Sunteți pe pagina 1din 11

International Journal of Heat and Fluid Flow 41 (2013) 1626

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Evolution of turbulence characteristics from straight to curved pipes


A. Noorani , G.K. El Khoury, P. Schlatter
Linn FLOW Centre and Swedish e-Science Research Centre (SeRC), KTH Mechanics, Royal Institute of Technology, SE-100 44 Stockholm, Sweden

a r t i c l e

i n f o

Article history:
Received 12 October 2012
Received in revised form 18 February 2013
Accepted 13 March 2013
Available online 11 April 2013
Keywords:
Wall turbulence
Pipe ow
Curvature effects
Reynolds-stress budgets
Coiled tube

a b s t r a c t
Fully developed, statistically steady turbulent ow in straight and curved pipes at moderate Reynolds
numbers is studied in detail using direct numerical simulations (DNS) based on a spectral element discretisation. After the validation of data and setup against existing DNS results, a comparative study of turbulent characteristics at different bulk Reynolds numbers Reb = 5300 and 11,700, and various curvature
parameters j = 0, 0.01, 0.1 is presented. In particular, complete Reynolds-stress budgets are reported
for the rst time. Instantaneous visualisations reveal partial relaminarisation along the inner surface of
the curved pipe at the highest curvature, whereas developed turbulence is always maintained at the
outer side. The mean ow shows asymmetry in the axial velocity prole and distinct Dean vortices as secondary motions. For strong curvature a distinct bulge appears close to the pipe centre, which has previously been observed in laminar and transitional curved pipes at lower Reb only. On the other hand, mild
curvature allows the interesting observation of a friction factor which is lower than in a straight pipe for
the same ow rate.
All statistical data, including mean prole, uctuations and the Reynolds-stress budgets, is available for
development and validation of turbulence models in curved geometries.
2013 Elsevier Inc. All rights reserved.

1. Introduction
Turbulent ow in curved pipes is frequently occurring in a variety of industrial applications. Typical prominent examples are heat
and mass transfer systems where straight, bent and helically coiled
pipes are encountered in heat exchangers, chemical reactors, pipeline systems as well as components of internal combustion engines
(e.g. exhaust manifolds). Similarly, biological systems such as the
blood ow in arteries or the air ow in the respiratory system,
are also occurring mostly in bent geometries. A better understanding of the physical mechanisms in action and improved ability to
accurately model the specic uid phenomena would help to improve the performance of such devices; e.g. improved heat and
mass transfer coefcients and enhanced cross-sectional mixing,
and reduced axial dispersion, as for instance discussed by Vashisth
et al. (2008).
Compared with other (canonical) internal ows such as ow in
straight pipes and ducts, the curved pipe conguration has been
studied in detail mainly for laminar ow. In general, the curvature
causes the appearance of centrifugal forces which deects the axial
maximum velocity away from the centre towards the outer side of
the curved pipe. At the same time, the curvature gives rise to a secondary motion in the cross-section of the bent pipe, essentially due
Corresponding author.
E-mail address: azad@mech.kth.se (A. Noorani).
0142-727X/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatuidow.2013.03.005

to the imbalance between the cross-stream pressure gradient and


centrifugal force. This inviscid process is usually referred to as Prandtls secondary ow of rst kind, which by itself leads to the formation of a pair of counter-rotating, axially-oriented vortices, socalled Dean vortices. These vortices are skew-induced, and as such
appear both in laminar and turbulent ow. Alternatively, mean
streamwise vortices can also be formed as a result of local variation
of the Reynolds stresses. These are then referred to as Prandtls secondary ow of second kind (or stress-induced vortices) and are observed, for instance, in the turbulent ow in straight ducts with
non-circular cross-section (see Bradshaw, 1987).
Curved pipe geometries can be classied into two major types:
(i) spatially developing bends and (ii) coiled tubes/conduits. In spatially developing bends such as U-bends or elbows (90 bends), the
entry ow passes through a straight inlet section before it reaches
the bend, whereas in coiled tubes, the uid ow is totally conned
within the curved geometry. An extensive literature review of early
research activities on ow in curved pipes has been carried out by
Berger et al. (1983) and later also by Ito (1987). In addition, a more
recent survey by Naphon and Wongwises (2006) also includes aspects of heat transfer.
It is a common approach among experimentalists to use helically coiled pipes in order to study the effect of curvature in bent
pipes. In this case, the pitch of the resulting coil induces an additional torsion force acting alongside the centrifugal force on the
uid ow. However, in most practical applications, the pitch angle

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

is small compared to the coil diameter and thus the inuence of


torsion becomes negligible with respect to curvature; see for instance the discussions on that topic in Manlapaz and Churchill
(1980), Germano (1982), and Yamamoto et al. (1995). The idealised
conguration is then reduced to that of a torus; i.e. an innitely
long bent pipe. This ow case provides a unique opportunity to isolate the effect of curvature on pipe turbulence, and also to study
the inuence of centrifugal forces and secondary motion on nearwall features.
In order to perform numerical simulations of uid ow in
curved pipes, on the other hand, the NavierStokes equation have
to be expressed in curvilinear or body-tted coordinates. Germano
(1982) proposed local orthogonal helical coordinate system to
study the laminar ow in a helical pipe. Despite these efforts, very
few data bases regarding turbulence statistics and near-wall mechanisms in bent pipes are available from simulations. Boersma and
Nieuwstadt (1996), Httl and Friedrich (2000), and Httl and
Friedrich (2001) applied Germanos coordinates system in order
to perform large-eddy simulation (LES) and direct numerical simulation (DNS) of the turbulent ow in curved pipes, respectively. In
both references innitely bent pipes using periodic boundary conditions were considered. Boersma and Nieuwstadt (1996) investigated the inuence of curvature on the mean ow and rms (rootmean-square) uctuations, and also tested the inuence of different initial conditions. Httl and Friedrich (2000) employed DNS
and studied the inuence of curvature and torsion on turbulence
at friction Reynolds number Res = 230; based on azimuthally averaged friction velocity and pipe radius. In their follow-up study,
Httl and Friedrich (2001) observed that the turbulent uctuations
in curved pipe are drastically reduced compared to ow in straight
conguration and also provided a useful database, albeit at low
Reynolds number, for ow modelling for a variety of congurations, including data for few selected terms of the Reynolds
stresses.
From an engineering point of view, one of the most crucial aspects of internal ow in pipes has always been the accurate determination and prediction of the relation between wall friction and
ow rate. A customary way to parameterise this relation in pipe
ow is given by the so-called Fanning friction factor (f) which
non-dimensionalises the wall shear stress sw, expressed as friction
velocity us = (sw/q)(1/2), using the bulk velocity ub and the uid density q. Alternatively, the hydraulic head-loss (i.e. pressure drop) can
be used as basis, yielding the DarcyWeisbach friction factor (k). It
can be simply deduced that the two factors are related by k = 4f.
Determining concrete values for either of the two coefcients in
various ow congurations has been the subjects of many studies:
Regarding bent pipes, in the early part of the twentieth century
Grindley and Gibson (1908) examined the viscosity of air in helically coiled pipes to measure the pressure drop. Since then there
have several reports been published on renements of the friction-factor diagram. For instance, Ito (1959) presented a series of
experiments to nd the correlation between k and the ow rate
for many curvature congurations. He also provided an empirical
equation indicating the critical Reynolds number for transition to
turbulence at each curvature conguration. It is generally accepted
that the ow in curved pipes has a higher pressure drop than the
straight ones at similar ow rates, which is believed to be related
to the existence of the mentioned secondary motion. Cioncolini
and Santini (2006) conducted a set of experiments in a wider range
of the curvature parameter to obtain the Fanning friction coefcient, and they conrmed, in general terms, the validity of Itos data.
In their work, Cioncolini and Santini (2006) observed an additional
interesting behaviour in the changeover from the laminar and turbulent ow for mildly curved pipes: When the ow rate is increased
(parametrised by the Reynolds number), the friction factor does not
monotonously increase away from the laminar value. Rather, it rst

