You are on page 1of 18

Building and Environment 114 (2017) 148e165

Contents lists available at ScienceDirect

Building and Environment


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

On the accuracy of CFD simulations of cross-ventilation flows for a


generic isolated building: Comparison of RANS, LES and experiments
T. van Hooff a, b, *, B. Blocken a, b, Y. Tominaga c
a
Department of Civil Engineering, KU Leuven, Kasteelpark Arenberg 40 e bus 2447, 3001 Leuven, Belgium
b
Department of the Built Environment, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
c
Department of Architecture and Building Engineering, Niigata Institute of Technology, 1719, Fujihashi, Kashiwazaki, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Accurate and reliable computational fluid dynamics (CFD) simulations are essential for the assessment of
Received 10 November 2016 cross-ventilation of buildings. To determine which CFD models are most suitable, validation studies are
Received in revised form required. A detailed review of the literature indicates that most CFD validation studies only employed the
9 December 2016
3D steady Reynolds-averaged Navier-Stokes (RANS) approach and/or focused on a limited set of flow
Accepted 12 December 2016
Available online 13 December 2016
parameters. Therefore, the objective of this paper is the validation of both 3D steady RANS simulations
and large eddy simulation (LES) of cross-ventilation in a generic isolated enclosure with wind-tunnel
measurements. The evaluation is based on five parameters: mean velocity, turbulent kinetic energy,
Keywords:
Natural cross-ventilation flow
ventilation flow rate, incoming jet angle and incoming jet spreading width. The RANS simulations are
Numerical simulation conducted with the standard k-ε (SKE), RNG k-ε, realizable k-ε (RLZ), SST k-u and RSM turbulence
Atmospheric boundary layer (ABL) flow models, whereas the LES is performed with the dynamic Smagorinsky subgrid-scale model. SST/RNG/
Building aerodynamics RSM reproduce the experimentally observed direction of the incoming jet, but all RANS models fail in
Turbulence model validation reproducing the turbulent kinetic energy, which is too low especially above and below the jet, because
Transient flow features steady RANS does not capture the vertical flapping of the jet. This transient feature is reproduced by LES,
resulting in a better reproduction of all three measured parameters (velocity, turbulent kinetic energy,
volume flow rate). It is concluded that choice of the model (RANS vs. LES) actually depends on which
parameter is the target parameter, noting that the use of LES entails an increase in computational de-
mand with a factor of z80e100.
© 2016 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction ventilation occurs through large openings, which makes the


resulting ventilation flow through an enclosure difficult to predict
1.1. Literature review using analytical or (semi-)empirical models. For example, Straw
et al. [4] showed in their cross-ventilation study that the use of
Cross-ventilation is an important ventilation method since it pressure coefficients in combination with the orifice equation did
can provide a fast and effective way to remove large amounts of not result in an accurate prediction of volume flow rates when
pollutants and to quickly release internal heat from a building or compared to the measured volume flow rates (deviations of
other enclosure (i.e. ventilative cooling). For detailed analysis of 28e32%). The observed deviations were attributed to either as-
natural ventilation in general, and cross-ventilation in particular, sumptions needed for the values of the discharge coefficient and/or
researchers are resorting more and more to the use of computa- values for the pressure coefficients obtained from a sealed body
tional fluid dynamics (CFD) (e.g. Refs. [1e14]). The increased use of that do not correspond with the pressure distribution on the same
CFD for such studies can be attributed to the strong interaction body when (large) openings are present [4]. The coupled CFD
between the outdoor wind flow and indoor airflow when natural simulation (indoor and outdoor airflow modeled simultaneously in
the same computational domain) reported in their publication did
provide far more accurate results, with a deviation of only 3e9%
from the measurements [4].
* Corresponding author. Building Physics Section, Department of Civil Engineer-
Most natural ventilation studies presented in the past used the
ing, KU Leuven, Kasteelpark Arenberg 40, Bus 2447, 3001 Leuven, Belgium.
E-mail address: twan.vanhooff@kuleuven.be (T. van Hooff). Reynolds-averaged Navier-Stokes (RANS) approach, in which the

http://dx.doi.org/10.1016/j.buildenv.2016.12.019
0360-1323/© 2016 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 149

Reynolds decomposition is used to split the flow variables in a experiments. A detailed review of the literature has been performed,
mean and a fluctuating component. In RANS CFD simulations, the resulting in Table 1 that provides an overview of CFD cross-
mean flow is resolved whereas the effects of turbulence (fluctu- ventilation studies with validation based on either reduced-scale
ating components) on the mean flow is modeled using a turbulence wind-tunnel or full-scale on-site measurements, or both. Table 1
model. Since assumptions have to be made to model these effects, a shows that the majority of these publications considered valida-
proper validation study is imperative to ensure an accurate solution tion for RANS turbulence models only and only a relatively limited
of the natural ventilation flow using RANS. In large eddy simulation number of these studies compared multiple turbulence models with
(LES), the mean flow and the large-scale turbulence with a length experimental data. Table 1 also indicates the type of study (generic
scale larger than the filter size (often taken as the grid size) are or applied) and the flow parameters for which the validation study
resolved and the small-scale turbulence is modeled using a was performed, i.e. validation based on velocities, turbulence levels,
subgrid-scale model. Although LES is an intrinsically more accurate volume flow rates, surface pressures, velocity vector fields, etc. Most
method for CFD simulations of wind flows (e.g. Refs. [15e18]), its studies listed in Table 1 included one to three parameters in the
use entails a (very) large increase in computational demand (e.g. validation. For example, Lee et al. [27] performed particle image
factor > 102 [1]) and it is therefore often not a viable option in velocimetry (PIV) measurements in a generic greenhouse and used
research and consultancy (e.g. Refs. [1,2]). these results for a comparison with results from CFD simulations
To assess the performance of RANS and LES for cross-ventilation using the standard k-ε, renormalization group (RNG) k-ε, and real-
flows, detailed validation studies are required using high-resolution, izable k-ε models, and a Reynolds stress model (RSM). Mean velocity
well-defined and well-documented reduced-scale or full-scale and turbulence intensity were compared along one vertical line in

Table 1
Non-exhaustive overview of wind-induced cross-ventilation validation studies with indication of the type of study (generic or applied), the numerical method used (RANS/LES/
DES), the turbulence models tested, the type of experiments used for the validation study (wind tunnel and/or full scale) and the parameters for which the validation has been
performed (i.e. velocity, velocity vector fields, turbulence levels, volume flow rates, temperatures, concentrations, surface pressures).

Authors (year) Ref. Gen./App. Method Turb. models Exp. Val. param.

Kato et al. (1992) [3] Generic LES Standard Smagorinsky WT V,VF,Q,P


Mistrotis et al. (1997) [19] Applied RANS SKE, CK, RNG WT/FS V,VF,TP
Mistrotis et al. (1997) [20] Applied RANS RNG WT VF,T
Kurabuchi et al. (2000) [21] Generic RANS SKE, LK, MMK WT VF,TP,Q,P
LES Standard Smagorinsky WT
Straw et al. (2000) [4] Generic RANS RNG FS V,Q
Jiang and Chen (2002) [22] Applied LES Standard Smagorinsky WT/FS V,P
Jiang et al. (2003) [6] Generic LES Standard Smagorinsky WT V,VF,TP,P
Bartzanas et al. (2004) [23] Applied RANS SKE FS V,Q,T
Kurabuchi et al. (2004) [24] Generic LES Standard Smagorinsky WT VF,TP,P
Shklyar & Arbel (2004) [25] Generic RANS SKE, RSM FS P
Hu et al. (2005) [26] Generic RANS SKE, RNG, SKO, SST WT VF,TP,Q
Lee et al. (2005) [27] Generic RANS SKE, RNG, RLZ, RSM WT V,VF,TP
Mochida et al. (2005) [28] Applied RANS DKE FS V
Mochida et al. (2006) [29] Applied RANS DKE FS V,Q
Evola and Popov (2006) [30] Generic RANS SKE, RNG WT V,Q,P
Fatnassi et al. (2006) [31] Applied RANS SKE FS Q
Yang et al. (2006) [9] Generic RANS SKE, RNG FS Q,P
Bartzanas et al. (2007) [32] Applied RANS SKE, RNG, RLZ, RSM FS Q
Hu et al. (2008) [33] Generic LES Standard Smagorinsky WT VF,TP,P
Stravakakis et al. (2008) [34] Applied RANS SKE, RNG, RLZ FS V,T
Teitel et al. (2008) [35] Applied RANS SKE WT/FS V
Kobayashi et al. (2009) [36] Generic RANS RSM WT Q,P
Kurabuchi et al. (2009) [37] Applied RANS DKE WT VF,Q,P
Meroney (2009) [38] Generic RANS SKE, RNG, RLZ, SKO, RSM WT V,VF
LES Standard Smagorinsky WT
DES Not reported WT
Norton et al. (2009) [10] Applied RANS SKE RS V,VF
Kobayashi et al. (2010) [39] Generic RANS RSM WT V,VF,P
Norton et al. (2010) [40] Generic RANS SKE FS V
Norton et al. (2010) [41] Applied RANS SKE SCV V,T
Nikas et al. (2010) [42] Generic RANS SKO WT/FS V,Q
van Hooff & Blocken (2010a,b) [11,43] Applied RANS RLZ FS V
Cheung and Liu (2011) [44] Generic RANS SKE WT V,VF,TP,Q
Larsen et al. (2011) [45] Generic RANS SKO WT/FS V
Lo & Novoselac (2011) [46] Applied RANS RNG FS TP,Q
Lo (2012) [47] Generic RANS SKE, RNG WT/FS Q,T,C
Ramponi & Blocken (2012a,2012b) [48,49] Generic RANS SKE, RNG, RLZ, SKO, SST, RSM WT V,VF
Lo & Novoselac (2013) [50] Applied RANS SKE FS Q,T,C
Mohamed et al. (2013) [51] Generic RANS SKE WT Q
van Hooff & Blocken (2013) [52] Applied RANS RLZ FS V,C
Peren et al. (2015a-c, 2016) [53e56] Generic RANS SKE, RNG, RLZ, SKO, SST, RSM WT V,VF
Martins & Carrilho da Graça (2016) [57] Applied RANS SKE, RNG WT P
Tong et al. (2016) [14] Generic LES Dynamic Smagorinsky WT V
Tong et al. (2016) [58] Generic LES Dynamic Smagorinsky WT V

