamwise velocity. Secondary flows typically
one order of magnitude smaller are also known to exist in
fully-developed turbulent flows in straight rectangular ducts.2
These weak secondary motions, known as Prandtl’s second-
ary flows of second kind, are due to the inhomogeneity of the
transverse Reynolds stresses near the corners of the duct and
result in a pair of counter-rotating streamwise vortices near
each corner. They have been the subject of several detailed
investigations, as they induce significant transport of mo-
mentum and heat within the duct cross section and also be-
cause their prediction represents a serious challenge for one-
point turbulence models.3During a joint experimental/
numerical investigation of buoyancy-induced turbulent
mixing in an inclined tube see Refs. 4and 5for more detail
on the experiments, we observed the existence of a second-
ary mean motion whose topology and sustaining mechanism
differ from those of the secondary flows reported so far,
since it arises in a straight circular pipe in presence of a
fully-developed primary flow. It is the purpose of this letter
to describe the structure of this, apparently undocumented,
secondary flow and elucidate its generating mechanism.
The physical configuration we consider is as follows. A
circular pipe of diameter dand length Lclosed at both ends
x=L/2is filled with two miscible i.e., with zero inter-
facial tensionlow-viscosity fluids of different densities
2
1in such a way that each fluid initially occupies half of
the pipe length, the two fluids being separated by a lock at
x=0. The pipe is tilted at an angle from the vertical in
such a way that the arrangement is unstable, owing to buoy-
ancy. After the lock is removed at time t= 0, the light heavy
fluid tends to flow along the upper lowerpart of the cross
section toward the upper lowerend wall, forming a stably
stratified turbulent countercurrent flow where turbulent mix-
ing takes place. As far as the fronts of both currents remain
far from the end walls, this relative motion results in a
turbulent shear flow which, in the region xL/2 on which
we focus in this letter, is statistically independent of both x
and t.
The results to be discussed below were obtained by solv-
ing the three-dimensional, time-dependent Navier–Stokes
equations for a variable-density incompressible flow with no
explicit reference to the Boussinesq approximation, together
with the density equation, assuming molecular diffusivity to
be negligibly small. The computations were carried out in a
176dlong circular pipe of diameter d, using a cylindrical
grid with 32642816 grid points in the radial, azimuthal,
and streamwise directions, respectively. Details regarding the
computational technique and the validation of the code may
be found elsewhere6and are not repeated here. In the ex-
ample discussed below, we select a tilt angle =15°, an
Atwood number At=
2
1/
2+
1=10−2 and a Reynolds
number Re=Vtd/
=790 with Vt=At gd1/2,gdenoting
gravity and
being the kinematic viscosity which we assume
to be identical in both fluids. The x,taveraging procedure
was performed throughout a 9d-long window centered at x
=0, starting at time TIAt g/d1/2=177 and continued until
the final time of the computation, TFAt g/d1/2=554, so that
the averaged quantities hereinafter denoted with angle
bracketsresult from averaging over 180 000 time steps and
384 grid cells.
Let the x-axis be along the streamwise ascending direc-
tion, the z-axis be such that the x,zplane is vertical with
z0 in the ascending direction, and the y-axis be horizontal
and such that the Cartesian x,y,zcoordinate system is
right-handed with y=z=0 on the tube centerline. The aver-
aged velocities in the vertical diametrical plane y=0 are
shown in Fig. 1. The streamwise velocity Uexhibits the
expected S-shape with an almost constant shear except in the
PHYSICS OF FLUIDS 21, 081704 2009
1070-6631/2009/218/081704/4/$25.00 © 2009 American Institute of Physics21, 081704-1
Downloaded 25 Aug 2009 to 193.48.203.101. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
vicinity of the wall. The puzzling feature of the flow is the
crosswise velocity Wwhich is seen to be nonzero even
though it is typically only 3%–5% of U. This secondary
velocity is negative positivein the upper lowerpart of the
vertical midplane, indicating that the fluid converges toward
the center of the pipe for y0. The complete spatial struc-
ture of the secondary velocity field is shown in Fig. 2, to-
gether with the so-called swirling strength ci which is com-
monly used as a criterion allowing the identification of
vortical structures in three-dimensional flows ci is defined
as the imaginary part of the conjugate eigenvalues of the
velocity gradient tensor7. This plot confirms the presence of
four stationary streamwise vortices within the pipe cross sec-
tion. Note that the secondary flow revealed by Fig. 2only
exhibits approximate left-right and top-bottom symmetries,
owing to the limited sampling time allowed by the computa-
tional run.
Obviously the nonzero secondary mean flow is directly
related to the existence of a nonzero mean component of the
streamwise vorticity we omit the brackets for simplicity
whose magnitude not shown hereis observed to be
O关共At g/d1/2, even though the mean secondary velocities
themselves are only a few percents of the primary velocity
U. We shall come back to this point later.
To elucidate the origin of the secondary mean flow, we
need to consider the balance equation for . Since the mean
flow is both stationary and homogeneous in the streamwise
direction, this equation reduces to
V
y+W
z=B+R+
yy
2+
zz
2,1
where
B=具共
y
zp
z
yp/
2,
2
R=
yz
2共具v2w2典兲
yy
2
zz
2兲具vw,
p,
,v, and wstanding for the pressure, density, and tur-
bulent velocity fluctuations along the y- and z-axes, respec-
tively. The first and last terms in Eq. 1represent the trans-
port of by the mean secondary flow and its diffusion by
viscous effects, respectively, so that both of them remain
zero if the mean flow is parallel. Therefore the generation of
is due to term Bresulting from baroclinic effects and to
term Rinduced by variations of the secondary Reynolds
stresses within the cross section.