17

reaches a minimum which is below the laminar Ito correlation. After


an inection point, the friction factor increases, and nally settles
on friction values pertaining to developed turbulent ow in bent
pipes. In addition, Cioncolini and Santini (2006) identied a small
range of Reynolds numbers in which the ow at mildly curved pipes
showed a lower hydraulic resistance compared to the straight pipe
ow with the same ow rate. The authors suggested that the
smoothing effect of curvature on turbulence may have balanced
the effect of the secondary motion, however no real explanation
or conrmation has been given yet.
As already discussed by e.g. Sreenivasan and Strykowski (1983),
curved pipes tend to have a higher critical Reynolds number for
transition to turbulence than straight pipes. A recent numerical
study of ow in the transitional regime in highly curved tori (with
radii of the pipe being 0.1 and 0.3 times the major radius of torus)
by Di Piazza and Ciofalo (2011) revealed an intermediate state between laminar and fully turbulent ow, denoted as the quasi-periodic state: When increasing the Reynolds number, the ow would
rst reach a chaotic state along the pipe centreline (see also Httl
and Friedrich, 2000), however without near-wall features characteristic of pipe turbulence. With a further increase of Re, these features appear and co-exist with the strong chaotic motion in the
pipe centre. Based on Itos correlation for the critical Reynolds
number in curved pipes, Di Piazza and Ciofalo (2011) further presented a tentative ow regime map in the form of Reynolds number against curvature parameter.
Even though the direct numerical simulation is the most accurate way of computing a solution of a given ow, the range of relevant (length and time) scales increases dramatically with the
Reynolds number. This makes DNS impractical for most engineering applications which occur at substantially high Reynolds numbers. Therefore, well validated turbulence models, be it for LES or
the Reynolds-averaged NavierStokes (RANS) equations, are necessary. In curved pipes, the presence of centrifugal forces due to the
inevitable streamline curvature affects not only the bulk ow
around the centre of the pipe but also the near-wall regions. The
resulting strong anisotropy makes the ow to be away from quasi-equilibrium suitable for traditional turbulence modelling (e.g.
via eddy-viscosity models) and has been a major issue for the Reynolds stress transport modelling community (see Wallin and
Johansson, 2002). Recently, Di Piazza and Ciofalo (2010) conducted
a number of marginally resolved DNS of turbulent ow in helically
coiled pipes with heat transfer. By comparing their DNS results, the
authors found that the SST kx and RSM-x performed reasonably
well in computing the friction and heat-transfer coefcient of the
ow even at high curvature. However, a more detailed analysis,
including comparison of the individual modelled terms in the budgets, seems appropriate for various curvatures and Re.
The present DNS study is aimed at investigating the evolution of
selected characteristics of the turbulent ow in straight to bent
pipes over a limited range of Reynolds numbers and curvatures.
The main purpose is to provide a validation of our simulation setup
and the specic way our statistics are extracted, together with a
documentation of the various simulation parameters and the
respective results. In a rst step we validate our data against existing DNS results by Httl and Friedrich (2001). Our newly obtained
results at higher Reynolds number and higher curvature will be
introduced later. The obtained statistical data (mainly Reynolds
stresses and turbulent kinetic energy budgets) will serve to investigate the observed near-wall and pipe-core turbulent features. The
performance of traditional eddy-viscosity models in capturing turbulent ow in curved pipes will be put to test applying the present
DNS data. Finally, the friction factor as an integral quantity will be
computed and compared to existing experimental results. Further
studies aiming at a deeper physical understanding of the various
ndings are certainly necessary.

18

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

2. Computational methodology
2.1. Flow conguration and governing equations
Fig. 1 (top) shows a schematic view of a curved pipe together
with its Cartesian (X, Y, Z) and toroidal (R, s, f) coordinate systems.
A relevant quantity in this conguration is the curvature parameter
j dened as Ra/Rc; Ra is the radius of the pipe cross-section and Rc
is the radius of curvature at the pipe centreline. This dimensionless
parameter distinguishes between mildly (j  0.01) and strongly
curved pipes (j  0.1).
The governing incompressible NavierStokes equations,

r  u 0;

@u
1 2
u  ru rp
r u
@t
Reb

are integrated in time for the velocity vector u and the pressure
pReb is the bulk Reynolds number based on the bulk velocity ub
and the pipe diameter D = 2Ra, i.e. Reb = 2Raub/m with the kinematic
viscosity m. The dimensionless Dean number is given by
p
Deb Reb j and is widely used in the literature to characterise
the secondary motion, generally in the laminar regime. Throughout
the present study, the horizontal and vertical cuts are referred to as
azimuthal in-plane positions of h = 0 and p/2, respectively; see the
sketch in Fig. 1 (bottom).
2.2. Numerical approach and parameters
The simulations are performed using the massively parallel
spectral-element method (SEM) solver nek5000. This code has
been developed by Fischer et al. (2008), and it solves the incom-

Fig. 1. (Top) embedded in-plane polar (r, h) and Cartesian (x, y) coordinates along
with side view of ow conguration and associated coordinate systems, (X, Y, Z) as
reference orthogonal coordinate triad and (R, s, f) as attached toroidal cordinate
system. (Bottom, left) Cut-away section of a curved pipe, with locations of the
horizontal and equatorial cross-sections indicated, h= 0 and h= p/2, respectively.
(Bottom, right) Front view with torodial coordinate system.