CK ¼ Chen-Kim k-ε model [59], LK ¼ Launder-Kato k-ε model [60], MMK ¼ Murakami-Mochida-Kondo k-ε model [61], DKE ¼ Durbin k-ε model [62], DES ¼ Detached Eddy
Simulation [63]. FS ¼ full scale, RS ¼ reduced scale, WT ¼ wind tunnel. V ¼ velocities, VF ¼ velocity vector fields, TP ¼ turbulence parameters, Q ¼ volume flow rates,
T ¼ temperatures, C ¼ concentrations, P ¼ surface pressures.
150 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

the middle of the greenhouse and they concluded that the best pressure coefficients on the windward and leeward facade and
agreement was obtained with the RNG k-ε model and the RSM mean surface pressure coefficients on the floor obtained from the
model, with average errors of 7.9% and 7.6%, respectively, for mean experiments with those from LES. LES was capable of reproducing
velocity. However, Lee et al. [27] also stated that the average errors of the pressures on the windward facade and ground surface, but
turbulence intensity for the RNG k-ε model and the RSM model were deviations were observed for the leeward facade [3]. Also Jiang et al.
much larger: 37.5% and 43.2%, respectively. A comparison of the [6], Kurabuchi et al. [24] and Hu et al. [33] compared LES simula-
measured velocity vector fields with those obtained from CFD with tions with the standard Smagorinsky subgrid-scale model with
the RNG k-ε model showed very similar internal flow patterns [27]. wind-tunnel measurements and found good qualitative agree-
Evola and Popov [30] compared the mean velocity along three ver- ments as well. Note that no specific quantification of the agreement
tical lines inside a cubical building as measured by Jiang et al. [6] in a between CFD and LES was provided at that time.
wind tunnel with the results from steady RANS CFD simulations Only a very limited number of publications reported validation
with both the standard k-ε model and the RNG k-ε model. The studies including both steady RANS and LES simulations. Kurabuchi
agreement was better with the RNG k-ε model than with the stan- et al. [21] compared split-film anemometer measurements with
dard k-ε model. In terms of volume flow rate, the deviation was 8.7% both steady RANS and LES simulations. In the steady RANS simula-
for the RNG k-ε model and 13.7% for the standard k-ε model. In tions, the standard k-ε model and two modified k-ε models were
addition, they compared the mean surface-averaged external pres- used, whereas the LES simulation used the standard Smagorinsky
sure coefficients for the roof and the four facades of the building. Hu subgrid-scale model. Based on velocity vector plots and contours of
et al. [26] used the experimental results presented by Kurabuchi turbulent kinetic energy they concluded that LES provided an
et al. [24] for a validation study of four RANS turbulence models excellent qualitative agreement with the measurements, and was
(standard k-ε, RNG k-ε, standard k-u, shear stress transport (SST) k- more accurate than the RANS simulations. Kurabuchi et al. [21] also
u), and focused on mean velocity vector fields and contours of tur- compared the predicted flow rates through the building with flow
bulent kinetic energy, which were qualitatively compared in the rates based on tracer gas (ethylene) measurements obtaining a very
vertical center plane. They concluded that both k-u models showed good agreement with LES (2.3% deviation) and a good agreement
a better agreement with the measurements than the k-ε models, with the RANS models (4.6e7% deviation). In addition, it was shown
mainly due to a too horizontal jet trajectory predicted by the k-ε that the modified k-ε models showed a better qualitative agreement
models [26]. In addition, Hu et al. [26] showed that the predicted than the standard k-ε model based on velocity vector plots and
values of turbulent kinetic energy inside the enclosure were lower contours of turbulent kinetic energy. Finally, they compared
than the measured ones, while they were higher than the measured measured mean external surface pressure coefficients on the
ones in the stagnation region outside the enclosure. Overall, they windward and leeward facade and on the roof with results obtained
concluded that the best agreement between the experimental and from steady RANS and LES and found a much better agreement with
numerical data with respect to the mean velocity vector fields and LES than with RANS, both at the windward and leeward facade.
turbulent kinetic energy contours was obtained using the SST k-u
model [26]. Finally, Hu et al. [26] compared the experimentally ob- 1.2. Introduction of current study
tained volume flow rates with those obtained using the SST k-u
model for different wind directions; the results were within z12% The overview of cross-ventilation validation studies in Section
for each wind direction with an average deviation of 5%. Ramponi 1.1 clearly demonstrates that there is currently no consensus in the
and Blocken [48,49] and Pere n et al. [53e56] used PIV measure- literature on which CFD method provides the best results and the
ments by Karava et al. [64] to validate steady RANS CFD models of best balance between accuracy and computational demand, which
cross-ventilation flow. Ramponi and Blocken [48,49] made a com- is at least partly due to the lack of validation studies based on
parison of the performance of the standard k-ε model, realizable k-ε detailed experimental data of mean velocity and turbulence in-
model, RNG k-ε model, standard k-u model, SST k-u model and a tensity and including both RANS simulation and LES simulations;
RSM model with respect to the mean streamwise velocity along a previous studies generally only assessed one of both methods. For
horizontal line through the center of the two symmetrically placed example, several studies have reported the inability of steady RANS
window openings (vertical location in middle of facades). Also mean to accurately predict turbulent kinetic energy inside the enclosure
velocity vector fields were compared with the measured ones. The (e.g. Refs. [26,27]), while the predictions of velocities are often
best agreement was obtained with the SST k-u model followed by stated to be sufficiently accurate. In addition, most validation
the RNG k-ε model [48]. These two models were the only ones that studies focused on the comparison of mean velocity vector fields,
accurately predicted the direction of the incoming jet through the mean velocity and turbulence kinetic energy or volume flow rates,
windward opening. Pere n et al. [53e56] also used the experimental or a combination of these. Although several studies pointed to
data by Karava et al. [64], but this time for asymmetrical window difficulties in correctly predicting the incoming jet, to the best
openings; i.e. windward opening in bottom part and leeward knowledge of the authors, no validation study has yet analyzed the
opening in upper part of facade. A comparison was made along a behavior of the incoming jet and its reproduction by CFD simula-
horizontal line through the windward opening and along a diagonal tions (either RANS or LES) in more detail.
line through both the windward and leeward opening. Their vali- Therefore, the objective of this paper is the validation of both 3D
dation study resulted in the same conclusions as by Ramponi and steady RANS simulations and LES of cross-ventilation in a generic
Blocken [48]; with the best agreement obtained by the SST k-u isolated enclosure with wind-tunnel measurements from literature
model and the RNG k-ε model [53e56]. Turbulent kinetic energy [65,66]. The RANS simulations are conducted with the standard k-ε
levels were not assessed in these studies. (SKE), RNG k-ε, realizable k-ε (RLZ), SST k-u and RSM turbulence
A few validation studies were conducted in which only LES was models, whereas the LES is performed with the dynamic Smagor-
assessed. The early and pioneering study by Kato et al. [3] insky subgrid-scale model. The evaluation is based on five param-
compared LES simulations with the standard Smagorinsky eters: mean velocity, turbulent kinetic energy, ventilation flow rate,
subgrid-scale model with their own wind-tunnel experiments incoming jet angle and incoming jet spreading width. The nearly
finding a good agreement with respect to the mean velocity along cubical enclosure is equipped with two windows, one on both the
four vertical lines, and also with respect to the volume flow rates windward and on the leeward side of the building. To the best
(within 13%). Kato et al. [3] also compared mean external surface knowledge of the authors, such a detailed validation study for both
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 151

RANS and LES, including an analysis of the jet dynamics, has not energy values were obtained by three-component measurement of
been reported before and can thus be regarded as novel. the variances in the velocity fluctuations. This distribution can be
The wind-tunnel experiments and the building geometry are approximated by the following equation [65]:
presented in Section 2. In Section 3 the computational settings are

described and the results of the grid-sensitivity analysis are pre- 0:32
kðzÞ z

¼ 0:033exp
H
sented. The comparison of the steady RANS simulations with the (2)
experiments is provided in Section 4. The results of the LES simu- U2H
lation are presented in Section 5 and a more detailed comparison
The velocity at building height H (UH) was equal to 4.3 m/s.
between the RANS and LES results is provided in Section 6. Section
The dimensions of the building model under study were
7 (Discussion) and Section 8 (Conclusions) conclude this paper.
0.2  0.2  0.16 m3 (D x W x H) (Fig. 2). An opening area of
3.3  103 m2 was present in both the windward and the leeward
2. Description of experiments facade. The thickness of the walls and ceiling was 3 mm (0.003 m).

