You are on page 1of 46

CFD inflow conditions, wall functions and turbulence

models for flows around obstacles


Alessandro Parente ∗
Université Libre de Bruxelles, Belgium
March 19, 2013

Contents
1 Introduction 5

2 Theory 6
2.1 Inlet conditions and turbulence model . . . . . . . . . . . . . . . . . . . . . 7
2.2 Wall treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Generalization of the ABL model for the case of obstacles immersed in the
flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3 Applications 19
3.1 Empty fetch at wind-tunnel and full scale . . . . . . . . . . . . . . . . . . . 19
3.2 Flow around a ground-mounted building . . . . . . . . . . . . . . . . . . . 23
3.3 Flow over complex terrains, wind-tunnel and full-scale hills . . . . . . . . . 34

4 Influence of stability classes 40


The present lecture notes are based on the research work performed by several Authors in the field
of ABL flows at the von Karman Institute of Fluid Dynamics and at the Université Libre de Bruxelles.
In particular, the contribution by Prof. Carlo Benocci, Prof. Jeroen van Beeck, Dr. Catherine Gorlé and
Dr. Miklos Balogh should be acknowledged.

VKI -1-
LIST OF FIGURES LIST OF FIGURES

List of Figures
1 Computational domain with building models for CFD simulation of ABL
flows and indication of different parts in the domain for roughness modelling
Blocken et al. [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Law of the wall for smooth and sand-grain roughened surfaces as a function
of the dimensionless sand-grain roughness height kS+ Blocken et al. [1]. . . . 12
3 Turbulent kinetic energy profiles at the inlet and outlet sections of an empty
computational domain (see Figure 2b), Dashes: cell value for turbulent
dissipation rate and turbulent kinetic energy averaged over the first cell.
Short dashes: cell value for turbulent dissipation rate and kinetic energy.
Dots: turbulent dissipation rate and kinetic energy averaged over the first
cell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4 Rough law of the wall implementation. . . . . . . . . . . . . . . . . . . . . 15
5 Configurations PS1 (a) and PS2 (b) for the definition of the building influ-
ence area (BIA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
6 Variation of the turbulence model parameter Cµ for an undisturbed ABL
(top), using a prescribed region Beranek [2] (middle) and using an auto-
matic switching function (bottom). . . . . . . . . . . . . . . . . . . . . . . 18
7 Computational domain and main boundary conditions applied for the nu-
merical simulation of the unperturbed ABL at wind tunnel (a) and full (b)
scale. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
8 Profiles of velocity, turbulent kinetic energy and turbulent dissipation rate
at inlet and outlet section of the computational domain (Figure 7), ob-
tained when applying inlet conditions given by Equations 10-12 (a-c) and
Equations 10,15 and 12 (d-f). STD WF = Standard Wall Function; MOD
WF = Modified Wall Function [3]. . . . . . . . . . . . . . . . . . . . . . . 21
9 Wall shear stress as a function of the axial coordinate. Solid black line:
theoretical value, ρu2∗ . Dashed red line: Richards and Hoxey [4] profile.
Blue dots: Yang et al. [5] profile [3]. . . . . . . . . . . . . . . . . . . . . . . 23
10 Profiles of velocity (a), turbulent kinetic energy (b), turbulent dissipation
rate (c) and non-dimensional velocity gradient (d) at inlet and outlet section
of the computational domain, obtained with Equations 10, (23) and 12 [6]. 24
11 Profiles of velocity, turbulent kinetic energy and turbulent dissipation rate
at inlet and outlet section of the computational domain (Figure 7), ob-
tained when applying inlet conditions given by Equations 10-12 (a-c) and
Equations 10, (29) and 12 (d-f). Results obtained with the wall function
formulation by [3] [3]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
12 Wall shear stress as a function of the axial coordinate. Solid black line:
theoretical value, ρu2∗ . Dashed red line: Richards and Hoxey [4] profile.
Blue dots: Yang et al. [5] profile. . . . . . . . . . . . . . . . . . . . . . . . 26
13 Building geometry and location of measurement planes for the flow around
the obstacle [7]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
14 Computational domain and grid for the flow around the obstacle [7]. . . . . 26

VKI -2-
LIST OF FIGURES LIST OF FIGURES

15 Contour plots of non-dimensional velocity on the planes y = 0 (left) and


z = 0.035m (right). Experimental measurements are compared to the
results obtained applying the PS1 and PS2 model [3]. . . . . . . . . . . . . 29
16 Contour plots of non-dimensional turbulence kinetic energy on the planes
y = 0(left) and z = 0.035m (right). Experimental measurements are com-
pared to the results obtained applying the PS1 and PS2 model [3]. . . . . . 29
17 Experimental and numerical profiles of non-dimensional velocity upstream,
over and downstream of the obstacle. Solid line: experimental data [3].
Dashes: PS1 configuration. Short dashes: PS2 configuration . . . . . . . . 31
18 Experimental and numerical profiles of non-dimensional turbulent kinetic
energy upstream, over and downstream of the obstacle [3]. Solid line: ex-
perimental data. Dashes: PS1 configuration. Short dashes: PS2 configura-
tion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
19 Experimental and numerical profiles of non-dimensional velocity over and
downstream of the obstacle. Solid line: experimental data [6]. Dashes:
UABL model. Short dashes: PS model. Dots: ASQ model. . . . . . . . . 31
20 Experimental and numerical profiles of non-dimensional turbulent kinetic
energy over and downstream of the obstacle. Solid line: experimental data
[6]. Dashes: UABL model. Short dashes: PS model. Dots: ASQ model. . . 32
21 Local hit rates for the non-dimensional turbulent kinetic energy, applying
the ASQ model (left) and the UABL model (right) [6]. . . . . . . . . . . . 33
22 Stream-wise velocity (top) and turbulent kinetic energy profiles (bottom)
in the symmetry plane against measurements obtained on the 3D hill at
laboratory scale using Fluent and OpenFOAM [8]. . . . . . . . . . . . . . . 35
23 Simulated wall shear stress along the symmetry of the domain against theo-
retical values, extracted from the inlet profile (Inlet τw) and against values
extracted from measured profiles (meas.).[8]. . . . . . . . . . . . . . . . . . 36
24 Building geometry and location of measurement planes for the flow around
the obstacle [7]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
25 Comparison of simulated and measured horizontal and vertical stream ve-
locity (Uh and W ) and turbulent kinetic energy (k) along line-A, using the
comprehensive approach [6] with α = 3 [8]. . . . . . . . . . . . . . . . . . 39
26 Comparison of simulated and measured vertical profiles (U and k) at the
hill summit [8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
27 Profiles of (a) velocity, (b) turbulent kinetic energy, (c) turbulent dissipa-
tion rate, (d) turbulent viscosity and e) temperature at the inlet and outlet
section of a 2D computational domain (60m high and 400m long ), and (f)
shear stress at the wall. Inlet conditions taken from [9]. . . . . . . . . . . . 42

VKI -3-
LIST OF TABLES LIST OF TABLES

List of Tables
1 Inlet conditions and turbulence model formulation. . . . . . . . . . . . . . 10
2 Fitting parameters for velocity and turbulent kinetic energy inlet profiles
according to Yang et al. [5], Parente et al. [6] and turbulent model parameters. 19
3 Fitting parameters for velocity and turbulent kinetic energy inlet profiles
according to Richards and Hoxey [4], Brost and Wyngaard [10] and turbu-
lent model parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4 Hit rate values for non-dimensional velocity and turbulent kinetic energy for
the prescribed BIA size approach., varying the turbulence model settings [3]. 28
5 Test cases and corresponding model settings for the numerical simulation
of the flow around a bluff-body [7], using the prescribed BIA size approach.
TM=Turbulence Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6 Test cases and corresponding model settings for the numerical simulation
of the flow around a bluff-body [7], using the automatic switch approach
BIA for the BIA. TM=Turbulence Model. . . . . . . . . . . . . . . . . . . 30
7 Hit rate values for non-dimensional velocity and turbulent kinetic energy
for the automatic switch approach, varying the turbulence model settings
[6]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
8 Hit rate values for non-dimensional velocity and turbulent kinetic energy
for the automatic switch approach, varying the turbulence model settings
[11]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
9 Fitting parameters for velocity and turbulent kinetic energy inlet profiles
according to Parente et al. [6] and turbulent model parameters for the 3D
hill simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
10 Hit rate values for non-dimensional velocity and turbulent kinetic energy
for the 3D hill simulation, varying the turbulence model settings [8]. . . . . 37
11 Normalized errors in the prediction of the separation point, SP , and wake
length, W L , for the 3D hill. Negative and positive values sign the under-
and overestimation respectively[8]. . . . . . . . . . . . . . . . . . . . . . . . 37
12 Fitting parameters for velocity and turbulent kinetic energy inlet profiles
according to Parente et al. [6] and turbulent model parameters for the
Askervein hill simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
13 Hit rate values for non-dimensional velocity and turbulent kinetic energy
for the Askervein hill simulation, varying the turbulence model settings
within the wake region [8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

VKI -4-
1 INTRODUCTION

1 Introduction
Computational Fluid Dynamics (CFD) is widely used to study flow phenomena in the
lower part of the atmospheric boundary layer (ABL), with applications to pollutant dis-
persion, risk analysis, optimization and siting of windmills and wind farms, and micro-
climate studies. Numerical simulations of ABL flows can be performed either by solving
the Reynolds-averaged Navier-Stokes (RANS) equations or by conducting large-eddy sim-
ulations (LES). It is generally acknowledged that LES, which explicitly accounts for the
larger spatial and temporal turbulent scales, can provide a more accurate solution for
the turbulent flow field, provided that the range of resolved turbulence scales is suffi-
ciently large and that the turbulent inflow conditions are well characterized Shah and
Ferziger [12], Lim et al. [13], Xie and Castro [14, 15]. For example, Xie and Castro [14]
presented a comparison of LES and RANS for the flow over an array of uniform height
wall-mounted obstacles: the Authors compared the results to available direct numerical
simulation (DNS) data, showing that LES simulations outperform RANS results within
the canopy. Dejoan et al. [16] compared LES and RANS for the simulation of pollutant
dispersion in the MUST field experiment and found that LES performs better in pre-
dicting vertical velocity and Reynolds shear stress, while the results for the stream-wise
velocity component are comparable. However, LES simulations are at least one order
of magnitude computationally more expensive than RANS [17] and the sensitivity to in-
put parameters such as inlet conditions, imply that, as for RANS, multiple simulations
are needed to quantify the resulting uncertainty in the output for realistic applications.
Hence, practical simulations of ABL flows are still often carried out solving the RANS
equations in combination with two-equation turbulence models. Consequently, investigat-
ing possible improvements to these models is worthwhile. In RANS simulations, the effect
of roughness on ABL flows is generally represented with the so-called sand-grain based
wall functions Cebeci and Bradshaw [18], based on the experiments conducted by Niku-
radze [19] for flow in rough, circular pipes covered with sand. Moreover, the upstream
turbulent characteristics of a homogeneous ABL flows are generally modeled using the
profiles suggested by Richards and Hoxey [4] for mean velocity, turbulent kinetic energy
and turbulent dissipation rate. However, this modelling approach can result in an unsat-
isfactory reproduction of the ABL for two main causes. The first cause of discrepancy
lies in the inconsistency between the fully developed ABL inlet profiles and the rough
wall function formulation Riddle et al. [20], Franke et al. [21], B. Blocken [22], Blocken
et al. [1], Hargreaves and Wright [23], Franke et al. [24]. Furthermore, the inlet profile
for the turbulence kinetic energy, k, proposed by Richards and Hoxey [4], assumes a con-
stant value with height, in conflict with wind-tunnel measurements Leitl [7], Xie et al.
[25], Yang et al. [5], where a variation of k with height is observed. A remedial measure
to solve the inconsistency between the sand-grain based rough wall function and the fully
developed inlet profiles was proposed by Blocken et al. [1]. It consists in the modifica-
tion of the wall law coefficients, namely the equivalent sand-grain roughness height ks
and the roughness constant Cs , to ensure a proper matching with the velocity boundary
conditions. This approach ensures the desired homogeneity of the velocity distribution
for the fully developed ABL, but it is code dependent and does not provide a general
solution to the problem. Moreover, the standard law of the wall for rough surfaces poses
limitations concerning the level of grid refinement that can be achieved at the wall. This