pressible NavierStokes equations on GaussLobattoLegendre


(GLL) nodes. It essentially divides the physical domain into a number of hexahedral local elements where the equations of motion
are solved by means of local approximations based on high-order
orthogonal polynomial basis. Along with its efcient parallelisation, this code provides spectral accuracy with geometrical exibility applicable to engineering problems. The velocity and pressure
spaces are represented by the same polynomial order in the formally known PN  PN method that is used along with over-integration and ltering stabilisation schemes. Time is advanced with a
3rd order mixed Backward Difference/Extrapolation (BDF3/EXT3)
scheme. Similar to the straight pipe conguration, the streamwise
direction is assumed to be homogeneous and thus periodic boundary conditions are applied in that direction of the curved geometry.
In practice, the streamwise periodicity for curved pipes is applied
in the local toroidal coordinate system. In the present case, however, the solution obtained from nek5000 for the NavierStokes
equations is in a Cartesian coordinate system and thus proper rotation of the velocity components is necessary in order to couple the
two ends of the domain; i.e. the velocity vector is the same at the
beginning and end of the computational domain. Note that similar
periodicity has also been used by e.g. Httl and Friedrich (2000)
and Httl and Friedrich (2001), but since their formulation was
in the toroidal system no additional transformations were
necessary.
To provide the necessary driving force to the ow, instead of xing the pressure gradient, the mass ow rate is held exactly constant by the time-integration scheme. The basic idea is to realise
that the mean-ow equation is linear in the pressure gradient.
Thus the instantaneous pressure gradient can be adapted in each
time step such that unit bulk velocity ub is obtained. All simulations were started from a laminar Poiseuille prole with superimposed low-amplitude pseudo-random noise acting as threedimensional perturbations. The ow is then evolved in time until
a statistically stationary turbulent state is established; this happened for each case after approximately 200 time units.
The cross-sectional grid is constructed by a decomposition for
cylindrical geometries in a Cartesian solver, as illustrated in
Fig. 2. This stencil is uniformly extruded in the axial direction to
generate a straight pipe mesh. Further grid renement for each element up to polynomial order 7 is used to achieve the desired
numerical resolution. The nal toroidal grid is obtained by means
of analytical morphing of the straight pipe mesh. A part of crosssectional and equatorial planes of the resulting grid is sketched
in Fig. 2.
Since the focus of the present study is on analysing the inuence of curvature on turbulence structures and statistics, various

Fig. 2. (Left) Mesh conguration in the equatorial mid-plane of the toroidal pipe for
the higher Reynolds number cases. Element boundaries and their GaussLobatto
Legendre (GLL) points are also shown in the darker region. (Right) A view of a
quarter of the cross-sectional plane pertaining to the mesh for the same case.

19

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

15000

Turbulent

Chaotic

Reb

10000

QuasiPeriodic

5000
Laminar

0.05

0.1

0.15

0.2

Table 2
Various simulation parameters for the present study. Case names are coded based on
the bulk Reynolds number and domain congurations (S, as straight, M, as mildly
curved, and H, for strongly curved pipes). The time t is normalised by Ra/ub, starting
from the point that the ow has reached a statistically stationary state.
Case

Reb,D

Res

Sampling period

R5300S
R5300M
R6900S
R6900M
R11700S
R11700M
R11700H

5300
5300
6926
6926
11,700
11,700
11,700

0.00
0.01
0.00
0.01
0.00
0.01
0.10

180
176
228
230
360
368
400

400
2800
400
2800
400
1000
1000

0.25

Fig. 3. Tentative ow regime map following Di Piazza and Ciofalo (2011), - - - transition from stationary to quasi-periodic regime, -- transition to chaotic state,
transition to fully developed turbulent ow. j Parameters for present DNS. h DNS
performed yielding quasi-periodic state but not further discussed here.

pipe congurations with xed Reb are considered such that the
ow rate is above the critical value given by Ito (1959) for laminar-turbulent transition. Fig. 3 shows the aforementioned tentative
Rebj map by Di Piazza and Ciofalo (2011). For our DNS, we chose
three Reynolds numbers, Reb = 5300, 6900 and 11,700; for each Reb
the curvature parameters are chosen such that the fully turbulent
regime is obtained, as indicated in Fig. 3. We have also performed a
simulation with Reb = 5300 and high curvature j = 0.1, which
reached, as expected, the quasi-periodic state. We believe that this
state is quite regular rather than genuinely turbulent and therefore
we do not further study this parameter range.
At a given Reynolds number, the grid resolution of the curved
pipe congurations is chosen to be the same as that for the straight
pipe. Apart from the validation case (Reb = 6900, see Section 3.2)
the length of the domain for all cases is set to be 25Ra along the
pipe centreline. The grid spacing, measured in wall units where
the viscous length scale is based on the straight pipe friction Reynolds number at each Reb, is set such that Dr
max 6 5 with four grid
points placed below Dr+ = 1, measured from the wall, and
DRa hmax 6 5 and Dsmax 6 10. Viscous scaling is indicated by the
superscript +. The details of the computational meshes at the various Reynolds numbers are presented in Table 1.
The ow is homogeneous in the axial direction (s) and statistically steady in time (t). Therefore, all statistical data are averaged
in these two directions, and the nal average is denoted by hi. In
the straight pipe, azimuthal averaging is also used. Sampling of statistics is started after the ow has settled in the developed turbulent regime approximately after 200Ra/ub in each case. Different
statistical averaging time periods are used for the various cases
depending on the observed convergence of the data. This and other
relevant simulation parameters are listed in Table 2. A brief outline
of the procedure on how to obtain statistical averages via tensor
transformations under rotation from Cartesian coordinates to
toroidal coordinates is given in Appendix A.
In the
p
present study the mean friction velocity us is dened as
sw;s =q where sw,s indicates the streamwise component of the
mean wall shear stress and an overbar denotes averaging along
the circumference of the pipe section. The azimuthal component

of the wall shear stress sw,h is not used as a normalisation parameter since its magnitude is small compared to the streamwise component. Based on that, the friction Reynolds number is dened as
Res = usRa/m. For the purpose of this paper, the friction factor refers
to the Fanning coefcient and is computed based on the following
denition:

2
sw;s 
Res
:
8
2
Reb
qub
2

f 1

3. Results and analysis


3.1. Instantaneous velocity elds
To give an overview of the ow at hand and the expected features, the instantaneous streamwise velocity is displayed in Fig. 4
for Reb = 11,700, corresponding to a nominal Res = 360 in the
straight case. Owing to the centrifugal force, which is absent in
the straight conguration, the bulk of the ow is deected from
the centre towards the outer wall for j = 0.01 and 0.1. The ow
subsequently loses its azimuthal homogeneity and it can be clearly
seen that turbulence is substantially damped along the inner side
of the pipe bend. The appearance of the Dean vortices and the
inuence of the curvature j on the asymmetry of the mean ow
and turbulence budgets among other quantities will be discussed
further down.
3.2. Validation
In a rst step the present results and simulation setup are validated against the existing DNS data by Httl and Friedrich
(2001). For that purpose the same geometrical conguration is
considered; i.e. a weakly curved pipe with j = 0.01 and domain
length 15.23Ra at Reb=6926. The simulation is performed on a computational mesh with 65 elements in the axial direction and 336
elements in each cross-sectional plane with the polynomial order
set to 7. The total number of grid points is 11.2 million, and the grid
spacing, based on the mean friction Reynolds number of Res=230,

is Dr
min ; Dr max 0:26; 6:74, DRa hmin ; DRa hmax 1:45; 7:31

and Ds
;
D
s

3:45;
11:27.
This
resolution
is very similar
max
min
to what will be used for the new cases, as discussed in Section 2.2.
The mean axial velocity and turbulence kinetic energy proles
(plotted in Fig. 5) show very good agreement with the literature
data of Httl and Friedrich (2001). As we xed the bulk velocity