The numerical data were used to evaluate the latter two
terms throughout the pipe cross section. Figure 3indicates
that both terms are concentrated near the wall and exhibit a
four-lobe structure which is antisymmetric with respect to
the horizontal and vertical axes. The baroclinic torque Bis
typically five times smaller than the turbulent term R.
Hence, while buoyancy effects drive the main flow, they are
not at the root of the secondary flow. The turbulent term R
appears to be the key of the generation of the streamwise
vorticity, suggesting that the present secondary flow is of
Prandtl’s second type. To elucidate the origin of the second-
ary flow, it is thus necessary to understand the spatial struc-
FIG. 1. Averaged velocity profiles in the vertical midplane. Bold line:
streamwise velocity; thin line: crosswise velocity. Lengths are normalized
by the pipe diameter dand velocities are normalized by Vt=At gd1/2.
0.5
0.4
0.3
0.2
0.1 0 0.1 0.2 0.3 0.4 0.5
0.5
0.4
0.3
0.2
0.1
0
0.1
0.2
0.3
0.4
0
.5
0.15
0.2
0.25
0.3
0.35
0.4
FIG. 2. Mean secondary flow and streamwise vortices identified by the
magnitude of the swirling strength ci criterion. The normalization is simi-
lar to that of Fig. 1.
0
.4
0
.2
0 0
.2
0
.4
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
0.4
0
.4
0
.2
0 0
.2
0
.4
−0.4
−0.3
−0.2
−0.1
0
0.1
0.2
0.3
0.4
FIG. 3. Map of the turbulent term Rleftand baroclinic term Brightin
Eq. 1, both normalized by Vt/d2. The contour levels of Rrange from
0.342 to 0.342 with a step of 0.049, while those of Brange from 0.085
to 0.085 with a step of 0.015. Solid dashedlines denote positive negative
values. The zero level has been removed for clarity.
081704-2 Y. Hallez and J. Magnaudet Phys. Fluids 21, 081704 2009
Downloaded 25 Aug 2009 to 193.48.203.101. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
ture of the secondary Reynolds stresses. The spatial distribu-
tion of Rand Brevealed by Fig. 3clearly suggests that the
lateral wall plays a key role in the mechanism, even though
no geometrical anisotropy exists here, in contrast with the
case of a rectangular duct. The other crucial ingredient,
which is specific to the present situation, is the existence of
the nonaxisymmetric mean shear
zUin the flow. This sa-
lient characteristic of the primary flow, which results directly
from the existence of the light ascending and heavy descend-
ing currents, does not exist in a usual Poiseuille flow, allow-
ing Uto be almost constant in the core in the latter case.
Several classical studies of the turbulent structure in homo-
geneous shear flows have revealed that, in presence of a
constant and nonzero
zU, the three main Reynolds stresses
are such that u2v2w2.810 This is because the
turbulent fluctuation, which is produced only along the
x-direction, is then unevenly redistributed along the other
two directions, owing to the anisotropic transport and
stretching induced by the shear. In particular Refs. 9and 10
found the ratio v21/2/w21/2to range from 1.15 to 1.5,
depending on the time ratio S=Tt
zU,Ttdenoting the
large-eddy turnover time. In presence of a stable stratifica-
tion characterized by the Richardson number Ri
=g
z
/
兲共
zU典兲−2,v21/2/w21/2has been found to in-
crease from about 1.28 for Ri=0 to about 1.44 for Ri=1,
which reflects the gradual inhibition of vertical motions as Ri
increases.11 Present results indicate v21/2/w21/2=1.3
with Ri=0.07 near the diametrical midplane z=0, Ri being
based on the effective gravity gsin in the z-direction.
Hence the anisotropy level determined in the present compu-
tations is consistent with known results and indicates that the
main cause of the difference in magnitude between the span-
wise fluctuation vand the crosswise fluctuation win the
core of the flow is the shear rather than the stratification.
Within a large part of the pipe cross section, the mean
shear is almost constant see Fig. 1so that the flow is close
to a homogeneous turbulent shear flow. Hence the anisotropy
level determined above does not vary much and the cross-
correlation vwis also almost zero since the y-axis is a
principal axis of the time-averaged flow. This allows us to
conclude that the whole term Ris negligibly small outside
the peripheral region directly influenced by the wall. Since
the main gradients lie along the radial direction in the near-
wall region, the analysis of this region is more easily
achieved in cylindrical coordinates. Defining the polar coor-
dinate system r,
so that r,
,xis right-handed and intro-
ducing the associated radial and azimuthal velocity
fluctuations vr,v
, it may be shown that the generation
term becomes R=1/r2
r
2r共具vr
2v
2典兲兴
rr
21/r2
␪␪
2
+3/r
r兴具vrv
. In a constant-density turbulent flow in a cir-
cular pipe, no average quantity depends on
and vrand v
are uncorrelated, so that the above expression of Rmakes it
clear that no secondary flow can exist in this case. Consid-
ering that the thickness
of the near-wall boundary layer
is much less than the pipe radius implies that, in the vicinity
of the wall, the radial derivative of any Reynolds stress in-
volved in Ris proportional to 1/
and thus dominates
over terms with a 1/rprefactor. Hence, although the relative
magnitude of vrv
and vr
2v
2is unknown, one has at
leading order in both contributions R⬇共1/r
r
2共具vr
2v
2典兲
rr
2vrv
.
