Sunteți pe pagina 1din 6

Evaluation of URANS Solvers for Cylindrical Structures in Tidal Flow

Robert Matthew Stringer and Jun Zang


Department of Architecture & Civil Engineering, University of Bath
Bath, UK







ABSTRACT

In tidal conditions Reynolds numbers (Re) around cylindrical structures
can range from zero up to supercritical values, with vortex shedding
causing significant vibration and undesirable forces throughout. Two
realistic options for investigating these phenomena in an engineering
environment include the Navier-Stokes solvers ANSYSCFX-13.0
and OpenFOAM. In this study both are considered at Reynolds
numbers ranging from 40 to 10
6
, representing a full range of tidal
conditions. A 2D Unsteady Reynolds Averaged Naver-Stokes
(URANS) approach is developed in an attempt to balance the speed and
accuracy of the solutions. To eliminate mesh dependency, low-Re
boundary layer meshing is employed along with Courant controlled
time stepping.
Numerical results obtained from both solvers are presented
for lift coefficient, drag coefficient and Strouhal number. Values from
both solvers are compared to literary sources and examined for
accuracy. Given that the turbulence for much of the Reynolds number
region explored is known to be highly three dimensional, both codes
offer a mix of good and poor results depending on parameters. Overall
low and subcritical Re values are modelled with success by
OpenFOAM, while force terms in upper critical and supercritical flows
are better predicted by CFX.

KEY WORDS: CFX, Cylinder; Low-Re; Marine Energy;
OpenFOAM; Renewables; Tidal; URANS; Vortex Induced Vibrations

INTRODUCTION

Historically, the circular cylinder form has been included in many
offshore structures, particularly in foundations and pipelines. The
emerging marine renewable energy industry already includes countless
examples of wave and tidal device designs featuring major structural
and active components in cylindrical form. With vortex induced
vibrations (VIV) already a common problem in offshore structures, the
addition of power extracting mechanisms makes the understanding and
control of the flow field ever more important. Both local to devices and
as part of an array, the ability to capture the forces and wake shed from
cylinders is a desirable asset, allowing suitable tolerances or intelligent
design of major framework to be factored into device development.
In this study two software packages are examined for the purpose of
making engineering predictions for offshore renewable energy
structures; first is the open source solver OpenFOAM 1.7.1, and second
is the commercial program ANSYSCFX-13.0 Academic Research
(CFX). The motivation of the study is to provide a comparative
evaluation of the two solvers over a range of Reynolds numbers (Re)
analogous to tidal flow conditions using realisable numerical models.
OpenFOAMis a C++based CFD software written for Linux and
released in 2004 under a general public license by OpenCFDLtd.
The commercial derivative of CFX is widely used in a number of
industries with an extensive history of validation.

Background

The flow around cylinders has been the subject of numerous
investigations initially through experimentation and more recently by
numerical methods. The development of turbulence is a highly
transient phenomenon with both location and intensity varying with
time and flow speed. In order to classify such complex behaviour
Lienhard (1966) proposes six distinct Reynolds number ranges
accompanied by a regularly cited diagram. The regimes of flow have
been incrementally refined, most notably Zdravkovich (1990) identifies
fifteen distinct ranges. Table 1 gives a summary of the key flow
characteristics for all known Reynolds number values.

Table 1. Flow regimes around a circular cylinder

Re Range* Flow Regime
Re <1 Creeping flow
3-5 <Re <30-40 Steady separation (Fppl vortices)
30-40 <Re <150-300 Laminar periodic shedding
150-200 <Re <1.4x10
5
Subcritical
1.4x10
5
<Re <1x10
6
Critical
1x10
6
<Re <5x10
6
Supercritical
5x10
6
<Re <8x10
6
Transcritical
8x10
6
<Re Postcritical
*Values from Zdravkovich (1990) and Raghaven & Bernitsas (2011)
703
Proceedings of the Twenty-second (2012) I nternational Offshore and Polar Engineering Conference
Rhodes, Greece, June 1722, 2012
Copyright 2012 by the I nternational Society of Offshore and Polar Engineers (I SOPE)
ISBN 978-1-880653-944 (Set); ISSN 1098-6189 (Set)
www.isope.org

