Sie sind auf Seite 1von 15

ARTICLE IN PRESS

Atmospheric Environment 41 (2007) 238–252


www.elsevier.com/locate/atmosenv

CFD simulation of the atmospheric boundary layer:


wall function problems
Bert Blockena,, Ted Stathopoulosb, Jan Carmelieta,c
a
Building Physics and Systems, Technische Universiteit Eindhoven, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
b
Department of Building, Civil and Environmental Engineering, Concordia University, 1455 de Maisonneuve Blvd West,
Montreal, Que., Canada, H3G 1M8
c
Laboratory of Building Physics, Department of Civil Engineering, Katholieke Universiteit Leuven, Kasteelpark Arenberg 40,
3001 Leuven, Belgium
Received 17 May 2006; received in revised form 14 August 2006; accepted 15 August 2006

Abstract

Accurate Computational Fluid Dynamics (CFD) simulations of atmospheric boundary layer (ABL) flow are essential
for a wide variety of atmospheric studies including pollutant dispersion and deposition. The accuracy of such simulations
can be seriously compromised when wall-function roughness modifications based on experimental data for sand-grain
roughened pipes and channels are applied at the bottom of the computational domain. This type of roughness modification
is currently present in many CFD codes including Fluent 6.2 and Ansys CFX 10.0, previously called CFX-5. The problems
typically manifest themselves as unintended streamwise gradients in the vertical mean wind speed and turbulence profiles
as they travel through the computational domain. These gradients can be held responsible—at least partly—for the
discrepancies that are sometimes found between seemingly identical CFD simulations performed with different CFD codes
and between CFD simulations and measurements. This paper discusses the problem by focusing on the simulation of a
neutrally stratified, fully developed, horizontally homogeneous ABL over uniformly rough, flat terrain. The problem and
its negative consequences are discussed and suggestions to improve the CFD simulations are made.
r 2006 Elsevier Ltd. All rights reserved.

Keywords: Computational Fluid Dynamics (CFD); Numerical simulation; Atmospheric boundary layer (ABL); Sustainable boundary
layer; Equilibrium vertical profiles; Horizontal homogeneity

1. Introduction sion and deposition, wind-driven rain, building


ventilation, etc. Recently, comprehensive literature
Computational Fluid Dynamics (CFD) is increas- reviews on the use of CFD for these applications
ingly being used to study a wide variety of processes have been published (Stathopoulos, 1997; Reichrath
in the lower parts of the atmospheric boundary and Davies, 2002; Blocken and Carmeliet, 2004;
layer (ABL) (0–200 m) including pollutant disper- Bitsuamlak et al., 2004; Meroney, 2004; Franke
et al., 2004).
Corresponding author. Tel.: +31 40 247 2138; Accurate simulation of ABL flow in the compu-
fax: +31 40 243 8595. tational domain is imperative to obtain accurate
E-mail address: b.j.e.blocken@tue.nl (B. Blocken). and reliable predictions of the related atmospheric

1352-2310/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.atmosenv.2006.08.019
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 239

Nomenclature u+ dimensionless mean streamwise wind


speed
B integration constant in the log law U mean streamwise wind speed, m s1
CS roughness constant x, y streamwise and height co-ordinates, m
Ce1, Ce2, Cm constants in the k–e model y0 aerodynamic roughness length, m
e inhomogeneity error, % yP distance from point P to the wall, m
E empirical constant for a smooth wall in y+ dimensionless wall unit
wall function (E9.793) e turbulence dissipation rate, m2 s3
k turbulent kinetic energy, m2 s2 k von Karman constant (0.40–0.42)
kS equivalent sand-grain roughness height, m dynamic molecular viscosity, kgm1 s1
m mt dynamic turbulent viscosity, kgm1 s1
k+S dimensionless equivalent sand-grain n kinematic molecular viscosity, m2 s1
roughness height r fluid density, kg m3
L, H length and height of computational se constant in the k–e model
domain, m tw wall shear stress, Pa
P centre point of wall-adjacent cell j flow variable
TI turbulence intensity o specific dissipation rate, s1
ut wall-function friction velocity, m s1 DB roughness function
u wall-function friction velocity, m s1 Dx control volume length, m
uABL ABL friction velocity, m s1

processes. In a CFD simulation, the flow profiles of ness height for the ABL, kS,ABL, which is typically
mean wind speed and turbulence quantities that are quite high (large-scale roughness; e.g. y0 in the range
applied at the inlet plane of the computational 0.03–2 m, (Wieringa, 1992), kS,ABL in the range
domain are generally fully developed, equilibrium 0.9–60 m). Note that in CFD simulations, often the
profiles. These profiles should be representative of upstream part of the domain and the terrain outside
the roughness characteristics of that part of the the domain upstream of the inlet plane are assumed
upstream terrain that is not included in the to be of the same roughness, implying that it is not
computational domain (i.e., the terrain upstream the intention to simulate the development of an
of the inlet plane). This is expressed by the presence internal boundary layer (IBL) starting from the inlet
of either the appropriate aerodynamic roughness plane. In the centre of the computational domain,
length y0 or the appropriate power-law exponent a where the actual obstacles are modelled explicitly,
of this terrain in the expressions of the inlet profiles additional roughness modelling is limited to the
(Davenport, 1960, 1961; Wieringa, 1992). Within surfaces of the obstacles themselves (walls, roofs,
the computational domain, generally three different etc.) and the surfaces between these obstacles
regions can be distinguished, as illustrated in Fig. 1: (streets, grass plains, etc.). This is often also done
(1) the central region of the domain where the actual with wall functions. The roughness of these surfaces
obstacles (buildings, trees, stacks, etc.) are modelled is most often expressed in terms of the roughness
explicitly with their geometrical shape; and (2) the height kS that is typically quite small (small-scale
upstream and downstream region of the domain roughness; e.g. kS in the range 0–0.1 m).
where the actual obstacles are modelled implicitly, The simulation of a horizontally homogeneous
i.e., their geometry is not included in the domain but ABL over uniformly rough terrain is often required
their effect on the flow can be modelled in terms of in the upstream and the downstream region of the
roughness, e.g., by means of wall functions applied computational domain. The term ‘‘horizontally
to the bottom of the domain. These wall functions homogeneous’’ refers to the absence of streamwise
replace the actual roughness obstacles but they gradients in the vertical profiles of the mean wind
should have the same overall effect on the flow as speed and turbulence quantities, i.e. these profiles
these obstacles. This roughness is expressed in terms are maintained with downstream distance. This flow
of the aerodynamic roughness length y0 or, less type occurs when the vertical mean wind speed and
often, in terms of the equivalent sand-grain rough- turbulence profiles are in equilibrium with the
ARTICLE IN PRESS
240 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

outlet plane

incident flow

approach flow
downstream part of
computational
inlet flow domain

central part of
computational
domain

upstream part of
computational
inlet plane domain

Fig. 1. Computational domain with building models for CFD simulation of ABL flow—definition of inlet flow, approach flow and
incident flow and indication of different parts in the domain for roughness modelling.

