You are on page 1of 14

Journal of Wind Engineering

and Industrial Aerodynamics 81 (1999) 145}158

Application of CFD to environmental #ows


Sung-Eun Kim*, Ferit Boysan
Fluent Incorporated, Centerra Resource Park, 10 Cavendish Court, Lebanon, NH 03766-1442, USA

Abstract

This paper is concerned with the major issues considered to be important for successful
application of computational #uid dynamics to environmental #ows. Among the issues of
primary concern in the present paper are meshing and turbulence modeling. As regards
meshing, we propose an approach that employs unstructured meshes in conjunction with
solution-adaptive mesh re"nement. Examples are presented to illustrate the e$cacy of the
unstructured mesh-based approach. The issue of turbulence modeling is discussed at length,
inasmuch as turbulence modeling determines the "delity of computational results for environ-
mental applications. Emphasis is laid upon the capability of engineering turbulence models to
capture the salient features of environmental #ows such as atmospheric boundary layer over
a smooth terrain and #ows around blu! bodies.  1999 Published by Elsevier Science Ltd. All
rights reserved.

Keywords: Unstructured mesh; Turbulence models; Blu! bodies; Separation; Vortex shedding

1. Introduction

The topics that fall in the realm of modern environmental #uid mechanics are
diverse. The examples are; indoor air #ows laden with contaminants, #ows around
in-land and o!-shore structures including buildings, bridges, stacks, towers, oil rigs,
and e%uents dispersion into the environment including rivers, lakes, estuaries, coastal
waters, and atmosphere. Among this broad range of topics, wind climate in urban
areas and its rami"cations to environmental, structural and architectural engineering
practices are of major interest to city planners, designers and builders. The whole issue
of wind engineering has numerous components to it, ranging from urban climatology

* Corresponding author.
E-mail address: sek@#uent.com (S.E. Kim)

0167-6105/99/$ - see front matter  1999 Published by Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 9 9 ) 0 0 0 1 3 - 6
146 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

to building aerodynamics to #uid}structure interaction. In this paper, we will focus


our discussion on the aerodynamics aspect.
Traditional approaches to investigating wind e!ects on buildings and structures in
the past and present days have employed #ow visualization and measurement using
scaled models in meteorological or environmental wind tunnels. Full-scale measure-
ments are sometimes carried out to provide data for wind loading on buildings and
structures and to determine the characteristics of atmospheric wind which is needed to
simulate the natural wind in wind tunnels. However, like in any other applications in
the discipline of #uid mechanics, computational #uid dynamics (CFD) backed up by
the tremendous computing power available today is greatly impacting the whole
practice.
Numerical modeling of the wind e!ects on buildings and structures poses several
challenges to CFD practitioners. First, computation of the #ows around buildings
and structures in cities requires knowledge of the characteristics of atmospheric
boundary layer. Besides mean wind speed data, atmospheric turbulence data (wind
gusts) are needed, in principle, to accurately represent atmospheric wind and its e!ects
on buildings and structures in computational models. Often times, however, urban
meteorological data for atmospheric wind do not provide the level of details needed
for CFD modeling. This directly translates to the di$culty of obtaining appropriate
boundary conditions that CFD modeling requires. Secondly, topography of the
con"gurations to be modeled is usually complex. Especially in urban areas, where
closely spaced building groups are commonplace, with di!erent individual topologies,
heights and orientations, great challenges are encountered when discretizing the
computational domain. Thirdly, the #ows are highly complex, having all the elements
modern #uid mechanics has not quite successfully resolved yet. The major challenge
lies in turbulence modelling. The di$culty is associated with the fact that the #ows are
highly three dimensional, being accompanied, almost without exception, by strong
streamline curvature, separation, vortices of various origins and unsteadiness. Lastly,
the #ows encountered in urban areas involve spatial and temporal scales ranging from
small gusts to vortices of size comparable to that of building or structure itself. There
is no doubt that resolving such widely varying scales requires a signi"cant amount of
computational resource.
We discuss in this paper the opportunities and challenges provided by CFD tools in
the area of building aerodynamics. The discussion will address what CFD can o!er
today and what di$culties CFD practitioners are faced with. The paper will focus,
among others, the two major issues alluded to in the foregoing, i.e., meshing and
turbulence modeling. For meshing, we propose an approach based on unstructured
meshes that enables CFD practitioners to economically model complex geometries
and complex #ow physics. The impacts of turbulence modeling on the "delity of CFD
predictions will also be discussed via a number of examples which may appear
simple but have direct relevances to practical applications in wind engineering
and building aerodynamics. The turbulence models to be discussed here include
a number of popular engineering turbulence models and some more recently pro-
posed models. Besides the engineering turbulence models based on Reynolds-aver-
aged Navier}Stokes (RANS) equation, we will also discuss large eddy simulation
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 147