Using realistic velocity data reported by The Carbon Trust (TCT, 2005)
and an estimated pile size ranging from 0.5m up to 3m, the larger being
equal to the Seagen turbine currently deployed in Northern Ireland, the
resulting maximum Reynolds number ranges from 0.810
6
to 12.910
6
.
The study here will consider low Re slack water values, up to a
maximum of Re =10
6
, this upper limit is chosen due to computational
requirements, representing the lower end of the calculated maxima.
Existing numerical contributions are now considered with
selected examples representing the general consensus throughout the
Re range. Beginning with low Reynolds numbers, up to 160, work by
Park et al. (1998) and Dehkordi and J afari (2009) both find a laminar
URANS model to be capable of achieving realistic values. Rahman et
al. (2008) compares a number of two-equation turbulence models at Re
values of 1000 and 3900, considered to be in the lower critical regime.
The results show a superior performance by the stress transport model
(SST) over the k - e and realizable k -e models. Increasing the
Reynolds number to supercritical values includes work by Ong et al.
(2009), where a k - e model is used at Re values of 10
6
and 3.6x10
6
.
Results are compared with 2D and 3D LES, and experimental data,
demonstrating strong correlation and the conclusion that the model is
suitable for engineering applications. Although some success is
demonstrated in this region, the work includes only a narrow band of
Reynolds number and uses standard wall functions with appropriately
coarse meshing at the cylinder surface. Benim et al. (2007) explores
the issue of near wall meshing by using the commercial code Fluent
to compute a cylinder at Re =10
4
using wall models and the standard
k - e turbulence model. Meshes vary in y
+
values from 10 to 1000
yielding a large range of diverging drag values. Consequently, the
author elects to test without wall functions using meshes that conform
to y
+
=1. Using this technique Benim finds similar correlation to Ong
in the supercritical regime but with rapid accuracy loss in the critical
transition region.
Briefly considering LES, Tutar and Hold (2001) compute
both RANS and LES models at an Re of 1.4x10
5
. Findings conclude
that k -e RANS model produced a general under-prediction of forces
while LES gave over-prediction. While LES in this case uses a fully
resolved boundary layer, the RANS uses wall models that have
previously been shown to be highly mesh dependant.
From the reviewed literature one can make a number of
observations; firstly that URANS methods can give satisfactory results
for the circular cylinder case. Secondly, that the results are highly mesh
and turbulence model dependant, and lastly that LES does not appear to
give significant improvement in the 2D case. Due to these findings a
URANS SST formulation and development of a low-Re meshing
scheme is employed throughout this study.

Solver Comparison

Both OpenFOAM and CFX utilise the finite volume method (FVM) to
approximate solutions to the Navier-Stokes equations. Table 2 gives an
outline of the critical mathematical features of the two solvers.

Table 2. Comparison of fundamental solver features

Attribute CFX OpenFOAM*
Solution matrix Coupled Segregated
Temporal control Implicit Implicit/Explicit
Discretisation Median-Dual
Cell-Vertex
Cell-Centred
Variable handling Collocated Collocated
Equation coupling Rhie-Chow (adapted) PISO
*Attributes specific to the pisoFoam module

Comparing the two solvers in Table 2, it is clear that they are distinctly
different, in both calculation method and discretisation. Note that all
future references to OpenFOAM are specifically in regard to the
implementation of the pisoFoam solver.
Turbulence modelling for both solvers, is split into globally
low-Re cases, using a Laminar model, while flows with Re>150
implement the SST model, as developed by Menter (1994).
In both cases solvers have been kept to what would be
considered largely default. While a comprehensive account of the
setup is provided in the next section, any omissions regarding
underlying constants should be assumed to be baseline values.

DEFINITION & METHOD

Defining the problem in a computational domain includes geometry,
boundary conditions, meshing and solution control. Each of these is
examined in this section beginning with the domain geometry in Fig. 1.