roughness characteristics of the ground surface. empty computational domain, especially for the
Concerning the upstream part of the domain, a turbulent kinetic energy. Problems in simulating a
distinction is made between inlet flow, approach horizontally homogeneous ABL flow were also
flow and incident flow (Fig. 1). The ‘‘approach reported by Miles and Westbury (2003), using
flow’’ profiles are those travelling towards the CFX-5, and by Franke et al. (2004), Franke and
building models, while the ‘‘incident flow’’ profiles Frank (2005) and Blocken and Carmeliet (2006)
are those obtained in a similar but empty computa- using Fluent 5 and 6.
tional domain, at the position where the buildings The unintended differences between inlet profiles
would be positioned. Horizontal homogeneity im- and incident profiles (i.e. the horizontal homogene-
plies that the inlet profiles, the approach flow ity problem) can be detrimental for the success of
profiles and the incident profiles are the same. CFD simulations given that even minor changes to
In the past, several authors have reported the incident flow profiles can cause significant
difficulties in simulating a horizontally homoge- changes in the flow field. Indeed, sensitivity studies
neous ABL flow in at least the upstream part of by Castro and Robins (1977), Miles and Westbury
computational domains. Richards and Younis (2003), Gao and Chow (2005) and Blocken et al.
(1990), discussing the work of Mathews (1987), (2006) have indicated the important influence of the
referred to a situation in which the approach flow shape of the vertical incident flow profiles on the
changed rapidly in the upstream region of the simulation results of flow around buildings.
computational domain. A particular observation Furthermore, the considerable problems in simulat-
was the considerable acceleration of the flow near ing the simple case of a horizontally homogeneous
the surface. Zhang (1994), using the k–e model and ABL flow suggest that similar or maybe even more
the standard wall functions (Launder and Spalding, serious problems can be expected when more
1974) without roughness modification, reported an complex cases of ABL flow have to be simulated,
unwanted change in the profiles of mean wind speed e.g. the development of IBLs over terrains with
and especially turbulent kinetic energy, which he roughness changes.
suggested to be responsible for some of the This paper addresses the problem of horizontal
discrepancies found between the CFD simulations homogeneity associated with the use of sand-grain
and the corresponding wind tunnel measurements. roughness wall functions. This is done by focusing
A similar problem for turbulent kinetic energy was on the CFD simulation of a neutrally stratified,
reported by Quinn et al. (2001) who used the k–e horizontally homogeneous ABL flow over uni-
model in CFX-4.1. Riddle et al. (2004), employing formly rough, flat terrain. The reasons for the
Fluent 6 with the k–e and the Reynolds stress model difficulties possibly encountered are explained, the
(RSM), observed significant profile changes in an negative consequences involved are discussed and
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 241

suggestions to handle them are made. First, in about the ground roughness (roughness of the
Section 2, the basic requirements for a CFD bottom of the computational domain) into the
simulation of ABL flow with sand-grain wall simulation to prevent streamwise gradients in
functions are set. Section 3 describes the commonly the flow in the upstream and downstream part of
used, fully developed ABL inlet profiles for mean the domain, i.e. outside the main disturbance of the
wind speed, turbulent kinetic energy and turbulence flow field by the explicitly modelled obstacles
dissipation rate. In Section 4, the so-called sand- (Richards and Hoxey, 1993). This generally requires
grain roughness wall-function modification is briefly the use of wall functions. The third requirement
described. Section 5 points to the inconsistency of implies that it is not physically meaningful to have
the basic requirements for ABL flow simulation grid cells with centre points within the physical
with these wall functions. In Section 6, the typical roughness height. This requirement is explicitly
negative consequences of this inconsistency are mentioned by several commercial CFD codes
discussed. Section 7 summarizes various remedial including Fluent 6.2 (Fluent Inc., 2005) and Ansys
measures. Finally, Section 8 concludes the paper. CFX 10.0 (Ansys Ltd., 2005). Both codes warn the
user to abide by the requirement yP4kS. In
2. Basic requirements for ABL flow simulation addition, Ansys Ltd. (2005) mentions that violation
of this requirement can lead to inaccuracies and to
In almost all CFD simulations of the lower part solver failure but it does not elaborate further on
of the ABL, an accurate description of the flow near this issue. The fourth requirement concerns a
the ground surface is required. In such cases, if the relationship that results from matching the ABL
wall roughness is expressed by an equivalent sand- mean velocity profile and the wall function in the
grain roughness kS in the wall functions, four CFD code and will be discussed later.
requirements should be simultaneously satisfied. All four requirements should be satisfied in the
This set of requirements has been distilled from upstream and downstream region of the computa-
various sources including CFD literature and CFD tional domain, while in the central part, only
software manuals (Richards and Hoxey, 1993; requirements (1) and (3) must be adhered to.
Franke et al., 2004; Fluent Inc., 2005; Ansys Ltd., However, it is generally impossible to satisfy all
2005): four requirements. This paper focuses on the
standard k–e model by Jones and Launder (1972)
(1) A sufficiently high mesh resolution in the used in combination with the standard wall func-
vertical direction close to the bottom of the tions by Launder and Spalding (1974). Note
computational domain (e.g. height of first cell however that the validity of the findings and the
o1 m); statements made in the paper is not limited to this
(2) A horizontally homogeneous ABL flow in the type of turbulence model and these wall functions.
upstream and downstream region of the do-
main; 3. Fully developed ABL profiles
(3) A distance yP from the centre point P of the
wall-adjacent cell to the wall (bottom of For the k–e model, Richards (1989) proposed
domain) that is larger than the physical rough- vertical profiles for the mean wind speed U,
ness height kS of the terrain (yP4kS); and turbulent kinetic energy k and turbulence dissipa-
(4) Knowing the relationship between the equiva- tion rate e in the ABL that are based on the Harris
lent sand-grain roughness height kS and the and Deaves (1981) model. Because the height of the
corresponding aerodynamic roughness length computational domain is often significantly lower
y0. than the ABL height, these profiles are generally
simplified by assuming a constant shear stress with
The first requirement is important for all compu- height (Richards and Hoxey, 1993):
tational studies of flow near the surface of the  
uABL y þ y0
Earth. For instance, for pedestrian wind comfort UðyÞ ¼ ln , (1)
studies, Franke et al. (2004) state that at least 2 or 3 k y0
control volume layers should be provided below
u2
pedestrian height (1.75 m). The second requirement kðyÞ ¼ pABL
ffiffiffiffiffiffi , (2)
implies the insertion of (empirical) information Cm
ARTICLE IN PRESS
242 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