VKI -5-
2 THEORY

restriction becomes particularly relevant for applications requiring a high resolution near
the wall boundaries. An additional complicating factor is the necessity to apply different
wall treatments when a combination of rough terrains and smooth building walls must
be simulated. Concerning the inlet profile for turbulent kinetic energy, Yang et al. [5]
derived a new set of inlet conditions, with k decreasing with height. However, the ap-
plication of such a profile at the inlet boundary only provides an approximate solution
for the system of equations describing a fully developed ABL. In a recent work, Gorlé
et al. [26] proposed a modification of the constant Cµ , and of the turbulent dissipation
Prandtl number, σ , to ensure homogeneity along the longitudinal ABL direction, when
the k profile of Yang et al. [5] is applied. Parente and Benocci [27], Parente et al. [3]
proposed a modification of the k −  turbulence model compatible with the set of inlet
conditions proposed by Yang et al. [5]. Such a modification consisted in the generalization
of the model coefficient Cµ , which becomes a local function of the flow variables, and in
the introduction of two source terms in the transport equations for k and , respectively.
The limitation of such an approach consisted in the inlet profile adopted for turbulent ki-
netic energy, which does not satisfy all the governing simulations involved in the problem
Parente et al. [6]. Parente et al. [6] addresses the aforementioned aspects by proposing
a comprehensive approach for the numerical simulation of the neutral ABL. First, a new
profile for turbulent kinetic energy was derived from the solution of the turbulent kinetic
energy transport equation, resulting in a new set of fully developed inlet conditions for the
neutral ABL, which satisfies the standard k −  model. This was accomplished through
the introduction of a universal source term in the transport equation for the turbulent
dissipation rate, , and the re-definition of the k −  model coefficient Cµ as a function of
the flow variables. Second, for the purpose of solving the flow around obstacles immersed
in the flow, the modelling approach derived for the homogeneous ABL was generalized
with an algorithm for the automatic identification of the building influence area (BIA).
As a consequence, the turbulence model formulation is gradually adapted moving from
the undisturbed ABL to the region affected by the obstacle. Parente et al. [28, 3] also
proposed a novel implementation of a wall function, which incorporates both smooth-
and rough-wall treatments, employing a screening algorithm to automatically select the
desired formulation, i.e. rough or smooth, depending on the boundary surface properties.
Balogh et al. [8] extended the approach by Parente et al. [28, 3] to the simulation of flows
above complex terrains, i.e. wind-tunnel scale 3D hill model and Askervein Hill.
The present notes are organized as follows. The modelling approach for the numerical
simulation of neutral ABL flows is presented, by discussing the turbulence model formu-
lation, the different inlet profiles and the wall function. Applications are presented and
discussed for the flow over flat terrain, around ground mounted bluff bodies and over hills.

2 Theory
The standard k− model remains the most common option for the numerical simulation of
the homogeneous ABL. Such a family of models solves a transport equations for turbulent
kinetic energy, k, and for turbulent dissipation rate, :
  
∂ ∂ ∂ µt ∂k
(ρk) + (ρkui ) = µ+ + Gk − Gb − ρ − YM (1)
∂ ∂xi ∂xj σk ∂xj

VKI -6-
2.1 Inlet conditions and turbulence model 2 THEORY

  
∂ ∂ ∂ µt ∂  2
(ρ) + (ρui ) = µ+ + C1 (Gk + C3 Gb ) − C2 ρ . (2)
∂ ∂xi ∂xj σ ∂xj k k
In Equations (1)-(2), ui is the ith velocity component, ρ is the density, C1 , C2 and C3 are
model constants, σk , and σ are the turbulent Prandtl numbers for k and , respectively, Gb
is the turbulent kinetic energy production due to buoyancy, YM represents the contribution
of the fluctuating dilatation in compressible turbulence to the overall dissipation rate, Gk is
the generation of turbulence kinetic energy due to the mean velocity gradients, calculated
from the mean rate-of-strain tensor, Sij , as:
 
2
p 1 ∂ui ∂uj
Gk = µt S S = 2Sij Sij Sij = + . (3)
2 ∂xj ∂xi
For a steady ABL, under the hypothesis of zero vertical velocity, constant pressure along
vertical (z) and longitudinal (x) directions, constant shear stress throughout the boundary
layer and no buoyancy effects, the transport equations for turbulent kinetic energy k, and
turbulent dissipation rate , simplify to:
 
∂ µt ∂k
+ Gk − ρ = 0 (4)
∂z σk ∂z
 
∂ µt ∂  2
+ C1 Gk − C2 ρ = 0 (5)
∂z σ ∂z k k
 2
∂u
Gk = µt . (6)
∂z
The model is completed by the momentum equations, which takes the form:
∂u
µt = τw = ρu2∗ (7)
∂z
where τw is the wall shear stress and u∗ is the friction velocity
r
τw
u∗ = . (8)
ρ
In Equations (4), (5) and (7) the laminar viscosity has been neglected with respect to the
turbulent one, µt , expressed as:
k2
µt = ρcµ . (9)


2.1 Inlet conditions and turbulence model


Fully developed inlet profiles of mean longitudinal velocity, turbulent kinetic energy
and dissipation rate under neutral stratification conditions are often specified following
Richards and Hoxey [4]:  
u∗ z + z0
U = ln (10)
κ z0
u2
k = p∗ (11)

VKI -7-
2 THEORY 2.1 Inlet conditions and turbulence model

u3∗
= (12)
κ (z + z0 )
where κ is the von Karman constant and z0 is the aerodynamic roughness length. It can
be shown that Equations (10)-(12) are analytical solutions of the standard k −  model if
the turbulent dissipation Prandtl number, σv , is defined as Richards and Hoxey [4]:

κ2
σ = p (13)
(C2 − C1 ) Cµ
or, equivalently Parente et al. [3, 6], Pontiggia et al. [29], if the following source term is
added to the dissipation rate equation:
p !
ρu4∗ (C2 − C1 ) Cµ 1
S (z) = − . (14)
(z + z0 )2 κ2 σ

A weakness of the formulation presented above is the assumption of a constant value for
the turbulent kinetic energy k in Equation (11). Indeed, experimental observations show
a decay of k with height Leitl [7], Xie et al. [25], Yang et al. [5]. Following this observation,
Yang et al. [5] analytically derived an alternative inlet condition for k:
p
k = C1 ln (z + z0 ) + C2 (15)

where C1 and C2 are constants determined via experimental data fitting. The profile for k
expressed by Equation (15) is obtained directly as solution of the turbulent kinetic energy
transport equation, under the assumption of constant value for Cµ and local equilibrium
between production and dissipation:
p du
 (z) = Cµ k . (16)
dz
Yang et al. [5] mentioned that the constant Cµ should be correctly specified in order to
ensure the correct level of turbulence kinetic energy throughout the domain. However, this
could be unnecessary, if the effect of a non-constant k profile on the momentum equation
is taken into account. Gorlé et al. [26] generalized the expression of Cµ as a function of
z, by substituting Equations (9) and (16) into Equation (7):

∂u k 2 ∂u k2 ∂u
µt = ρu2∗ → ρcµ = ρu2∗ → ρcµ p ∂u
= ρu2∗ (17)
∂z  ∂z Cµ k ∂z ∂z

and, then,

u4∗
Cµ = 2 (18)
k
Equation (18) is simply the relation proposed by Richards and Hoxey [4] inverted, to
ensure consistency between the turbulence model and the k profile throughout the ABL
domain. From the point of view of the physical interpretation, the non-uniform k profile
and the definition of Cµ can be related to the large-scale turbulence present in ABL flows,

VKI -8-
2.1 Inlet conditions and turbulence model 2 THEORY

which can vary significantly with height. Bottema [30] indicated the relevance of large-
scale turbulence to several RANS models, pointing out the necessity for case and location
dependent model constants.
Using the k inlet profile by Yang et al. [5], together with Equations (10) and (12) for u
and , and employing Equation (18) for Cµ does not allow to close the system of Equations
(4)-(7), with the definition of an appropriate expression for σ . Only an approximate
solution Gorlé et al. [26] can be found, using the constant value of Cµ obtained at the
wall adjacent cell. In alternative, the functional form of Cµ ((18)), by introducing an an
additional source term for the k transport equation Parente et al. [3], in addition to the
one expressed by Eq. 8 for the ε transport equation:
 
ρu∗ κ ∂ (z + z0 ) ∂k
∂z
Sk (z) = . (19)
σk ∂z
As a consequence, an arbitrary set of inlet conditions, including the ones by Yang et al.
[5] can be adopted at the inlet boundary, ensuring their conservation throughout the
computational domain.
An alternative approach is that of repeating the exercise by Yang et al. [5] considering
the functional variation of Cµ ((18)). In particular, assuming local equilibrium between
turbulence production and dissipation Equation (16), Equation (4) becomes:
 
∂ µt ∂k
=0 (20)
∂z σk ∂z
Substituting Equations (9), (16) and (18) into Equation (20), we get:
 4 
  u∗ k2
! 2 ρ 2
ρc √ k
r
2
∂ ρcµ k ∂k ∂  µ Cµ k du ∂k ∂  k u24∗ k du ∂k 
dz =  k dz 
=  .
∂z σk ∂z ∂z σk ∂z ∂z  σk ∂z 

Employing the analytical expression of the inlet velocity profile, du


dz
= uκ∗ (z+z
1
0)
, (Equation
(10)):
! !  
∂ ρu2∗ ∂k ∂ ρu2∗ ∂k ∂ ρu∗ κ ∂k
= = (z + z0 ) =0 (21)
∂z σk dudz
∂z ∂z σk uκ∗ (z+z1
0)
∂z ∂z σk ∂z

which gives:
∂k
(z + z0 )= const (22)
∂z
By integrating Equation (22), the following general solution for turbulent kinetic energy
profile is obtained:

k (z) = C1 ln (z + z0 ) + C2 (23)
which differs from Equation (15) since the square root operator disappears. Similarly
to Equation (15), C1 and C2 are constants determined by fitting the equations to the
measured profile of k. For what concerns the profile of turbulent dissipation rate, the

VKI -9-
2 THEORY 2.1 Inlet conditions and turbulence model

Table 1: Inlet conditions and turbulence model formulation.