Table 1
Resolution details for the present pipe ow meshes used for the two Reynolds numbers under consideration.
Reb

# Of elements

# Grid points

Dr +

DRah+

Ds+

5300
11,700

36,480
237,120

18.67  106
121.4  106

(0.14, 4.44)
(0.16, 4.70)

(1.51, 4.93)
(1.49, 4.93)

(3.03, 9.91)
(3.03, 9.91)

20

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

+
()
w,s

1.5

0.5

/2

3/2

Fig. 6. Circumferential distribution of the local axial wall shear stress normalised
by its averaged value.      Straight pipe, - - - - R11700M, -- R5300M,
R11700H. Note that the equatorial symmetry is not being used.

Fig. 4. Pseudocolours of streamwise velocity at (left) generic axial cross-section,


(right) generic azimuthal cross-section of straight pipe and equatorial mid-plane of
mildly and highly curved pipes (Reb = 11,700, j = 0, 0.01, 0.1). Colours range from 0
(blue) to 1.25 (red). (For interpretation of the references to colour in this gure
legend, the reader is referred to the web version of this article.)

0.2<us> ,k

+ +

4
3
2
1
0
1

0.5

0.5

r/R

Fig. 5. Mean axial velocity component husi+ normalised by us and turbulent kinetic
energy k+ at the equatorial mid-plane of the toroidal pipe, with r/Ra = 1 denoting
the inner side and r/Ra = +1 indicating the outer side.  DNS data by Httl and
Friedrich (2001), - scaled mean streamwise velocity, - - - - turbulent kinetic
energy case R6900M.

for the present simulation, the same frictional Reynolds number as


given by Httl and Friedrich (2001) is recovered. In addition, comparisons with other statistical quantities, such as local mean friction velocity and Reynolds shear stress component hu0s u0r i were in
good correspondence. We therefore can conclude that our methodology to perform DNS of turbulent ow in an innitely bent pipe is
correctly implemented, and thus we can proceed to perform new
simulations in longer domains at higher Reb and j.

part (h = 3p/2) with s


w;s being substantially higher along the outer
surface. The effect of increasing Reb for xed curvature j is that
the variation of sw along the circumferential direction is slightly reduced and thus a more uniform distribution of the wall stress can
be expected at even higher Reynolds numbers. On the other hand,
by xing Reb and increasing j it is clear that the variation of sw with
h is enhanced. Interestingly, the case with highest Reb and j
(R11700H) shows a plateau at the inner part of the curved pipe indicating possible relaminarisation. Furthermore, small but distinct
oscillations of sw near the outer side are clearly visible, which are
symmetric with respect to the vertical mid-plane. The same oscillations were visible also in the validation case with short pipe (not
shown). The physical mechanism related to this feature is still not
clear, but has not been further investigated in the present context.
The effect of passing from mild to strong curvature on the mean
axial velocity at Reb = 11,700 is shown in Figs. 7 and 8, (right section). The peak in the streamwise velocity deects more towards
the outer wall for strong curvature (j = 0.1) and is located at
0.7Ra in comparison to 0.5Ra for mild curvature (j = 0.01). At the
same time the secondary motion of the ow, i.e. the mean in-plane
velocities (huri, huhi) intensify from about 5% of ub at j = 0.01 to almost 15% at j = 0.1. This intensication in the cross-stream motion
is expected as the imbalance of the centrifugal force with the axial
pressure gradient gets more pronounced at higher j. The same effect can be seen in the left panel of Figs. 7 and 8 in which vector
plots of the mean secondary ow are displayed. Here, each case
develops a distinct cross-ow boundary layer at the side walls.
For mild curvature, the layer is attached along the whole pipe circumference all the way down to the stagnation point. From this
point, the secondary motion attains a vertical velocity component

3.3. Mean ow statistics and turbulent uctuations


The presence of a centrifugal force due to the streamline curvature affects not only the bulk ow around the pipe centre but also
the near-wall regions. This can be clearly seen in Fig. 6 where the
circumferential distribution of the local axial shear stress at the
wall is plotted. Here, s
w;s has been normalised by its averaged value
along the pipe circumference. For all the cases presented in Table 2
with j 0, the axial wall shear stress attains a maximum at the
outer side of the curved pipe (h = p/2), and a minimum in the inner

Fig. 7. (Left) vector plot of mean secondary ow at R11700M. (Right) (left section)
contour plot of mean in-plane velocity at R11700M, (right section) Contour plot of
mean axial velocity at R11700M.

21

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

25
20
15
10
5
25
Fig. 8. (Left) Vector plot of mean secondary ow at R11700H. (Right) (left section)
Contour plot of mean in-plane velocity at R11700H, (right section) Contour plot of
mean axial velocity at R11700H.

15

<u >+

20

that accelerates in the radial direction towards the pipes centre. In


the case of higher j, on the other hand, the cross-stream boundary
layer at the side walls separates at about 20 away from the symmetry plane at the inner bend. This generates a large radial velocity
decit just below the pipe centre. This region is the major qualitative difference of the in-plane motion between low and high curvature, and is readily visible in the stream function (W) map of
secondary motion in Fig. 9. Contour lines of spectrally computed
W display a pair of stable counter-rotating Dean vortices in the
mean of the in-plane motion. The core of the vortices is indicated
with the symbol 0 0 . Even though only one pair of Dean vortices exists in both cases, the difference between mild and strongly curved
pipes is quite obvious from this gure; namely the appearance of a
bulge near the pipe core in the case of higher curvature. The
source and consequences in the secondary ow are non-trivial,
and will be discussed later in the context of turbulent budgets.
In order to obtain a more detailed picture of the effect of curvature on the turbulent ow eld, we focus our attention in the rest
of this section on the turbulence statistics in the vertical cut presented in Fig. 1. It is worthwhile to note that the statistics are symmetric with respect to the equatorial mid-plane, and thus the
vertical plane gives a better sense on the asymmetry caused by
the streamline curvature. Fig. 10 shows the mean axial velocity
component of the various cases in this cut. The proles are symmetric around the vertical plane for zero curvature (j = 0) and they
become increasingly asymmetric as j is increases from 0.01 to 0.1.
For the mild curvature case it is observed that the asymmetry increases with Reb and within a certain region husi shows an approximate linear behaviour. It is interesting to recall that in rotating
channel ows a similar constant-slope region is observed in the
channel centre (see for instance Kristoffersen and Andersson,
1993). In the case of high curvature R11700H, it can be seen that
from the centre down to r/Ra = 0.5, towards the inner bend, the
streamwise velocity gradient surges and separates a low-speed