Within the viscous sublayer adjacent to the wall which
given the modest Reynolds number extends approximatively
up to r/d=0.4 here,v
vrvaries linearly quadratically
with the distance d/2−r, owing to the no-slip condition and
the divergence-free constraint. Therefore we may write vr
2
v
2典⬇
兲共d/2−r2, with
0, by which we con-
clude that the radial derivative
r共具vr
2v
2典兲 is positive close
to the wall, whatever the azimuthal location
. Since the
mean velocity Ureaches its maximum at some radial loca-
tion r=rM
and returns to zero for r=d/2, a negative mean
shear exists in the range rM
rd/2. This negative shear
is maximum in the vertical midplane see Fig. 1and is zero
in the diametrical midplane z=0 where the mean velocity is
almost zero whatever the spanwise location. Hence, owing to
the generic shear-induced anisotropy of the Reynolds
stresses discussed above, we expect v
2to be larger near the
top and bottom of the cross section, where it corresponds to
the spanwise fluctuation, than near the midplane z=0 where
it corresponds to the crosswise fluctuation. This behavior,
which is confirmed by Fig. 4left, implies
00in
the first and third second and fourthquadrants, so that the
same conclusion holds for the whole term 1/r
r
2共具vr
2
v
2典兲.
Let us now consider the contribution to Rof the cross-
correlation vrv
which, according to the near-wall behavior
of vrand v
, evolves like
兲共d/2−r3within the viscous
sublayer. To determine the sign of
, it is useful to note
the geometrical property vrv
=−1
2共具v2w2典兲sin 2
+vwcos 2
. We may then take advantage of the left-right
and top-bottom symmetries of all second-order moments
keeping in mind that in a homogeneous turbulent shear flow
with a mean shear
zUthe y-axis is a principal directionto
conclude that vrv
is zero for both y=0 and z= 0. Hence we
expect vrv
to reach its maximum strength along the two
diagonals
=
/4, where vrv
=⫿共具v2w2典兲/2.
These various conclusions are confirmed by Fig. 4right.As
we saw above, the mean shear results in a positive difference
v2w2throughout the core of the pipe, so that
is
negative positivealong the first seconddiagonal. As the
sign of the second derivative
rrvrv
follows that of
,we
conclude that the contribution
rrvrv
to Ris positive
negativewithin the first and third second and fourthquad-
rants.
Coming back to Eq. 1it turns out that, whatever their
relative magnitude, both leading terms in Rhave the same
four-lobe structure and the same sign in each quadrant, i.e.,
Ris 00within the first and third second and fourth
quadrants, in agreement with the results displayed in Fig. 3
left.
The final point to be addressed is that of the scaling of
the secondary flow obtained through the above mechanism.
The first estimate we need is that of R. In the present flow,
the wall shear stress is responsible for the turbulent velocity
fluctuations in the region where Ris nonzero. The mean
axial velocity goes from zero at the wall to its maximum
081704-3 Turbulence-induced secondary motion Phys. Fluids 21, 081704 2009
Downloaded 25 Aug 2009 to 193.48.203.101. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
OVtvalue within the boundary layer of thickness
=ORe−1/2d, so that the wall shear stress scales like
ORe1/2
␳␯
Vt/d. Thus the magnitude uof the velocity fluc-
tuations obeys
u2=Re1/2
␳␯
Vt/d, implying u2/Vt
2=Re−1/2
note, however, that within the viscous sublayer vrv
is
smaller than v
2by ORe−1/2, owing to its cubic variation
with the distance to the wall as compared to the quadratic
variation of v
2典兴. It is then straightforward to determine the
magnitude of Rby noting that the radial variation of the
turbulent stresses arises within the viscous sublayer of thick-
ness O
, whereas their azimuthal variation arises over one-
fourth of the pipe perimeter, i.e., over an Oddistance. Ob-
viously this azimuthal variation and the correlation between
vrand v
also depends on the dimensionless shear rate S
through some function FSthat vanishes for S=0. Here
the turnover time of the large-scale turbulent motions is gov-
erned by the mean shear, so that S=O1. The aforemen-
tioned results on homogeneous sheared turbulence911 indi-
cate that the Reynolds stress anisotropy does not vary much
with Sin this range, so that here Fmay simply be regarded
as a nonzero constant. All this implies that Ris of
O关共Vt/d2, as indicated by Fig. 3left. Very close to the
wall, Ris balanced by the viscous diffusion of the stream-
wise vorticity last term in Eq. 1. This diffusion arises
within the boundary layer, so that the diffusion term is of
O
/
2, which implies =OVt/d. Finally, the second-
ary velocities reach their typical magnitude VSnear r=d/2
, so that the transport term in Eq. 1is of OVSVt/
d兲兴
and this term balances Rat this radial location, thus imply-
ing VS=OVt
/d=ORe−1/2Vt. These conclusions are con-
sistent with the features displayed by the computational re-
sults. In particular we recover the already noticed feature that
is of the order of Vt/d, even though the mean secondary
velocities are only a few percents of Vthere Re=790, so that
Re−1/20.036.
The above qualitative analysis allows the spatial struc-
ture and sign of the source term Rin Eq. 1and the scaling
of the resulting streamwise vorticity and secondary velocities
to be correctly predicted. This supports the idea that the sec-
ondary flow observed in our computations and recently con-
firmed in experiments5is a special case of Prandtl’s second-
ary flow of second kind which results from the combined
effect of the mean shear and the lateral wall on the turbu-
lence structure.
This research has been funded by the ANR through
Grant No. ANR-07-BLAN-0181. We are grateful to F. Moisy
for stimulating discussions on the generation mechanism of
the secondary flow.