u3
ABL the RSM, etc. In these cases, the profiles for k and e
ðyÞ ¼ , (3) are converted into profiles for either the specific
kðy þ y0 Þ
dissipation rate o (o ¼ =C m k), the turbulent
where y is the height co-ordinate, uABL the ABL viscosity ratio mt/m or the Reynolds stresses.
friction velocity, k the von Karman constant
(E0.40–0.42) and Cm a model constant of the
standard k–e model. It can be easily shown that 4. Standard wall functions with a sand-grain-based
Eqs. (1)–(3) are an analytical solution to the roughness modification
standard k–e model if the model constants Ce1,
Ce2, se, Cm are chosen in such a way that Eq. (4) is Due to the importance of the surface roughness
satisfied (Richards and Hoxey, 1993): and the high Reynolds numbers associated with
pffiffiffiffiffiffi ABL flow, the use of wall functions is generally
k2 ¼ ðC 2  C 1 Þs C m . (4)
required for near-wall modelling. The wall functions
Similarly, it can be shown that the set of Eqs. (2) in CFD codes are generally based on the universal
and (5)–(6) are also an analytical solution to the near-wall velocity distribution (law of the wall) that
same model under the same condition (Durbin and can be modified for the effects of rough surfaces. In
Petterson Reif, 2001): this section, the often used roughness modification
  that is based on experiments with sand-grain
uABL y
UðyÞ ¼ ln , (5) roughness is described.
k y0

u3
ABL
ðyÞ ¼ . (6) 4.1. Law of the wall for fully rough surfaces
ky
These profiles are commonly used as inlet profiles The universal law of the wall for a smooth surface
for CFD simulations when measured profiles of U is plotted in Fig. 2 (dashed line) using the
and k are not available. It should be noted that dimensionless variables uþ ¼ U=u and yþ ¼
these profiles are not only used for simulations with u y=n, where U is the mean velocity tangential to
the standard k–e model but also with other types of the wall, u is a wall-function friction velocity and n
turbulence models: RNG k–e, realizable k–e, stan- is the kinematic viscosity. Note that u can be
dard k–o, SST k–o, the Spalart–Allmaras model, different from uABL .

30
ΔB = 0
+ <2
.25
25 +) kS
(k s
-ΔB ΔB =7.4
+ +B
20 1 lny += 9
0
+ = κ kS ΔB =10.3
u
u+ = y+ e
+ =3
00
gim Δ B = 13.1
u+

e
15 al r kS 0
tion + =1
00 ΔB = 15.8
nsi
Tra kS
00
0
10 e
im + = 3 0 ΔB = 18.6
re g k S 000
gh + =1
r ou kS
ll y
5 Fu

0
1 10 100 1000 10000
y+
linear buffer logarithmic
sublayer layer layer

y+=5 y+=30

Fig. 2. Law of the wall for smooth and sand-grain roughened surfaces with the dimensionless sand-grain roughness height kþ
S as a
parameter.
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 243

The near-wall region consists of three main parts: This is the logarithmic law of the wall for fully
the laminar layer or linear sublayer, the buffer layer rough surfaces based on sand-grain roughness. Eq.
and the logarithmic layer. In the linear sublayer, the (9) is illustrated in Fig. 2 with kþ
S as a parameter.
laminar law holds (u+ ¼ y+) while in the log layer,
the logarithmic law is valid: (u+ ¼ ln(y+)/k+B) 4.2. Wall functions for fully rough surfaces
where the integration constant BE5.0–5.4 (e.g.
Schlichting, 1968; White, 1991). The laminar law In this paper, the term ‘‘sand-grain-roughness wall
is valid below about y+ ¼ 5 and the logarithmic law functions’’ refers to standard wall functions modified
above about y+ ¼ 30 up to y+ ¼ 500–1000. The for roughness based on experiments with sand-grain
modification of the log law for rough surfaces is roughness, also called kS-type wall functions. The kS-
mainly based on the extensive experiments by type wall function for mean velocity is obtained by
Nikuradse (1933) for flow in rough, circular pipes replacing U and y in Eq. (9) by their values in the
that were covered on the inside as tightly as possible centre point P of the wall-adjacent cell: UP and yP.
with sand grains (sand-grain roughness kS). The Several commercial CFD codes use slightly different
experiments indicated that the mean velocity kS-type wall functions than Eq. (9).
distribution near rough walls, when plotted in a
semi-logarithmic scale, as in Fig. 2, has the same 4.2.1. Fully rough kS-type wall function for mean
slope (1/k) but a different intercept. The shift of the velocity in Fluent 6.2
intercept, DB, as shown in Fig. 2, is a function of the The wall function in Fluent 6.2 is given by (Fluent
dimensionless sand-grain roughness height kþ S ¼ Inc., 2005):
u kS =n; also called ‘‘dimensionless physical rough-  
U P u 1 Eu yP
ness height’’ or ‘‘roughness Reynolds number’’. The ¼ ln , (10)
logarithmic law for a rough wall is (Cebeci and u2t k nð1 þ C S kþ

Bradshaw, 1977): where the factor (1 þ C S kþ S ) represents the rough-
   ness modification, E the empirical constant for a
U 1 u y 1=2

¼ ln þ B  DBðkþs Þ. (7) smooth wall (E9.793) and u ¼ C 1=4 m kP and
u k n 1/2
ut ¼ (tw/r) are two different wall-function fric-
The roughness function DB takes different forms tion velocities. kP is the turbulent kinetic energy in
depending on the kþ S value. Three regimes are the centre point P, tw is the wall shear stress and r
distinguished: aerodynamically smooth (kþ S o2:25), the fluid density. CS, the roughness constant, is an
transitional (2:25pkþ S o90) and fully rough attempt to take into account the type of roughness.
(kþS X90). ABL flow over rough terrain classifies as However, due to the lack of specific guidelines, it is
fully rough because the roughness elements (ob- generally set at its default value for sand-grain
stacles) are so large that the laminar sublayer is roughened pipes and channels: 0.5. The user inputs
eliminated and the flow is considered to be in the code are the values kS and CS, in Fluent 6.1
independent of the molecular viscosity. Note that and 6.2 with the restriction that CS should lie in the
this is the case for flow in the upstream and interval [0;1]. For an equilibrium boundary layer
downstream part of the computational domain but (u ¼ ut ) and when C S kþ S b1 (fully rough regime
not necessarily for the flow over the explicitly with about CS40.2), Eq. (10) can be simplified and
modelled surfaces with a small-scale roughness in takes a form similar to Eq. (9):
the central part of the domain. For the fully rough   
UP 1 u yP
regime, Cebeci and Bradshaw (1977) report the ¼ ln þ 5:43. (11)
u  k nC S kþ
S
following analytic fit to the sand-grain roughness
data of Nikuradse (1933), which was originally
provided by Ioselevich and Pilipenko (1974): 4.2.2. Fully rough kS-type wall function for mean
velocity in Ansys CFX 10.0
1
DB ¼ lnðkþ
S Þ  3:3. (8) Ansys CFX 10.0 provides a similar wall function,
k
however with a fixed value for the roughness
Combining Eqs. (7) and (8), with B ¼ 5:2, yields: constant: C S ¼ 0:3 (Ansys Ltd., 2005):
    