10
5
0
1

0.5

0.5

r/R

Fig. 10. Mean streamwise velocity component normalised by us along the


equatorial mid-plane of the toroidal pipe (vertical mid-plane), with r/Ra = 1
denoting the inner side and r/R = +1 the outer side.  R5300S,      R5300M, - - - R11700S, -- R11700M, - R11700H. (top) Effect of Reynolds number, (bottom)
effect of curvature.

core region found near the inner side of the pipe from a high-speed
region close to the outer bend. This sudden change is in qualitative
agreement with the data by Boersma and Nieuwstadt (1996), Httl
and Friedrich (2000), and Httl and Friedrich (2001) at similar curvature. The experimental results by Webster and Humphrey
(1993) also show a similar quantitative behaviour in this region.
However, it should be noted that the mentioned literature data
are all at lower Reynolds numbers, or even in the transitional regime. Furthermore, similar structures are commonly observed in
the pipe centre in laminar curved pipe ows at higher Dean numbers. The existence of such a feature at higher Re in turbulence is
thus remarkable, and is clearly linked to the bulge in the streamlines discussed previously. Meanwhile, the mean radial velocity
component is displayed in Fig. 11. For low curvature parameters,
the radial velocity increases monotonically from the walls towards
a local maximum that is located in the lower section of the pipe
whereas the main effect of high curvature is the presence of a local

0.7
0.6

<ur>+

0.5
0.4
0.3
0.2
0.1
0
1

0.5

0.5

r/R

Fig. 9. Iso-contours of stream function W at (left) R11700M, (right) R11700H. The


symbol 0 0 indicates the centre of counter-rotating mean Dean vortices.

Fig. 11. Normalised mean radial velocity component at the vertical cut. For legend
see Fig. 10.

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

0.5

1
1

k
2
1
0
1

0.5

0.5

r/R

Fig. 12. Normalised turbulent kinetic energy along the vertical cut of the pipe
section. For legend see Fig. 10.

0.5

Fig. 13. Normalised Reynolds shear stress component hu0s u0r i prole along the
vertical mid-plane. For legend see Fig. 10.

7
6

5
4
3
2
1
0
1

0.5

0.5

r/R

Fig. 14. Normalised rms uctuations of the pressure along the vertical mid-plane.
For legend see Fig. 10.

In addition, the peak in the core region observed in k can be seen in


the prms plot characteristic of the aforementioned vortical motion
in that region.
3.4. Reynolds stresses and turbulent kinetic energy budgets
The transport equation for the Reynolds-stress tensor reads

@ 0 0
hu u i C ij Pij  eij Psij Pdij Dij T ij
@t i j

where the turbulent convection, production, dissipation, pressure


strain, pressure diffusion, molecular diffusion and turbulent diffusion are dened, respectively, as:

C ij huk i

0.5

r/R

5
4

<u u >

0.5

rms

minimum in addition to a local maximum implying strong vortical


motion. The signicant differences observed in the radial velocity
for low and high j indicate fundamental differences in the crossstream features between mildly and strongly curved pipes.
By considering the turbulent kinetic energy k hu0i u0i i=2 in
Fig. 12, the prime indicating the uctuating part of the variable, it
can be seen that each of the straight and mildly curved pipes exhibit
a peak close the inner and outer walls. In the case of R11700H, however, the peak near the inner wall diminishes and only a very low
level of k remains. Subsequently, it can be concluded that turbulence is signicantly inhibited in this region and nearly laminarised
though low-frequency oscillations remain as evidenced by visualisations such as Fig. 4. The ow in that region is very intermittent as
can be quantied by the atness factor of the axial velocity component which reaches a value of 15.55 compared to 6.31 for the mildly
curved pipe and 4.36 for the straight pipe at the same Reb.
The proles of the mean shear component hu0s u0r i in Fig. 13 show
similar peaks near the inner and the outer bend of the various
cases. Interestingly, in the case of R11700H this Reynolds-stress
component is weak or almost zero in a majority of the pipe section
except close to the outer bend and near the pipe core at r/R = 0.4.
This place close to the core is at the same location as where the k
prole exhibits a at peak. This is in correspondence with the position of the bulge-like structure in the pipe core as previously discussed. In fact, in the previous study by Webster and Humphrey
(1997) on helical pipe ow in the transitional stage (j = 0.05,
Reb = 5060) such a location is also observed and described as the
region with high rms. Httl and Friedrich (2000) have also observed a single peak in the k prole in the same position for similar
curvature but lower Reynolds number. However, they could not
establish whether this structure would persist even at higher Reynolds number when clear near-wall turbulence features appear. In
Section 3.4 the energy distribution of the uctuations in this region
will be examined in more detail.
The distribution of pressure uctuations (Fig. 14) conrms the
preceding description in terms of the effect of the curvature on turbulent uctuations. In a comparison between straight and mildly
curved pipes, the pressure uctuations at the outer and inner sides
of the mildly curved pipe are higher and lower than that of a
straight pipe at these two locations, respectively. The discrepancy
observed in prms at the outer side of the pipe is partly due to the
friction velocity employed in normalisation; azimuthal average of
us rather than the local value at h = p/2. At the inner side, on the
other hand, the mean pressure is clearly lower in the mildly curved
pipe (see e.g. Httl and Friedrich, 2000) which contributes to lower
pressure uctuations. At high curvature (j = 0.1), even the pressure uctuations are considerably reduced in the inner side of
the pipe without near-wall peak, implying intermittent turbulence.

22

D
E
@ u0i u0j

@xk
 0 0  @huj i D 0 0 E @hui i
Pij  ui uk
 uj uk
@xk
@xk
0
0

@ui @uj
eij 2m
@xk @xk



@u0j
@u0
1
Psij
p0 i p0
q
@xj
@xi


1 @
@  0 0
Pdij 
hp0 u0j i
p ui
@xj
q @xi
2 D
E
@
Dij m 2 u0i u0j
@xk
@ D 0 0 0E
uuu :
T ij 
@xk i j k

5a
5b
5c
5d
5e
5f
5g

23

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

The budget terms for the various simulations in this study are
calculated according to the scheme presented in Appendix A. It is
based on tensor transformation and thus the (cumbersome)
expression of these terms in orthogonal helical coordinate system
is avoided. To our knowledge, the complete Reynolds-stress budgets have not been presented before for periodic bent pipes. These,
however, are of great importance to address three-dimensional effects in turbulence models, e.g. how empirical constants in turbulence models might be inuenced by mean streamwise vorticity
(Bradshaw, 1987). Note that the budget terms are inherently
two-dimensional, and all Reynolds-stress components are nonzero. Our results are based on long averaging times, such that
the residuals in the various components are less than 104 in viscous scaling.