1L. Prandtl, “Über die ausgebildete turbulenz,” Verfahren diese Zweite In-
ternationale Kongress für Technische Mechanik, Zürich 1926兲关“Turbu-
lent flow,” NACA Technical Memo 435 1927,p.62.
2P. Bradshaw, “Turbulent secondary flows,” Annu. Rev. Fluid Mech. 19,
53 1987.
3A. O. Demuren and W. Rodi, “Calculation of turbulence-driven secondary
motion in non-circular ducts,” J. Fluid Mech. 140, 189 1984.
4T. Séon, J. P. Hulin, D. Salin, B. Perrin, and E. J. Hinch, “Buoyant mixing
of miscible fluids in tilted tubes,” Phys. Fluids 16, L103 2004.
5J. Znaien, “Étude locale du mélange de deux fluides dans la géométrie
confinée d’un tube incliné,” Ph.D. thesis, Université Paris 11, 2009.
6Y. Hallez and J. Magnaudet, “Effects of channel geometry on buoyancy-
driven mixing,” Phys. Fluids 20, 053306 2008.
7J. Zhou, S. Balachandar, and T. M. Kendall, “Mechanisms for generating
coherent packets of hairpin vortices in channel flow,” J. Fluid Mech. 387,
353 1999.
8V. G. Harris, J. A. H. Graham, and S. Corrsin, “Further experiments in
nearly homogeneous turbulent shear flow,” J. Fluid Mech. 81,6571977.
9S. Tavoularis and S. Corrsin, “Experiments in nearly homogenous turbu-
lent shear flow with a uniform mean temperature gradient. Part 1,” J. Fluid
Mech. 104,3111981.
10S. Tavoularis and U. Karnik, “Further experiments on the evolution of
turbulent stresses and scales in uniformly sheared turbulence,” J. Fluid
Mech. 204, 457 1989.
11H.-J. Kaltenbach, T. Gerz, and U. Schumann, “Large-eddy simulation of
homogeneous turbulence and diffusion in stably stratified shear flow,” J.
Fluid Mech. 280,11994.
FIG. 4. Color onlineMap of the secondary anisotropy vr
2v
2典共leftand cross correlation vrv
典共right, both normalized by Vt
2.
081704-4 Y. Hallez and J. Magnaudet Phys. Fluids 21, 081704 2009
Downloaded 25 Aug 2009 to 193.48.203.101. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
... Prandtl's secondary currents of the second kind typically occur at channel corners or at transverse bed roughness transitions. Recently, it has also been shown that this kind of secondary current may be formed in buoyancy-driven flows even in straight circular pipes (Hallez and Magnaudet, 2009), where normally this feature does not exist. ...
... This experimentally guided conjecture was later replaced by the alternative idea that the observed suspended sediment streaks are generated by the preexisting secondary currents. A recent study by Hallez and Magnaudet (2009), however, points out the possibility of generating secondary currents, even by weak density stratification, which could have been created in the Vanoni experiments by the addition of suspended sediments. In light of Hallez and Magnaudet's (2009) study, it is beneficial to reconsider Vanoni's (1946) abandoned hypothesis, which may well be correct, representing a specific form of Prandtl's second kind of secondary current. ...
... A recent study by Hallez and Magnaudet (2009), however, points out the possibility of generating secondary currents, even by weak density stratification, which could have been created in the Vanoni experiments by the addition of suspended sediments. In light of Hallez and Magnaudet's (2009) study, it is beneficial to reconsider Vanoni's (1946) abandoned hypothesis, which may well be correct, representing a specific form of Prandtl's second kind of secondary current. ...
... In the following analysis the results will be shown as scalar views in the XY section of the pipe as it gives a better representation of the complex mixing phenomena occurring within the pipe cross section. The a b [49], are well captured by the present simulation with a very good symmetry of the vortices It is worth emphasising that these secondary flows correspond to only a few percents of the axial velocity, which makes them insignificant for practical purposes. More will be said about these secondary flows at the end of the paper when comparing the selected uRANS modelling approaches. ...
... As mentioned earlier, the existence of secondary flows was already discussed in the literature. Hallez and Magnaudet [49] specifically looked at mechanism generating these a b secondary flows. They showed that it can be explained by looking at the mean axial vorticity equation written as follows: ...
... This equation for mean axial vorticity was simplified by taking into account the fact the mean flow is stationary and homogeneous in the streamwise direction. Hallez and Magnaudet [49] proved that the main source of axial vorticity is due to the term denoted RHS2 in the above Eq. 57. ...
Article
Full-text available
In the present paper, both Large Eddy Simulation (LES) and unsteady-RANS (uRANS) CFD studies of single-phase buoyancy-driven counter-current flow in a pipe are presented for an Atwood number of 1.15×10−2 and an inclination angle of 15 degrees from the vertical. The basic flow phenomena involved are fundamentally the same as those encountered in various industrial applications including passive coolant flow in the loops of a nuclear reactor. Earlier work on this type of flow was focused on its physical aspects using both experimental measurements and Direct Numerical Simulation (DNS). The present work investigates the performance of several commonly used LES eddy viscosity subgrid-scale models. The results show that LES is able to reproduce accurately the experimental results and that the dynamic Smagorinsky subgrid-scale model gives the best predictions. The results of those calculations were then used to obtain more information on the physics of such flow. As regards uRANS simulations, a range of classic two-equation linear eddy viscosity models, based on low-y + formulation at the wall and the single gradient diffusion hypothesis for the turbulent density fluxes were compared against experimental data available in the literature. An elliptic blending Reynolds stresses model (EBRSM) with generalised gradient diffusion hypothesis (GGDH) for the scalar fluxes was then implemented and compared with experiments. The latter showed very good agreement for the first order moment with the experimental data whereas all the two-equation eddy viscosity based models showed rather high levels of discrepancy. The main cause of discrepancy was found to be the underprediction of the axial turbulent buoyancy production effect, which has a more detrimental impact on the eddy viscosity models than the second moment closure one.