(LES) in lieu of the increasingly important role it plays these days in blu!-body
aerodynamics.
For the computations illustrated here, the commercial CFD software FLUENT
[1}3] was used.

2. Domain discretization for complex topography

Any numerical procedure to compute #uid #ows in the context of CFD requires
discretization of the #ow domain under consideration. Meshing the #ow domain
around an isolated building hardly poses a grand challenge. However, it is not an easy
task to design meshes for buildings groups or urban structures involving complex
topography. Especially when employing structured mesh, complex geometries with all
the signi"cant details frequently encountered in industrial applications often make
adequate meshing very di$cult, if not impossible, limiting the usability of CFD.
Furthermore, employing structured meshes with local clustering to resolve large
solution gradients often results in a considerable portion of the mesh being wasted in
areas where "ne mesh resolution is not really needed. These predicaments encoun-
tered in using structured meshes for complex geometries are carried over to building
aerodynamics.
In recent years, unstructured mesh technology has attracted a great deal of atten-
tion from the CFD community and has been quite successfully applied to industrial
#ows. It allows one to employ computational cell shapes of arbitrary topology
including quadrilaterals, hexahedra, triangles, tetrahedra, prisms, and combinations
of all these. One immediate advantage o!ered by unstructured meshes over structured
meshes is the #exibility in dealing with complicated #ow con"gurations which directly
translates into dramatic reduction in the amount of time to generate meshes.
Fig. 1 illustrates two unstructured hexahedral meshes generated for typical building
groups using commercial preprocessors. These meshes could be generated in a small
fraction of time that would have taken if block-structured meshes were adopted. The
signi"cant time saving is possible only because unstructured meshes are much more
amenable to automated meshing.
Another advantage of using unstructured meshes is that they naturally provide
a convenient framework for local mesh adaptation. This approach has a great
potential to bene"t CFD predictions of building aerodynamics, especially #ows over
building groups with a wide range of length scales to be resolved. The mesh adapta-
tion a!ects the mesh locally, yielding a mesh re"ned only where necessary. In the
adaptations using hanging nodes (locally embedded mesh), those computational cells
which satisfy a prescribed adaptation criterion are subdivided into smaller cells by
using midpoint nodes on edges and/or at cell and face centers. At the simplest level,
a region adaptation can be employed where the cells inside a prescribed region can be
marked for re"nement. More sophisticated adaptations make use of numerical solu-
tions themselves, yielding `solution-adapteda meshes. In the solution-adaptive re"ne-
ment, any solution variables or quantities derived therefrom, or the gradients of such
quantities can be used to specify adaptation criteria.
148 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

Fig. 1. Unstructured hexahedral meshes around typical building con"gurations.

Fig. 2. Examples of mesh adaptation for a triangular bump mounted on a wall left: coarse mesh and
predicted streamlines right: adapted mesh and predicted streamline.