Fig. 1, Geometry of the computed domain (not to scale)

Boundary Conditions

The boundary conditions for all cases were defined in OpenFOAM and
CFX with similarity an objective.
A uniform flow is specified at the inlet, calculated by
rearrangement of the Reynolds number Eq. 1 for flow velocity u;
where p is density, I is characteristic length (cylinder diameter) and p
is dynamic viscosity.

Rc =
puI
p
(1)

The outlet is set as a pressure boundary for both solvers, in CFX a
relative pressure of zero is given; P
cI
= u, In OpenFOAM the
equivalent is a free stream pressure outlet. The sides assigned as
free-slip boundaries, shown in Fig 1, provide the condition of zero
velocity normal to the wall and zero shear stress; u

= u,
wuII
= u.
The boundaries in the X-Y plane were set as symmetry planes; here the
geometry and variables are considered to be mathematically symmetric,
with zero diffusion or flow across the boundary. Velocity is the one
exception, where parallel flow is symmetric, but normal flow is reduced
to zero. The cylinder boundary is set to a no-slip condition, where
pressure is set to zero gradient and velocities are set to zero; u
x
=
u

= u. For cases implementing the SST turbulence model, values are


required for turbulence kinetic energy k and turbulent frequency , see
Eqs. 2~3 (ANSYSAcademic Research, 2010). Turbulence intensity
I is 2.5%, where turbulence length l = u.u7L and C

is the non-
dimensional constant 0.09. Values of k and at the cylinder wall are
20L
20L
20L
40L
u


Velocity
Inlet
Free Slip
Pressure
Outlet
L
y
x
Free Slip
704

given in Eqs. 4~5 (Menter, 1994), where : is kinematic viscosity,
[ = u.u7S, and y
1
is the distance to the first node.

k =
S
2
(u

I)
2

(2)
=
k
3
2
kC

l

(3)
k
wuII
= u (4)

wuII
=
6u:
[y
1
2
(5)

Meshing

To assure a high level of grid independence, a low-Re approach to
surface meshing is taken. The term low-Re is not to be confused with
global Reynolds number, but indicates the low turbulent Reynolds
number that exists in the viscous sublayer. The y
+
value represents a
non-dimensional distance of the first node from a no-slip wall, see Eqn.
6. where
o
is shear stress. In order to utilise low-Re boundary
properties it is generally accepted that the mesh must achieve first layer
cell thicknesses equivalent to a y
+
< 1.

y
+
=
_

o
p
y
1
:

(6)

An iterative refinement of the boundary mesh was conducted until the
boundary mesh achieved a maximum y
+
of 1, and a minimum 5 mesh
points located in the boundary layer, as defined by a velocity thickness
of u.99u

..



Fig. 2, Mesh images for Re =1000 flow case, with near field structured
hexahedral mesh and far field unstructured prism mesh.

A typical mesh in the near field of the cylinder is shown in Fig. 2. The
lack of any symmetry is theoretically unimportant given sufficient grid
resolution, a positive aspect being that it aids the development of vortex
shedding. An expansion ratio of 1.1 was used with a total of 30 layers.
Furthermore, the aspect ratio of near wall cells was kept to below 20:1
for all cases, with the exception of Re=10
6
. An exact mesh match
between the two solvers was maintained by generating all meshes in
ANSYSMeshing 13.0 then converting into OpenFOAM format for
each Reynolds number.

Solver Control

Both CFX and OpenFOAM are essentially both explicit solvers;
however, the pisoFoam solver does not include outer loop corrections,
i.e. recalculation of the N-S equations at any given timestep. The result
is that a low Courant number (Cr) is required to maintain numerical
stability, calculation of which is given in Eqn. 7, where t is the
timestep and x is the minimum cell width. As a consequence the
timestep decreases significantly as the mesh is refined for greater
Reynolds numbers and hence processing time increases
disproportionally. It should be noted that pisoFoam does not include
Courant controlled timesteping by default; therefore modifications to
the source code were completed.