8
6
4
2
0
2
4
2

8
6
4
2
0
2
4
2

1
0

2
2

2
2

2
4

2
4

0
1

2
4

2
4

0.5

0.5

r/Ra

rr

2
2

2
2

ss

0
1

8
6
4
2
0
2
4
2

Bk

8
6
4
2
0
2
4
2

rr

ss

3.4.1. Outer layer


The Reynolds-stress budget in the outer layer is pre-multiplied
by the distance from the wall to emphasise the outer-layer behaviour of various terms. As customary the basic scaling parameters
are chosen as us3/Ra and r/Ra. Figs. 15 and 16 show budgets of
the normal Reynolds stress and the turbulent kinetic energy at
the vertical mid-plane for both mildly and strongly curved pipes.
For R11700M, the generated energy by streamwise uctuations is
re-distributed through turbulent diffusion and pressure work to
the other components whereas the remaining streamwise uctuations are dissipated on the molecular level. In the region away from
the buffer layer (0.4 < r/Ra < 0.7) production goes to zero due to the
presence of a maximum in the streamwise velocity prole, i.e. zero
velocitystrain rate. In the same region, turbulent diffusion balances dissipation until it is levelled off and does not perform any
signicant contributions. The whole activity in the outer bend is
similar to the one in a straight pipe, except in the curved conguration it seems the process is deected towards the outer side of
the toroidal pipe due to the centrifugal force.

Approaching

 the pipe centre from the inner bend, the production of u0s u0s is compensated by pressurestrain and dissipation.
The process for the other normal components is similar except that
the balance is solely among pressurestrain and dissipation.
Although the level of turbulent uctuations are stronger at the outer bend, it is interesting to note that the mentioned terms are
decreasing inversely proportional to the distance from the wall
mimicking an ideal logarithmic layer at high Reynolds number
(see Hoyas and Jimnez, 2008).
The contribution of the various budget terms at the outer side of
the strongly curved pipe (Fig. 16 right section) is very similar to that
at mild curvature, Fig. 15, except that now the cross-ow budget
terms are about a factor of two lower, while the terms in the axial
budget are similar in magnitude. The pressurestrain in the axial
direction withdraws energy from streamwise production to supply
Psrr ; Pshh . These terms are the main source of energy for the lateral
and azimuthal components since their own production is insignificantly small. This process occurs in straight pipes and likewise in
mildly curved ones. However, for the strongly curved pipe, the production in the Brr budget has the same magnitude as the pressure
strain. This may indicate that the streamwise production is predominantly converted into azimuthal uctuations.
On the other hand, the budget in the inner side of the highly
curved pipe signicantly differs from that in the straight or mild
curvature congurations. Streamwise velocity uctuations energise the ow between r/Ra = 0.3 and r/Ra = 0.5. Close to the wall,
the ow is intermittent and even partially laminarised, the peak of
production is occurring at about r/Ra = 0.35. This is essentially the
position where the mean radial and streamwise velocity exhibits


high gradients and u0s u0r declines. The location of this process
coincides with the bulge in the middle of the secondary motion
map. Being located far away from the wall, the process is more
homogeneous similar to turbulence in free shear ows, rather than
comparable to near-wall turbulence.

2
1

2
0.5

0.5

r/R

Fig. 15. Pre-multiplied budgets for the normal Reynolds-stress components and the
turbulent kinetic energy k at the equatorial mid-plane of the case R11700M, with r/
Ra = 1 indicating the inner side and r/R = +1 the outer side. + Cij,      Pij, - - - eij ; Psij ; Pdij , -- Dij, Tij.

Fig. 16. Pre-multiplied budgets for the normal Reynolds-stress components and the
turbulent kinetic energy k at the equatorial mid-plane of the case R11700H, with r/
Ra = 1 indicating the inner side and r/R = +1 the outer side. For legend see Fig. 15.

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

gain

24

0.4

0.4

0.2

0.2

loss

0.2

0.4

0.2

10

20

30

40

50 50

40

30

20

10

0.4

(1r/Ra)

(1r/Ra)

Fig. 17. Turbulent kinetic energy budget (major components) along vertical cut. Black represents R11700S, blue as R11700M, and red indicates R11700H case. (Left) is the
inner side of the curved pipe, and (right) is the outer side. For line caption see Fig. 15. (For interpretation of the references to colour in this gure legend, the reader is referred
to the web version of this article.)

3.4.2. Inner layer


We now turn our attention to the budgets close to the outer
wall as shown in the Fig. 17. In the turbulent kinetic energy clear
similarities between the different cases are evident. Close to the
wall almost all terms are active in balancing the budget in the
usual way, whereas with increasing distance almost exclusively a
balance between production and dissipation is established. As in
canonical wall turbulence, in the viscous sub-layer dissipation is
mainly balanced by both Dk and Tk while in the buffer layer these
two change sign in order to contribute balancing production which
is the dominant term there. It can be deduced that with increasing
j the contribution of molecular diffusion is more dominant in the
buffer layer due to the more active secondary motion and larger
wall gradients.
At the inner bend, the turbulent kinetic energy budget for the
mildly curved pipe shares similarities with the outer bend location
although the magnitude is smaller. This is in agreement with a
generally reduced turbulence intensity at that location. The various
peaks move away from the wall for increasing curvature indicating
lower friction. However, this is not the case for the highly curved
conguration as the near-wall turbulence practically ceased.

shows the computed Cl for turbulence in straight and curved pipes


with different Reynolds numbers. The value of Cl for the straight
pipe is quite constant along the radius and with increasing Reb
reaches closer to the standard value of Cl = 0.09. The value monotonically decreases towards zero at the walls where the value for
Cl is known to be lower due to strong anisotropy. Conversely, this
scalar quantity in the curved pipe uctuates at the core. In mildly
curved pipes with increasing Reynolds number Cl is distributed
more homogeneously. This and the previous observation of standard value of Cl in straight pipe suggest that a traditional eddy-viscosity model such as the ke model might be capable of performing
turbulence in straight and mildly curved pipe. However, the case
with high curvature exhibits strong deviations from the standard
constant and thus implies that eddy-viscosity models may not be

3.5. Eddy viscosity


The modelling ofDeddy
E viscosity instead of the complete Reynolds-stress tensor u0i u0j , in fact, results in a large reduction in
the complexity of closing the RANS equation. This process is widely
used in modelling of the generic turbulent and particularly industrial ows. Based on the Boussinesq approximation, the standard
eddy-viscosity formulation for incompressible turbulence reads