Tominaga and Blocken [65] reported reduced-scale measure-


ments of wind-induced cross-ventilation in the atmospheric
3. CFD simulations: computational settings and parameters
boundary layer (ABL) wind tunnel at Niigata Institute of Technology
in Japan. The wind tunnel has a test section of 13  1.8  1.8 m3 (L x H
3.1. Computation geometry, domain and grid
x W). An atmospheric boundary layer (ABL) velocity profile was
created using a combination of spires and surface roughness ele-
The computational model represents the reduced-scale model
ments. The velocity was measured using a split fiber probe (SFP)
used in the wind-tunnel measurements. The thickness of the walls
(Dantec Dynamics; 55R55) and a constant temperature anemom-
and ceiling of 3 mm is explicitly included in the computational
etry (CTA) module (Dantec Dynamics; 90C10) to identify the three-
model.
dimensional components of the velocity vector [65]. The measured
The computational domain is constructed based on the best
vertical approach flow profiles of mean streamwise velocity and
practice guidelines by Franke et al. [67], Tominaga et al. [68] and
turbulent kinetic energy are depicted in Fig. 1. The mean streamwise
Blocken [69]; i.e. a distance of 5H from the building to the top and
velocity can be fitted to a power law with an exponent of 0.25 [64]:
sides of the computational domain and a distance of 15H between
UðzÞ  z 0:25 the building and the outlet boundary downstream of the building.
¼ (1) The upstream length of the domain is reduced to 3 times the height
UH H
of the building to limit the occurrence of unintended streamwise
with UH the velocity at building height H. The turbulent kinetic gradients in the approach flow profiles [69,70]. The resulting di-
mensions of the domain are 5.32  1.8  0.96 m3 (L x W x H)
(Fig. 3a).
The computational grid is created using the surface-grid
extrusion technique by van Hooff and Blocken [11] and is shown
in Fig. 3bed. The computational grid consists of hexahedral cells
only, with a high spatial resolution near the building and its
openings. The grid resolution results from a grid-sensitivity anal-
ysis using three different grids, created by refining and coarsening
the basic grid with a factor of √2 in each direction. The results of
the grid-sensitivity analysis are presented in Section 3.4.

3.2. Boundary conditions

The boundary conditions reproduce the conditions during the


wind-tunnel experiments as much as possible. A logarithmic inlet
Fig. 1. Approach flow profiles of dimensionless mean velocity and turbulent kinetic velocity profile is constructed based on a fit with the power law
energy [65]. profile described in Eq. (1). The resulting log-law equation is:

Fig. 2. Building geometry. (a) Front view. (b) Vertical cross-section. Dimensions in m.
152 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

Fig. 3. (a) Computational domain. Dimensions in m. (b) Computational grid on ground and building surfaces for basic grid (5,321,588 cells) (RANS simulation). (c) Computational
grid on building surfaces and on part of bottom of computational domain for basic grid (5,321,588 cells). (d) Detail of the computational grid. Note that no internal obstructions are
present.

  roughness height is set to zero (kS ¼ 0 m) for the bottom of the


uABL z þ z0 domain (to represent a very smooth turntable) and the building
UðzÞ ¼ ln (3)
k z0 surfaces (smooth walls). Fig. 4 shows a check of the horizontal
homogeneity for two RANS simulations (RNG k-ε and SST k-u
with u*ABL the ABL friction velocity, k the von Karman constant model) and the LES simulation [70]. It compares the dimensionless
(0.42) and z the height coordinate. The log law is preferred over the mean velocity magnitude (jVj/UH) profile at the inlet (x/D ¼ 3)
power law because it yields the value of u*ABL that is needed for the with the incident profile at the building location (x/D ¼ 0). There is
profile of the turbulence dissipation rate ε. The aerodynamic a very small acceleration near the ground due to the absence of a
roughness length z0 is determined based on a fitting procedure surface roughness. Note however that this acceleration will also be
using the measured velocity profile yielding z0 ¼ 0.0009 m at present in the wind tunnel when the air flows over the smooth
reduced scale. The corresponding value of u*ABL is 0.348 m/s for a turntable of the wind tunnel (no roughness elements present at
reference velocity at building height of UH ¼ 4.3 m/s (H ¼ 0.16 m). turntable [65]) as also shown in previous experimental studies [81].
The turbulent kinetic energy k is given by Eq. (2). The turbulence The incident profiles obtained with the three other RANS turbu-
dissipation rate ε is calculated using Eq. (4) [71]: lence models are identical to those of the RNG k-ε and SST k-u
 * 3 models and are therefore not shown for the sake of brevity. At the
uABL outlet plane, zero static gauge pressure is applied and at the top and
εðzÞ ¼ (4)
kðz þ z0 Þ lateral sides of the domain, zero normal velocities and zero normal
gradients of all variables are imposed.
The specific dissipation rate u for the SST k-u model is calcu-
lated using Eq. (5):
3.3. Solver settings
εðzÞ
uðzÞ ¼ (5)
Cm kðzÞ The finite volume method in the commercial CFD code ANSYS
Fluent 15 [78] is used to solve the approximate forms of the gov-
where Cm is an empirical constant taken equal to 0.09. For the LES erning equations. The 3D steady RANS equations are solved in
computations, a time-dependent inlet profile is generated by using combination with five turbulence models; 4 two-equation eddy-
the vortex method [72] with a number of vortices Nv ¼ 190. As viscosity models and one second-order closure model: (1) standard
shown by Sergent [72], the influence of small changes in the k-ε model (SKE) [79]; (2) realizable k-ε model (RLZ) [82]; (3) RNG k-
number of vortices on the generated velocity fluctuations is ε model [83,84]; (4) SST k-u model [85]; (5) low-Re stress-omega
negligibly small. In addition, this setting was successfully used in RSM model [86]. The SIMPLE algorithm is used for pressure-
previous LES validation studies by the authors for wind flows velocity coupling, pressure interpolation is second order and
around buildings [73e76] and indoor airflows [77]. second-order discretization schemes are used for both the con-
At the ground and building surfaces, either an automated wall vection terms and the viscous terms of the governing equations.
treatment is used [78] (for the k-u model) or the standard wall Convergence is assumed to be obtained when all the scaled re-
functions by Launder and Spalding [79] (for the other models) are siduals level off and reach a minimum value. The minimum values
used in conjunction with the sand-grain based roughness (kS) of the residuals vary for the different turbulence models, for
modification defined by Cebeci and Bradshaw [80]. The sand-grain example the minimum values for the SST k-u model are 107 for x,
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 153

Fig. 4. (a) Comparison of inlet and incident dimensionless mean velocity magnitude (jVj/UH) profiles for two RANS simulations and one LES simulation. (b) Schematic cross-section
of the domain (not to scale) with location of inlet profile (x/D ¼ 3) and incident profile (x/D ¼ 0).

y, z momentum, 106 for k, and 105 for continuity and u. Note that h . i
oscillatory convergence is observed in the simulation with the RSM rp Ubasic  Ufine UH 

model, which required averaging over a sufficient number of iter- GCIbasic ¼ Fs   (6)
ations (e.g. Refs. [48,69]). The LES simulation in this study is per-
1  rp
formed using the dynamic Smagorinsky subgrid-scale model with r the linear grid refinement factor (r ¼ √2), p the formal order
[87e89]. The filtered momentum equation is discretized with a of accuracy which is assigned a value of 2 due to the second-order
bounded central-differencing scheme and pressure interpolation is discretization schemes used for the simulations, and Fs is the safety
second order. Time integration is second-order implicit. Pressure- factor, taken to be 1.25, which is the recommended value when
velocity coupling is taken care of by the fractional step algorithm. three or more grids are considered in the grid-sensitivity analysis
The time step Dt is based on a maximum CFL number of 1 and is [90]. The GCI provides an estimate of the error in the solution of the
equal to Dt ¼ 0.0002 s. The results of the LES computation pre- basic grid by comparing it to that of the fine grid. The calculated GCI
sented here are averaged over 375,000 time steps, which corre- along the three vertical lines is shown in Fig. 5def. The average
spond to a flow time of 75 s, i.e. 60 times the flow-through time values of GCI along the three lines are: 2.4%, 1.7% and 1.0%,
(through the computational domain). It is verified that the aver- respectively. It is concluded that the basic grid provides nearly grid-
aging time is sufficient to obtain statistically-steady results by independent results and it is therefore used in the remainder of the
monitoring the evolution of the mean velocity at several locations study for the RANS simulations. The LES simulation is conducted on
inside the enclosure with time (moving average). The simulation is the coarse grid to limit the computational cost, as explained later.
terminated when the moving average of the mean velocity at all Note that the definition “coarse” with respect to the mesh resolu-
location remains within 0.5%. tion is a bit misleading, since the minimum cell size is still as small
All simulations are conducted on the HPC cluster (x86_64 ar- as 0.001 m, 54  78 x 56 cells are present inside the enclosure over
chitecture) at Eindhoven University of Technology in the the building depth, width and height, and the cell count for the
Netherlands. For each simulation 14 cores on a node are used building including its direct vicinity (within H) is z 1.5 million,
(Intel(R) Xeon(R) CPU - X5650 @ 2.67 GHz), and 64 or 96 GB of resulting in a high-resolution grid in the area of interest.
system memory is available depending on the node used.