 
U = uκ∗ ln z+z z0
0
Equation (10)
Inlet conditions k (z) = C1 ln (z + z0 ) + C2 Equation (23)
u3∗
 = κ(z+z0 ) Equation (12)
k2
µt = ρcµ   Equation (9)

Turbulence Model 4
ρu∗ (C 2 −C 1 ) Cµ
S (z) = (z+z )2 κ2
− σ1 Equation (14)
0
u4∗
Cµ = k2
Equation (18)

equilibrium assumption (Equation (16)) and the relation for Cµ (Equation (18)) ensure
that Equation (12) remains valid. The full set of inlet conditions, the turbulence model
formulation and the wall function implementation are summarized in Table 1. The set of
inlet boundary conditions provided by Equations (10), (23) and (12) for velocity, turbulent
kinetic energy and dissipation rate, respectively, represents a consistent extension of the
formulation proposed by Richards and Hoxey [4] to the case of a non-constant turbulent
kinetic energy profile. Indeed, if Equation (23) for and Equation (18) for Cµ are used, the
transport equation for the turbulent dissipation rate is identically satisfied by the source
term S (z) (Equation (14)), which is independent of the specific form of the inlet profile.
In fact, the equilibrium assumption and the generalization of Cµ make the first term of
Equation (5) (which is the only one affected by the functional variation of k) universal
and equal to:
   q 4 q 
ρcµ √ k du ∂ pC k du
2
ρ u∗
k ∂ u4∗ ku∗  4

∂  C k
µ dz µ dz
= ∂  k 2 2
k κ(z+z0 )
= ∂ ρu ∗ 1
u∗ −
∂z σ ∂z ∂z σ κ(z+z 0)
∂z ∂z σ (z + z0 )
 
∂ ρu4∗ 1 ρu4 1
− = ∗ (24)
∂z σ (z + z0 ) σ (z + z0 )2
It should be noted that, when employing the novel turbulent kinetic energy profile (Equa-
tion (23)), the source term in Equation (19) reduces to zero. Finally, it can be observed
that, assuming a constant profile for k, i.e. C1 = 0 in Equation (23), the proposed ap-
proach reduces to a formulation equivalent to the one proposed by Richards and Hoxey
[4], with the difference that the proper value of Cµ is automatically selected via Equation
(18).
The profile proposed by Yang et al. [5] requires the availability of experimental data to
determine the parameters C1 and C2 of Equation (23). This is not always guaranteed, es-
pecially for full-scale measurements. In this case, semi-empirical parameterizations avail-
able in the literature Brost and Wyngaard [10] can be applied for the turbulent quantities.
provided the following expressions for the mean squared fluctuating velocity components:

02
u z
2
=5−4 (25)
u∗ h

02
v z
2
=2− (26)
u∗ h

VKI - 10 -
2.2 Wall treatment 2 THEORY

0
w2 z
2
= 1.7 − (27)
u∗ h
where h is the ABL height. For neutral stratification conditions the value of h can be
deduced from the following relation Bechmann [31]:
hfc
≈ 0.33 (28)
u2∗
where a typical mid-latitude value for the Coriolis parameter, fc = 10−4, can be con-
sidered. The variation of turbulent kinetic energy with height can be then expressed
as:
1 D 0 2 E D 0 2 E D 0 2 E u2∗  z
k (z) = u + v + w = 8.7 − 6 (29)
2 2 h
As stated earlier, the use of a specific inlet profile for k does not result in inconsistencies in
the turbulence model formulation, as long as the source term for turbulent kinetic energy
is employed (Equation (19)). Such a term automatically vanishes when Equation (23) for
k is employed.

2.2 Wall treatment


Due to the importance of the surface roughness and the high Reynolds numbers associated
with ABL flow, the use of wall functions is generally required for near-wall modelling.
The use of wall functions is particularly important for those regions of the computational
domain where the actual elements are not modeled explicitly but in terms of roughness
(Figure 1). The wall functions in CFD codes are generally based on the universal law of
the wall, that can be modified to account for the near-surface roughness. The universal
law of the wall for a smooth surface is plotted in Figure 2 using the black solid line,
using the dimensionless variables u+ = U/u∗ and y + = u∗ y/ν . where U is the mean velocity
tangential to the wall, u is a wall-function friction velocity and ν is the kinematic viscosity.
The near-wall region consists of three main parts: the laminar layer or linear sublayer,
the buffer layer and the logarithmic layer. In the linear sublayer, the laminar law holds
(u+ = y + ) while in the log layer, the logarithmic law is valid: The standard smooth law
of the wall takes the form:
Up ln (y + )
= +B (30)
u∗ κ
where the integration constant B = 5.0–5.4.  The effect of roughness is modeled by
+
introducing a shift in the intercept, ∆B kS . Moving the integration constant inside the
logarithm, the following expression is obtained for the logarithmic law for a rough wall:
Up 1  
= ln Ey + − ∆B kS+ . (31)
u∗ κ
where κ is the von Karman
 constant, E is the wall function constant (E = 9 − 9.7935.
The function ∆B kS . depends of the dimensionless roughness height, kS+ = u∗ kS/ν ,
+

and measures the departure of the wall velocity from smooth conditions. It can assume
different forms, depending on the equivalent sand grain roughness values Nikuradze [19],
Cebeci and Bradshaw [18]. In particular, when kS+ > 90:

VKI - 11 -
2 THEORY 2.2 Wall treatment

Figure 1: Computational domain with building models for CFD simulation of ABL flows
and indication of different parts in the domain for roughness modelling Blocken et al. [1].

Figure 2: Law of the wall for smooth and sand-grain roughened surfaces as a function of
the dimensionless sand-grain roughness height kS+ Blocken et al. [1].

VKI - 12 -
2.2 Wall treatment 2 THEORY

 1 
∆B kS+ = ln CS kS+ (32)
κ
which gives:  
Up 1 Ey +
= ln . (33)
u∗ κ CS kS+
where Cs is a roughness constant. By comparing Equations (10) and (33), it becomes
clear that the two treatment are inconsistent and they may lead to discrepancies in the
prediction of the near wall velocity, if a proper selection of the roughness constants is
not performed. To avoid that, Blocken et al. [1] proposed a first order match between
the rough wall of the wall and the velocity inlet profile at the first cell centroid, zp , to
appropriately select CS :

Ey + zp + z0 E u∗νzp z0 Ez0 Ez0


+ = → CS = u∗ kS
∼ ∼ (34)
CS kS z0 ν
(z0 + zp ) kS zp

In Equation (34), a common requirement of ABL simulations has been made explicit,
namely that the distance zp between the centroid of the wall-adjacent cell and the wall
should be larger than the sand-equivalent roughness ks of the terrain. This requirement
is generally translated into an upper limit for ks , taken equal to zp , zp ≥ kS Blocken
et al. [1]. Different implementations of Equation (33) can be found in commercial CFD
codes. Currently, the value of CS can be freely set in the commercial software StarCCM+.
In Ansys CFX, a fixed value for the roughness constant is adopted, CS = 0.3, whereas
FLUENT allows defining custom profiles for CS through user defined functions (UDF).
However, even when the value of the velocity at the first cell matches the one provided by
Equation (10), the standard rough wall function suffers from two main drawbacks. First,
it poses strong limitations on the maximum size of the wall adjacent cell, the maximum
allowable value for CS being limited by the wall function constant E. In fact, at the first
cell centroid, zp , Equations (33) gives:
 
Up 1 E
= ln
u∗ κ CS

taking kS = zp . This implies that CS cannot be higher than the value of the parameter
E. Moreover, the standard wall function does not imply any direct effect of the roughness
properties on the turbulence quantities at the wall. This can be shown by taking the
squared derivative of the velocity profile (proportional to the production of k, using the
sand-grain law of the wall and the ABL velocity inlet profile:
 2    2  u 2
∂U ∂ u∗ Ez ∗
= ln = (35)
∂z wf, std ∂z κ CS kS κz
 2    2  2
∂U ∂ u∗ z + z0 u∗
= ln = (36)
∂z in, ABL ∂z κ z0 κ (z + z0 )
To overcome these drawbacks, an alternative boundary condition could be imple-
mented with the same functional form of the Richards and Hoxey [4] inlet profiles. To

VKI - 13 -
2 THEORY 2.2 Wall treatment

that purpose, velocity, turbulent dissipation rate and turbulent kinetic energy production
are specified as follows Richards and Hoxey [4]:
 
u∗ z + z0
Up = ln (37)
κ z0
Cµ0.75 k 1.5
p = (38)
κ (z + z0 )
τw2
Gk = (39)
ρκCµ0.25 k 0.5 (z + z0 )

Richards and Hoxey [4] proposed a different expression ´ 2zfor Gk , by integrating the pro-
1 p
duction term over the first cell height, i.e. Gk = 2zp 0 Gk dz,. Such formulation was
found to produce a peak in the k profile close to the wall Parente et al. [3, 6] and the
evaluation of Gk at the cell centroid may be preferable (Equation (39)). Figure 3 shows
the turbulent kinetic energy at the inlet and outlet sections of an empty domain, where
fully developed profiles of velocity and turbulence are specified at the domain inlet. The
test case is presented to show how the proposed wall function formulation (green dots)
provides the best agreement (deviation below 4%) between the profile specified at the
inlet and the one retrieved at the outlet. By averaging both Gk and  (blue short-dashes),
the deviation at the wall is about 15%, but it remains higher than the error obtained
using the cell values of Gk and . Averaging Gk over the first cell, while keeping the cell
value for  (red dashes), leads to an overestimation of the turbulent kinetic energy at the
wall of about 100%. Richards and Norris [32] have shown that such a peak in the profile
for k can also be avoided by changing the discretization of the production term, to ensure
equilibrium between production and dissipation.
Parente et al. [3, 6]proposed an implementation of the rough wall function which
preserves the the form of the universal law of the wall, through the introduction of a new
wall function constant and non-dimensional wall distance:

Up 1  e +
= ln Ee
y (40)
u∗ κ
with
u∗ (z + z0 ) e= ν
ye+ = E . (41)
ν z0 u∗

The non-dimensional distance, ye+ , is simply a y + shifted by the aerodynamic roughness,


whereas the new wall function constant, E e depends on the roughness characteristics of the
surface. In Equation (40) the friction velocity, u∗ , is not kept constant in the longitudinal
direction but calculated locally as u∗ = Cµ0.25 k 0.5 . The present approach removes the
drawbacks of the standard wall function, without limiting its flexibility. In particular, it
is easily extendable to mixed rough and smooth surface through a redefinition of the law
of the wall constants. Moreover, it allows full flexibility from the point of view of mesh
generation, as the wall function parameters do not impose any limitation on the first cell
height. A schematic of the rough wall implementation is shown in Figure 4.