Cr =
ut
x

(7)

After testing initially at a low Reynolds number using a laminar model,
it was established that a time accurate stable result was achievable with
a courant number of 0.8.
The second important aspect of solver control is the
convergence criteria. A key judge of solution convergence is the mass
and momentum solution residuals. These were set to 10
-6
, the
recommended value for both CFX and the pisoFoam solver in
accompanying guidance for tight convergence; this value is selected for
all cases.
Specifying total time is calculated from a non-dimensional
time value t
-
as given in Eqn. 8. All simulations are solved to 150 non-
dimensional time units.

t
-
=
tu
I

(8)

Post-Processing

Data from CFX and OpenFOAM were post-processed using ANSYS
CFD-Post 13.0 and ParaView 3.12.0-RC2 respectively. Instantaneous
values of drag coefficient C

and root mean square of lift coefficient


C
Lms
are calculated using Eqns. 9~10, where F

and F
L
are the
corresponding unit forces. The Strouhal number (St) represents a
normalised value of shedding frequency; see Eqn. 11, where is the
shedding frequency in Hz.
C

=
F

u.Spu
2
I
(9)
C
Lms
=
_
1
n
(F
L
1
2
+ F
L
2
2
++ F
L
n
2
)
u.Spu
2
I

(10)
St =
I
u

(11)
705

RESULTS & DISCUSSION

To represent the full range of conditions expected in the case of a
cylinder in tidal flow, computations have been performed at Re =40,
100, 10
3
, 10
4
, 10
5
and 10
6
. The following results serve to evaluate a
number of objectives, namely, the accuracy of URANS simulation
using the Menter SST turbulence model combined with low-Re
boundary layer meshing. Secondly, the comparability of the
commercial code CFX with the open-source code OpenFOAM (using
the pisoFoam solver), given nominally identical cases. The following
results presented for CFX and OpenFOAM were calculated fromdata
representing a period of 5 shedding oscillations, equivalent to a solution
time period of t*=140 ~150, with the associated value having reached
a quasi-steady transient state in all cases.

Coefficient of Drag

The mean drag coefficient for both solvers are plotted in Fig. 3 against
published work for an infinitely long smooth cylinder by Zdravkovich
(1990) and Massey (1989). Comparing initially the results from CFX
and OpenFOAM, there is full agreement at low Re values, up to 100.
However, as Re increases above this point the two codes increasingly
differ. In comparison with published data, OpenFOAM continues to
give values within 0.1C
D
up to Re =10
4
. CFX over-predicts C
D
at Re
values of 10
3
and 10
4
, but shows some recovery in the critical and
supercritical regions, that of Re = 10
5
and 10
6
. At these values
OpenFOAM rapidly loses accuracy, under predicting C
D
. The wide
variance in the critical region, around, 10
5
<Re <3.5x10
6
, is a known
region of high instability with experimental values varying significantly
as discussed by Shih et al (1993) and identified specifically by
Zdravkovich in Fig. 3. To investigate the variance between the solvers,
one must examine the lift coefficient.

Lift Coefficient

A significant contribution to the pressure force distribution is that of
vortex shedding, an expected feature of the flow at the Reynolds
numbers tested (with the exception of 40). The fluctuating lift force F
L

provides an accessible record of vortex shedding; the rms values of lift
coefficient are plotted in Fig. 4. Data from Norberg (2003) are used for
comparison due to the impressive determination of the given curves
resulting from analysis of around 50 published sources. In parallel to
the trends in C
D
, OpenFOAM performs reasonably up to Re =10
4
with
results falling within the scatter of the original data points presented by
Norberg (not shown in Fig. 4). At Re =10
5
OpenFOAM is seen to
generate almost zero lift, suggesting the absence of any significant
shedding, followed by a final value at Re =10
6
recovering to closely
match Norbergs prediction. The CFX results differ significantly from
OpenFOAM in the subcritical region with highly over predicted values
at Re =10
3
and Re =10
4
. These results are indicative of the limitations
imposed by the 2-dimensional constraint of the model when attempting
to resolve a shedding region containing highly 3D phenomenon. Issues
including a lack of spanwise flow, and a low resolution of the
turbulence spectra, contribute to this error. At supercritical Re the CFX
results for C
L
rms return to values with less than a 15% error from
Norbergs result. One may postulate that the dramatic decline to more
realistic values is due to the transition from a strongly 3D regime with
distinct laminar shedding modes at the point of separation.
Generally the lift forces can be shown to correlate very well
with drag predictions in terms of magnitude. However, the dramatic
differences between the two solvers tested are somewhat surprising.
Analysis of the vortices shows a distinctly lower pressure at vortex
centres in CFX than in OpenFOAM, a disagreement that would account
for the additional forces acting on the cylinder.