4. CFD simulations: RANS results


3.4. Grid-sensitivity analysis
Fig. 6 shows a comparison of x-velocity (UX/UH) by all five RANS
A grid-sensitivity analysis is conducted for RANS with the SST k- turbulence models with the corresponding measurements along
u model. The coarse, basic and fine grid contain 1,897,416 cells, seven vertical lines in the vertical center plane (y/W ¼ 0) of the
5,321,588 cells, and 14,335,040 cells, respectively. 2, 3, and 4 cells enclosure. The main observations are:
are used across the wall thickness for the coarse, basic and fine grid,
respectively. Fig. 5 shows the results of the grid-sensitivity analysis  The different RANS models provide very different results. The
by means of the dimensionless mean streamwise velocity (UX/UH) RNG, SST and RSM models provide about the same vertical
along three vertical lines in the vertical center plane (y/W ¼ 0). The location of maximum UX/UH in the jet region along each vertical
basic and the fine grid provide almost the same results, while the line. It is remarkable that the RNG and the SST models provide
results obtained with the coarse grid show clear deviations with almost identical velocity profiles. On the contrary, the SKE and
the results of the basic grid at all three lines. Differences between RLZ models provide a too high location of maximum velocity.
the basic and the fine grid at x/D ¼ 0.125 below z/H ¼ 0.3 go up to  SKE and RLZ models also provide a larger width of the incoming
D(UX/UH) ¼ 0.03. jet and the related smaller velocity gradients in the shear layer
The grid-convergence index (GCI) as proposed by Roache [90] is of the incoming jet.
used to provide an estimate of the error in UX/UH obtained on the  The RSM model provides substantially higher values for the
basic grid: maximum velocity in the jet region compared to the RNG and
154 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

Fig. 5. Grid-sensitivity analysis: (aec) Comparison of dimensionless mean x-velocity (UX/UH) obtained with steady RANS CFD simulations on three different grids (coarse, basic,
fine) along three vertical lines inside the building in the vertical center plane (y/W ¼ 0). (def) Grid-convergence index (GCI) along same three vertical lines. (a,d) x/D ¼ 0.125; (b,e)
x/D ¼ 0.5; (c,f) x/D ¼ 0.875.

SST models; e.g. at x/D ¼ 0.375 UX/UH z 0.50 for the RNG and  The direction of the incoming jet is very different for SKE and
SST models, while UX/UH ¼ 0.58 for the RSM model (Fig. 6c). RLZ on the one hand, versus RNG, SST and RSM on the other
 The general flow pattern in the lower part of the enclosure is hand. The SKE and RLZ models provide an incoming jet that is
better reproduced by RNG and SST (better prediction of vertical more horizontal, in line with the observations from Fig. 6.
location of maximum velocity), and to a lesser extent by the RSM  Fig. 7 also clearly shows the larger jet spreading with SKE and
model, while the flow in the upper part (z/H > 0.6) is more RLZ compared to RNG, SST and RSM. The jet spreading param-
accurately reproduced by SKE, RLZ and RSM. eter (d0.5;upperþd0.5;lower)/HO is largest for RLZ and SKE, with
values of 1.7 and 1.6, respectively, followed by SST (1.2), RNG
Fig. 7 displays the contours of the dimensionless mean velocity (1.1) and RSM (1.0).
magnitude (jVj/UH) in the vertical center plane including the jet  Fig. 7e shows the higher velocity in the incoming jet predicted
half-widths d0.5 for each model (visualized with white dots) on by the RSM model, also visible in Fig. 6, which is directly related
each side of the jet centerline (d0.5;upper and d0.5;lower). The value of to the lesser spreading rate.
(d0.5;upperþd0.5;lower)/HO at x/D ¼ 0.625, with HO the inlet opening
height, is also reported. The jet half-width d0.5 is defined as the Fig. 8 provides a comparison of the dimensionless turbulent
vertical linear distance between the local maximum jet velocity kinetic energy (k/U2H) along the seven vertical lines in the vertical
magnitude (jVjMAX) and the location where the velocity magnitude center plane. The most important observations are:
at is equal to half the local jet centerline velocity jVjMAX; i.e.
0.5jVjMAX (see Fig. 7f). Note that due to asymmetry of the incoming  Turbulent kinetic energy is underpredicted in large areas above
jet the jet half-width below the jet centerline (d0.5;lower) is not and below the incoming jet by all RANS turbulence models. This
necessarily equal to the one above the jet centerline (d0.5;upper). The can be explained by the fact that steady RANS does not correctly
main observations are: model the effect of transient flow features on the resolved flow
pattern, while the experiments [65] clearly indicate the pres-
ence of pronounced transient effects, such as the jet flapping
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 155

Fig. 6. Comparison of dimensionless mean x-velocity (UX/UH) obtained with steady RANS CFD simulations with different turbulence models (solid and dashed lines), and with
wind-tunnel measurements (A), along seven vertical lines inside the building in the vertical center plane (y/W ¼ 0). (a) x/D ¼ 0.125; (b) x/D ¼ 0.25; (c) x/D ¼ 0.375; (d) x/D ¼ 0.5;
(e) x/D ¼ 0.625; (f) x/D ¼ 0.75; (g) x/D ¼ 0.875.

Fig. 7. Contours of dimensionless mean velocity magnitude (jVj/UH) obtained with steady RANS CFD simulations with different turbulence models in the vertical center plane (y/
W ¼ 0). (a) SKE. (b) RNG. (c) RLZ. (d) SST. (e) RSM. (f) Schematic indication of location of maximum jet velocity (jVMAXj), 0.5jVjMAX and jet half-widths d0.5;upper and d0.5;lower. The
white dots in (aee) indicate the location of jet half-widths d0.5;upper and d0.5;lower.
156 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

Fig. 8. Comparison of dimensionless mean turbulent kinetic energy (k/U2H) obtained with steady RANS CFD simulations with different turbulence models (solid and dashed lines),
and with wind-tunnel measurements (A), along seven vertical lines inside the building in the vertical center plane (y/W ¼ 0). (a) x/D ¼ 0.125; (b) x/D ¼ 0.25; (c) x/D ¼ 0.375; (d) x/
D ¼ 0.5; (e) x/D ¼ 0.625; (f) x/D ¼ 0.75; (g) x/D ¼ 0.875.

(vertical motion of the jet) and the Kelvin-Helmholtz type  The higher values obtained with the three k-ε eddy-viscosity
instability in the shear layers of the incoming jet. RANS models, and especially with SKE and RLZ, are due to the
 Along the first two vertical lines (x/D ¼ 0.125 and x/D ¼ 0.25), excessive production of turbulent kinetic energy in stagnation
the RANS models predict values of k/U2H z 0.0025, while the zones, i.e. the stagnation point anomaly (e.g. Refs. [17,62,91,92]).
measured values are around k/U2H z 0.015 and 0.02, depending
on the measurement location. Four validation metrics are used to obtain an overall and
 Turbulent kinetic energy in the jet region (0.35 < z/H < 0.6) is quantitative evaluation of the performance of all five RANS turbu-
strongly overpredicted by SKE and RLZ at the lines x/D ¼ 0.125 lence models; the factor of 2 of the observations (FAC2), the factor
and x/D ¼ 0.25, which is related to the larger jet spreading of 1.3 of the observations (FAC1.3), the fractional bias (FB), and the
observed in Fig. 7a,c. normalized mean square errors (NMSE). For the streamwise ve-
 RSM predicts the lowest values of turbulent kinetic energy in locity only FAC2 and FAC1.3 are used, while for turbulent kinetic
the jet region at x/D ¼ 0.125 and x/D ¼ 0.25 and provides a very energy FAC2, FB, and NMSE are calculated. The metrics are calcu-
close agreement with the measurements. lated using Eq. (7) until Eq. (10):

Fig. 9 depicts contours of k/U2H. The main observations are:


8 9
>
<1 Pi >
 Turbulent kinetic energy in the incoming jet region is higher 1 XN for 0:5   2=
FAC2 ¼ n with ni ¼ Oi (7)
with SKE and RLZ (Fig. 9a,c). The higher values seem to originate N i¼1 i >
: >
;
from outside the enclosure, i.e. from the stagnation region, 0 else
where a large area with high values of turbulent kinetic energy
(k/U2H > 0.08) is present. This area with high values of k/U2H is
much larger for SKE and RLZ than for RNG, SST and RSM. Tur- 8 9
>
<1 Pi >
bulent kinetic energy is transported to the interior volume and 1 XN for 0:77   1:3 =
FAC1:3 ¼ n with ni ¼ Oi
leads to higher indoor values of k/U2H for the SKE and RLZ N i¼1 i >
: >
;
models, especially near the windward opening (see Figs. 8 and 0 else
9). (8)
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 157

Fig. 9. Contours of dimensionless turbulent kinetic energy (k/U2H) obtained with steady RANS CFD simulations with different turbulence models in the vertical center plane (y/
W ¼ 0). (a) SKE. (b) RNG. (c) RLZ. (d) SST. (e) RSM.