VKI - 14 -
2.2 Wall treatment 2 THEORY

Figure 3: Turbulent kinetic energy profiles at the inlet and outlet sections of an empty
computational domain (see Figure 2b), Dashes: cell value for turbulent dissipation rate
and turbulent kinetic energy averaged over the first cell. Short dashes: cell value for
turbulent dissipation rate and kinetic energy. Dots: turbulent dissipation rate and kinetic
energy averaged over the first cell.

Richards and Hoxey (1993) Implementation


(Fluent, OpenFOAM, Fine Open)

✓ ◆
u⇤ z + z0 centroid Up 1 ⇣ e +⌘
Up = ln = ln Ee
y
 z0 u⇤ 
Cµ0.75 k 1.5 UP, Gk, ϵP
Smooth wall
✏p = e=E
 (z + z0 ) ye = y +
+
E

⌧w2 Rough wall


Gk =
⇢Cµ0.25 k 0.5 (z + z0 ) u⇤ (z + z0 ) e= ⌫
ye+ = E
⌫ z0 u ⇤

Figure 4: Rough law of the wall implementation.

VKI - 15 -
2 THEORY
2.3 Generalization of the ABL model for the case of obstacles immersed in the flow

2.3 Generalization of the ABL model for the case of obstacles


immersed in the flow
The proposed methodologies for neutral ABL flows is valid when the ABL is undisturbed.
When an obstacle is immersed in the flow field or for the flow on more complex terrains, the
turbulence model formulation (including the definition of the turbulence model parameter
Cµ (Equation (18)) and the turbulent dissipation rate source term source term S (z)
(Equation (14))Pare no longer valid, since the governing equations are no longer given by
Equations (4)-(7).
Gorlé et al. [26]proposed a modification of the turbulence model parameters Cµ and
σv in a building influence area (BIA) defined following Beranek [2], as a half sphere with
radius r = 1.6H and centered 0.16H downstream of the building, where H is the building
height. Parente et al. [3] further investigated the effect of the BIA size and shape, by
limiting the BIA area to the region downstream of the building. The investigated config-
urations, indicates as PS1 and PS2 respectively, are shown in Figure for the flow around
a bluff-body. Parente et al. [6] proposed an approach for the automatic determination of
the BIA, to allow a gradual transition of the turbulence model parameters from the for-
mulation proposed for the undisturbed ABL to one more suitable for wake flow regions. A
local deviation from the undisturbed ABL conditions is introduced, which automatically
identifies the extent of the flow region affected by the obstacle. The relative deviation, δ,
of the actual local velocity profile, u, with respect to the inlet logarithmic one, uABL , is
taken as blending parameter between the two formulations:
  
u − uABL
δ = min , 1 (42)
uABL
The coefficient Cµ and the source term S (z) are then weighted as a function of δ, to
allow a gradual transition between different flow regions:

φ = δ α φwake + (1 − δ α ) φABL = φwake + (1 − δ α ) (φABL − φwake ) (43)

where wake and ABL indicate the wake and undisturbed ABL values for the turbulence
model parameters. If the standard k −  is employed in the wake, as proposed by Gorlé
et al. [26], Parente et al. [3, 6], Balogh et al. [8], then S,wake = 0 and Cµ,wake = 0.09.
The parameter α determines the shape of the transition function between the different

Figure 5: Configurations PS1 (a) and PS2 (b) for the definition of the building influence
area (BIA)

VKI - 16 -
2.3 Generalization of the ABL model for the case of obstacles immersed in2 the
THEORY
flow

formulations. Different values of α have been tested by Parente et al. [6], Balogh et al. [8],
showing that quadratic blending is the most appropriate choice. Figure 6 shows the varia-
tion of the model parameter Cµ for the ABL model, for the prescribed switch approach and
for an undisturbed ABL (top), using a prescribed region Beranek [2] (middle) and using
an automatic switching function (bottom). To improve the performances of the standard
k −  model in wakes, different corrections may be considered such as the ones proposed
by Kato and Launder [33], Yap [34], Tsuchiya et al. [35] to improve the simulation of
separation and reattachment regions, as reported in Parente et al. [3, 6], Balogh et al. [8].
The Kato-Launder (KL) approach is based on the reformulation of the production term
written as:
Gk,KL = µt SΩ (44)
 
p 1 ∂ui ∂uj
Ω = 2Ωij Ωij Ωij = −
2 ∂xj ∂xi
 
p 1 ∂ui ∂uj
S = 2Sij Sij Sij = +
2 ∂xj ∂xi
where S is the strain tensor (symmetric part of the velocity gradient tensor) and Ω is the
vorticity (antisymmetric part of the velocity gradient tensor). The approach proposed by
Yap is a correction for separated flows; it corrects the turbulent kinetic energy prediction
in the separated regions through an additive source term in the  transport equation,
defined as:    1.5 2
2 k 1.5 k
S, Y AP = 0.83 −1 (45)
k le le
where le = Cµ−0.75 κdw , and dw is the nearest wall distance. Following Launder [36], the
Yap correction is applied together with the Kato-Launder approach. The modification
proposed by Tsuchiya et al. [35] (usually reported as MMK model in the literature)
differs from the standard k −  model for the evaluation of the eddy viscosity µt . In this
approach, the turbulent viscosity if modified with a multiplier expressed as a constrained
ratio between the vorticity and the modulus of the rate of strain tensor:
 
k2 Ω
µt, M M K = FM M K ρCµ FM M K = min , 1 . (46)
 S

In alternative to linear closures, nonlinear eddy viscosity models the Reynolds stress
relation can be used to improve the predictions in wake regions. For instance, a quadratic
closure can be employed [37]:
 
2 k3 1
Rij = ui uj = kδij − νt Sij + c1 Cµ 2 Sik Sjk − Skl Skl δij +
3  3
 
k3 k3 1
+ c2 Cµ 2 (Ωik Sjk + Ωjk Sik ) + c3 Cµ 2 Ωik Ωjk - Ωkl Ωkl δij (47)
  3
where Sij = Ui,j + Uj,i and Ωij = Ui,j −Uj,i , according to Craft et al. [37].

VKI - 17 -
2 THEORY
2.3 Generalization of the ABL model for the case of obstacles immersed in the flow

Figure 6: Variation of the turbulence model parameter Cµ for an undisturbed ABL (top),
using a prescribed region Beranek [2] (middle) and using an automatic switching function
(bottom).

VKI - 18 -
3 APPLICATIONS

3 Applications
The methodologies for the simulation of the ABL are tested and presented in this section.
The first objective is that of verifying that the proposed approaches can provide the
desired uniformity of velocity and turbulent quantities for an empty domain case. Then,
the simulation over more complex terrains and around obstacles will be considered

3.1 Empty fetch at wind-tunnel and full scale


The first test case is a wind tunnel scale ABL Leitl [7], fully characterized in terms of
velocity and turbulence intensity measurements, investigated by several authors. The
results presented here refer to Parente et al. [3, 6]. The computational domain consists
of a two-dimensional box of 4 m length and 1 m height, with a grid of 400x71 cells.
The domain (Figure 7a) is discretized with a grid uniform in the longitudinal direction
and stretched in the vertical one to have the centre point of the wall adjacent cell at a
height of 0.0025 m. At the inlet boundary profiles of velocity, turbulent kinetic energy
and dissipation rate are specified. The inlet conditions proposed by Richards and Hoxey
[4], Yang et al. [5] and Parente et al. [6] are tested for turbulent kinetic energy (Table
2), and the turbulence model parameters are modified accordingly. A pressure outlet
condition is applied for the outlet section. Both standard and modified formulations for
the wall function are applied at the lower boundary whereas a constant stress is applied
to the upper boundary, following the recommendation of Richards and Hoxey [4]. The
second test case is the full scale ABL adopted for the blind-test exercise on the CFD
modelling of wind loading on the full-scale Silsoe cube Richards et al. [38]. Results from
Parente et al. [3] are presented for a 2D domain (Figure 7b) of 5000 m length and 500
m height, with a grid of 500x50 cells. The generated grid is uniform in the longitudinal
direction and stretched in the vertical direction to have the centre point of the wall
adjacent cell at a height of 1 m. The velocity profile is taken from the blind-test exercise
Richards et al. [38] whereas the k profile is specified using both the relations proposed
by Richards and Hoxey [4] and the one obtained using the semi-empirical correlation by
Brost and Wyngaard [10]. As for the wind tunnel scale test case, a constant stress is
applied to the upper boundary, while only the modified wall treatment is used to model

Table 2: Fitting parameters for velocity and turbulent kinetic energy inlet profiles accord-
ing to Yang et al. [5], Parente et al. [6] and turbulent model parameters.
Parameter Richards and Hoxey [4] Yang et al. [5] Parente et al. [6]
u∗ 0.374
z0 0.00075
C1 0 -0.11 -0.04
u4∗
C2 Cµ
0.53 0.52
u4∗
Cµ 0.09 k2
∂k
ρu∗ κ ∂ [(z+z0 ) ∂z ]
Sk ∂z
-
σk √ 
ρu4∗ (C2 −C1 ) Cµ 1
S (z+z0 )2 κ2
− σ

VKI - 19 -
3 APPLICATIONS 3.1 Empty fetch at wind-tunnel and full scale

Figure 7: Computational domain and main boundary conditions applied for the numerical
simulation of the unperturbed ABL at wind tunnel (a) and full (b) scale.

the rough terrain properties, following the conclusions of the results obtained from the
wind-tunnel scale simulations. The presented results were replicated using both FLUENT
and OpenFOAM codes, with a double precision, pressure based solver. The standard
discretization scheme is applied to pressure, while second order schemes are adopted for
momentum and turbulence quantities, and the SIMPLE algorithm is chosen for pressure-
velocity coupling. The simulations are considered converged when the residuals level out,
resulting in a decrease of at least six orders of magnitude.
Figure 8 shows the profiles of velocity, turbulent kinetic energy and dissipation rate at
the inlet and outlet section of the computational domain in Figure 7, for the inlet condi-
tions specified by Richards and Hoxey [4] (Figure 8a-c) and Yang et al. [5]. (Figure 8d-f).
For the case of a constant profile for turbulent kinetic energy Richards and Hoxey [4], the
results obtained with the standard wall function with a constant Cs defined according to
Equation (18) are compared to those provided by the modified law of the wall Parente
et al. [3]. The turbulence model parameters used in combination with the different sets of
boundary conditions are summarized in Parle 2. The results (Figure 8a-c) indicate that
the two approaches are comparable; however, the modified wall approach better preserves
the homogeneity of velocity and turbulence throughout the domain. In particular, Figure
8b-c confirms that the modification of Cs only affects the velocity profile, whereas the
turbulent dissipation rate at the wall is overestimated, as a result of omitting the aero-
dynamic roughness in the denominator of Equation (38). Consequently, the turbulent
kinetic energy at the wall adjacent cell is slightly underestimated. Therefore, the formu-
lation of the wall function alone already has a non-negligible impact on the simulation
results, for what concerns the reproduction of the turbulent quantities. For the case of a
non constant k profile, the comparison between the profiles of velocity, turbulent kinetic
energy and dissipation rate at the inlet and outlet sections of the domain indicate that the
proposed source terms and the implemented wall function ensure the desired longitudinal
homogeneity. The highest differences are observed for the k profile, but they are below
4% in all cases. Figure 9 shows the evolution of the wall shear stress along the longitu-
dinal coordinate for the two cases of constant Richards and Hoxey [4] and variable inlet
k profiles Yang et al. [5]. It can be remarked that the modified wall function produces a

VKI - 20 -
3.1 Empty fetch at wind-tunnel and full scale 3 APPLICATIONS

Figure 8: Profiles of velocity, turbulent kinetic energy and turbulent dissipation rate at
inlet and outlet section of the computational domain (Figure 7), obtained when applying
inlet conditions given by Equations 10-12 (a-c) and Equations 10,15 and 12 (d-f). STD
WF = Standard Wall Function; MOD WF = Modified Wall Function [3].