As an example of local mesh adaptation, let us consider here the #ow over
a triangular bump [4]. The incoming boundary layer #ow separates at the apex of the
ridge, reattaching the bottom plane at a distance in the downstream. A strong shear
layer is formed in the downstream of the apex between the recirculation region and
the freestream. From the #ow physics' point of view, the ideal computational mesh
would be the one that can resolve the incoming boundary layer and the recirculation
region, the shear layer and the redeveloping boundary layer after the reattachment, all
at the same time.
Fig. 2 shows two meshes (original and re"ned) and the corresponding streamlines
computed using the two meshes. The "ne mesh was generated using the original
coarse mesh by re"ning the computational cells where the vorticity magnitude is
larger than a prescribed value. We made use of the fact that vorticity magnitude is
usually large in boundary layers and wakes, in order to demarcate the region where
viscous e!ects are important, i.e. the boundary layer and the recirculation region in
this example. The reattachment points predicted with these two meshes are 8.4 and 9.0
bump heights, respectively. Still underpredicting the size of the recirculation (note that
the standard k}e model was used here), the result obtained with the locally re"ned
mesh matches the experimental data [4] (9.8 bump heights) quite closely.
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 149

In summary, unstructured meshes are anticipated to become increasingly popular


in the coming years because of the mesh #exibility and the adaptation capability, and
will enhance the usability of CFD and the quality of CFD predictions for complex
environmental #ows.

3. Turbulence modeling for environmental 6ows

The "delity of CFD solutions for turbulent #ows is dictated by turbulence


modeling, especially when it comes to the #ows around building and structures,
because of the complex features of the #ows as alluded to earlier. One of the
challenging features of the #ow is that the #ows hardly remain at or near equilibrium.
The #ows around and in the near-wakes of buildings and structures keep changing
relentlessly. It is this non-equilibrium nature of the #ows that makes turbulence
modeling for blu!-body #ows di$cult.
Surveying recent literature shows that there is no shortage of turbulence models.
There are, indeed, numerous engineering turbulence models ranging from simple
algebraic models to highly sophisticated models like Reynolds-stress transport
models (RSTM) adopting second-moment closure. Among them, the original k}e
model proposed by Launder and Spalding [5] (referred to `standarda k}e model
hereafter) almost three decades ago seems to be still the most widely used in the
building aerodynamics community for their simplicity, robustness and reasonable
accuracy [6,25]. Despite their e$cacy, however, the conventional k}e models have
inherent drawbacks originating from the underlying hypotheses (isotropic eddy-
viscosity hypothesis). The conventional k}e models have been found by many investi-
gators to give mediocre results in situations where the #ows are dominated by strong
anisotropy in turbulence and nonequilibrium e!ects. Murakami et al. [7], in particu-
lar, made a careful study of a number of engineering turbulence models including the
standard k}e model, attempting to shed light on what causes the poor performance of
the isotropic eddy-viscosity-based k}e models for blu!-body #ows. The mediocre
performance of the traditional k}e models has prompted pursuit of better alternatives.
A number of alternatives have been proposed [8}10] including several anisotropic k}e
models [11}13].
In the last few years, one increasingly encounters industrial #ow simulations
employing the RSTM wherein the transport equations for individual Reynolds
stresses are solved. Anisotropy and transport of the Reynolds stresses are rigorously
accounted for in the RSTM. Although these more elaborate models "nd their frequent
usages in speci"c applications where conventional two-equation models awfully fail,
e.g. cyclones, swirl combustors, etc., the RSTM has not been widely used for building
aerodynamics applications, although Murakami et al. [7] have employed an algebraic
Reynolds stress model where the transport of stresses are approximated using an
equilibrium argument. The primary reason for the limited use of the full RSTM for
industrial applications including environmental #ows probably lies in its higher
computational cost, although it is not quite justi"able these days. Besides, there are
several issues yet to be resolved in the second-moment closure, such as the modeling of
150 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

pressure}strain correlation, wall}echo e!ects, and e-equation, among others. Never-