Table 2 and RSM the worst performance. The other three models score very
Validation metrics for dimensionless x-velocity (UX/UH) and turbulent kinetic en- differently in the ranking for each of the three validation metrics
ergy (k/U2H): FAC2: factor of two observations; FAC1.3: factor of 1.3 observations;
used; i.e. RLZ scores better than the RNG and SST for FAC2 and FB,
NMSE: normalized mean square error; FB: fractional bias.
while it scores worse than RNG and SST for NMSE. The results from
Method Turbulence model UX/UH k/U2H Table 2 might seem to differ from the findings extracted from Fig. 6,
FAC2 FAC1.3 FAC2 FB NMSE however, it must be noted that in the calculation of the validation
RANS SKE 0.63 0.33 0.52 0.25 0.46
metrics all locations are given the same weight, where the locations
RNG 0.56 0.27 0.38 0.67 1.02 in the jet region e often regarded as the most important region e
RLZ 0.52 0.24 0.46 0.49 1.45 are considered equally important as the locations above and below
SST 0.56 0.32 0.38 0.69 1.38 the jet region. The overall best performance of SKE compared to the
RSM 0.59 0.25 0.32 0.76 1.99
more sophisticated models is noteworthy. It is attributed to the
LES Dynamic Smagorinsky 0.83 0.48 0.95 0.20 0.30
Ideal values 1 1 1 0 0 observed overprediction of turbulent kinetic energy in the stag-
nation region outside the building (Fig. 9), which counteracts the
lack of turbulence above and below the incoming jet region due to
the use of the steady RANS approach. Since the excessive genera-
tion of turbulent kinetic energy was most pronounced for SKE (see
½O  ½P
FB ¼ (9) Fig. 9), the results obtained with this model show a good agreement
0:5ð½O þ ½PÞ with experimental results, especially above and below the
incoming jet region.
h i Finally, the volume flow rates through the building are
ðOi  Pi Þ2 compared with the experimentally obtained value [65]. The volume
NMSE ¼ (10)
½Oi ½Pi  flow rate is made dimensionless using the velocity at building
height (UH) and the window area (Ainlet); i.e. Q/UHAinlet [65]. Table 3
with Pi the predicted (CFD) value, Oi the observed (measured) value shows the values obtained from the steady RANS simulations and
and n the number of measurement locations, which is equal to 63
(7  9). The square brackets indicate averaging over all data points.
The results are shown in Table 2, with an indication of the ideal
Table 3
values at the bottom row; i.e. 1 for FAC2 and FAC1.3, 0 for FB and Dimensionless volume flow rate (Q/UHAinlet) obtained from the experiments and the
0 for NMSE. CFD simulations.
FAC2 for UX/UH for all RANS models lies between 0.52 and 0.63,
Method Turbulence model Q/UHAinlet [-] Deviation [%]
with the highest value for SKE (0.63), followed by RSM (0.59), and
Experiments [65] 0.50
the lowest value for RLZ (0.52). However, the calculation of FAC1.3
RANS SKE 0.483 3.5
results in a different ranking with the best performance with SKE RNG 0.506 1.2
(0.33), closely followed by SST (0.32), and then by RNG (0.27), RSM RLZ 0.503 0.6
(0.25), RLZ (0.24). For turbulent kinetic energy, the FAC2 values are SST 0.516 3.1
highest for SKE (0.52) and lowest for RSM (0.32). The same holds for RSM 0.543 8.6
LES Dynamic Smagorinsky 0.538 7.6
FB and NMSE; for both metrics SKE shows the best performance
158 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

Fig. 10. Comparison of dimensionless mean x-velocity (UX/UH) obtained with LES (dashed line), with RANS with the SKE model (dotted line) and SST k-u model (solid line), and
with wind-tunnel measurements (A), along seven vertical lines inside the building in the vertical center plane (y/W ¼ 0). (a) x/D ¼ 0.125; (b) x/D ¼ 0.25; (c) x/D ¼ 0.375; (d) x/
D ¼ 0.5; (e) x/D ¼ 0.625; (f) x/D ¼ 0.75; (g) x/D ¼ 0.875.

Fig. 11. Contours of dimensionless mean velocity magnitude (jVj/UH) obtained with steady RANS CFD simulation and LES simulation in the vertical center plane (y/W ¼ 0). (a) SKE.
(b) SST. (c) LES. The white dots indicate the location of jet half-widths d0.5;upper and d0.5;lower.

the experimental one (Q/UHAinlet ¼ 0.5). The largest deviation is 5. CFD simulations: LES results
present for RSM (8.6%) followed by SKE and SST, with deviations of
3.5% and 3.1%, respectively. The best agreement is provided by RLZ Figs. 10e13 show the dimensionless mean velocity and turbu-
(0.6% deviation) followed by RNG (1.2% deviation). lent kinetic energy obtained by LES using the dynamic Smagorinsky
subgrid-scale model on the coarse grid. The results of SKE and SST
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 159

Fig. 12. Comparison of dimensionless mean turbulent kinetic energy (k/U2H) obtained with LES (dashed line), with RANS with the SKE model (dotted line) and SST k-u model (solid
line), and with wind-tunnel measurements (A), along seven vertical lines inside the building in the vertical center plane (y/W ¼ 0). (a) x/D ¼ 0.125; (b) x/D ¼ 0.25; (c) x/D ¼ 0.375;
(d) x/D ¼ 0.5; (e) x/D ¼ 0.625; (f) x/D ¼ 0.75; (g) x/D ¼ 0.875.

Fig. 13. Contours of dimensionless turbulent kinetic energy (k/U2H) obtained with steady RANS CFD simulations and LES simulation in the vertical center plane. (a) SKE. (b) SST. (c)
LES.

on the basic grid are added to the figures for comparison. The  Especially in the areas below and above the incoming jet, LES
following main observations are made for Fig. 10 (UX/UH in the outperforms RANS. For example, at x/D ¼ 0.5 (Fig. 10d) the LES
vertical center plane): results closely follow the measured values below z/H ¼ 0.3 and
above z/H ¼ 0.5, while clear discrepancies are present at these
 A very good agreement is present between LES and experiments, locations between RANS and the experimental results. Also
clearly better than the best performing RANS models (SKE and further downstream, at x/D ¼ 0.625, x/D ¼ 0.75 and x/D ¼ 0.875,
SST). the results obtained with LES are much better, with the
160 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

strongest improvements above the incoming jet for RANS SST Table 2 shows that the agreement between LES and the measure-
and below the jet for RANS SKE. ment data is much better than for the RANS simulations. The FAC2
 In general, the velocity gradients are smaller in the profiles and FAC1.3 values for the x-velocities are equal to 0.83 and 0.48,
obtained with LES compared to RANS SST, which can be attrib- and FAC2 is equal to 0.95 for turbulent kinetic energy, while the
uted to the transient flow features that are captured with LES highest FAC2 value for turbulent kinetic energy for the RANS was
and which lead to a stronger redistribution of momentum and 0.52. Also the FB and NMSE scores are clearly better for LES; 0.2
turbulence levels. The smaller velocity gradients and the and 0.3, respectively.
broader jet (more spread) are for example clearly visible in Table 3 shows the mean volume flow rate through the building
Fig. 10d. The velocity gradients with RANS SKE are similar to predicted by LES, together with the steady RANS results. The vol-
those of LES, however, in RANS SKE they are the result of an ume flow rate deviates with 7.6% from the experimentally obtained
overestimation of turbulent kinetic energy in the stagnation value. This deviation is larger than the deviation with the best
region outside the enclosure. performing steady RANS model; i.e. RLZ (deviation of 0.6%). Note
that the use of tracer gas measurements to obtain volume flow
Fig. 11 shows contours of jVj/UH in the vertical center plane. The rates has its limitations with respect to accuracy and it is therefore
most important observations are: difficult to draw strong conclusions on the performance of each of
the tested models.
 The LES results provide a slightly higher velocity magnitude in
the incoming jet compared to RANS. 6. Comparison RANS vs. LES
 The incoming jet is slightly more horizontal in LES than with
RANS SST. The results from Section 4 (RANS) and Section 5 (LES) show clear
 The spreading of the jet is more pronounced in LES. For example, differences in the predicted flow fields with respect to mean ve-
at x/D ¼ 0.625 the jet predicted by LES is wider compared to the locity, turbulent kinetic energy and general flow pattern. In this
jet predicted by SST, as quantified in Fig. 10. The calculated section, a more detailed analysis is provided with respect to the
spreading of the incoming jet in the LES simulation based on the observed differences and their physical background.
jet half-widths ((d0.5;upperþd0.5;lower)/HO) at x/D ¼ 0.625 is about Comparison of Figs. 9 and 13c shows that the interior distribu-
equal to those predicted by SKE and RLZ, i.e. (d0.5;upper- tion of turbulent kinetic energy obtained with LES is clearly
þd0.5;lower)/HO ¼ 1.7 (see Fig. 7a,c). different from that of the steady RANS simulations. There is a
different origin of turbulence inside the enclosure between RANS
Fig. 12 shows values of k/U2H, for which the following main ob- and LES. In the case of SKE and RLZ, the indoor turbulent kinetic
servations are made: energy levels can largely be attributed to the excessive generation
of turbulent kinetic energy in the stagnation zone outside the
 The turbulent kinetic energy levels above and below the jet are enclosure (Fig. 9a,c), which is subsequently transported indoors. In
generally much higher in the LES simulation compared to RANS LES, this excessive generation of turbulent kinetic energy in the
and generally show a better agreement with the experimentally stagnation region is not present at all (Fig. 13c) and the increased
obtained values. levels of turbulence are the result of velocity gradients and local
 LES overpredicts turbulent kinetic energy in the jet region at x/ unsteadiness of the flow inside the enclosure; i.e. the jet flapping
D ¼ 0.25 until x/D ¼ 0.875. (vertical motion of the jet) and the instabilities in the shear layer of
 At x/D ¼ 0.125 and x/D ¼ 0.25, LES predicts a turbulent kinetic the jet flow. To illustrate the presence of one of the unsteady flow
profile with two clear peaks in the shear layer of the incoming features inside the enclosure (i.e. vertical jet flapping), Fig. 14 shows
jet (at z/H ¼ 0.38 and z/H ¼ 0.55 at x/D ¼ 0.125; at z/H ¼ 0.35 instantaneous images from the flow visualizations and the LES
and z/H ¼ 0.5 at x/D ¼ 0.25), which are not visible in both RANS simulation. Vertical jet flapping is present in both the experiments
results. and the LES simulation; the jet direction varies considerably in
time, from an upward direction over a nearly horizontal direction
Fig. 13 shows contours of k/U2H by RANS SKE, RANS SST and LES with a small downward bend when the jet enters the enclosure to a
in the vertical center plane. The following main observations are downward direction. The absence of this distinct vertical flapping
made: motion in the steady RANS simulations results in deviations of the
simulation results when compared to the measurement data.
 As opposed to SKE and SST, LES provides a region of high tur- Fig. 14aec also shows Kelvin-Helmholtz type instabilities in the
bulence in front of the building, which can be attributed to the shear layer of the incoming jet, which is not visible in the LES
standing vortex in front of the building which is a highly tran- simulation. The accurate prediction of this type of instability in the
sient flow feature (moving back and forward along the ground shear layer requires a much higher grid resolution, which for the
surface). present study exceeded the available computational resources. The
 The interior distribution of k/U2H is very different. As indicated computational cost for the LES simulation on the coarse grid is
before, RANS SKE predicts very high values near the windward already about 80e100 times higher than for steady RANS on the
window opening due to the overprediction of turbulent kinetic basic grid. LES on the basic grid would add a factor of 2e4, resulting
energy in the stagnation zone outside the enclosure. RANS SST in 160e400 times the computation time of a steady RANS simu-
mainly predicts higher indoor values near the upstream window lation on the same grid.
due to transport of turbulent kinetic energy from the stagnation Fig. 15a shows the predicted jet trajectory for all evaluated
zone outdoors to the indoor environment and in the shear layer models based on the jet centerline velocity (i.e. jVjMAX). The sym-
of the incoming jet. LES on the other hand provides higher bols represent the vertical location (zMAX) of the local maximum jet
values downstream of the opening and in a far larger region velocity magnitude (jVjMAX) along all seven vertical lines inside the
around the jet. enclosure. There is no RANS model that predicts the same trajectory
as the LES model; SKE and RLZ predict a more horizontal jet di-
Also for the LES results the agreement with the experimental rection (as discussed earlier), while RNG, SST and RSM predict a
data is quantified using four validation metrics (Eqs. (7)e(10)). more downward direction of the incoming jet (Fig. 15a).
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 161