... Moving from the Hele-Shaw paradigm into regimes of significant inertia, Hallez et al. looked into exchange flows, i.e., flows with zero imposed velocity in channels 21,22 and pipes 21,23 using Direct Numerical Simulations (DNSs). In inertial regimes, they discovered that there are essential differences between 2D and 3D flows which arise from different dynamic vorticity fields. ...
... In comparing with pipe flows, there are clear limitations. For example, in the turbulent pipe exchange flow simulations in Ref. 23, non-zero radial and azimuthal components of velocity were found that govern the mixing mechanisms. These (anisotropic) secondary motions are induced by a combined effect of the lateral walls and turbulent velocity fluctuations. ...
Article
We numerically study the displacement flow of two iso-viscous Newtonian fluids in an inclined two-dimensional channel, formed by two parallel plates. The results are complementary to our previous studies on displacement flows in pipes and channels. The heavier displacing fluid moves the lighter displaced fluid in the downward direction. Three dimensionless groups largely describe these flows: the densimetric Froude number (Fr), the Reynolds number (Re), and the duct inclination (β). As a first order approximation, we are able to classify different flow regimes phenomenologically in a two-dimensional (Fr; Re cos β/Fr)-plane and provide leading order expressions for the transitions between different regimes. The stabilizing and/or destabilizing effects of the imposed mean flow on buoyant exchange flows (zero imposed velocity) are described for a broad range of dimensionless parameters. C 2014 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4903822]
... Nezu and Rodi (1985) experimentally showed that lateral variations in bed topography and roughness can lead to the formation of secondary currents, which are independent of the side-wall effect or the corner induced secondary current Nakagawa 1993, Wang andCheng 2005). Besides these reasons, also it has been shown that this kind of secondary current may be formed in buoyancy-driven flow even in straight circular pipes (Hallez and Magnaudet 2009), where bed roughness and bed topography do not play any role. Further, Kinoshita (1967) postulated the form of streamwise secondary currents in straight rivers. ...
Article
Cellular secondary flows are inevitably present in turbulent flows through ducts, natural or artificial channels, and compound channels. Secondary currents significantly modify the characteristics of turbulent quantities, the pattern of primary flow velocity by causing dip-phenomenon. To understand the detailed mechanism and hidden cause, modelling of secondary flow velocities is crucial. In this study, proper mathematical models of secondary flow velocities along vertical and transverse directions are proposed for steady and uniform turbulent flow through wide open channels with equal smooth and rough bed strips. Starting from the continuity and the Reynolds averaged Navier-Stokes equations, governing equation for secondary velocity is derived first and then using appropriate boundary conditions (no-slip boundary conditions at channel bottom and free surface, and maximum vertical velocity in magnitude at the interface of two cellular secondary cells and at mid-depth of the channel. All these conditions are consistent with several experimental observations). A new model of the streamwise Reynolds shear stress is proposed for the entire cross-sectional plane and using it, the analytical solutions are obtained. Proposed models include the effects of viscosity of the fluid and the eddy viscosity model of turbulence. All suggested models are validated with existing experimental data in rectangular open-channel flows, compound open channel flows, and duct flows, and satisfactory results are obtained. Furthermore, models are also compared with existing empirical models from literature to show the effectiveness and superiority of proposed models. Apart from these, the obtained results from this study are used to investigate the effects of vertical and transverse secondary flow velocities on the settling velocity vector in a cross-sectional plane. Effective alternative models for the settling velocity vector are suggested. The model of settling velocity vector is also compared with the existing model. Finally, all results are justified from physical viewpoints.
... The displacement of a light fluid by a heavy fluid is of great importance in many geophysical and industrial applications. In the context of geophysical settings, a sudden or sustained release of a heavy fluid in a less dense or stratified environment, called gravity currents, has been studied by many authors in the past, e.g., Refs. 1 and 2. In a confined environment, such as a channel or pipe, the problem is termed as a lock-exchange flow which has been studied experimentally in vertical tubes, 3,4 tilted tubes, [5][6][7][8][9][10] and nearly horizontal tubes. 11 Our interest however is mainly the primary-cementing stage of oil/gas well production, where the drilling mud has to be displaced by a cement slurry and various preflushes downwards in the well, along the casing and then upward in the annulus, formed between the casing and surrounding earth; see Ref. 12. Successful displacement of the mud is crucial to ensure mechanical integrity and sealing of the well. ...
Article
We investigate experimentally the density-unstable displacement flow of two miscible fluids along an inclined pipe. This means that the flow is from the top to bottom of the pipe (downwards), with the more dense fluid above the less dense. Whereas past studies have focused on iso-viscous displacements, here we consider viscosity ratios in the range 1/10–10. Our focus is on displacements where the degree of transverse mixing is low-moderate, and thus a two-layer, stratified flow is observed. A wide range of parameters is covered in order to observe the resulting flow regimes and to understand the effect of the viscosity contrast. The inclination of the pipe (β) is varied from near horizontal β = 85° to near vertical β = 10°. At each angle, the flow rate and viscosity ratio are varied at fixed density contrast. Flow regimes are mapped in the (Fr, Re cos β/Fr)-plane, delineated in terms of interfacial instability, front dynamics, and front velocity. Amongst the many observations, we find that viscosifying the less dense fluid tends to significantly destabilize the flow. Different instabilities develop at the interface and in the wall-layers.