U 1 u y UP 1 u yP
¼ ln þ 8:5. (9) ¼ ln þ 5:2. (12)
u  k nkþ
S u k nð1 þ 0:3kþ

ARTICLE IN PRESS
244 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

In the fully rough regime (kþS 490) it can be


rewritten as follows:
  
UP 1 u yP ABL log law
¼ ln þ 5:2. (13)
u k 0:3nkþS

P UP
4.2.3. Wall functions for k and e
Irrespective of the value of kþ
S , the wall functions wall function
yP
for the turbulent quantities are generally given by
u2
kP ¼ pffiffiffiffiffiffi , (14)
Cm Fig. 3. Graphical representation of fitting the mean-velocity ABL
3 log-law inlet profile to the wall function for mean velocity in the
u centre point P of the wall-adjacent cell.
P ¼ . (15)
kyP
Note that in these equations u carries the effect of the
depends to some extent on the value of k
roughness. In Fluent 6, Eq. (14) is not used but instead
(0.40–0.42). In all cases, kS,ABL is clearly much
the k equation is solved in the wall-adjacent cells.
larger than the corresponding aerodynamic rough-
ness length y0. As an example, for some terrain
kS-type wall functions are not only used for types in the updated Davenport roughness classifi-
simulations of flow over sand-grain roughened cation, the following values are obtained: for rough
surfaces. Indeed, many CFD codes including Fluent open terrain: y0 ¼ 0:1 m, kS,ABLE3 m; for very
6.2 and Ansys CFX 10.0 only provide this type of rough terrain: y0 ¼ 0:5 m, kS,ABLE15 m; for city
roughness modification. Consequently, CFD simu- centres: y0 ¼ 2 m, kS,ABLE60 m. Clearly, kS,ABL will
lations over other types of rough surfaces are also often be very large in CFD simulations in built
made with kS-type wall functions. In that case, the environments. Note that a perfect match between
actual roughness is characterized by an ‘‘equiva- Eq. (1) and Eqs. (9), (11), (13) cannot be achieved
lent’’ sand-grain roughness height kS. since in this case uABL is different from u .
To satisfy all four requirements mentioned in
5. Inconsistency in the requirements for ABL flow Section 2 is generally impossible with the kS-type wall
simulation functions outlined above. The main reason is that the
fourth requirement, expressed by Eq. (16), (17) or
The fourth requirement for ABL flow simulation (18), in combination with the third requirement
mentioned in Section 2 concerns the relationship (yP4kS,ABL) implies that very large (high) control
between kS and y0. It provides the equivalent sand- volumes should be used, which is in conflict with the
grain roughness height for the ABL, kS,ABL. It can be first requirement (high mesh resolution, hence small
derived by first-order matching (continuity of function yP). Note that in Fluent, the required cell height can
and its first derivative) of the ABL velocity profile (Eq. be limited to some extent by maximizing CS (C S ¼ 1;
(5)) with the wall-function velocity profile in the centre see Eq. (17)) but that this will often not be satisfactory.
point P of the wall-adjacent cell, as indicated in Fig. 3. The discussion in the remainder of this paper will
focus on—but not be limited to—the relationship
For Eqs. (9), (11), (13) a perfect match with Eq. (5) kS,ABLE30y0. Then, as an example, for a grass-
can be obtained, yielding, respectively, covered plain with a low aerodynamic roughness
length y0 ¼ 0:03 m, kS,ABL is about 0.9 m and yP
kS;ABL ¼ 30y0 , (16) should be at least equal to this value, yielding cells of
minimum 1.8 m height. For larger values of y0, much
9:793y0 larger (higher) cells are needed. It is clear that this
kS;ABL ¼ ðFluentÞ, (17)
Cs requirement conflicts with the need for a high grid
resolution near the bottom of the computational
kS;ABL ¼ 29:6y0 ðAnsys CFXÞ (18)
domain and that no accurate solutions for near-
and uABL 
¼ u for all three cases. In Eqs. (16)–(18), ground flow can be obtained with cell sizes so large—
the actual value of the constants (30, 9.793 and 29.6) see also Franke et al. (2004).
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 245

6. Discussion are taken equal to Eqs. (2), (5), (6) with y0 ¼ 0:1 m
and uABL ¼ 0:912 m s1 . At the bottom of the
A typical consequence of not adhering to all four domain, the standard wall functions (Eqs. (10),
ABL flow requirements is the occurrence of unin- (15)) in the code are used, with kS,ground ¼ 0.24 m
tended streamwise gradients in the vertical profiles (as large as possible while satisfying kS,groundoyP)
of the mean wind speed and turbulence quantities and with the default roughness constant C S ¼ 0:5.
(horizontal inhomogeneity) as the flow travels Note that kS,groundokS,ABL ¼ 1.959 m. As also
through the computational domain. The extent of indicated by Richards and Hoxey (1993), specific
the streamwise gradients depends on the shape of attention is needed for the boundary condition at
the vertical inlet profiles, the downstream flow the top of the domain. Along the length of this top
distance, the turbulence model, the type of wall boundary, the values from the inlet profiles of U, k
function, the grid resolution (yP), the roughness and e at this height are imposed (U ¼ 18:5 m s1 ,
height (kS), the roughness constant (CS) and the k ¼ 2:77 m2 s2 ,  ¼ 0:0036 m2 s3 ). This is done by
boundary conditions at the top and outlet of the fixing these constant values in the top layer of cells
domain. A scenario that can explain—at least in the domain. The application of this particular
partly—the problems observed in some previous type of top boundary condition is important
studies is discussed below. In the scenario the ABL because other top boundary conditions (symmetry,
flow profiles at the inlet of the domain are Eqs. slip wall, etc) can themselves cause streamwise
(1)–(3) or Eqs. (2), (5), (6), which is common gradients, in addition to those caused by the wall
practice in CFD simulations of ABL flow. functions. At the outlet, an ‘‘outflow’’ boundary is
Given the requirement yP4kS, the most straight- used, which assumes no streamwise gradients at this
forward choice might be to apply the required high location. The 2D Reynolds-averaged Navier–Stokes
mesh resolution near the bottom of the domain and (RANS) equations and the continuity equation are
to insert a kS value (kS,ground) that is low enough to solved using the control volume method. Closure is
satisfy kS,groundoyP. Generally, this value will be obtained using the standard k–e model. Pressure–
significantly lower than that required for ABL flow velocity coupling is taken care of by the SIMPLE
simulation (kS,ABLE30y0); e.g. sometimes kS ¼ y0 algorithm. Pressure interpolation is second order.
has been used. In this case, the change in roughness Second-order discretization schemes are used for
between the inlet profile that is representative of the both the convection terms and the viscous terms of
terrain upstream of the domain inlet (corresponding the governing equations. Fig. 4 illustrates the
to kS,inlet ¼ kS,ABLE30y0) and the actual smaller simulation results. The figures on the left show the
ground roughness in the upstream part of the profiles from ground level up to 500 m height, while
computational domain (kS,groundo30y0) will intro- the figures on the right only show the lowest 50 m.
duce an IBL, in which the mean wind speed and As the profiles travel downstream, the streamwise
turbulence profiles will rapidly adapt to the new and gradients become very pronounced, particularly
smaller roughness, yielding, amongst others, a near the ground surface and they exhibit the typical
considerable acceleration of the flow near the characteristics of a developing IBL. After a con-
surface. This can explain—at least partly—the siderable distance (about x ¼ 5000 m, not shown in
unintended streamwise gradients observed by sev- figure) the profiles attain a new equilibrium with the
eral authors mentioned in the Introduction. current simulation parameters (i.e. turbulence mod-
To illustrate this scenario, CFD simulations of el, wall functions, values of yP, kS, CS and top and
ABL flow were performed in a 2D, empty computa- outlet boundary condition). Note that changes in
tional domain with Fluent 6.1.22. Note that a 2D these parameters will lead to similar observations
simulation does not truly represent 3D turbulence but to different profiles. Fig. 5 illustrates the relative
but that it serves the purpose of illustrating the change, i.e. inhomogeneity error e relative to inlet
problem using the standard k–e model and econom- profile, for each of the variables U, k, e and TI, as a
ically evaluating possible remedial measures. The function of downstream distance at two heights
domain has dimensions L  H ¼ 10,000  500 m. A (y ¼ 2 and 20 m):
structured mesh was generated based on grid-
 