theless, due to the ever-increasing computing power, the RSTM will become more
widely used in the coming years for building aerodynamics.
With so many choices of turbulence models with varying sophistication and
associated computational e!orts, one would naturally ask what one particular model
o!ers vis-à-vis other models, and what level of turbulence modeling is needed for the
#ows around buildings and structures. Unfortunately, there seems to be no indisput-
able answer to the question. In this paper, we present a number of examples that may
provide some hints to the question.
The "rst example is the two-dimensional steady boundary layer #ow over a curved
hill, which was studied experimentally by Baskaran et al. [14]. This #ow can be
considered to represent atmospheric wind blowing over a smooth terrain. The
incoming #ow is modeled with a typical #at-plate turbulent boundary layer. As the
boundary layer passes over the hill, it is subjected to streamline curvature of alternat-
ing signs (convex and concave) and strong favorable and adverse pressure gradients.
The experimental data indicate that there is an incipient separation at about 1.1 m
downstream of the leading-edge on the leeward side of the hill. The freestream
Reynolds number per unit length (; /l) is approximately 1.33;10 m\. The compu-

tations were performed on a 170;90 quadrilateral mesh using three di!erent k}e
models (standard k}e [5], realizable k}e [10], renormalization-group-based k}e [8]
models), and a RSTM [15]. The standard wall functions [5] were employed.
Fig. 3 shows the predicted pressure and skin-friction distributions. Note that x"0
and 1.284 m correspond to the leading-edge and the trailing-edge, respectively. All
four turbulence models yield a practically identical pressure distribution over the
upstream half of the hill. However, the predictions depart from one another in the
downstream of the hill crest. The k}e models are shown to signi"cantly overpredict the
pressure recovery in the recirculation region, although the renormalization group
(RNG)-based k}e model and the realizable k}e model slightly reduce the discrepancy.
The best prediction is given by the RSTM. The same ranking applies for the
skin-friction distributions. The RNG k}e and the realizable k}e models are seen to
reproduce the onset of the #ow reversal (x"1.1 m) quite closely. The RSTM again
gives the best result of all. The separation bubble shown in the "gure (top-right in
Fig. 3) is the one predicted by the RSTM in FLUENT.
As indicated in the early part of this paper, most con"gurations of interest in wind
engineering have blu!-body shapes. The implication is that the #ows around build-
ings and structures are subjected to strong nonequilibrium e!ects. The usual sequence
of phenomena that the #ows around blu! bodies go through are; (i) #ow impingement
(stagnation), (ii) rapid acceleration of the #ow in the downstream of the impingement,
(iii) #ow dettachment (separation) near the maximum body width or at sharp edges,
(iv) recirculation in the downstream of the separation point, (v) strong shear layer
between the recirculation region and the main #ow, and (vi) #ow reattachment on the
body surface or in the body wake.
As brie#y discussed earlier, traditional k}e models perform poorly for these situ-
ations. Among other things, the standard k}e model has been found to be inadequate
in dealing with impinging (stagnation) #ows, which has undesirable rami"cations for
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 151

Fig. 3. Flow over the curved two-dimensional hill } predictions using four di!erent turbulence models,
top-left: mesh, top-right: separation bubble, bottom-left: pressure distributions, bottom-right: skin-friction
distributions.

blu!-body aerodynamics. A typical symptom observed when using the standard k}e
model is the spuriously large generation of k (turbulent kinetic energy), which results
in, the framework of eddy-viscosity model, too large turbulent viscosity in the vicinity
of stagnation points. The spuriously large turbulent kinetic energy is convected
downstream, eventually a!ecting the mean #ow, suppressing or reducing #ow separ-
ation in the downstream. Fig. 4 aptly shows what typically happens in the vicinity of
the stagnation point of a blunt body when the standard k}e model is employed. This
con"guration was studied experimentally and numerically by Djilali et al. [16]. The
Reynolds number based on the body height is about Re "50 000. The "lled contours

in Fig. 4 depict the rate of production of turbulent kinetic energy given by the
standard k}e model (left) and the RSTM (right), respectively. It is clearly seen that the
standard k}e yields much larger generation of k than the RSTM that supposedly
predicts the production of k correctly.
It is worthwhile to examine this matter in some more detail. The exact expression
for the production of k includes a contribution from normal straining (*;/*x in this
example) which dominates the budget of k near the stagnation point. It can be written
as
*;
P"!(u!v) , (1)
I *x
152 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