VKI - 21 -
3 APPLICATIONS 3.1 Empty fetch at wind-tunnel and full scale

constant and realistic wall shear stress, after a short adaptation length. The wind tunnel
analysis is completed by the assessment of the approach proposed by Parente et al. [6].
Figure 10 presents shows the comparison between inlet and outlet profiles of velocity, tur-
bulent kinetic energy and turbulent dissipation rate when the approach by Parente et al.
[6] is employed. It can be observed how the overall modelling approach is able to ensure
longitudinal and vertical homogeneity for the simulated boundary layer. The highest dif-
ferences are observed for the k profile; however, these are safely below 2% in all cases.
The results shown in Figure 10 a-c are further  confirmed by Figure 10d, which shows the
κz ∂u
non-dimensional velocity gradient Φ = u∗ ∂z , which allows assessing the deviation of the
simulated velocity profile from the ideal logarithmic one. Indeed, the Φ profile obtained
at the outlet section of the domain closely follows the theoretical non-dimensional velocity
gradient. It can be concluded that the new set of boundary conditions, together with the
rough wall function formulation and the modification of the turbulence model provide a
comprehensive and consistent modelling approach for the neutral ABL. Figure 11 shows
the profiles of velocity, turbulent kinetic energy and dissipation rate at the inlet and outlet
section of the full-scale computational domain shown in Figure 7b, obtained when apply-
ing inlet conditions by Richards and Hoxey [4] (a-c) and Brost and Wyngaard [10] (d-f).
Table 3 shows the fitting parameters for velocity and turbulent kinetic energy , as well as
the turbulent model parameters for the two sets of inlet conditions employed. Similarly
to the wind-tunnel scale, it can be concluded that the modelling approach summarized in
Table 3 can be successfully applied also to the numerical simulation of full scale, unper-
turbed ABL flows. As far as the longitudinal velocity component is concerned, maximum
deviations of about 3% are observed for both sets of inlet conditions tested. The turbulent
kinetic energy profile specified at the inlet section is also very well maintained throughout
the computational domain: maximum differences of about 5% and 3% are observed when
applying the Richards and Hoxey [4], Brost and Wyngaard [10] boundary condition for
k, respectively. It is also interesting to observe that the semi-empirical correlations by
Brost and Wyngaard [10] result in a consistent level of turbulent kinetic energy, although
slightly higher (˜30%), than the one by Richards and Hoxey [4]. Finally, the wall shear
stress along the longitudinal coordinate (Figure 12) indicates that a constant and realistic
shear stress is retrieved also for the full-scale ABL simulation.

Table 3: Fitting parameters for velocity and turbulent kinetic energy inlet profiles accord-
ing to Richards and Hoxey [4], Brost and Wyngaard [10] and turbulent model parameters.
Parameter Richards and Hoxey [4] Brost and Wyngaard [10]
u∗ [m/s] 0.325
z0 [m] 0.01
u2∗ u2∗ 
k √ 2
8.7 − 6 hz

u4∗
Cµ 0.09 k2
ρu∗ κ ∂ [ (z+z0 ) ∂k
∂z ]
Sk -
 √ σk  ∂z
ρu4∗ (C2 −C1 ) Cµ 1
S (z+z0 )2 κ2
− σ

VKI - 22 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

Figure 9: Wall shear stress as a function of the axial coordinate. Solid black line: theo-
retical value, ρu2∗ . Dashed red line: Richards and Hoxey [4] profile. Blue dots: Yang et al.
[5] profile [3].

3.2 Flow around a ground-mounted building


The simulation of the flow around a single, rectangular building [7] is considered to further
assess the modelling proposed in [3, 6]. The building (Figure 13) has width 0.1m, length
0.15m, height H = 0.125m and 4 source elements on the leeward building side. The
computational domain inlet boundary is set 1m upstream of the building, where ABL
profiles are measured in the wind tunnel, whereas the outlet boundary is located 4m
downstream of the building. As the model is symmetrical with respect to the plane
y = 0m, only half of the domain is represented. The width and height of the domain are
0.65m and 1m respectively, corresponding to the wind tunnel size. A structured mesh
consisting of 1.7 million cells (200 x 114 x 107 elements) is applied. The height of the
ground adjacent cell is 0.00075m, which correspond to one cell between the ground surface
and the lower edge of the source elements. The wall function formulation (Section 2.2)
based on the use of the aerodynamic roughness [3, 6] allows maintaining such resolution
in the entire computational domain. Previous investigations by Gorlé et al. [39] employed
the standard rough wall formulation with a modification of the roughness constant CS
[1]. As a consequence, the computational grid had to be non-conformal, with coarser
wall adjacent cells in the far field, to allow the correct reproduction of the incoming
ABL without exceeding the requirement CS ≤ E (Section 2.2). Hence the test case is
well suited to demonstrate the advantage of a wall function based on the aerodynamic
roughness, which does not require the definition of CS and kS and allows preserving the
same high near-wall resolution throughout the entire domain (Figure 14). The presence
of an obstacle immersed in the ABL flow has important modelling implications. The
turbulence model formulation presented in Section 1 is valid for an unperturbed ABL.
As indicated in Section 2.3, the application of turbulence model modifications derived
for unperturbed ABL flows in regions of the flow affected by obstacles can negatively
affect model predictions. The problem can be alleviated by dividing the computational
domain in two regions: a region unaffected by the building, where the modified turbulence
model parameter Cµ and the source terms Sk (if needed) and S are applied, and a region

VKI - 23 -
3 APPLICATIONS 3.2 Flow around a ground-mounted building

Figure 10: Profiles of velocity (a), turbulent kinetic energy (b), turbulent dissipation rate
(c) and non-dimensional velocity gradient (d) at inlet and outlet section of the computa-
tional domain, obtained with Equations 10, (23) and 12 [6].

VKI - 24 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

Figure 11: Profiles of velocity, turbulent kinetic energy and turbulent dissipation rate at
inlet and outlet section of the computational domain (Figure 7), obtained when applying
inlet conditions given by Equations 10-12 (a-c) and Equations 10, (29) and 12 (d-f).
Results obtained with the wall function formulation by [3] [3].

VKI - 25 -
3 APPLICATIONS 3.2 Flow around a ground-mounted building

Figure 12: Wall shear stress as a function of the axial coordinate. Solid black line:
theoretical value, ρu2∗ . Dashed red line: Richards and Hoxey [4] profile. Blue dots: Yang
et al. [5] profile.

Figure 13: Building geometry and location of measurement planes for the flow around the
obstacle [7].

Figure 14: Computational domain and grid for the flow around the obstacle [7].

VKI - 26 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

influenced by the building, called building influence area (BIA), where other turbulence
model formulations of the k −  model are used (Section 2.3). The results in this section
were obtained with Fluent 6.3 using the steady, 3d, double precision, pressure based solver.
The standard discretization scheme was applied to pressure, while second order schemes
were adopted for momentum and turbulence quantities, and the SIMPLE algorithm for
pressure-velocity coupling.
Results are first presented for the case of a BIA with a pre-defined size, as discussed
in Section 2.3 and shown in Figure 5. The quality of the results is assessed using the hit
rate metric, defined for both velocity and turbulent kinetic energy on the basis of the N
available measurements, δI :
XN
q= δi (48)
i=1

with (
VCF D −VT EST
1 if VT EST ≤ 0.25 or |VCF D − VT EST | ≤ W
δi = (49)
0 else
The hit rate indicates the fraction of N measurement points at which the CFD results are
within 25% of the measurement data or within the uncertainty interval, W , of the data
Franke et al. [24]. The value used for the uncertainty interval W is 0.012 for the non-
dimensional velocity and for the non-dimensional turbulence kinetic energy. The selected
values affect the absolute value of the hit rate, but they are verified not to change the
observed model performance trends.
Table 4 shows both the total hit rate for all measurement points and local hit rates
for selected points (upstream, on the side, on top and downstream of the building). In
addition to the two prescribed BIA size models (PS1 and PS2), the result presented in
Gorlé et al. [39] for settings similar to the present PS1 and the result obtained using
the classic Richards and Hoxey [4] boundary conditions and the ABL model everywhere
(indicated as UABL). The test conditions for the different cases are summarized in Table
5. For velocity, an improvement in the hit rate is obtained using any of the modified
approaches as compared to the Richards and Hoxey [4] boundary conditions. Moreover,
we observe no major differences in the hit rate for u/Uref between the PS1 or PS2 settings.
For the turbulence kinetic energy, however, the PS2 setting produces a significant increase
of the hit rate (59%) with respect to the PS1 settings (52%). An important improvement
is also observed with respect to the reference case from Gorlé et al. [39], which provides
a hit rate of 47%, and the results obtained using the Richards and Hoxey [4] boundary
conditions (51%), in all regions of the flow field.
Figures 15 and 16 present the contour plots for the non-dimensional velocity and
turbulent kinetic energy, respectively, at planes y = 0 and z = 0.035m. The experimental
measurements are compared to the results obtained applying the PS1 and PS2 model.
It can be observed that the main difference between the PS1 and PS2 settings is found
in front of the building, with the PS2 setting resulting in a slightly better prediction of
the upstream separation bubble and a smaller over-prediction of the turbulent kinetic
energy. Both settings over-predict the size of the wake region by about 30% ( 2.2H
versus 1.7H). However, a larger overestimation is observed when applying the standard
model and the Richards and Hoxey [4] boundary conditions ( 2.8H versus 1.7H). To

VKI - 27 -
3 APPLICATIONS 3.2 Flow around a ground-mounted building

Table 4: Hit rate values for non-dimensional velocity and turbulent kinetic energy for the
prescribed BIA size approach., varying the turbulence model settings [3].
U
Uref
upstream size/top downstream all
UABL 0.9 0.75 0.49 0.66
Gorlé et al. [39] 0.87 0.81 0.59 0.71
PS1 0.89 0.78 0.56 0.7
PS2 0.9 0.79 0.58 0.71
k
2
Uref
upstream size/top downstream all
UABL 0.36 0.51 0.59 0.51
Gorlé et al. [39] 0.54 0.31 0.49 0.47
PS1 0.54 0.4 0.54 0.52
PS2 0.58 0.54 0.62 0.59

Table 5: Test cases and corresponding model settings for the numerical simulation of the
flow around a bluff-body [7], using the prescribed BIA size approach. TM=Turbulence
Model.
Test case Inlet conditions Wall function BIA BIA TM
RH Richards and Hoxey [4] Richards and Hoxey [4] - -
Gorlé et al. [39] Yang et al. [5] Blocken et al. [1] Half sphere KL
PS1 Yang et al. [5] Parente et al. [3] Half sphere KL
PS2 Yang et al. [5] Parente et al. [3] Cut half sphere KL

VKI - 28 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

Figure 15: Contour plots of non-dimensional velocity on the planes y = 0 (left) and
z = 0.035m (right). Experimental measurements are compared to the results obtained
applying the PS1 and PS2 model [3].