Fig. 3, Drag coefficient versus Reynolds number; correlation between experimental and numerical results. Published values for smooth cylinder:
Zdravkovich (1990) and Massey (1989).
0.1
1
10
1 10 100 1000 10000 100000 1000000 10000000
C
D
Re
Zdravkovich
Massey
CFX
OpenFOAM
706







Fig. 4, Graph of the rms of lift coefficient versus Reynolds number. Published experimental data: Norberg (2003)






Fig. 5, Graph of Strouhal number versus Reynolds number. Experimental data: Norberg (2003); Achenbach & Heinecke (1981)
-0.1
0.1
0.3
0.5
0.7
0.9
1.1
1.3
10 100 1000 10000 100000 1000000 10000000
C
L
r
m
s
Re
Norberg
CFX
OpenFOAM
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
0.55
10 100 1000 10000 100000 1000000 10000000
S
t
Re
Norberg
Achenbach & Heinecke
CFX
OpenFOAM
707

Strouhal Number

The final part of the analysis presented here considers the Strouhal
number of simulated vortex shedding. The Strouhal number is an
important indicator of the transient accuracy of the simulation. The
results in Fig. 5 for CFX and OpenFOAM are compared with an
experimental best fit curve from Norberg (2003) and experimental
results at high Re from Achenbach & Heinecke (1981). Success in the
laminar range continues as the Strouhal number is accurately captured
by both codes to 3.d.p. of accuracy. Both codes are within 9% of
findings by Norberg at Re = 1000. Above this point the CFX
simulations begin to shed at steadily increasing rates, failing to predict
the drop at intermediate Re values. OpenFOAM provides a matching
Strouhal frequency at Re =10
4
, before a sharp drop at 10
5
and failure
to shed at 10
6
. With both solvers failing to predict the sharp rise in the
supercritical region, it is clear that the SST model is no longer able to
produce a realistic flow field. The sharp rise indicates the transition of
the boundary layer to a turbulent state, this dramatically reduces the
length scale of eddies significantly below the resolution of URANS
method. However, although the shedding can no longer be realistically
captured, we have previously seen the drag and lift coefficients being
predicted with reasonable accuracy. One may postulate that the RANS
averaging of the more highly turbulent flow is more suited to the CFX
model than the structured shedding at lower Re values. Paradoxically
OpenFOAM fails to shed, with the previous C
L
value at 10
6
being
accurate by chance rather than realistic flow conditions. In an attempt
to increase the transient resolution of the OpenFOAM result, a Courant
number of 0.1 was assigned, showing little improvement. Although
this differs fromCFX, the PISO SST combination of OpenFOAM is
the likely cause of this behavior, with time averaging of such of high
frequency oscillation resulting in the failure to capture shedding.