Fig. 14. Comparison of instantaneous images from flow visualization [65] (aec) with images from LES simulation (def). The figures indicate the observation and prediction of jet
flapping inside the enclosure. (a,d) Upwards directed jet (bended jet). (b,e) Horizontally directed jet with small downward bend after jet entry. (c,f) Downward directed jet.

Fig. 15b compares the incoming jet angle a. It is defined based on velocity vs. turbulence levels), etc. A comparison with conclusions
the vertical location of maximum velocity magnitude (jVjMAX) at x/ from previous validation studies from literature also becomes
D ¼ 0 (windward opening) and at x/D ¼ 0.5 (middle of enclosure) difficult due to the issues mentioned above and it can be compli-
and Eq. (11): cated even more due to different geometries used, different loca-
  tions at which velocities and turbulence levels are compared, etc.
zMAX;x=D¼0  zMAX;x=D¼0:5 Despite all difficulties mentioned above, the performance of
a ¼ tan1 (11)
x different turbulence models in predicting volume flow rates as
found in literature is compared with the performance found in the
The value of a ranges from 13.6 to 24.6 for the five RANS
current study. Table 3 showed that all models predict volume flow
models and the LES model. The SST, RSM, RNG models predict a
rates within 8.6% of the measured value, with the RLZ model
value for a (22.7 e24.6 ) close to that of the LES model (a ¼ 23.0 ).
showing the best performance (0.6% deviation). One must note that
SKE and RLZ provide low values of a due to the more horizontal jet
the tracer gas method has its limitations with respect to accuracy
direction. Note that a not only depends on the location of jVjMAX at
and that strong conclusions are difficult to make. In literature,
x/D ¼ 0.5 as depicted in Fig. 15a, but also on the location of jVjMAX at
Straw et al. [4] for example found higher differences in volume flow
x/D ¼ 0, which differs considerably for all models due to different
rate using RANS with RNG (3e9%), however, they did not test other
velocity distributions over the height of the windward window
turbulence models in their study. Evola and Popov [30] found de-
opening.
viations of 8.7% for RNG and 13.7% for SKE when they compared
Fig. 15c compares UX/UH obtained with RANS and LES with the
steady RANS CFD with wind-tunnel measurements of the volume
measurement data along a horizontal line through the middle of
flow rate by Jiang et al. [6]. The better performance of RNG
the two window openings. The RANS models either underpredict or
compared to SKE in Evola and Popov [30] is found in the current
overpredict the mean velocity along this line, which can directly be
study as well. Hu et al. [26] found deviations within z12% for each
attributed to the jet being too horizontal (SKE and RLZ) or too in-
wind direction with an average deviation of 5%, using SST, which is
clined (RNG, SST, RSM) compared to the measurements. The results
higher than the deviation in the current study (3.1%). As mentioned
of the LES simulation provide a good agreement with the mea-
before, the results from different studies are very difficult to
surement data, with an overprediction of UX/UH along the line.
compare, due to the different experimental techniques used, the
Fig. 15d shows k/U2H along the same horizontal line (Fig. 15e). LES
wide range of computational settings that all affect the outcome of
overpredicts the values of k/U2H in the middle of the enclosure
the simulations, etc.
(0.25 < x/D < 0.875), while RNG, SST and RSM underpredict k/U2H
along this line. SKE and RLZ strongly overpredict k/U2H in the up-
stream part of the enclosure (x/D < 0.25), but subsequently predict 7.2. Limitations and future work
similar levels of k/U2H as measured in the wind tunnel between
0.375 < x/D < 0.875. This paper presented the results of a validation study of cross-
ventilation flow in a generic enclosure. Both RANS and LES simu-
7. Discussion lations were compared with experimental data from wind-tunnel
measurements by Tominaga and Blocken [65]. The following
7.1. Comparison with other studies points of discussion and of future work can be raised:

A comparison of the performance of the five different RANS  Due to computational limitations in the present study it was not
turbulence models is not straightforward due to the different possible to run a LES simulation on the basic grid. Future
assessment criteria that can be used; e.g. different validation research might include this simulation and also include a more
metrics, emphasis on different parts of the flow (e.g. jet region vs. detailed parameter study for LES for cross-ventilation flow. For
entire enclosure), emphasis on one parameter over another (e.g. example, a more uniform grid distribution could be tested, as
162 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

Fig. 15. (a) Vertical location (zMAX/H) of maximum velocity (jVjMAX) along seven vertical lines in the vertical center plane. (b) Definition of incoming jet angle a based on vertical
location of jVjMAX at x/D ¼ 0 and x/D ¼ 0.5 (zMAX;x/D¼0-zMAX;x/D¼0.5) (not to scale). (c,d) Dimensionless mean x-velocity (UX/UH) (c), and dimensionless mean turbulent kinetic energy
(k/U2H) (d) obtained with RANS, LES and wind-tunnel measurements (A), along a horizontal line through the middle of the window openings (z/H ¼ 0.5; see (e)) in the vertical
center plane (y/W ¼ 0).

well as other subgrid-scale models, inflow turbulence genera- unsteady motion of the flow (e.g. Ref. [93]), and it is worthwhile
tors, etc. The use of the coarse grid among others resulted in the to study it since the computational demand is less than that of
absence of Kelvin-Helmholtz type instabilities in the LES simu- LES.
lation, while this phenomenon could be observed in the flow  In this study, five commonly used turbulence models were
visualizations by Tominaga and Blocken [65]. Note that the grid- included. The study can be extended by including for example
resolution for the “coarse” grid was still very high with around other second-order closure models, eddy-viscosity models, or
1.5 million hexahedral cells for the building and its direct vi- hybrid RANS/LES models (e.g. Ref. [94]).
cinity (within H of the building) and that LES on this grid pro-
vided a very good agreement with the experimental results.
8. Conclusions
 Future work will focus on the use of unsteady RANS (URANS) to
capture the flow field in the cross-ventilated building. Although
The objective of the study presented is the validation of cross-
the use of URANS is not straightforward, past research has also
ventilation flow through a generic enclosure for five different
shown the potential of URANS with respect to obtaining the
RANS turbulence models and for a LES model in combination with
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 163