Fig. 4. Production of turbulent kinetic energy over the front face a blunt rectangular section, left: standard
k}e model, right: RSTM.

where P is the production of k due to the normal straining, u and v are the #uctuating
I
velocity components in the x and y directions, respectively.
Note that this term does not require modeling when the RSTM model is employed
because the normal Reynolds stresses are known. However, in k}e models, the normal
stresses needs modeling (using the Boussinesq's hypothesis), and the modeled term
"nally reads

 
*; 
P"4l . (2)
I R *x

As noted, this term always makes a positive contribution to the production of k,


while the exact term indicates otherwise. It should be noted that, as the #ow
approaches the stagnation point, u becomes smaller than v due to the wall damping,
and since the #ow is decelerated (*;/*x(0) the contribution by the normal strains to
the production of k, according to Eq. (1), can become negative near the stagnation
point.
In conclusion, all computations using the isotropic eddy-viscosity-based k}e models
for impinging #ows are inherently prone to overprediction of k and eddy viscosity. As
mentioned earlier, this can lead to reduction or complete suppression of #ow separ-
ation in the downstream. This is shown by the predicted streamlines in Fig. 5 for the
same blunt rectangular body as in Fig. 4. The experiment [16] shows that the #ow
separates at the front edge of the blunt face and reattaches to the wall at 4.7 body
heights. As shown in the "gure, the standard k}e model severely underpredicts the size
of the recirculation, predicting the reattachment point at 2.0 body heights. The RSTM
remedies this problem, predicting the reattachment point quite closely at 4.3 body
heights.
As brie#y mentioned in the foregoing, more elaborate k}e models have been
proposed in recent years. These more recently proposed models all aim at improving
the mediocre performance of the standard k}e model, yet still in the framework of
two-equation model. A legitimate question would be then what all those models can
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 153

Fig. 5. Predicted separation bubbles over a blunt rectangular section, left: standard k}e model, right:
RSTM.

o!er to blu!-body aerodynamics. It seems that this question has been rarely raised,
particularly in the building aerodynamics community. We employed here the RNG
k}e model [8] and the realizable k}e model [10] to see how they work for the blunt
rectangular body. Still adopting the isotropic eddy viscosity, the RNG k}e model and
the realizable k}e model cannot fundamentally resolve the predicament of anomal-
ously large generation of k. But both models have certain features in them which
greatly alleviate the predicament. In the RNG-based k}e model, an additional term in
the e equation augments the production of e in response to the rapid strain in the
neighborhood of the stagnation point, reducing spuriously high level of turbulent
kinetic energy. The realizable k}e model, on the other hand, is based on the idea that
the model constants, which were calibrated for simple equilibrium #ows, need to be
sensitized to the mean #ow (it rate of deformation) to ensure `realizabilitya of the
Reynolds stresses, i.e., positivity of normal Reynolds stresses and Schwarz' inequality
for shear stresses [10].
In Fig. 6 are shown the pressure and skin-friction distributions predicted by the
three di!erent k}e models. It is seen that the RNG k}e model and the realizable k}e
model signi"cantly improve the results. In particular, the realizable k}e model repro-
duces both the pressure and skin friction distributions remarkably well. This example
clearly demonstrates that there is a lot, indeed, that more recently proposed models
can o!er.
The #ows around actual buildings and structures in cities are far more complicated
than those in the examples shown thus far. In addition to all the features of blu!-body
aerodynamics, buildings are immersed in atmospheric boundary layer, being sub-
jected to ground or surfaces e!ects. The interaction between incoming atmospheric
boundary layer and buildings produces a highly complex #ow pattern in the up-
stream, around, and downstream of the body and on the ground surface. Fig. 7 depicts
the computed pressure distribution on the surface of a surface-mounted cube [17]. It
quite closely reproduces the characteristic features of the surface pressure distribution
observed in the experiment.
The #ows around buildings and structures in urban areas are inherently transient.
Besides atmospheric turbulence occurring in the form of gusts, buildings and struc-
tures themselves are responsible for contributing to #ow unsteadiness. In many #ows
around buildings, organized, large-scale motions are observed. One example is the
alternate shedding of vortices behind blu! bodies with a clearly identi"able frequency.
The periodic shedding of vortices is often mainly responsible for #uctuating wind
154 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