sensitivity analysis, with yP ¼ 0:25 m and a total of j  j 
 ðxÞ ðx¼0Þ 
46,000 cells, equidistantly spaced in the horizontal e ¼ 100 , (19)
 jðx¼0Þ 
direction (cell length Dx ¼ 10 m). The inlet profiles
ARTICLE IN PRESS
246 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

500 50
inlet inlet
x=100m x=100m
400 x=500m
40 x=500m
x=1000m x=1000m
300 x=10000m 30 x=10000m

y (m)
y (m)
200 20

100 10

0 0
8 12 16 20 2 4 6 8 10 12 14 16
(a) U (m/s) U (m/s)

500 50
inlet inlet
x=100m x=100m
400 x=500m 40 x=500m
x=1000m x=1000m
x=10000m x=10000m
300 30

y (m)
y (m)

200 20

100 10

0 0
0 1 2 3 4 0 1 2 3 4
(b) k (m2/s2) k (m2/s2)

500 50
inlet inlet
x=100m x=100m
400 x=500m 40 x=500m
x=1000m x=1000m
300 x=10000m 30 x=10000m
y (m)
y (m)

200 20

100 10

0 0
0 0.1 0.2 0.3 0.4 0 0.2 0.4 0.6 0.8 1
(c) ε (m2/s3) ε (m2/s3)

500 50
inlet inlet
400 x=100m 40 x=100m
x=500m x=500m
x=1000m x=1000m
300 30
y (m)

x=10000m
y (m)

x=10000m

200 20

100 10

0 0
0.05 0.1 0.15 0.2 0.05 0.15 0.25 0.35 0.45
(d) TI TI

Fig. 4. CFD simulation results illustrating the streamwise gradients in the vertical profiles of (a) mean wind speed U; (b) turbulent kinetic
energy k; (c) turbulence dissipation rate e and (d) turbulence intensity TI at different downstream distances in the empty domain (x co-
ordinate). The left column shows the lowest 500 m of the boundary layer; the right column shows the lowest 50 m.
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 247

60 60
y =2 m y =20 m
50 U 50 U
k k
40 ε 40 ε

error (%)
error (%) TI TI
30 30

20 20

10 10

0 0
1 10 100 1000 10000 1 10 100 1000 10000
(a) x (m) (b) x (m)

Fig. 5. CFD simulation results: relative changes (inhomogeneity errors) of the values of mean wind speed U, turbulent kinetic energy k,
turbulence dissipation rate e and turbulence intensity TI at two heights (y ¼ 2 and 20 m) and at various downstream positions in the empty
domain (x co-ordinate). All changes are expressed as positive percentage values relative to the values at the inlet of the domain (x ¼ 0).

where j represents one of the flow variables. The central region of the domain (small-scale roughness,
errors in all variables reach (very) high values with small kS); e.g. Fig. 6. The construction of such
increasing downstream distance. The error curve for meshes is a feasible option when the approach flow
e at y ¼ 2 m shows an overshoot around 100 m. This corresponds to smooth terrain (e.g. rural area,
overshoot is only present near ground level (below y0 ¼ 0:03 m). For rough terrain however (e.g.
3 m). It can be attributed to the combined effect of urban, y0 ¼ 1 m), Franke et al. (2004) correctly
the transition from the inlet profile to a new state that reduction of the cell height from 60 m
equilibrium profile, the vertical discretization of upstream of the model (needed for an inlet flow with
the inlet profiles—to be explained in the next y0 ¼ 1 m) to generally less than 0.5 m close to the
section—and the very steep gradient of the e profile building model is not evident, even for unstructured
near the ground surface. Given the sensitivity of meshes. Nevertheless, with this option, a good
CFD simulation results to the incident flow profiles, horizontal homogeneity can be obtained. An
the observed gradients can be detrimental for the additional drawback however is the coarse
accuracy of CFD simulations of atmospheric near-wall mesh distribution at the inlet. This is
processes in the ABL. important because it determines how the analytical,
continuous inlet profiles imposed at the inlet are
7. Remedial measures and other wall functions discretized for input into the simulation. This
discretization occurs because each cell only contains
7.1. Remedial measures for kS-type wall functions one value of each flow variable (in a cell-centred
with kSoyP scheme). In case of high near-wall cells, the
continuous inlet profiles will be converted into
Various remedial measures to rectify or at least rough discrete profiles. Fig. 7 illustrates this
address the errors discussed in the previous section, problem. Figs. 7a and b display two near-wall mesh
are presented and discussed here. Some have been distributions for the simulation case outlined in
suggested previously, others are new. None of them Section 6 with y0 ¼ 0:1 m and kS,ABL ¼ 1.959 m
however can be considered totally satisfactory. (Eq. (17) with C S ¼ 0:5). The fine near-wall mesh
(a) Variable height of wall-adjacent cells: Franke distribution (with yP ¼ 0:25 m) is the one used in the
et al. (2004) have mentioned abandoning the usual previous simulation. The coarse near-wall mesh
approach of using a constant vertical height of all distribution (with yP ¼ 2 m) satisfies the require-
the cells adjacent to the bottom of the computa- ment yP4kS,ABL, by having wall-adjacent cells
tional domain (first layer of cells near the ground with a height of 4 m. Fig. 7c and d show the
surface). Instead different initial vertical cell heights analytical inlet profiles together with the
can be used in different areas of the domain to corresponding discrete profiles of U and TI, all at
satisfy Eq. (16), (17) or (18) at every position in the the domain inlet. The discrepancies that are
domain. This implies using higher cells in the introduced in case of the coarse mesh, especially
upstream and downstream regions (large-scale near ground level, are very large (up to 100% and
roughness; kS,ABLE30y0) and lower cells in the more), and motivate the use of a fine near-ground
ARTICLE IN PRESS
248 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

H
y

upstream region central region downstream region

domain length L

Fig. 6. Computational mesh with variable height of the wall-adjacent cells along the length of the domain.