CONCLUSIONS

The aim of the research presented was to bring together a number of
potentially optimal numerical attributes in an attempt to predict the
forces and flow field around a circular cylinder over a realistic
operational range of flow velocities analogous to tidal conditions. To
achieve this, two software options have been employed and compared;
ANSYSCFX-13.0 and OpenFOAM. A key element of the study
has been to establish a specific methodology with regard to geometric,
numerical and discrestisation practices and apply them to all cases.
Specifically this includes:

Application of URANS calculation using SST turbulence
Domain size/Cylinder ratio avoiding edge effects
Surface meshing to specified y
+
and cells in boundary layer
Utilising fully adaptive Courant controlled timestepping
Maintaining maximum commonality between solvers

The results have shown some success for both solvers. In the case of
OpenFOAM, low and subcritical flows have been modelled with all
values falling within the scatter of published experimental values.
CFX produces a more typical response, over predicting subcritical
flows but succeeding in the estimation of forces in the supercritical
region. While additional data points are required to consolidate the
results from OpenFOAM, results suggest that it is capable of providing
values suitable for engineering purposes up to Re = 10
4
. This
conclusion is supported by a more comprehensive study of the
simulations including transient vortex growth, boundary layer analysis,
viscous and inertial force correlation, and turbulence quantities.



REFERENCES

Achenbach, E. & Heinecke, E. (1981). On Vortex Shedding from
Smooth and Rough Cylinders in the Range of Reynolds-Numbers
6x103 to 5x106. Journal of Fluid Mechanics, 109, 239-251.
ANSYSAcademic Research (2010). Help System, 3.2.12.3 Manual
Specification, ANSYS, INC. . 2010.
Benim, A. C., Cagan, M., Nahavandi, A. & Pasqualotto, E. (2007).
RANS predictions of turbulent flow past a circular cylinder over the
critical regime. Proceedings of the 5th Iasme/Wseas International
Conference on Fluid Mechanics and Aerodynamics (Fma '07), 235-
240.
Dehkordi, B. G. & Jafari, H. H. (2009). Numerical simulation of flow
through tube bundles in in-line square and general staggered
arrangements. International Journal of Numerical Methods for Heat
and Fluid Flow, 19, 1038-1062.
Lienhard, J . H. (1966). Synopsis of lift, drag, and vortex frequency
data for rigid circular cylinders. Washington State University
College of Engineering -- Research Division Bulletin, 32.
Massey, B. S. (1989). Mechanics of fluids, Van Nostrand Reinhold.
Menter, F. R. (1994). 2-Equation Eddy-Viscosity Turbulence Models
for Engineering Applications. Aiaa Journal, 32, 1598-1605.
Norberg, C. (2003). Fluctuating lift on a circular cylinder: review and
new measurements. Journal of Fluids and Structures, 17, 57-96.
Ong, M. C., Utnes, T., Holmedal, L. E., Myrhaug, D. & Pettersen, B.
(2009). Numerical simulation of flow around a smooth circular
cylinder at very high Reynolds numbers. Marine Structures, 22,
142-153.
Park, J ., Kwon, K. & Choi, H. (1998). Numerical solutions of flow past
a circular cylinder at Reynolds numbers up to 160. Journal of
Mechanical Science and Technology, 12, 1200-1205.
Raghavan, K. & Bernitsas, M. M. (2011). Experimental investigation
of Reynolds number effect on vortex induced vibration of rigid
circular cylinder on elastic supports. Ocean Engineering, 38, 719-
731.
Rahman, M. M., Karim, M. M. & Alim, M. A. (2008). Numerical
investigation of unsteady flow past a circular cylinder using 2-D
finite volume method. Journal of Naval Architecture and Marine
Engineering, 4.
Shih, W. C. L., Wang, C., Coles, D. & Roshko, A. (1993).
Experiments on Flow Past Rough Circular-Cylinders at Large
Reynolds-Numbers. Journal of Wind Engineering and Industrial
Aerodynamics, 49, 351-368.
TCT (2005). Tidal streamenergy resource and technology summary
report. 4 ed.: The Carbon Trust.
Tutar, M. & Holdo, A. E. (2001). Computational modelling of flow
around a circular cylinder in sub-critical flow regime with various
turbulence models. International Journal for Numerical Methods in
Fluids, 35, 763-784.
Zdravkovich, M. M. (1990). Conceptual Overview of Laminar and
Turbulent Flows Past Smooth and Rough Circular-Cylinders.
Journal of Wind Engineering and Industrial Aerodynamics, 33, 53-
62.


708

S-ar putea să vă placă și