Fig. 6. Comparison of three di!erent k}e turbulence models for a blunt rectangular section, left:
C predictions, right: C predictions.
 D

Fig. 7. Surface pressure distribution on a surface-mounted cube and the bottom wall.

loading on buildings and structures. Since the periodic vortex shedding has non-
turbulent origin (yet being a!ected by turbulence), numerical solutions based on the
ensemble-averaged Navier}Stokes equations with proper modeling of turbulence
should be able to reproduce it. Surveying various computational studies [18], how-
ever, indicates that the attempts to numerically simulate vortex shedding have not
been entirely successful.
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 155

Fig. 8. Periodic vortex-shedding in turbulent #ow over a square cylinder left: time history of C , right:
"
contours of instantaneous e!ective viscosity.

Here we present computational results for the #ow around a square cylinder
experimentally investigated by Lyn [19]. The Reynolds number (Re ) is approxim-
"
ately 21 400, based on the freestream velocity, ; , and the width of the square, D. The

experiment shows that the Strouhal number ( fD/;, where f is the shedding frequency)
is in the range of 0.13}0.14. The time-mean drag has not been reported, but Rodi [18]
quotes the range of 2.05}2.23 for the Reynolds number of 22 000.
A quadrilateral mesh with 13 600 cells was used for the computations. A second-
order accurate temporal discretization scheme was used along with a second-order
upwind scheme for convection term. The time step in dimensionless unit (*t; /D)

chosen for the computations is 0.04. The computations were made with the RNG k}e
model. Fig. 8 shows the results. As can be deducted from the time history of the drag
coe$cient (the left "gure in Fig. 8), the Strouhal number predicted by the computation
is approximately 0.135, falling in the range observed in the experiment [19]. The mean
drag is also found to agree fairly well with the values quoted by Rodi [18].
In cities where building groups are commonplace, aerodynamic interferences by
adjacent buildings are of great concern, inasmuch as the interference modi"es the
#ow "eld and eventually the wind loading on individual buildings. The most
common phenomenon observed around closely spaced buildings is the ampli"cation
of wind speed due to blockage e!ects. The concern for the consequences of the wind
ampli"cation e!ects has prompted investigations of the #ows around bodies in
proximity.
Two square cylinders placed side-by-side normal to a uniform stream can be
considered as an idealized case which we can learn from about the interaction of the
#ows around bodies in proximity. This #ow was the subject of the recent experimental
study by Kolar et al. [20]. The Reynolds number (Re ) based on the cylinder width
"
and the freestream velocity is about 23 100, which is quite close to the single cylinder
case [19]. The experiment reveals that, depending on the separation of the cylinders,
the interaction produces di!erent #ow regimes in terms of structure of the vortex
streets. In this paper, we considered the case where two square cylinders are separated
by 2.0 cylinder widths. According to the experimental study [20], the #ow for the
156 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

Fig. 9. Vortex-shedding behind two cylinders placed side-by-side. left: time history of C , right: contours of
"
instantaneous vorticity magnitude.