20 20
analytical analytical
CFD (fine) CFD (fine)
15 15
CFD (coarse) CFD (coarse)

y (m)
y (m)

10 10

y y 5 5

x x
0 0
0 2 4 6 8 10 12 0.1 0.2 0.3 0.4 0.5
(a) (b) (c) U (m/s) (d) TI

Fig. 7. (a–b) Fine and coarse vertical near-ground mesh distribution at the inlet plane. (c–d) Corresponding mean wind speed and
turbulence intensity profiles at the inlet of the domain: the analytical profile (imposed boundary condition) and the CFD profiles (one
variable value per cell) for the fine and the coarse mesh distribution.

mesh along the entire length of the computational from previous wind tunnel tests. Additional draw-
domain. backs are the increased number of cells and the
(b) Explicit modelling of roughness elements: subsequent increase in required computing power
Another option is to explicitly model the upstream and time.
and downstream roughness in the computational (c) Minimization of upstream domain length:
domain as rectangular blocks in an attempt to fully Reducing as much as possible the upstream length
reproduce the actual roughness effects on the flow. of the computational domain will limit the devel-
In this case, sand-grain roughness wall functions opment of streamwise gradients (Blocken et al.,
can be used to model the small-scale roughness of 2006), as indicated in Figs. 4 and 5. Nevertheless,
the surfaces of these blocks. This possibility has the minimum upstream length, which is the extent
been pursued by Miles and Westbury (2003) and to of the upstream disturbance of the flow, should
some extent by Moonen et al. (2006). Although this always be provided. This option can be useful in
is an interesting option, its practical use can be those studies where the upstream flow character-
hindered by the time-consuming iterative study istics are of importance rather than the downstream
needed to obtain the correct configuration of characteristics (e.g. wind-driven rain impact on the
roughness blocks for a certain set of ABL profiles, windward facades of isolated buildings) but may be
which is required if this configuration is not known insufficient for other cases.
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 249

50 30
y=2m y =20 m
U(NC) U(NC)
40
U(ALK) U(ALK)
U(WSS) 20
error (%)
U(WSS)

error (%)
30 TI(NC) TI(NC)
TI(WSS) TI(WSS)
20
10
10

0 0
1 10 100 1000 10000 1 10 100 1000 10000
(a) x (m) (b) x (m)

Fig. 8. CFD simulation results: relative changes (inhomogeneity errors) of the values of mean wind speed U and turbulence intensity TI at
two heights (y ¼ 2 and 20 m) and at various downstream positions in the empty domain (x co-ordinate), for three options: NC ¼ no
correction, ALK ¼ artificially lower turbulent kinetic energy, WSS ¼ wall shear stress imposed at the bottom of the domain. All changes
are expressed as positive percentage values relative to the values at the inlet of the domain (x ¼ 0).

(d) Generation of ABL profiles based on the kS- shows that quite a good horizontal homogeneity of
type wall functions: Instead of using the ABL inlet U can be obtained, at least up to x ¼ 500 m (see
profiles mentioned in Section 3, artificial ABL inlet Fig. 8, where ALK stands for ‘‘artificially lower k’’).
profiles can be generated by first performing a Reduction of k by a factor of 4 was needed to
simulation in an empty computational domain with achieve this level of homogeneity for U. The use of
identical parameters (grid, roughness, etc.) as the this option, in combination with option (c), has
intended simulation, with the required high mesh been successfully validated for the wind speed
resolution and with kS,groundoyP. This simulation conditions upstream of single buildings and has
can be performed either by using periodic boundary proven to be valuable for CFD simulations of wind-
conditions at the inlet and outlet of a short domain driven rain deposition on the windward facade of
or with a very long domain (typically L410,000 m). isolated buildings where the upstream flow is of
The resulting profiles at the outlet of the domain are main importance (Blocken and Carmeliet, 2002,
the equilibrium profiles of mean wind speed and 2004, 2006). Clearly, however, because of the
turbulence quantities that will yield horizontal change of inlet turbulent kinetic energy, erroneous
homogeneity when applied at the inlet of a similar predictions will be obtained downstream of the
domain. By this calculation, these profiles are windward building face and for turbulence proper-
actually adjusted to fit the wall functions. An ties in general.
example are the profiles obtained at x ¼ 10; 000 m (f) Wall shear stress boundary condition: A final
in Fig. 4. While this may be a good solution in some option to attempt horizontally homogeneous ABL
cases (Blocken et al., 2004), changing the ABL inlet flow simulation is to explicitly specify the wall shear
profiles is not the most straightforward option and stress tw ¼ rðuABL Þ2 associated with the ABL
should be performed with great care. The obtained profiles at the bottom of the upstream and down-
profiles should still be realistic and representative stream region in the domain. This forces the ABL
for the situation to be simulated but they can be friction velocity (uABL ) to be equal to the wall-
quite different from the traditional equilibrium function friction velocity (u and/or ut). The result is
profiles as obtained by e.g. measurements. a very good horizontal homogeneity for both the
(e) Artificial reduction of turbulent kinetic energy: mean wind speed and turbulence profiles when Eqs.
In case the upstream mean velocity field is of main (1)–(3) or Eqs. (2), (5), (6) are used. Fig. 8 shows
importance, rather than the upstream turbulence that the errors are very small and remain below 5%
quantities and the downstream flow, a suitable for the entire length of the domain. Fig. 9 compares
option might be to artificially lower the turbulent the inlet profiles and the downstream profiles at x ¼
kinetic energy at the inlet to decrease the momen- 10; 000 m without corrections and with wall shear
tum transfer between the fluid layers which in turn stress imposed. The latter option clearly provides a
decreases the acceleration of the wind speed near the very good horizontal homogeneity. Note however
surface (Blocken and Carmeliet, 2006). Application that this boundary condition should only be used in
of this option for the simulation case outlined above the region outside the disturbance by the building
ARTICLE IN PRESS
250 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

50 50
inlet (x =0) inlet (x = 0)
x= 10000 m; 40 x = 10000 m;
40 no correction
no correction
x= 10000 m; x = 10000 m;
wall shear stress 30 wall shear stress
30

y (m)
y (m)

20 20

10 10

0 0
2 4 6 8 10 12 14 16 0.05 0.15 0.25 0.35 0.45
(a) U (m/s) (b) TI

Fig. 9. CFD simulation results in terms of inlet (x ¼ 0 m) and downstream (x ¼ 10; 000 m) vertical profiles of (a) streamwise wind speed U
and (b) turbulence intensity TI for two cases: without correction and with wall shear stress imposed at the bottom of the empty domain.