the dynamic Smagorinsky subgrid-scale model. A comparison with References


a range of RANS models as well as LES is not common, as can be
seen in the overview of validation studies earlier in this introduc- [1] Q. Chen, Ventilation performance prediction for buildings: a method overview
and recent applications, Build. Environ. 44 (4) (2009) 848e858.
tion section, although such comparisons are of primary importance [2] B. Blocken, 50 years of computational wind engineering: past, present and
when deciding which numerical method to use and to aid in future, J. Wind Eng. Ind. Aerodyn. 129 (2014) 69e102.
drafting guidelines on the use of CFD for the analysis and optimi- [3] S. Kato, S. Murakami, A. Mochida, S. Akabayashi, Y. Tominaga, Veloc-
ityepressure field of cross ventilation with open windows analyzed by wind
zation of cross-ventilated buildings. Mean velocity, turbulent ki- tunnel and numerical simulation, J. Wind Eng. Ind. Aerodyn. 44 (1992)
netic energy, and volume flow rate are compared with 2575e2586.
experimental data from literature. In addition, an evaluation of the [4] M.P. Straw, C.J. Baker, A.P. Robertson, Experimental measurements and
computations of the wind-induced ventilation of a cubic structure, J. Wind
jet half-widths (d0.5;upper and d0.5;lower), incoming jet angle (a) and Eng. Ind. Aerodyn. 88 (2e3) (2000) 213e230.
transient flow features (i.e. jet flapping) is presented, which is [5] C. Allocca, Q. Chen, L.R. Glicksman, Design analysis of single-sided natural
regarded as an innovative aspect compared to previous studies ventilation, Energy Build. 35 (8) (2003) 785e795.
[6] Y. Jiang, D. Alexander, H. Jenkins, R. Arthur, Q. Chen, Natural ventilation in
from literature. From this study, the following main conclusions can
buildings: measurement in a wind tunnel and numerical simulation with
be made: large-eddy simulation, J. Wind Eng. Ind. Aerodyn. 91 (3) (2003) 331e353.
[7] P. Heiselberg, Y. Li, A. Andersen, M. Bjerre, Z. Chen, Experimental and CFD
evidence of multiple solutions in a naturally ventilated building, Indoor Air 14
 The five different steady RANS models provide very different
(2003) 43e54.
results. The RNG, SST and RSM models provide about the same [8] M.J. Cook, Y. Ji, G.R. Hunt, CFD modeling of natural ventilation: combined wind
location of maximum velocity (and thus jet direction) along and buoyancy forces, Int. J. Vent. 1 (3) (2003) 169e180.
each vertical line. SKE and RLZ provide a too high location of [9] T. Yang, N.G. Wright, D.W. Etheridge, A.D. Quinn, A comparison of CFD and
full-scale measurements for analysis of natural ventilation, Int. J. Vent. 4 (4)
maximum velocity (too horizontal direction compared to ex- (2006) 337e348.
periments). In addition, SKE and RLZ predict a larger width of [10] T. Norton, J. Grant, R. Fallon, D.-W. Sun, Assessing the ventilation effectiveness
the incoming jet (larger value for ((d0.5;upperþd0.5;lower)/HO) and of naturally ventilated livestock buildings under wind dominated conditions
using computational fluid dynamics, Biosyst. Eng. 103 (1) (2009) 78e99.
smaller velocity gradients in the shear layer of the incoming jet [11] T. van Hooff, B. Blocken, Coupled urban wind flow and indoor natural venti-
than RNG, SST and RSM. lation modelling on a high-resolution grid: a case study for the Amsterdam
 The general flow pattern in the lower part of the enclosure is ArenA stadium, Environ. Modell. Softw. 25 (1) (2010) 51e65.
[12] Z.T. Ai, C.M. Mak, Determination of single-sided ventilation rates in multistory
better reproduced by RNG and SST, and to a lesser extent by buildings: evaluation of methods, Energy Build. 69 (2014) 292e300.
RSM, while the flow in the upper part is more accurately [13] Z.T. Ai, C.M. Mak, Modeling of coupled urban wind flow and indoor air flow on
reproduced by SKE, RLZ and RSM. a high-density near-wall mesh: sensitivity analyses and case study for single-
sided ventilation, Environ. Model. Softw. 60 (2014) 57e68.
 Turbulent kinetic energy is underpredicted in large areas above
[14] Z. Tong, Y. Chen, A. Malkawi, Defining the Influence Region in neighborhood-
and below the incoming jet by all five RANS turbulence models. scale CFD simulations for natural ventilation design, Appl. Energy 182 (2016)
 SKE and RLZ predict too high levels of turbulent kinetic energy 625e633.
[15] S. Murakami, A. Mochida, Y. Hayashi, Examining the k-ε model by means of a
in the incoming jet region, which originate from outside the
wind tunnel test and large-eddy simulation of the turbulence structure
enclosure (stagnation region). This stagnation region with high around a cube, J. Wind Eng. Ind. Aerodyn. 35 (1990) 87e100.
values of k/U2H is much larger for SKE and RLZ than for RNG, SST [16] S. Murakami, A. Mochida, Y. Hayashi, S. Sakamoto, Numerical study on
and RSM. velocity-pressure field and wind forces for bluff bodies by keε, ASM and LES,
J. Wind Eng. Ind. Aerodyn. 44 (1e3) (1992) 2841e2852.
 SKE provides the best agreement with the experiments in terms [17] S. Murakami, Comparison of various turbulence models applied to a bluff
of the validation metrics (FAC2, FAC1.3, FB, NMSE) for x-velocity body, J. Wind Eng. Ind. Aerodyn. 46e47 (1993) 21e36.
and turbulent kinetic energy; the worst agreement is obtained [18] C.J. Baker, Wind engineering e past, present and future, J. Wind Eng. Ind.
Aerodyn. 95 (9e11) (2007) 843e870.
with RLZ and RSM. [19] A. Mistriotis, G. Bot, P. Picuno, G. Scarascia-Mugnozza, Analysis of the effi-
 The LES simulation provides a very good agreement with the ciency of greenhouse ventilation using computational fluid dynamics, Agric
experimental data, both with respect to the mean velocities, For Meteorol. 85 (1997) 217e228.
[20] A. Mistriotis, C. Arcidiacono, P. Picuno, G.P.A. Bot, G. Scarascia-Mugnozza,
flow pattern (jet direction), and turbulent kinetic energy, Computational analysis of ventilation in greenhouses at zero-and low-wind-
although the latter is overestimated in the jet region in the LES speeds, Agric For Meteorol. 88 (1997) 121e135.
simulation. [21] T. Kurabuchi, M. Ohba, A. Arashiguchi, T. Iwabuchi, Numerical study of airflow
structure of a cross-ventilated model building, in: Air Distribution in Rooms,
 LES performs better than the steady RANS models, which is
2000 (ROOMVENT 2000). Edited by Awbi HB.
supported by the results of the four validation metrics (FAC2, [22] Y. Jiang, Q. Chen, Effect of fluctuating wind direction on cross natural venti-
FAC1.3, FB, NMSE). The better performance of LES can largely be lation in buildings from large eddy simulation, Build. Environ. 37 (4) (2002)
379e386.
attributed to the unsteady flow features (e.g. vertical jet flap-
[23] T. Bartzanas, T. Boulard, C. Kittas, Effect of vent arrangement on windward
ping) present in the flow, as also observed in the experiments, ventilation of a tunnel greenhouse, Biosyst. Eng. 88 (2004) 479e490.
which is reproduced by the LES model and not by the steady [24] T. Kurabuchi, M. Ohba, T. Endo, Y. Akamine, F. Nakayama, Local dynamic
RANS models. similarity model of cross-ventilation Part 1-Theoretical framework, Int. J.
Vent. 2.4 (2004) 371e382.
 The use of LES entails an increase in computational demand [25] A. Shklyar, A. Arbel, Numerical model of the three-dimensional isothermal
with a factor of z80e100, which is a strong increase, especially flow patterns and mass fluxes in a pitched-roof greenhouse, J. Wind Eng. Ind.
when considering that LES is performed on the coarse grid and Aerodyn. 92 (2004) 1039e1059.
[26] C.-H. Hu, T. Kurabuchi, M. Ohba, Numerical study of cross-ventilation using
RANS on the basic grid. The majority of the RANS simulations two-equation RANS turbulence models, I J. Vent. 4 (2) (2005) 123e132.
took a bit less than 12 h, while the LES simulation lasted about [27] I. Lee, S. Lee, G. Kim, J. Sung, S. Sung, Y. Yoon, PIV verification of greenhouse
40 days. ventilation air flows to evaluate CFD accuracy, Trans. ASAE Am. Soc. Agric.
Eng. 48 (2005) 2277e2288.
[28] A. Mochida, H. Yoshino, T. Takeda, T. Kakegawa, S. Miyauchi, Methods for
controlling airflow in and around a building under cross-ventilation to
Acknowledgements improve indoor thermal comfort, J. Wind Eng. Ind. Aerodyn. 93 (2005)
437e449.
[29] A. Mochida, H. Yoshino, S. Miyauchi, T. Mitamura, Total analysis of cooling
Twan van Hooff is currently a postdoctoral fellow of the effects of cross-ventilation affected by microclimate around a building, Sol.
Research Foundation Flanders (FWO) and acknowledges its finan- Energy 80 (2006) 371e382.
[30] G. Evola, V. Popov, Computational analysis of wind driven natural ventilation
cial support (project FWO 12R9715N). The authors also gratefully
in buildings, Energy Build. 38 (5) (2006) 491e501.
acknowledge the academic partnership with ANSYS CFD.
164 T. van Hooff et al. / Building and Environment 114 (2017) 148e165