Figure 16: Contour plots of non-dimensional turbulence kinetic energy on the planes
y = 0(left) and z = 0.035m (right). Experimental measurements are compared to the
results obtained applying the PS1 and PS2 model [3].

VKI - 29 -
3 APPLICATIONS 3.2 Flow around a ground-mounted building

allow a more quantitative comparison between experiments and numerical simulations,


the calculated and measured non-dimensional vertical profiles of velocity and k on the
symmetry plane are shown in Figure and Figure , respectively, for different locations
upstream, over and downstream of the building. The different configurations (PS1 and
PS2) used to prescribe the BIA are compared to the experimental measurements. It can
be observed how the difference between the two configurations mainly affects the turbulent
kinetic energy field (Figure ), in the region upstream of the building. In particular, the
application of the ABL model at the axial location x = −0.075m, allows significantly
reducing the overestimation of turbulent kinetic energy. This effect remains at the first
location on top of the building (x = −0.04m). On the other hand, on the roof further
downstream (x = 0m and x = 0.04m) and behind of the building (x = 0.105m and
x = 0.3m), the two model predictions become comparable, almost collapsing onto a single
line.
The approach based on the a priori prescription of the BIA can be modified by consid-
ering the approach discussed in Section 2.3, using the blending function given by Equation
(43) to achieve a gradual transition between the different flow regions, as shown in Figure
6. Depending on the value of the exponent α controlling the transition between the two
regions, different models can be obtained. If α = 1 and α = 2 are chosen, a linear and a
quadratic switch are obtained, indicated by the acronyms ASL and ASQ, respectively. A
comparison between PS and ASQ is shown in the following for the flow around the bluff-
body. As a reference, a test case without separation between flow regions is also included
and denoted Undisturbed ABL (UABL). The test conditions for the different cases are
summarized in Table 5. The calculated and measured non-dimensional vertical profiles of
velocity and k are shown in Figures 19 and 20, respectively, for different axial locations
over and downstream of the building. Results were obtained applying the UABL, ASQ
and PS models. It can be observed how the difference between the models becomes more
pronounced downstream of the building, in the wake region (x = 0.105m and x = 0.3m).
In particular, the automatic quadratic switch (ASQ) allows reducing the under-prediction
of velocity (Figure 19) and the over prediction of turbulent kinetic energy (Figure 19)
downstream of the building, with respect to the UABL model. It can be also remarked
that the results obtained with ASQ are very similar to those provided by the prescribed
switch model (PS) for what concerns the velocity distribution. On the other hand, the
ASQ model results in a better prediction of turbulent kinetic energy (Figure 19), in par-
ticular at location x = 0.3m, where the prescribed switch model (PS) shows a very strong
over-prediction of the k values. Nevertheless, the peak of k at x = 0.105m is better cap-
tured by the PS model. Finally, Table 4 lists the hit rate values for the non-dimensional
velocity and turbulent kinetic energy.

Table 6: Test cases and corresponding model settings for the numerical simulation of
the flow around a bluff-body [7], using the automatic switch approach BIA for the BIA.
TM=Turbulence Model.
Test case Inlet conditions Wall function BIA BIA TM
UABL Parente et al. [6] Parente et al. [3] - -
PS Parente et al. [6] Parente et al. [3] Half sphere Kato and Launder [33]
ASL, ASQ Parente et al. [6] Parente et al. [3] Automatic Kato and Launder [33]

VKI - 30 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

Figure 17: Experimental and numerical profiles of non-dimensional velocity upstream,


over and downstream of the obstacle. Solid line: experimental data [3]. Dashes: PS1
configuration. Short dashes: PS2 configuration .

Figure 18: Experimental and numerical profiles of non-dimensional turbulent kinetic en-
ergy upstream, over and downstream of the obstacle [3]. Solid line: experimental data.
Dashes: PS1 configuration. Short dashes: PS2 configuration .

Figure 19: Experimental and numerical profiles of non-dimensional velocity over and
downstream of the obstacle. Solid line: experimental data [6]. Dashes: UABL model.
Short dashes: PS model. Dots: ASQ model.

VKI - 31 -
3 APPLICATIONS 3.2 Flow around a ground-mounted building

Figure 20: Experimental and numerical profiles of non-dimensional turbulent kinetic en-
ergy over and downstream of the obstacle. Solid line: experimental data [6]. Dashes:
UABL model. Short dashes: PS model. Dots: ASQ model.

Table 7 shows the hit rates for all investigated conditions. The first remark is that the
hit rates obtained applying the UABL and ASQ models are very similar. In other words,
it is not possible to distinguish which of the two model settings provides the best results
by just comparing the hit rate values. This is obviously due to the global nature of the hit
rate metric, which does not provide any information regarding the local prediction of flow
features. To better clarify this aspect Figure 21 shows the local hits for the turbulent

kinetic energy field, obtained applying the ASQ and UABL models, respectively. It is
clear that the quadratic automatic switch (ASQ) provides better results in the wake,
which is confirmed by the increased number of hits. However, the global hit rate values
for the two cases are very similar due to the poorer performances of the ASQ model
at the border of the wake, where the turbulent kinetic energy is over-estimated. As far
as the other approaches are concerned, both the ASQ and the PS models optimize the
reproduction of the velocity field, but penalize significantly the prediction of turbulent
kinetic energy. To conclude, the ASQ model can be regarded as the best performing
option for the present investigation. In particular, it ensures the best compromise for
what concerns the prediction of velocity and turbulent kinetic energy within the BIA,
providing hit rate values higher than 60% for both fields.
Finally, to improve the predictions in the wake, a non-linear closure by Craft et al.
[37] (Section 2.3) has been tested [11] and benchmarked against the results provided by
the k −  model using the Kato-Launder correction, as summarized in Table 8. It can be
observed that the mean stream-wise velocity and turbulent kinetic energy hit rates do not
improve much with respect to the ASQ model. For linear models, the increase in turbulent
kinetic energy hit rates had the consequence of worsening flow field results. In the present
case, it appears that an improvement in both direction is achieved. The approach has the
potential of providing satisfactory results, but the transition between the two flow regions
appears not to be currently optimal. It can be observed that a sharper transition (α = 5)
improves the prediction of the velocity field as well as of turbulent kinetic energy, both
upstream and downstream of the building. On the building top, the increase of α has
the effect of worsening the predictions, suggesting that a multi-zone approach should be
probably pursue. This is currently under investigation.

VKI - 32 -
3.2 Flow around a ground-mounted building 3 APPLICATIONS

Table 7: Hit rate values for non-dimensional velocity and turbulent kinetic energy for the
automatic switch approach, varying the turbulence model settings [6].
U
Uref
upstream size/top downstream all
UABL 0.87 0.75 0.45 0.63
ASL 0.87 0.75 0.49 0.65
ASQ 0.85 0.77 0.45 0.62
PS1 0.9 0.79 0.58 0.71
k
2
Uref
upstream size/top downstream all
UABL 0.60 0.51 0.65 0.61
ASL 0.58 0.50 0.51 0.53
ASQ 0.59 0.55 0.63 0.61
PS1 0.53 0.28 0.28 0.35

Figure 21: Local hit rates for the non-dimensional turbulent kinetic energy, applying the
ASQ model (left) and the UABL model (right) [6].

VKI - 33 -
3 APPLICATIONS 3.3 Flow over complex terrains, wind-tunnel and full-scale hills

Table 8: Hit rate values for non-dimensional velocity and turbulent kinetic energy for the
automatic switch approach, varying the turbulence model settings [11].
U
Uref
upstream size/top downstream all
ASQ 0.85 0.77 0.45 0.62
Craft, α = 2 0.90 0.76 0.50 0.66
Craft, α = 5 0.91 0.76 0.52 0.68
k
2
Uref
upstream size/top downstream all
ASQ 0.59 0.55 0.63 0.61
Craft, α = 2 0.62 0.40 0.61 0.57
Craft, α = 5 0.63 0.38 0.66 0.61

3.3 Flow over complex terrains, wind-tunnel and full-scale hills


In this section, the comprehensive modelling approach for ABL flows is assessed on com-
plex terrains such rough hills. The approach by Parente et al. [6] was tested against
the measurements obtained in the thermally stratified wind tunnel of The University of
Tokyo, using three-dimensional laser doppler anemometry [40]. The study was reported
by Balogh et al. [8].The model is an axisymmetric hill, whose shape is given by:
(  
hmax 12 1 + cos r2πr if r < rmax
h (r) = max

0 else

with the radius at the hill base of rmax = 0.42m and with the height at the hill-top
hmax = 0.2m. The hill model was positioned 2m downstream of the test section inlet
of the 6 x 2.2 x 1.8m wind tunnel. The computational domain contains the entire test
section, with the origin set x = 0, y = 0. The hill and wind tunnel floor are modeled
as rough walls with the same aerodynamic roughness of z0 = 0.00122m, while smooth
wall boundary conditions are applied on the ceiling and side walls. At the downwind
side, a pressure outlet boundary condition was used. The fitting parameters for velocity
and turbulent kinetic energy [6] inlet profiles are shown in Table 9. The computational
mesh is composed of 200 x 87 x 60 cells, resulting in 1.044 million hexahedral elements
refined horizontally at the near-field of the hill. Numerical simulations were performed
with Fluent and OpenFOAM solver [8]. To allow a meaningful comparison between Flu-
ent and OpenFOAM, numerical settings were selected to be as similar as possible (see [8]
for more details). Figure 22 shows the measured and computed velocity and turbulent
kinetic energy profiles along the longitudinal direction, provided by the comprehensive
approach [6] with a blending exponent α = 3 for the automatic switch. Such a value of
α was chosen as it provided the best compromise between velocity and turbulent kinetic
energy predictions. It can be observed that OpenFOAM simulations always present a
large underestimation of k downstream of the hill, whereas Fluent provides calculated
values more in agreement with the measurements. However, this effect is due to the large
overestimation of the size of the separation bubble by Fluent, which results in erroneous
velocity predictions but turbulent kinetic energy levels closer to the measured ones. A

VKI - 34 -
3.3 Flow over complex terrains, wind-tunnel and full-scale hills 3 APPLICATIONS

Table 9: Fitting parameters for velocity and turbulent kinetic energy inlet profiles accord-
ing to Parente et al. [6] and turbulent model parameters for the 3D hill simulation.
Parameter Value/Expression
u∗ [m/s] 0.0923
z0 [m] 0.00122
C1 [m2 /s2 ] -0.0053
2 2
C2 [m /s ] 0.050
u4∗
Cµ k2
Sk  - 

4
ρu∗ (C2 −C 1 ) Cµ 1
S (z+z )2 κ2
− σ
0

Figure 22: Stream-wise velocity (top) and turbulent kinetic energy profiles (bottom) in
the symmetry plane against measurements obtained on the 3D hill at laboratory scale
using Fluent and OpenFOAM [8].