present Reynolds and the separation between the cylinders exhibits a stable `in-
antiphasea model of vortex-shedding pattern. The experiment also indicates that the
vortex-shedding frequency remains close to that of a single cylinder, although there
are some experimental evidence that indicates otherwise. Again the RNG k}e model
was employed for the computation. Fig. 9 shows the computational results. It was
found that the vortex-shedding frequency for the present double-cylinder case is
roughly 0.15, which is slightly larger than that for the single cylinder.
Despite their e$cacy for the simulation of practical engineering #ows, the phenom-
enological turbulence models based on Reynolds-averaged Navier}Strokes (RANS)
equations have not been completely satisfactory in dealing with blu!-body aerody-
namics. Partly being driven by the dissatisfaction with the RANS-based turbulence
models, yet perhaps more likely encouraged by the explosive increase in computing
power and its a!ordability, CFD practitioners, developers and users alike, have
come to seriously think about high-level simulation of turbulent #ows such as LES.
LES for blu!-body aerodynamics has been recently attempted by many investigators
[21,22].
The governing equations for LES contain terms that represent the e!ects of the
"ltered-out small eddies on the resolved eddies. These terms are called `subgrid scalea
(SGS) stresses, and should be modeled. The most popular and simplest SGS stress
model is the Smagorinsky model. In the Smagorinsky model, the SGS stresses are
related to the resolved scales via eddy-viscosity which is computed in terms of the
resolved velocity "eld (rate-of-strain) and the "lter width, which are all known. The
Smagorinsky model, despite its simple SGS stress modeling, has been successfully
applied to building aerodynamics. Recently, more elaborate SGS models have been
proposed. Particularly, worthy mentioning among others are the SGS eddy-viscosity
model derived by Yakhot and Orszag [23] and the dynamic model proposed by
Germano et al. [24] and its variants.
We present here a preliminary result for a surface-mounted cube [17]. The compu-
tation was made using the RNG-based SGS eddy viscosity model [23] on a 460 000
cell hexahedral mesh locally re"ned near the cube. Fig. 10 shows iso-contour of
S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158 157

Fig. 10. Iso-contours of instantaneous vorticity magnitude for the #ow around a surface-mounted cube
(LES result).

instantaneous vorticity magnitude, colored by static pressure. The characteristic


features of surface-mounted blu!-bodies #ows, including horseshoe vortex and arch
vortex, can be gleaned from the "gure.

4. Closure

The major issues that determine successful application of CFD to building aerody-
namics were discussed. For complex geometries encountered in the modeling of
buildings and structures, unstructured mesh has a great potential to signi"cantly save
time and e!ort for mesh generation. It was demonstrated that locally re"ned meshes
can economically resolve the wide range of length and time scales. Some of the more
recently proposed turbulence models were shown to signi"cantly improve the accu-
racy of numerical solutions for turbulent #ows around blu! bodies. Finally, large eddy
simulation will play an increasingly more important role in the coming years,
especially in dealing with turbulence modeling issues that RANS equations-based
turbulence models cannot resolve.

References

[1] J.M. Weiss, W.A. Smith, Preconditioning applied to variable and constant density #ows, AIAA J. 33
(11) (1995).
[2] S.R. Mathur, J.Y. Murthy, A pressure-based method for unstructured meshes, Numer. Heat Transfer
31 (1997) 195}215.
[3] S.-E. Kim, S.R. Mathur, J.Y. Murthy, D. Choudhury, A Reynolds-averaged Navier}Stokes solver
using an unstructured mesh based "nite-volume scheme, AIAA-98-0231, 1998.
158 S.-E. Kim, F. Boysan / J. Wind Eng. Ind. Aerodyn. 81 (1999) 145}158