... 19 Due to certain process constraints, inclined columns may be preferred over the vertical ones. [19][20][21][22][23] The exchange flows have largely been studied in literature experimentally, [24][25][26][27][28][29][30][31] computationally, [32][33][34][35][36][37][38] and analytically [39][40][41][42] assuming fluids with equal temperature, i.e., isothermal. The non-isothermal flows can be exceptionally distinct from those of isothermal as revealed in the recent computational study of Ref. 43. ...
Article
We study non-isothermal buoyancy-driven exchange flow of two miscible Newtonian fluids in an inclined pipe experimentally. The heavy cold fluid is released into the light hot one in an adiabatic small-aspect-ratio pipe under the Boussinesq limit (small Atwood number). At a fixed temperature, the two fluids involved have the same viscosity. Excellent qualitative and quantitative agreement is first found against rather recent studies in literature on isothermal flows where the driving force of the flow comes from salinity as opposed to temperature difference. The degree of flow instability and mixing enhances as the pipe is progressively inclined towards vertical. Similar to the isothermal limit, maximal rate of the fluids interpenetration in the non-isothermal case occurs at an intermediate angle, β. The interpenetration rate increases with the temperature difference. The degree of fluids mixing and diffusivity is found to increase in the non-isothermal case compared to the isothermal one. There has also been observed a novel asymmetric behavior in the flow, never reported before in the isothermal limit. The cold finger appears to advance faster than the hot one. Backed by meticulously designed supplementary experiments, this asymmetric behavior is hypothetically associated with the wall contact and the formation of a warm less-viscous film of the fluid lubricating the cold more-viscous finger along the pipe. On the other side of the pipe, a cool more-viscous film forms decelerating the hot less-viscous finger. Double diffusive effects associated with the diffusion of heat and mass (salinity) are further investigated. In this case and for the same range of inclination angles and density differences, the level of flow asymmetry is found to decrease. The asymmetric behaviour of the flow is quantified over the full range of experiments. Similar to the study of Salort et al. [“Turbulent velocity profiles in a tilted heat pipe,” Phys. Fluids 25(10), 105110-1–105110-16 (2013)] for tilted heat pipes, a small Richardson number of Ri≈0.05 is found, above which flow laminarization occurs. In terms of the dimensionless numbers of the problem, it is found that the interpenetrative speeds of the heavy and light fluid layers in non-isothermal and double-diffusive cases increase with the dimensionless temperature difference, rT, Atwood number, At, Grashof number, Gr, Reynolds number, Re, Nahme number, Na, and Péclet number, Pe but decreases with Prandtl number, Pr, and Brinkman number, Br.
... Hallez et al. found stable and unstable exchange flows while quantifying the effects of the flow geometry in channels 13,14 and pipes. 13,15 Many flow features were studied in detail. For example, they found that vortices in 2D geometries are strongly present, coherent and long persistent, leading to a periodic cut of the channels of pure fluid feeding the front; see Ref. 13. Sahu et al. 16 also computationally studied displacement flows of miscible fluids in 2D channels. ...
Article
The displacement flow of two miscible Newtonian fluids is investigated experimentally in a vertical pipe of long aspect ratio (δ ⁻¹ ≈ 210). The fluids have a small density difference and they have the same viscosity. The heavy displacing fluid is initially placed above the light displaced fluid. The displacement flow is downwards. The experiments cover a wide range of the two dimensionless parameters that largely describe the flow: the modified Reynolds number (0 ≤ Ret⪅800) and the densimetric Froude number (0 ≤ Fr ≤ 24). We report on the stabilizing effect of the imposed flow and uncover the existence of two main flow regimes at long times: a stable displacement flow and an unstable displacement flow. The transition between the two regimes occurs at a critical modified Reynolds number R e t Critical, as a function of Fr. We study in depth the stable flow regime: First, a lubrication model combined with a simple initial acceleration formulation delivers a reasonable prediction to the time-dependent penetrating displacing front velocity. Second, we find two sub-regimes for stable displacements, namely, sustained-back-flows and no-sustained-back-flows. The transition between the two sub-regimes is a marginal stationary interface flow state, which is also well predicted by the lubrication model. The unstable regime is associated to instabilities and diffusive features of the flow. In addition, particular patterns such as front detachment phenomenon appear in the unstable flow regime, for which we quantify the regions of existence versus the dimensionless groups.
... These long-time flows can be either viscous or inertial, with the transition occurring at Re t cos β ≈ 50. In pipe exchange flow simulations, 13 non-zero radial and azimuthal components of velocity were found to govern the mixing mechanism. Although studying different mechanisms of mixing in the diffusive flow regime is not in the scope of the current work, we do characterize the boundaries of viscous/inertial regimes through the important parameter Re t cos β. ...
Article
Full-text available
Vorticity fluxes are analyzed in fully-developed turbulent flow through rectangular ducts with a width-to-height ratio of 3, and both straight and semi-cylindrical side-walls, at a centerplane friction Reynolds number Re τ,c 180. The transport of secondary Reynolds stresses by the secondary flow of Prandtl's second kind is analyzed from a vorticity-flux perspective. This analysis reveals that the in-plane transport of viscous stresses locally counteracts the inhomogeneous distribution of the turbulent shear-stress gradient in the spanwise direction. A relationship is established between the mean and fluctuating transport terms that can be useful to improve turbulence models and their ability to accurately predict the secondary flow. Finally, quadrant analysis is used to evaluate the contribution from the different types of bursting events to the fluctuating transport terms.