VKI - 35 -
3 APPLICATIONS 3.3 Flow over complex terrains, wind-tunnel and full-scale hills

comparison of the wall shear stress as a function of the non-dimensional longitudinal coor-
dinate is given in Figure 23, where the circles denote the measurements. The comparison
shows that the simulations are in fair accordance with the observations in the separated
region, whereas an important deviation can be remarked at the first measurement location
at the top of the hill. The latter can be caused by the difference between the roughness
elements mounted on the flat part of the wind tunnel and the hill surface. The wall shear
stress obtained by Fluent better reproduces the experimental data, especially at higher
distances from the hilltop. Table 10 reports the hit rate values for both the velocity and
turbulent kinetic energy prediction with OpenFOAM and Fluent. It can be observed that
the modification of the turbulence model within the wake does not have a significant im-
pact on the results. Instead, the use of MMK results in slightly lower values of hit rates in
most cases (with the exception of the it rate of k for Fluent simulations). Moreover, the
numerical hit rate values consistently show significant differences between OpenFOAM
and Fluent, concerning the reproduction of the separation bubble. Table 11 shows the
normalized errors in the prediction of the separation point, SP , and wake length, W L , for
the 3D hill, using OpenFOAM and Fluent. Results show that, when the comprehensive
approach is applied in OpenFOAM, both the location of the separation point and the
size of the wake show a fairly good agreement with the measurements, regardless of the
turbulence model applied in the wake region, as indicated by the error values listed in
Table 11. Fluent results show, on the other hand, much larger discrepancies. The causes
for that, which may be related to the limited access to the source code granted n Fluent,
are still under investigation.
In practical atmospheric applications the surface is generally not as regular as in the
previous example; therefore, simulations were performed on more complex geometry to
validate the approach and, possibly, propose modifications. In particular, the full scale
measurements obtained over the Askervein hill [41, 42] were chosen to this purpose. This
is a popular case study for validating CFD models for ABL simulations [43, 44, 45, 46, 47].
The Askervein hill has a nearly elliptical form with major and minor axis of 2000m and
1000m. The height of the hill is 116m, and its slopes range from 12 to 25%. The
surrounding area is flat at the upwind side of the hill, and it is hilly at the downwind
side. For describing the surface coverage, the aerodynamic roughness was taken as z0 =
0.0353m, based on the data measured at the reference mast. The computational domain
(Figure 24) has dimensions 6000 x 6000 x 1000m with an origin located at the hill top

Figure 23: Simulated wall shear stress along the symmetry of the domain against theoret-
ical values, extracted from the inlet profile (Inlet τw) and against values extracted from
measured profiles (meas.).[8].

VKI - 36 -
3.3 Flow over complex terrains, wind-tunnel and full-scale hills 3 APPLICATIONS

Table 10: Hit rate values for non-dimensional velocity and turbulent kinetic energy for
the 3D hill simulation, varying the turbulence model settings [8].
OpenFOAM
Wake model Velocity Kinetic energy
STD 0.81 0.49
MMK 0.79 0.47
Fluent
Wake model Velocity Kinetic energy
STD 0.71 0.47
MMK 0.70 0.52

Table 11: Normalized errors in the prediction of the separation point, SP , and wake
length, W L , for the 3D hill. Negative and positive values sign the under- and overestima-
tion respectively[8].
OpenFOAM
Wake model SP [%] W L [%]
STD -6 -13
MMK -9 -5
Fluent
Wake model SP [%] W L [%]
STD -26 -23
MMK -29 32

VKI - 37 -
3 APPLICATIONS 3.3 Flow over complex terrains, wind-tunnel and full-scale hills

(HT). Similarly to a previous study [47], the mesh was generated with 97 x 111 x 30
hexahedral elements using 15 m maximum resolution on the hill and 1m for the first cell
height. The refinement the grid at the hill is shown in Figure 24. Measurements for inlet
conditions are obtained at almost neutral condition, stable wind direction and relatively
high wind speed [48]. The wind direction was 210 degrees, which determines the mesh
orientation. Namely, the inlet is fixed on the western side of the computational domain and
the outlet is the eastern side. On its southern and northern side the symmetry boundary
condition are defined. On the top of the domain, the uniform values, corresponding to
the fitted profiles are imposed, as summarized in Table 12. The measurement campaign
was carried out using sonic anemometers at 10 meters above the surface, in every must
along a given line (denoted as line-A in Figure 24) and at the hill top, thus the velocity
components, such as the turbulent fluctuations for the three directions are available for
this locations.. The comparison between computed and measured horizontal profiles is
shown in Figure 25. The horizontal component (Uh ), vertical velocity (W ) and turbulent
kinetic energy are presented, and the RMS of the velocity is used to characterize the
uncertainty. The agreement between measured and computed velocity can be considered
sufficiently satisfactory. Although the approach overestimates the horizontal velocity at
the far upwind of the hill summit (19.4% for OpenFOAM and 12.1% for Fluent), the
prediction at the top of the hill (Figures 25 and 26) and along the downwind of the
hill is fairly good (underestimation of 2.1% for OpenFOAM and 5.5% for Fluent). As
for turbulent kinetic energy, the results obtained with the comprehensive approach are
significantly better than the ones given by the original approach (resulting in hit rates
below 30%). The differences between the CFD solvers are significant: OpenFOAM results
are better for velocity, whereas Fluent provides a more accurate prediction of turbulent
kinetic energy. Table 13 shows the hit rate of the horizontal velocity and the turbulent
kinetic energy for the OpenFOAM and Fluent codes. A hit-rate value > 99% for velocity
is obtained for the OpenFOAM simulation using the comprehensive approach with the
MMK correction in the wake region, with a corresponding hit rate of about 44% for
turbulent kinetic energy. A comparable HR value for k is obtained only using Fluent.
However, for this case, the hit-rate value for velocity is below 99%, i.e. 81%.

Figure 24: Building geometry and location of measurement planes for the flow around the
obstacle [7].

VKI - 38 -
3.3 Flow over complex terrains, wind-tunnel and full-scale hills 3 APPLICATIONS

Table 12: Fitting parameters for velocity and turbulent kinetic energy inlet profiles accord-
ing to Parente et al. [6] and turbulent model parameters for the Askervein hill simulation.
Parameter Value/Expression
u∗ [m/s] 0.66
z0 [m] 0.0353
C1 [m2 /s2 ] -0.351
2 2
C2 [m /s ] 2.61
u4∗
Cµ k2
Sk  - 

4
ρu∗ (C2 −C 1 ) Cµ 1
S (z+z )2 κ2
− σ
0

Figure 25: Comparison of simulated and measured horizontal and vertical stream veloc-
ity (Uh and W ) and turbulent kinetic energy (k) along line-A, using the comprehensive
approach [6] with α = 3 [8].

Figure 26: Comparison of simulated and measured vertical profiles (U and k) at the hill
summit [8].

VKI - 39 -
4 INFLUENCE OF STABILITY CLASSES

Table 13: Hit rate values for non-dimensional velocity and turbulent kinetic energy for the
Askervein hill simulation, varying the turbulence model settings within the wake region
[8].
OpenFOAM
Wake model Velocity Kinetic energy
STD 0.94 0.31
MMK >0.99 0.44
Fluent
Wake model Velocity Kinetic energy
STD 0.81 0.44
MMK 0.81 0.44

4 Influence of stability classes


The models proposed above (Section 2) are derived under the hypothesis of neutral strat-
ification, i.e. when the the heat flux from the ground is equal to zero. Recently [29],
attempts have been made to extend ABL models for non-neutral conditions. From the
Monin-Obukhov similarity theory, the universal profiles of velocity and temperature can
be expressed as:    
u∗ z z
U= ln + φm −1 (50)
κ z0 L
   z 
T∗ z g
T = TW + ln + φm − 1 − (z − z0 ) (51)
κ z0 L cp
2
where the Monin-Obukhov length, L = uκgT ∗ Tw

, is an estimate of the height where the
turbulent dissipation due to the buoyancy is comparable with the shear stress production
of turbulence, Tw is the surface temperature, qw the surface heat flux, g the gravitational
acceleration module and Φm is a function that depends on z and L. For neutral and stable
stratification Φm = 1 + 5 Lz . However, for neutral stratification, the heat flux from the
ground is zero and the Monin-Obukhov length is infinite; the parameter Φm tends to 1,
thus simplifying the form of the inlet profiles. In fact, the friction temperature T∗ tends
to zero and temperature is constant along z. In stable conditions Φm is different than 1
and the temperature profile is not homogeneous. In such conditions, the following profiles
for k and  become [29]:
s 
u2∗ Φe Lz
k=p  (52)
Cµ Φm Lz

u3∗
= Φe (53)
κz
with
z
Φe = 1 + 4 . (54)
L
Similarly to what discussed for the neutral ABL (Section 2), two source terms need to be
added to the k −  turbulence model, to ensure that the inlet profile satisfy the transport

VKI - 40 -
4 INFLUENCE OF STABILITY CLASSES

equations. h  i
µT ∂k
∂ µ+ σk ∂z
Sk (z) = − (55)
∂z
" p r  #
ρu4∗ (C2 − C1 ) Cµ Φm 1 2 1 T∗
S (z) = 2 Φe − − 2 + . (56)
(z + z0 ) κ2 Φe σ Φm Φm κT
The implementation of such approach has been recently performed in OpenFOAM [49].
Figure 27 shows the profiles of (a) velocity, (b) turbulent kinetic energy, (c) turbulent
dissipation rate, (d) turbulent viscosity and e) temperature at the inlet and outlet section
of a 2D computational domain (60m high and 400m long ), as well as the (f) shear stress
at the domain wall. The inlet conditions are taken from [9]. Results indicate that the
modification of the turbulence model by means of the source terms given by Equations
(55) and (56) is able to provide the desired uniformity of velocity and turbulent quantities
throughout the domain.

VKI - 41 -
4 INFLUENCE OF STABILITY CLASSES

(a) (b)

(c) (d)

(e) (f)

Figure 27: Profiles of (a) velocity, (b) turbulent kinetic energy, (c) turbulent dissipation
rate, (d) turbulent viscosity and e) temperature at the inlet and outlet section of a 2D
computational domain (60m high and 400m long ), and (f) shear stress at the wall. Inlet
conditions taken from [9].

VKI - 42 -
REFERENCES REFERENCES

References
[1] B. Blocken, T. Stathopoulos, J. Carmeliet, CFD simulation of the atmospheric
boundary layer: wall function problems, Atmospheric Environment 41 (2007) 238–
252.

[2] W. Beranek, General rules for the determination of wind environment, Proceedings
of the 5th international conference on wind engineering (Fort Collins, Colorado, USA,
1979) 225–234.

[3] A. Parente, C. Gorlé, J. van Beeck, C. Benocci, Improved k-epsilon model and wall
function formulation for the rans simulation of abl flows, Journal of Wind Engineering
and Industrial Aerodynamics 99 (2011) 267–278.