models and that therefore the effect might not be profiles themselves: Eqs. (1)–(3) or Eqs. (2), (5), (6),
sufficient. and they provide horizontal homogeneity when the
standard k– e model constants satisfy Eq. (4)
7.2. The (yP4kS) requirement and other wall because then these equations form an analytical
functions solution to this model. Unfortunately, in many
existing CFD codes such wall functions are neither
Another possibility to deal with this problem is implemented nor can the user always implement
the use of wall functions for which not all four them easily at present.
requirements mentioned in Section 2 need to be It is important to note that, strictly speaking,
satisfied. Several commercial CFD codes including horizontal homogeneity can only be obtained if the
Fluent 6.2 and Ansys CFX 10.0 at present only inlet profiles of mean wind speed and turbulence
provide kS-type wall functions for which these quantities are consistent with the turbulence model
requirements have to be satisfied. However, the (including its constants) and with the wall functions
(yP4kS) requirement might need to be reconsid- and their roughness modification. This is the case
ered. At present, it is indeed not possible to perform for the combination of the inlet profiles in Eqs.
calculations with codes such as Fluent 6.2 or Ansys (1)–(3) or Eqs. (2), (5), (6), the standard k–e model
CFX 10.0 with kS4yP. As stated previously, this with its constants satisfying Eq. (4) and wall
seems logical because it is not physically meaningful functions based on the same equations as the inlet
to have grid cells with centre points within the profiles. In other cases, horizontal homogeneity will
physical roughness height. From a mathematical/ not be obtained, e.g. when measured inlet profiles
numerical point of view however the reason for this different from Eqs. (1)–(3) or Eqs. (2), (5), (6) are
requirement is less clear and it should actually be used. In this case a simulation in an empty domain,
possible to have yPokS,ground ¼ kS,ABLE30y0. This prior to the actual simulation with the building
would solve to a large extent the problems of models present, is required to assess the extent of
horizontal homogeneity reported in this paper. the horizontal inhomogeneity and its acceptability.
Note that this implies a code modification. The kS-type and y0-type wall functions are
Instead of kS-type wall functions, y0-type wall generally quite similar. Indeed, wall functions in
functions can be used in which the roughness is Eqs. (1)–(3), (5), (6) and those in Eqs. (9), (10), (12),
expressed as a function of the aerodynamic rough- (14), (15) have a very similar or even identical shape
ness length y0 instead of kS. In this case, the and were matched earlier to find the kS value
requirements for CFD simulation of ABL flow corresponding to a certain y0. The y0-type wall
relating to kS are no longer applicable. For the functions are reported to perform very well in
standard k–e turbulence model, an appropriate set producing horizontally homogeneous ABL flow
of inlet profiles and y0-type wall functions was when yP is much smaller than kS,ABLE30y0
provided by Richards and Hoxey (1993). These wall (Richards et al., 2002; Mochida et al., 2002;
functions have a similar form as the ABL inlet Bitsuamlak et al., 2006). This supports the idea
ARTICLE IN PRESS
B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252 251

that violating ‘‘yP4kS’’ should also be possible for simulation in an empty computational domain prior
kS-type wall functions. However, due to current to the actual simulation with the obstacle models
restrictions in several CFD codes including Fluent present. Sensitivity tests in an empty computational
6.2 and Ansys CFX 10.0, the requirement yP4kS domain are of critical importance. In addition, for
still has to be satisfied, and to limit horizontal every CFD simulation it is advisable to always
inhomogeneity one has to resort to one or several of report not only the inlet profiles but also the
the remedial measures reported in Section 7.1. incident flow profiles obtained from the simulation
in the empty domain because they characterize the
8. Summary and conclusions real flow to which the building models are subjected.

The accuracy of CFD simulations for atmo- Acknowledgements


spheric studies such as pollutant dispersion and
deposition can be seriously compromised when This research has been conducted while the first
wall-function roughness modifications based on author was a post-doctoral research fellow of the
experimental data for sand-grain roughened pipes FWO-Flanders (Research Fund—Flanders), an
and channels (kS-type wall functions) are applied at organization that supports and stimulates funda-
the bottom of the computational domain. This type mental research in Flanders (Belgium). Their
of roughness modification is currently present in financial contribution is gratefully acknowledged.
many CFD codes including Fluent 6.2 and Ansys The authors also wish to express their gratitude for
CFX 10.0. The problems typically manifest them- the valuable discussions with Dr. Jörg Franke, Dr.
selves as unintended changes (streamwise gradients) Alan Huber, Dr. Bob Meroney, Dr. David Banks
in the vertical mean wind speed and turbulence and Ir. David Dooms.
profiles as they travel through the computational
domain (horizontal homogeneity problem). These
gradients can—at least partly—be held responsible References
for the discrepancies sometimes found between
seemingly identical CFD simulations performed Ansys Ltd., 2005. Ansys CFX-Solver, Release 10.0: Theory.
with different CFD codes and between CFD Canonsburg.
simulations and measurements. Bitsuamlak, G.T., Stathopoulos, T., Bedard, C., 2004. Numerical
evaluation of wind flow over complex terrain: review. Journal
The problems are caused because it is generally of Aerospace Engineering 17 (4), 135–145.
impossible to simultaneously satisfy all four require- Bitsuamlak, G., Stathopoulos, T., Bedard, C., 2006. Effects of
ments for ABL flow simulations when kS-type wall upstream two-dimensional hills on design wind loads: a
functions are used. The extent of the streamwise computational approach. Wind and Structures 9 (1), 37–58.
Blocken, B., Carmeliet, J., 2002. Spatial and temporal distribu-
gradients depends on various simulation character-
tion of driving rain on a low-rise building. Wind and
istics and settings, including the inlet profiles, the Structures 5 (5), 441–462.
turbulence model, the type of wall functions, the Blocken, B., Carmeliet, J., 2004. A review of wind-driven rain
near-wall grid resolution, the roughness height and research in building science. Journal of Wind Engineering and
the roughness constant. Industrial Aerodynamics 92 (13), 1079–1130.
The best solution to this problem is to alleviate Blocken, B., Carmeliet, J., 2006. The influence of the wind-
blocking effect by a building on its wind-driven rain exposure.
the requirement yP4kS. This can be done either by Journal of Wind Engineering and Industrial Aerodynamics 94
using y0-type wall functions or using kS-type wall (2), 101–127.
functions in CFD code formulations that allow Blocken, B., Roels, S., Carmeliet, J., 2004. Modification of
violating this requirement. However, at the time of pedestrian wind comfort in the Silvertop Tower passages by
writing this paper, several CFD codes including an automatic control system. Journal of Wind Engineering
and Industrial Aerodynamics 92 (10), 849–873.
Fluent 6.2 and Ansys CFX 10.0 allow neither of Blocken, B., Carmeliet, J., Stathopoulos, T., 2006. CFD
these options, in which case one or more of the evaluation of the wind speed conditions in passages between
remedial measures mentioned in Section 7.1 should buildings—effect of wall-function roughness modifications on
be considered. the atmospheric boundary layer flow. Journal of Wind
Engineering and Industrial Aerodynamics, accepted for
Irrespective of the type of simulation, the inlet
publication.
profiles, turbulence model, wall functions and near- Castro, I.P., Robins, A.G., 1977. The flow around a surface
wall grid resolution used, it is advisable to always mounted cube in a uniform and turbulent shear flow. Journal
assess the extent of horizontal inhomogeneity by a of Fluid Mechanics 79 (2), 307–335.
ARTICLE IN PRESS
252 B. Blocken et al. / Atmospheric Environment 41 (2007) 238–252