Article
We investigate the inertial flows found in buoyant miscible displacements using a two-layer model. From displacement flow experiments in inclined pipes, it has been observed that for significant ranges of Fr and Re cos β/Fr, a two-layer, stratified flow develops with the heavier fluid moving at the bottom of the pipe. Due to significant inertial effects, thin-film/lubrication models developed for laminar, viscous flows are not effective for predicting these flows. Here we develop a displacement model that addresses this shortcoming. The complete model for the displacement flow consists of mass and momentum equations for each fluid, resulting in a set of four non-linear equations. By integrating over each layer and eliminating the pressure gradient, we reduce the system to two equations for the area and mean velocity of the heavy fluid layer. The wall and interfacial stresses appear as source terms in the reduced system. The final system of equations is solved numerically using a robust, shock-capturing scheme. The equations are stabilized to remove non-physical instabilities. A linear stability analysis is able to predict the onset of instabilities at the interface and together with numerical solution, is used to study displacement effectiveness over different parametric regimes. Backflow and instability onset predictions are made for different viscosity ratios.
Article
Full-text available
Buoyant mixing of two fluids in tubes is studied experimentally as a function of the tilt angle theta from vertical, the density contrast and the common viscosity mu. At high contrasts and low theta, longitudinal mixing is macroscopically diffusive, with a diffusivity D increasing strongly with theta and mu. At lower contrasts and higher theta, a counterflow of the two fluids with little transverse mixing sets in. The transition occurs at an angle increasing with density contrast and decreasing with mu. These results are discussed in terms of the dependence of transerse mixing on theta and an analogy with the Boycott effect.
Article
Full-text available
The evolution of a single hairpin vortex-like structure in the mean turbulent field of a low-Reynolds-number channel flow is studied by direct numerical simulation. The structure of the initial three-dimensional vortex is extracted from the two-point spatial correlation of the velocity field by linear stochastic estimation given a second-quadrant ejection event vector. Initial vortices having vorticity that is weak relative to the mean vorticity evolve gradually into omega-shaped vortices that persist for long times and decay slowly. As reported in Zhou, Adrian & Balachandar (1996), initial vortices that exceed a threshold strength relative to the mean flow generate new hairpin vortices upstream of the primary vortex. The detailed mechanisms for this upstream process are determined, and they are generally similar to the mechanisms proposed by Smith et al. (1991), with some notable differences in the details. It has also been found that new hairpins generate downstream of the primary hairpin, thereby forming, together with the upstream hairpins, a coherent packet of hairpins that propagate coherently. This is consistent with the experimental observations of Meinhart & Adrian (1995). The possibility of autogeneration above a critical threshold implies that hairpin vortices in fully turbulent fields may occur singly, but they more often occur in packets. The hairpins also generate quasi-streamwise vortices to the side of the primary hairpin legs. This mechanism bears many similarities to the mechanisms found by Brooke & Hanratty (1993) and Bernard, Thomas & Handler (1993). It provides a means by which new quasi-streamwise vortices, and, subsequently, new hairpin vortices can populate the near-wall layer.
Article
Full-text available
The evolution of the concentration and flow fields resulting from the gravitational mixing of two interpenetrating miscible fluids placed in a tilted tube or channel is studied by using direct numerical simulation. Three-dimensional (3D) geometries, including a cylindrical tube and a square channel, are considered as well as a purely two-dimensional (2D) channel. Striking differences between the 2D and 3D geometries are observed during the long-time evolution of the flow. We show that these differences are due to those existing between the 2D and 3D dynamics of the vorticity field. More precisely, in two dimensions, the strong coherence and long persistence of vortices enable them to periodically cut the channels of pure fluid that feed the front. In contrast, in 3D geometries, the weaker coherence of the vortical motions makes the segregational effect due to the transverse component of buoyancy strong enough to preserve a fluid channel near the front of each current. This results in three different regimes for the front velocity (depending on the tilt angle), which is in agreement with the results of a recent experimental investigation. The evolution of the front topology and the relation between the front velocity and the concentration jump across the front are investigated in planar and cylindrical geometries and highlight the differences between 2D and 3D mixing dynamics.
Article
The development status of characterizations of conventional three-dimensional boundary layers and of the secondary flows with embedded streamwise vortices that are encountered in turbomachinery is evaluated. Attention is given to flows with strong skew-induced streamwise vorticity or dominated by stress-induced vorticity. A lack of basic physical insight into the effect of mean-flow three-dimensionality on turbulence structure is identified, with the primary area of difficulty being the behavior of the pressure-strain term in the Reynolds stress transport equations. It is recommended that engineering calculations be conducted by Reynolds-averaged methods, but with computer simulations of eddy motion furnishing detailed statistics on such matters as pressure fluctuation that cannot be adequately measured.
Article
Experiments on and calculation methods for flow in straight non-circular ducts involving turbulence-driven secondary motion are reviewed. The origin of the secondary motion and the shortcomings of existing calculation methods are discussed. A more refined model is introduced, in which algebraic expressions are derived for the Reynolds stresses in the momentum equations for the secondary motion by simplifying the modelled Reynolds-stress equations of Launder, Reece & Rodi (1975), while a simple eddy-viscosity model is used for the shear stresses in the axial momentum equation. The kinetic energy k and the dissipation rate ε of the turbulent motion which appear in the algebraic and the eddy-viscosity expressions are determined from transport equations. The resulting set of equations is solved with a forward-marching numerical procedure for three-dimensional shear layers. The model, as well as a version proposed by Naot & Rodi (1982), is tested by application to developing flow in a square duct and to developed flow in a partially roughened rectangular duct investigated experimentally by Hinze (1973). In both cases, the main features of the mean-flow and the turbulence quantities are simulated realistically by both models, but the present model underpredicts the secondary velocity while the Naot-Rodi model tends to overpredict it.