D
E
2
 u0i u0j  kdij 2mT hsij i
3
mT C l k2 =e

6a
6b

with dij being the Kroneker delta, hsiji the mean strain rate tensor
dened as 1/2(@huii/@xj + @huji/@xi), and mT the turbulent eddy viscosity.D The turbulent
dissipation rate is dened as
E
e 2m s0ij du0i =dxj , and Cl denotes the turbulent viscosity constant
in the standard ke model. Equating production and dissipation
yields an expression which can be solved for the eddy viscosity,

Pk 2mT hsij ihsij i:

The present DNS data can now be used to evaluate these terms to
obtain Cl. This scalar value can be essentially considered as nondimensionalised mt and serves as an indicator for examining the
validity of a traditional eddy-viscosity model such as ke. Fig. 18

Fig. 18. Computed Cl from the DNS data for turbulent pipe ow at straight and
curved congurations. (Top) from left to right R11700H, R11700S, R5300S. (Bottom)
from left to right R11700M, R6900M, R5300M. Cl in the straight pipe is circumferentially averaged.

25

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

2.05

log10(f)

able to predict the correct behaviour of the turbulence. This can be


expected since no explicit representation of turbulence anisotropy
is present in such a simplied model. In that case more sophisticated models such as the explicit algebraic Reynolds stress model
(EARSM) are expected to be a better option as they potentially are
able to incorporate even strong curvature effects (see Wallin and
Johansson, 2002). A systematic study on modelling of turbulence
in bent pipes is certainly an interesting future work.

2.1

2.15
3.6. Sub-straight drag
As nal aspect in the present paper, we study the effect of
increasing Reynolds number at a xed curvature, j = 0.01 on the
Fanning friction factor f. Considering the friction factor for straight
(cases R5300S, R6900S and R11700S) and curved pipes with mild
curvature (cases R5300M, R6900M and R11700M) shows the surprising behaviour that f in the bent pipe is not always higher than
in the straight pipe despite the existence of secondary motion. As
illustrated in Fig. 19 (and its blow-up in Fig. 20), with increasing
Reynolds number, all three cases both in straight or curved conguration show a decrease in the friction factor. In cases R6900M and
R11700M the friction coefcient is larger in the curved pipe compared to their straight peer. However, for cases R5300M and
R5300S, the reversed situation can be seen. This is the same observation made by Cioncolini and Santini (2006) based on their pressure-drop experiments. Fig. 20 further shows an excellent
agreement between the DNS data and these experimental results.
To our knowledge this is the rst time that an accurate numerical
computation is conrming this effect. This opens the possibility to
further study this peculiar behaviour, as a sound physical explanation is still not given.
4. Summary and conclusions
Direct numerical simulations have been performed in order to
study the effect of curvature and Reynolds number in fully developed turbulent pipe ows. The newly obtained data has been carefully validated against previous work by Httl and Friedrich
(2001), yielding excellent agreement. The present results extend
the previous studies in both Reynolds number and curvature
parameter, reaching up to a nominal Res = 400 maintaining turbulence at a curvature of up to j = 0.1. We chose a pipe with length
L = 25Ra, which can be considered sufcient to obtain accurate statistics for the various cases.

1.7

log10(f)

1.8
1.9
2
2.1
2.2
3

3.5

4.5

log (Re )
10

Fig. 19. Fanning friction factor (f),  straight pipe,  coiled pipe at various Reynolds
numbers and j = 0.00964 (data obtained from Cioncolini and Santini, 2006), j DNS
data at different Reynolds numbers for straight pipe (R5300S, R6900S, R11700S), h
DNS data at different Reynolds numbers for curved pipe with j = 0.01 (R5300M,
R6900M, R11700M), - - - - and laminar and turbulent correlations by Ito (1987).

3.7

3.8

3.9

4.1

log10(Reb)
Fig. 20. Close up view of the Fig. 19.

Due to the curvature of the pipe, centrifugal forces are generated in the ow, which lead to a shift of the mean axial ow towards the outer side of the bent pipe. In consequence, the curved
streamlines may develop secondary motion perpendicular to the
primary ow. One pair of counter-rotating vortices are observed
and commonly denoted as Dean vortices. Depending on the values
of Reynolds number, here in terms of the bulk Reynolds number
Reb, and the curvature parameter j, i.e. the ratio of pipe and bend
radii, different regimes can be identied, ranging from intermittent
relaminarisation at low Reb and high j, to weakly modied turbulent pipe ow at high Reb and low j. In this study we concentrate
on those parameters which yield developed turbulent ow,
whereas transitional stages such as the quasi-periodic regime (Di
Piazza and Ciofalo, 2011) are not discussed. Nevertheless we have
observed for our highest curvature partial relaminarisation at the
inner side of the bend, characterised by high atness values of
the velocity signal, but the outer side of the pipe has always been
fully turbulent. Further studies of these relaminarisation processes
and in particular the laminar-turbulent interface might be interesting extensions of the present work.
The effect of the curvature on the secondary motion allows us to
distinguish three rather distinct regimes; zero curvature with vanishing secondary ow, mild curvature with two large counterrotating Dean cells, and strong curvature which causes the core
of these stable vortices to move towards the side walls and a distinct bulge region appears in the pipe centre. Straight and mildly
curved pipes share many similarities in most ow statistics. For
these cases, the mean axial velocity in a vertical cut is close to linear, similar to rotating channel ows.
For the strongly curved pipes, the aforementioned bulge visible
in the secondary ow has also an effect on the mean axial velocity,
which shows a distinct change of slope just below the centre of the


pipe. Together with the relevant shear stress component u0s u0r an
increased production in this region is observed, leading to
enhancement of turbulent kinetic energy in the core of the bent
pipe. This feature has already been reported in the experiments
by Webster and Humphrey (1997) and Httl and Friedrich (2000)
in the transitional (quasi-periodic) state. Here however we observe
both developed turbulence at the outer wall together with the turbulent bulge. Further studies are necessary with regards to what
extent near-wall turbulence is modied by the presence of the
core-turbulence.
The present data base also includes full Reynolds-stress budgets
for the considered cases, not available previously. These statistics
can be helpful for tuning and improving turbulence models. In particular models based on the eddy-viscosity assumption are known
to have problems in the presence of secondary motions and
stream-line curvature. It turns out that for straight and mildly