[4] D.K. Heist, F.C. Gouldin, Turbulent #ow normal to a triangular cylinder, J. Fluid Mech. 331 (1997)
107}125.
[5] B.E. Launder, D.B. Spalding, The numerical computation of turbulent #ows, Comput. Meth. Appl.
Mech. Eng 3 (1974) 269}289.
[6] D. Lakehal, Application of the k}e model to #ow over a building placed in di!erent roughness
sublayers, J. Wind Eng. Ind. Aerodyn. 49 (1998) 59}77.
[7] S. Murakami, A. Mochida, Y. Hayashi, Scrutinizing k}e EVM and ASM by means of LES and wind
tunnel for #ow"eld around a cube, Eighth Symposium on Turbulent Shear Flows, 17-1-1, 1991.
[8] V. Yakhot, S.A. Orazang, S. Thangam, T.B. Gatski, C.G. Speziale, Development of turbulence models
for shear #ows by a double expansion technique, Phys. Fluids A 4 (7) (1992) 1510}1520.
[9] D.C. Wilcox, Comparison of two-equationc turbulence models for boundary layers in pressure
gradient, AIAA J. 31 (8) (1993) 1414}1424.
[10] T.-H. Shih, W.W. Liou, A. Shabbir, J. Zhu, A new k}e eddy-viscosity model for high Reynolds number
turbulent #ows } model development and validation, Comput. Fluids 24 (3) 227}238.
[11] T.J. Craft, B.E. Launder, K. Suga, Extending the applicability of eddy viscosity models through the
use of deformation invariants and non-linear elements, Proc. Fifth International Symposium on
Re"ned Flow Modeling and Turbulence Measurements, Paris, 1993, pp. 125}132.
[12] T.B. Gatski, C.G. Speziale, On explicit algebraic stress models for complex turbulent #ows, J. Fluid
Mech. 254 (1993) 47}83.
[13] T.-H. Shih, J. Zhu, J.L. Lumely, A new Reynolds stress algebraic equation model, NASA Technical
Memorandum 106644, ICOMP-94-15; GMOTT-94-8, 1994.
[14] V. Baskaran, A.J. Smits, P.N. Joubert, A turbulent #ow over a curver hill } Part 1. Growth of an
internal boundary layer, J. Fluid Mech. 182 (1987) 47}83.
[15] M.M. Gibson, B.E. Launder, Ground e!ects on pressure #uctuations in the atmospheric boundary
layer, J. Fluid Mech. 86 (1978) 491}511.
[16] N. Djilali, I.S. Gartshore, M. Salcudean, An experimental and numerical study of the #ow around
a blunt rectangular section } a test case for computational method, Sixth Symposium on Turbulent
Shear Flows, 19-3-1, Toulouse, France, 1987.
[17] I.P. Castro, A.G. Robins, The #ow around a surface-mounted cube in uniform and turbulent streams,
J. Fluid Mech. 79 (Part 2) (1977) 307}335.
[18] W. Rodi, Experience with two-layer models combining the k}e model with a one-equation model near
the wall, AIAA Paper 91-0216, 1991.
[19] D.A. Lyn, Ensemble-averaged measurements in the turbulent wake of a square cylinder: a guide to the
data, Report No. CE-HSE-92-6, School of Civil Engineering, July 1992.
[20] V. Kolar, D.A. Lyn, W. Rodi, Ensemble-averaged measurements in the turbulent near wake of two
side-by-side square cylinders, J. Fluid Mech. 346 (1997) 201}238.
[21] S. Murakami, Blu! body aerodynamics and turbulence, J. Wind Eng. Ind. Aerodyn. 73 (1993) 65}78.
[22] T. Kogaki, T. Kobayashi, N. Taniguchi, Large eddy simulation of #ow around a rectangular cylinder,
Fluid Dyn. Res. 20 (1997) 11}24.
[23] V. Yakhot, S. Orszag, Renormalization group analysis of turbulence: I basic theory, J. Sci. Comput.
1 (1986) 1}64.
[24] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, A dynamic subgrid scale eddy viscosity model, Phys.
Fluids A 3 (1991) 1760.
[25] A.K. Mikkelsel, F.M. Livesey, Evaluation of the use of numerical k}e model Kameleon II, for
predicting wind pressure on building surfaces, J. Wind Eng. Ind. Aerodyn. 57 (1995) 375}389.

You might also like