Cebeci, T., Bradshaw, P., 1977. Momentum Transfer in Mochida, A., Tominaga, Y., Murakami, S., Yoshie, R., Ishihara,
Boundary Layers. Hemisphere Publishing Corporation, New T., Ooka, R., 2002. Comparison of various k–e models and
York. DSM applied to flow around a high-rise building—report on
Davenport, A.G., 1960. Rationale for determining design wind AIJ cooperative project for CFD prediction of wind environ-
velocities. Journal of the Structural Division, Proceedings of ment. Wind and Structures 5 (2–4), 227–244.
the American Society of Civil Engineers 86, 39–68. Moonen, P., Blocken, B., Roels, S., Carmeliet, J., 2006.
Davenport, A.G., 1961. The application of statistical concepts to Numerical modeling of the flow conditions in a closed-circuit
the wind loading of structures. In: Proceedings of the low-speed wind tunnel. Journal of Wind Engineering
Institution of Civil Engineers, August. and Industrial Aerodynamics, in press. doi:10.1016/
Durbin, P.A., Petterson Reif, B.A., 2001. Statistical Theory and j.jweia.2006.02.001.
Modelling for Turbulent Flows. Wiley, Chichester, UK. Nikuradse, J., 1933. Strömungsgesetze in rauhen Rohren.
Fluent Inc., 2005. Fluent 6.2 User’s Guide. Fluent Inc., Lebanon. Forschung Arb. Ing.-Wes. No. 361.
Franke, J., Frank, W., 2005. Numerical simulation of the flow Quinn, A.D., Wilson, M., Reynolds, A.M., Couling, S.B., Hoxey,
across an asymmetric street intersection. In: Náprstek, J., R.P., 2001. Modelling the dispersion of aerial pollutants from
Fischer, C. (Eds.), Proceedings of the 4EACWE, 11–15 July agricultural buildings—an evaluation of computational fluid
2005, Prague, Czech Republic. dynamics (CFD). Computers and Electronics in Agriculture
Franke, J., Hirsch, C., Jensen, A.G., Krüs, H.W., Schatzmann, 30, 219–235.
M., Westbury, P.S., Miles, S.D., Wisse, J.A., Wright, N.G., Reichrath, S., Davies, T.W., 2002. Using CFD to model the
2004. Recommendations on the use of CFD in wind internal climate of greenhouses: past, present and future.
engineering In: Proceedings of the International Conference Agronomie 22 (1), 3–19.
on Urban Wind Engineering and Building Aerodynamics. In: Richards, P.J., 1989. Computational modelling of wind flows
van Beeck JPAJ (Ed.), COST Action C14, Impact of Wind around low rise buildings using PHOENIX. Report for the
and Storm on City Life Built Environment. von Karman ARFC Institute of Engineering Research Wrest Park, Silsoe
Institute, Sint-Genesius-Rode, Belgium, 5–7 May 2004. Research Institute, Bedfordshire, UK.
Gao, Y., Chow, W.K., 2005. Numerical studies on air flow Richards, P.J., Hoxey, R.P., 1993. Appropriate boundary
around a cube. Journal of Wind Engineering and Industrial conditions for computational wind engineering models using
Aerodynamics 93 (3), 115–135. the k–e turbulence model. Journal of Wind Engineering and
Harris, R.I., Deaves, D.M., 1981. The structure of strong winds. Industrial Aerodynamics 46&47, 145–153.
Wind engineering in the eighties. In: Proceedings of the Richards, P.J., Younis, B.A., 1990. Comments on ‘‘Prediction of
CIRIA Conference, Construction Industry Research and the wind-generated pressure distribution around buildings’’
Information Association, London, 12–13 November 1980 by E.H. Mathews. Journal of Wind Engineering and
(paper 4). Industrial Aerodynamics 34, 107–110.
Ioselevich, V.A., Pilipenko, V.I., 1974. Logarithmic velocity Richards, P.J., Quinn, A.D., Parker, S., 2002. A 6 m cube in an
profile for flow of a weak polymer solution near a rough atmospheric boundary layer flow. Part 2. Computational
surface. Soviet Physics-Doklady 18, 790. solutions. Wind and Structures 5 (2–4), 177–192.
Jones, W.P., Launder, B.E., 1972. The prediction of laminariza- Riddle, A., Carruthers, D., Sharpe, A., McHugh, C., Stocker, J.,
tion with a 2-equation model of turbulence. International 2004. Comparisons between FLUENT and ADMS for
Journal of Heat and Mass Transfer 15, 301. atmospheric dispersion modelling. Atmospheric Environment
Launder, B.E., Spalding, D.B., 1974. The numerical computation 38 (7), 1029–1038.
of turbulent flows. Computer Methods in Applied Mechanics Schlichting, H., 1968. Boundary-Layer Theory, sixth ed.
and Engineering 3, 269–289. McGraw-Hill, New York.
Mathews, E.H., 1987. Prediction of the wind-generated pressure Stathopoulos, T., 1997. Computational wind engineering: past
distribution around buildings. Journal of Wind Engineering achievements and future challenges. Journal of Wind
and Industrial Aerodynamics 25, 219–228. Engineering and Industrial Aerodynamics 67–68, 509–532.
Meroney, R.N., 2004. Wind tunnel and numerical simulation of White, F.M., 1991. Viscous Fluid Flow, second ed. McGraw-
pollution dispersion: a hybrid approach. Working paper, Hill, New York.
Croucher Advanced Study Insitute on Wind Tunnel Model- Wieringa, J., 1992. Updating the Davenport roughness classifica-
ing, Hong Kong University of Science and Technology, 6–10 tion. Journal of Wind Engineering and Industrial Aerody-
December 2004, 60pp. namics 41–44, 357–368.
Miles, S., Westbury, P., 2003. Practical tools for wind engineering Zhang, C.X., 1994. Numerical prediction of turbulent recirculat-
in the built environment. QNET-CFD Network Newsletter 21 ing flows with a k–e model. Journal of Wind Engineering and
(2), 11–14. Industrial Aerodynamics 51, 177–201.

Das könnte Ihnen auch gefallen