26

A. Noorani et al. / International Journal of Heat and Fluid Flow 41 (2013) 1626

curved pipe the standard k-e model as a traditional eddy-viscosity


model might give acceptable results, which is however not the case
for larger curvatures.
Another interesting feature of the present simulations is revealed by considering the Fanning friction factor f for the pipes
with mild curvature, j = 0.01 at various Reynolds numbers. As observed experimentally by Cioncolini and Santini (2006), our data
conrms that for a specic range of ow rates mildly curved pipes
exhibit a reduced friction factor compared to straight pipes with
the same mass ow. The physical reasons for this sub-straight drag
are still unknown, and should be addressed in forthcoming studies.
All simulations have been performed using the spectral-element code nek5000, which provides high-order polynomial
approximations of the ow variables together with extremely high
scalability on parallel computers. All statistics including high-order
moments and Reynolds-stress budgets have been computed using
tensor rotations, avoiding cumbersome evaluations of the components in orthogonal toroidal coordinate systems.
Acknowledgements
Dr. Paul Fischer is gratefully acknowledged for his help with
nek5000. We also thank the Swedish National Infrastructure for
Computing (SNIC) for providing the computer time.
Appendix A. Computation of the budget terms in toroidal
coordinate system
In this Appendix the procedure of coordinate transformation
from the Cartesian to toroidal coordinate system used for the computation of statistical moments of different order is discussed. As
described in Section 2, the simulation code solves the governing
equation in Cartesian coordinates (X, Y, Z). In order to obtain averaged statistics in the desired toroidal system, an appropriate transformation needs to be applied. Alternatively, the Reynolds-stress
budget equations in orthogonal helical coordinates were derived
by Httl et al. (2004). Despite the generality of the latter approach
the amount of calculations to construct all different budget terms
based on these equations is very large and potentially memory
consuming. Therefore, the following algorithm is used, suitable to
simulation methods based on a Cartesian representation.
First, a snapshot of the pressure, velocities and velocity gradients is obtained in each necessary timestep in Cartesian coordinates (X, Y, Z); see Fig. 1. Based on these quantities, the required
tensor components (up to order three) of all necessary terms for
velocities, gradients, stresses and budgets are computed, still in
Cartesian coordinates. Using convenient rotation matrices all tensors are rotated about the Z-coordinate axis rst to account for
the pipe bend; this step is naturally omitted when considering a
straight pipe. In the resulting system, (R, s, f), the axial average of
the component can be computed, reducing all terms from 3D
dependency to a 2D plane. A running time-average of these
cross-sectional planes is stored during the simulation. Afterwards,
at the post-processing stage a spectrally accurate scheme is used to
interpolate the data onto a plane polar grid. This new grid is

equally distributed in the azimuthal direction and non-uniformly


spaced in the wall-normal/radial direction. The maximum and
minimum spacings are chosen similar to the original SEM grid.
The required derivatives in the azimuthal and radial directions
are obtained using Fourier and 6th order compact nite differences
(Lele, 1992), respectively. These procedures preserve the high-order accuracy of the simulation data. Finally, the terms are re-constructed again in fully Cartesian form and rotated about the axial
axis (s) to the in-plane polar system (r, h).
The procedure in the straight pipe conguration is similar except the rst coordinate transformation is not required. Also to exploit homogeneity along the circumference of the pipe section an
azimuthal average is performed.
References
Berger, S.A., Talbot, L., Yao, L.S., 1983. Flow in curved pipes. Annu. Rev. Fluid Mech.
15, 461512.
Boersma, B.J., Nieuwstadt, F.T.M., 1996. Large-eddy simulation of turbulent ow in a
curved pipe. J. Fluids Eng. 118, 248.
Bradshaw, P., 1987. Turbulent secondary ows. Annu. Rev. Fluid Mech. 19, 5374.
Cioncolini, A., Santini, L., 2006. An experimental investigation regarding the laminar
to turbulent ow transition in helically coiled pipes. Exp. Therm. Fluid Sci. 30,
367380.
Di Piazza, I., Ciofalo, M., 2010. Numerical prediction of turbulent ow and heat
transfer in helically coiled pipes. Int. J. Therm. Sci. 49, 653663.
Di Piazza, I., Ciofalo, M., 2011. Transition to turbulence in toroidal pipes. J. Fluid
Mech. 687, 72117.
Fischer, P.F., Lottes, J.W., Kerkemeier, S.G., 2008. nek5000 Web page.
Germano, M., 1982. On the effect of torsion on ow. J. Fluid Mech. 125, 18.
Grindley, J.H., Gibson, A.H., 1908. On the frictional resistance to the ow of air
through a pipe. Proc. R. Soc. London Ser. A 80, 114139.
Hoyas, S., Jimnez, J., 2008. Reynolds number effects on the Reynolds-stress budgets
in turbulent channels. Phys. Fluids 20, 101511.
Httl, T., Chaudhuri, M., Wagner, C., Friedrich, R., 2004. Reynolds-stress balance
equations in orthogonal helical coordinates and application. ZAMM 84, 403
416.
Httl, T.J., Friedrich, R., 2000. Inuence of curvature and torsion on turbulent ow in
helically coiled pipes. Int. J. Heat Fluid Flow 21, 345353.
Httl, T.J., Friedrich, R., 2001. Direct numerical simulation of turbulent ows in
curved and helically coiled pipes. Comput. Fluids 30, 591605.
Ito, H., 1959. Friction factors for turbulent ow in curved pipes. J. Basic Eng. 81,
123134.
Ito, H., 1987. Flow in curved pipes. JSME Int. J. 30, 543552.
Kristoffersen, R., Andersson, H., 1993. Direct simulations of low-Reynolds-number
turbulent ow in a rotating channel. J. Fluid Mech. 256, 163-163.
Lele, S., 1992. Compact nite difference schemes with spectral-like resolution. J.
Comput. Phys. 103, 1642.
Manlapaz, R., Churchill, S., 1980. Fully developed laminar ow in a helically coiled
tube of nite pitch. Chem. Eng. Commun. 7, 5778.
Naphon, P., Wongwises, S., 2006. A review of ow and heat transfer characteristics
in curved tubes. Renew. Sust. Energy Rev. 10, 463490.
Sreenivasan, K., Strykowski, P., 1983. Stabilization effects in ow through helically
coiled pipes. Exp. Fluids 1, 3136.
Vashisth, S., Kumar, V., Nigam, D.P.K., 2008. A review on the potential application of
curved geometries in process industry. Ind. Eng. Chem. Res. 47, 32913337.
Wallin, S., Johansson, A., 2002. Modelling streamline curvature effects in explicit
algebraic Reynolds stress turbulence models. Int. J. Heat Fluid Flow 23, 721
730.
Webster, D.R., Humphrey, J.A.C., 1993. Experimental observations of ow instability
in helical coil. J. Fluids Eng. 115, 436443.
Webster, D.R., Humphrey, J.A.C., 1997. Traveling wave instability in helical coil ow.
Phys. Fluids 9, 407418.
Yamamoto, K., Akita, T., Ikeuchi, H., Kita, Y., 1995. Experimental study of the ow in
a helical circular tube. Fluid Dyn. Res. 16, 237249.

S-ar putea să vă placă și