[4] P. J. Richards, R. P. Hoxey, Appropriate boundary conditions for computational


wind engineering models using the k-epsilon turbulence model, Journal of Wind En-
gineering and Industrial Aerodynamics 46-47 (1993) 145 – 153.

[5] Y. Yang, M. Gu, S. Chen, X. Jin, New inflow boundary conditions for modelling the
neutral equilibrium atmospheric boundary layer in computational wind engineering,
Journal of Wind Engineering and Industrial Aerodynamics 97 (2) (2009) 88 – 95.

[6] A. Parente, C. Gorlé, J. van Beeck, C. Benocci, A comprehensive modelling approach


for the neutral atmospheric boundary layer: Consistent inflow conditions, wall func-
tion and turbulence model, Boundary-Layer Meteorology 140 (2011) 411 – 428.

[7] B. Leitl, Cedval at hamburg university.


URL http://www.mi.uni-hamburg.de/cedval.

[8] M. Balogh, A. Parente, C. Benocci, Rans simulation of abl flow over complex terrains
applying an enhanced k-epsilon model and wall function formulation: Implementation
and comparison for fluent and openfoam, Journal of Wind Engineering and Industrial
Aerodynamics 104 - 106 (2012) 360 – 368.

[9] M. Magnusson, A. S. Smedman, Influence of atmospheric stability on wind turbine


wakes, Wind Engineering 18 (1994) 139 – 151.

[10] R. A. Brost, J. C. Wyngaard, A model study of the stably stratified planetary bound-
ary layer, Journal of the Atmospheric Sciences 35 (8) (1978) 1427–1440.

[11] A. Rakai, Simulation of flow around a building-shaped obstacle with openfoam, Tech.
rep., von Karman Institute for Fluid Dynamics (2012).

[12] K. B. Shah, J. H. Ferziger, A fluid mechanicians view of wind engineering: Large eddy
simulation of flow past a cubic obstacle, Journal of Wind Engineering and Industrial
Aerodynamics 67 (1997) 211–224.

[13] H. C. Lim, T. Thomas, I. P. Castro, Flow around a cube in a turbulent boundary


layer: Les and experiment, Journal of Wind Engineering and Industrial Aerodynam-
ics 97 (2) (2009) 96–109.

VKI - 43 -
REFERENCES REFERENCES

[14] Z.-T. Xie, I. P. Castro, Les and rans for turbulent flow over arrays of wall-mounted
obstacles, Flow, Turbulence and Combustion 76 (3) (2006) 291–312.

[15] Z.-T. Xie, I. P. Castro, Large-eddy simulation for flow and dispersion in urban streets,
Atmospheric Environment 43 (13) (2009) 2174–2185.

[16] A. Dejoan, J. Santiago, A. Martilli, F. Martin, A. Pinelli, Comparison between large-


eddy simulation and reynolds-averaged navier–stokes computations for the must field
experiment. part ii: Effects of incident wind angle deviation on the mean flow and
plume dispersion, Boundary-Layer Meteorology 135 (1) (2010) 133–150.

[17] W. Rodi, Comparison of les and rans calculations of the flow around bluff bodies,
Journal of Wind Engineering and Industrial Aerodynamics 69-71 (0) (1997) 55
– 75, <ce:title>Proceedings of the 3rd International Colloqium on Bluff Body
Aerodynamics and Applications</ce:title>.
URL http://www.sciencedirect.com/science/article/pii/
S0167610597001475

[18] T. Cebeci, P. Bradshaw, Momentum Transfer in Boundary Layers, Hemisphere Pub-


lishing Corporation, Washington, 1977.

[19] J. Nikuradze, Stromungsgesetze in rauhen rohren, Forschungsheft 361 (1933).

[20] A. Riddle, D. Carruthers, A. Sharpe, C. McHugh, J. Stocker, Comparisons between


fluent and adms for atmospheric dispersion modelling, Atmospheric Environment
38 (7) (2004) 1029 – 1038.

[21] J. Franke, C. Hirsch, A. G. Jensen, H. W. Krus, M. Schatzmann, P. S. Westbury,


S. D. Miles, J. A. Wisse, N. G. Wright, Recommendations on the use of cfd in
wind engineering, in: C. A. C14 (Ed.), Proceedings of the International Conference
on Urban Wind Engineering and Building Aerodynamics, von Karman Institute,
Rhode-Saint-Genèse, Belgium, 2004, pp. C1.1–C1.11.

[22] T. S. B. Blocken, J. Carmeliet, Cfd evaluation of wind speed conditions in passages


between parallel buildings - effect of wall-function roughness modifications for the
atmospheric boundary layer flow, Journal of Wind Engineering and Industrial Aero-
dynamics 95 (2007) 941–962.

[23] D. Hargreaves, N. Wright, On the use of the k-epsilon model in commercial cfd soft-
ware to model the neutral atmospheric boundary layer, Journal of Wind Engineering
and Industrial Aerodynamics 95 (5) (2007) 355 – 369.

[24] J. Franke, A. Hellsten, K. H. Schlunzen, B. Carissimo, The cost 732 best practice
guideline for cfd simulation of flows in the urban environment: a summary, Interna-
tional Journal of Environment and Pollution 44 (1-4) (2011) 419–427.
URL http://www.ingentaconnect.com/content/ind/ijep/2011/00000044/
F0040001/art00047

[25] Z. Xie, P. R. Voke, P. Hayden, A. G. Robins, Large-eddy simulation of turbulent flow


over a rough surface, Boundary-Layer Meteorology 111 (3) (2004) 417–440.

VKI - 44 -
REFERENCES REFERENCES

[26] C. Gorlé, J. van Beeck, P. Rambaud, G. V. Tendeloo, Cfd modelling of small particle
dispersion: The influence of the turbulence kinetic energy in the atmospheric
boundary layer, Atmospheric Environment 43 (3) (2009) 673 – 681.
URL http://www.sciencedirect.com/science/article/pii/
S1352231008009084

[27] A. Parente, C. Benocci, On the rans simulation of neutral abl flows, in: Proceed-
ings of the Fifth International Symposium on Computational Wind Engineering
(CWE2010), Chapel Hill, North Carolina, USA, 2010, pp. 1–8.

[28] A. Parente, C. Gorle, J. V. Beck, C. Benocci, Rans simulation of abl flows: application
of advanced wall boundary conditions to configurations with mixed rough and smooth
surfaces, in: Proceedings of the Fifth International Symposium on Computational
Wind Engineering (CWE2010), Chapel Hill, North Carolina, USA, 2010, pp. 1–8.

[29] M. Pontiggia, M. Derudi, V. Busini, R. Rota, Hazardous gas dispersion: A cfd model
accounting for atmospheric stability classes, Journal of Hazardous Materials 171 (1 -
3) (2009) 739 – 747.

[30] M. Bottema, Turbulence closure model constants and the problems of inactive atmo-
spheric turbulence, Journal of Wind Engineering and Industrial Aerodynamics 67-68
(1997) 897 – 908.

[31] A. Bechmann, Large-eddy simulation of atmospheric flow over complex terrain, Ph.D.
thesis, Technical University of Denmark (2006).

[32] P. J. Richards, S. E. Norris, Appropriate boundary conditions for computational wind


engineering models revisited, Journal of Wind Engineering and Industrial Aerody-
namics 99 (2011) 257–266.

[33] M. Kato, B. E. Launder, The modelling of turbulent flow around stationary and
vibrating square cylinders, in: Proceedings of 9th symposium on turbulence and
shear flows, Kyoto, Japan, 1993, pp. 1–6.

[34] C. J. Yap, Turbulent heat and momentum transfer in recirculating and impinging
flows, Ph.D. thesis, University of Manchester (1987).

[35] M. Tsuchiya, S. Murakami, A. Mochida, K. Kondo, Y. Ishida, Development of a


new k-epsilon model for flow and pressure fields around bluff body, Journal of Wind
Engineering and Industrial Aerodynamics 67-68 (0) (1997) 169 – 182.

[36] B. E. Launder, On the computation of convective heat transfer in complex turbulent


flows, Journal of Heat Transfer 110 (4b) (1988) 1112–1128.

[37] T. J. Craft, B. E. Launder, K. Suga, Development and application of a cubic eddyvis-


cosity model of turbulence, International Journal of Heat and Fluid Flow 17 (1995)
108–115.

[38] P. J. Richards, A. D. Quinn, S. Parker, A 6m cube in an atmospheric boundary layer


flow. part 2. computational solutions, Wind and Structures 5 (2) (2002) 177–192.

VKI - 45 -
REFERENCES REFERENCES

[39] C. Gorlé, J. Beeck, P. Rambaud, Dispersion in the wake of a rectangular building:


Validation of two reynolds-averaged navier–stokes modelling approaches, Boundary-
Layer Meteorology 137 (2010) 115–133.
URL http://dx.doi.org/10.1007/s10546-010-9521-0

[40] T. Takahashi, S. Kato, S. Murakami, R. Ooka, M. F. Yassin, R. Kono, Wind tunnel


tests of effects of atmospheric stability on turbulent flow over a three-dimensional
hill, Journal of Wind Engineering and Industrial Aerodynamics 93 (2) (2005) 155 –
169.

[41] P. A. Taylor, H. W. Teunissen, Askervein ’82: Report on the september/october 1982


experiment to study boundary layer flow over askervein, Technical Report MSRS-83-
8, Meteorological Services Research Branch Atmospheric Environment Service (1983).

[42] P. A. Taylor, H. W. Teunissen, The askervein hill project: Report on the septem-
ber/october 1983, main field experiment, Technical Report MSRB-84-6, Meteorolog-
ical Services Research Branch Atmospheric Environment Service (1985).

[43] G. D. Raithby, G. D. Stubley, P. A. Taylor, The askervein hill project: A finite


control volume prediction on three-dimensional flows over the hill, Boundary-Layer
Meteorology 39 (1987) 107–132.

[44] H. Kim, V. Patel, Test of turbulence models for wind flow over terrain with separation
and recirculation, Boundary-Layer Meteorology 94 (2000) 5–21.

[45] F. A. Castro, J. M. L. M. Palma, A. S. Lopes, Simulation of the askervein flow. part 1:


Reynolds averaged navier stokes equations (k-epsilon turbulence model), Boundary-
Layer Meteorology 107 (3) (20003) 501–530.

[46] F. A. Castro, J. M. L. M. Palma, A. S. Lopes, Simulation of the askervein flow. part


2: Large-eddy simulations, Boundary-Layer Meteorology 125 (1) 85–108.

[47] C. A. V. Rodrigues, Analysis of the atmospheric boundary layer flow over mountain-
ous terrain, Master’s thesis, von Karman Institute for Fluid Dynamics. (2005).

[48] R. Mickle, N. Cook, A. Hoff, N. Jensen, J. Salmon, P. Taylor, G. Tetzlaff, H. Teunis-


sen, The askervein hill project: Vertical profiles of wind and turbulence, Boundary-
Layer Meteorology 43 (1988) 143–169.

[49] B. Nagy, Simulation of stratified atmospheric flows with openfoam, Tech. rep., von
Karman Institute for Fluid Dynamics (2012).

VKI - 46 -

You might also like