Article
A reasonably uniform mean temperature gradient has been superimposed upon a nearly homogeneous turbulent shear flow in a wind tunnel. The overheat is small enough to have negligible effect on the turbulence. Away from the wind-tunnel entrance, the transverse statistical homogeneity is good and the temperature fluctuations and their integral scales grow monotonically like the corresponding velocity fluctuations (Harris, Graham & Corrsin 1977). Measurements of several moments, one- and two-point correlation functions, spectra, integral scales, microscales, probability densities, and joint probability densities of the turbulent velocities, temperature fluctuations, and temperature-velocity products are reported. The heat-transport characteristics are much like those of momentum transport, with the turbulent Prandtl number nearly 1. The temperature fluctuation is better correlated with the streamwise than the transverse velocity component, and the cross-component D12 of the turbulent diffusivity tensor has sign opposite to and about twice the magnitude of the diagonal component D22. Some resemblance of directional properties (relative magnitudes of correlation functions, integral scales, microscales) of the temperature with those of the streamwise velocity is also observed. Comparisons of the present data with measurements in the inner part of a heated boundary layer and a fully turbulent pipe flow (x2/d = 0·25) show comparable magnitudes of temperature-velocity correlation coefficients, turbulent Prandtl numbers and ratios of turbulent diffusivities, and show similar shapes of two-point correlation functions.
Article
The experiment of Champagne, Harris & Corrsin in generating and studying a nearly homogeneous turbulent shear flow has been extended to larger values of the dimensionless downstream time or strain by the use of a larger mean velocity gradient in the same wind tunnel. The system appears to reach an asymptotic state in which scales and turbulent energy grow monotonically. Two-point covariances and tensor structure of one-point ‘Reynolds stress’ and ‘pressure/strain-rate covariance’ agree with the earlier case. However, the linear intercomponent energy exchange hypothesis due to Rotta, very roughly confirmed by the earlier experiment, is contradicted by the present data.
Article
Measurements of the Reynolds stresses, integral lengthscales and Taylor microscales are reported for several cases of uniformly sheared turbulent flows with shear values in a range substantially wider than those of previous measurements. It is shown that such flows demonstrate a self-preserving structure, in which the dimensionless Reynolds stress ratios and the dissipation over production ratio, ε/P, remain essentially constant. Flows with sufficiently large $k_{\rm s} = (1/\overline{U_{\rm c}}){\rm d}\overline{U_1}dx_2$ have exponentially growing stresses and ε/P [approximate] 0.68; a linear relationship between the coefficient in the exponentiallaw and ks is shown to be compatible with measurements having ks > 3. The possibility of a self-preserving structure with asymptotically constant stresses and ε/P [approximate] 1.0 is also compatible with measurements, corresponding to flows with small values of ks. The integral lengthscales appear to grow according to a power law with an exponent of about 0.8, independent of the mean shear, while the Taylor microscales, in general, approach constant values. Various attempts to scale the stresses and to predict their evolution are discussed and the applicability of Hasen's theory is scrutinized. Finally, an ‘exact’ expression for the pressure-strain rate covariance is derived and compared to some popular models.
Article
By means of large-eddy simulation, homogeneous turbulence is simulated for neutrally and stably stratified shear flow at gradient-Richardson numbers between zero and one. We investigate the turbulent transport of three passive species which have uniform gradients in either the vertical, downstream or cross-stream direction. The results are compared with previous measurements in the laboratory and in the stable atmospheric boundary layer, as well as with results from direct numerical simulations. The computed and measured flow properties agree with each other generally within the scatter of the measurements. At strong stratification, the Froude number becomes the relevant flow-controlling parameter. Stable stratification suppresses vertical overturning and mixing when the inverse Froude number based on a turn-over timescale exceeds a critical value of about 3. The turbulent diffusivity tensor is strongly anisotropic and asymmetric. However, only the vertical and the cross-stream diagonal components are of practical importance in shear flows. The vertical diffusion coefficient is much smaller than the cross-stream one at strong stratification. This anisotropy is stronger than predicted by second-order closure models. Turbulence fluxes in downstream and cross-stream directions follow classical mixing-length models.
Article
Nous avons, dans cette thèse, étudié, à l'échelle locale, le mélange induit par gravité de deux fluides miscibles de densités différentes dans un long tube incliné. Nous avons mesuré, sur des expériences indépendantes, les champs de vitesse (PIV) et les champs de concentration (LIF) dans un plan diamétral vertical. Avec l'aide des profils de vitesse et de concentration moyenne nous avons pu établir une carte des régimes d'écoulement observés en fonction des paramètres de contrôles : angle d'inclinaison du tube $\theta$ (repéré par rapport à la verticale) variant de 15°à 60°, et contraste de densité caractérisé par le nombre d'Atwood At allant de $10^{-3}$ à $10^{-2}$. Nous avons étudié le transport de quantité de mouvement dans les régimes d'écoulement quasi-stationnaires en comparant les résultats expérimentaux à des modèles simples et à des simulations numériques. Nous nous sommes également intéressés et la stationnarité et la stabilité des écoulements.