[31] H. Fatnassi, T. Boulard, C. Poncet, M. Chave, Optimisation of greenhouse insect Turbulence Closure Model, NASA STI/Recon Tech Rep N 88, 1987, p. 11969.
screening with computational fluid dynamics, Biosyst. Eng. 93 (2006) [60] M. Kato, B.E. Launder, The modeling of turbulent flow around stationary and
301e312. vibrating square cylinders, in: 9th Symposium on Turbulent Shear Flows,
[32] T. Bartzanas, C. Kittas, A.A. Sapounas, C. Nikita-Martzopoulou, Analysis of Tokyo, 1993, pp. 1041e1046.
airflow through experimental rural buildings: sensitivity to turbulence [61] M. Tsuchiya, S. Murakami, A. Mochida, K. Kondo, Y. Ishida, Development of a
models, Biosyst. Eng. 97 (2007) 229e239. new k-ε model for flow and pressure fields around bluff body, J. Wind Eng.
[33] C.-H. Hu, M. Ohba, R. Yoshie, CFD modelling of unsteady cross ventilation Ind. Aerodyn. 67 (1997) 169e182.
flows using LES, J. Wind Eng. Ind. Aerodyn. 96 (10e11) (2008) 1692e1706. [62] P. Durbin, On the k-ε stagnation point anomaly, Int. J. Heat. Fluid Flow. 17
[34] G.M. Stavrakakis, M.K. Koukou, M.G. Vrachopoulos, N.C. Markatos, Natural (1996) 89e90.
cross-ventilation in buildings: building-scale experiments, numerical simu- [63] P.R. Spalart, W.H. Jou, M. Strelets, S.R. Allmaras, Comments on the Feasibility
lation and thermal comfort evaluation, Energy Build. 40 (2008) 1666e1681. of LES for Wings and on the Hybrid RANS/LES Approach, 1997. Advances in
[35] M. Teitel, G. Ziskind, O. Liran, V. Dubovsky, R. Letan, Effect of wind direction on DNS/LES. In: Proceedings of the first AFOSR international conference on DNS/
greenhouse ventilation rate, airflow patterns and temperature distributions, LES; 1997.
Biosyst. Eng. 101 (2008) 351e369. [64] P. Karava, T. Stathopoulos, A.K. Athienitis, Airflow assessment in cross-
[36] T. Kobayashi, K. Sagara, T. Yamanaka, H. Kotani, S. Takeda, M. Sandberg, ventilated buildings with operable facade elements, Build. Environ. 46
Stream tube based analysis of problems in prediction of cross-ventilation rate, (2011) 266e279.
Int. J. Vent. 7 (2009) 321e334. [65] Y. Tominaga, B. Blocken, Wind tunnel experiments on cross-ventilation flow
[37] T. Kurabuchi, M. Ohba, T. Nonaka, Domain decomposition technique applied of a generic building with contaminant dispersion in unsheltered and shel-
to the evaluation of cross-ventilation performance of opening positions of a tered conditions, Build. Environ. 92 (2015) 452e461.
building, I J. Vent. 8 (3) (2009) 207e217. [66] Y. Tominaga, B. Blocken, Wind tunnel analysis of flow and dispersion in cross-
[38] R.N. Meroney, CFD prediction of airflow in buildings for natural ventilation, in: ventilated isolated buildings: impact of opening positions, J. Wind Eng. Indus
Proceedings 11th Americas Conference on Wind Engineering, San Juan, Puerto Aerodyn. 155 (2016) 74e88.
Rico, 2009, pp. 1e11. [67] J. Franke, A. Hellsten, H. Schlünzen, B. Carissimo (Eds.), Best Practice Guideline
[39] T. Kobayashi, M. Sandberg, H. Kotani, L. Claesson, Experimental investigation for the CFD Simulation of Flows in the Urban Environment, COST Office
and CFD analysis of cross-ventilated flow through single room detached house Brussels, 2007. ISBN 3-00-018312-4.
model, Build. Environ. 45 (2010) 2723e2734. [68] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa,
[40] T. Norton, J. Grant, R. Fallon, D.W. Sun, Optimising the ventilation configura- T. Shirasawa, AIJ guidelines for practical applications of CFD to pedestrian
tion of naturally ventilated livestock buildings for improved indoor environ- wind environment around buildings, J. Wind Eng. Ind. Aerodyn. 96 (2008)
mental homogeneity, Build. Environ. 45 (2010) 983e995. 1749e1761.
[41] T. Norton, J. Grant, R. Fallon, D.W. Sun, Improving the representation of [69] B. Blocken, Computational Fluid Dynamics for urban physics: importance,
thermal boundary conditions of livestock during CFD modelling of the indoor scales, possibilities, limitations and ten tips and tricks towards accurate and
environment, Comput. Electron Agric. 73 (2010) 17e36. reliable simulations, Build. Environ. 91 (2015) 219e245.
[42] K.S. Nikas, N. Nikolopoulos, A. Nikolopoulos, Numerical study of a naturally [70] B. Blocken, T. Stathopoulos, J. Carmeliet, CFD simulation of the atmospheric
cross-ventilated building, Energy Build. 42 (2010) 422e434. boundary layer: wall function problems, Atmos. Environ. 41 (2) (2007)
[43] T. van Hooff, B. Blocken, On the effect of wind direction and urban sur- 238e252.
rounding on natural ventilation of a large semi-enclosed stadium, Comput. [71] P.J. Richards, R.P. Hoxey, Appropriate boundary conditions for computational
Fluids 39 (2010) 1146e1155. wind engineering models using the k-ε turbulence model, J. Wind Eng. Ind.
[44] J.O.P. Cheung, C.H. Liu, CFD simulations of natural ventilation behaviour in Aerodyn. 46&47 (1993) 145e153.
high-rise buildings in regular and staggered arrangements at various spacings, [72] E. Sergent, Vers une me thode de couplage entre la simulation des grandes
Energy Build. 43 (2011) 1149e1158. chelles et les mode
e les statistiques (in French), PhD thesis at Ecole doctoral
[45] T.S. Larsen, N. Nikolopoulos, A. Nikolopoulos, G. Strotos, K.S. Nikas, Charac- MEGA, L'Ecole Centrale de Lyon, France, 2002.
terization and prediction of the volume flow rate aerating a cross ventilated [73] P. Gousseau, B. Blocken, T. Stathopoulos, G.J.F. van Heijst, Near-field pollutant
building by means of experimental techniques and numerical approaches, dispersion in an actual urban area: analysis of the mass transport mechanism
Energy Build. 43 (2011) 1371e1381. by high-resolution Large Eddy Simulations, Comput. Fluids 114 (2015)
[46] L.J. Lo, A. Novoselac, CFD simulation of cross-ventilation using fluctuating 151e162.
pressure boundary conditions, ASHRAE Trans. 117 (1) (2011). LVe11eC075. [74] P. Gousseau, B. Blocken, G.J.F. van Heijst, Quality assessment of Large-Eddy
[47] L.J. Lo, Predicting Wind Driven Cross Ventilation in Buildings with Small Simulation of wind flow around a high-rise building: validation and solu-
Openings, PhD thesis, University of Texas, USA, 2012. tion verification, Comput. Fluids 79 (2013) 120e133.
[48] R. Ramponi, B. Blocken, CFD simulation of cross-ventilation for a generic [75] P. Gousseau, B. Blocken, G.J.F. van Heijst, CFD simulation of pollutant disper-
isolated building: impact of computational parameters, Build. Environ. 53 sion around isolated buildings: on the role of convective and turbulent mass
(2012) 34e48. fluxes in the prediction accuracy, J. Hazard Mater 194 (2011) 422e434.
[49] R. Ramponi, B. Blocken, CFD simulation of cross-ventilation flow for different [76] P. Gousseau, B. Blocken, T. Stathopoulos, G.J.F. van Heijst, CFD simulation of
isolated building configurations: validation with wind tunnel measurements near-field pollutant dispersion on a high-resolution grid: a case study by LES
and analysis of physical and numerical diffusion effects, J. Wind Eng. Ind. and RANS for a building group in downtown Montreal, Atmos. Environ. 45 (2)
Aerodyn. 104e106 (2012) 408e418. (2011) 428e438.
[50] L.J. Lo, A. Novoselac, Effect of indoor buoyancy flow on wind-driven cross [77] T. van Hooff, B. Blocken, P. Gousseau, G.J.F. van Heijst, Counter-gradient
ventilation, Build. Simul. 6 (2013) 69e79. diffusion in a slot-ventilated enclosure assessed by LES and RANS, Comput.
[51] M.F. Mohamed, S. King, M. Behnia, D. Prasad, Coupled outdoor and indoor Fluids 96 (2014) 63e75.
airflow prediction for buildings using Computational Fluid Dynamics (CFD), [78] ANSYS, Fluent 12 User's Guide, Fluent Inc, Lebanon, 2009.
Build 3 (2013) 399e421. [79] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows,
[52] T. van Hooff, B. Blocken, CFD evaluation of natural ventilation of indoor en- Comput. Method Appl. M. 3 (1974) 269e289.
vironments by the concentration decay method: CO2 gas dispersion from a [80] T. Cebeci, P. Bradshaw, Momentum Transfer in Boundary Layers, Hemisphere
semi-enclosed stadium, Build. Environ. 61 (2013) 1e17. Publishing Corporation, 1977.
[53] J.I. Peren, T. van Hooff, R. Ramponi, B. Blocken, B.C.C. Leite, Impact of roof [81] B. Blocken, T. Stathopoulos, J. Carmeliet, Wind environmental conditions in
geometry of an isolated leeward sawtooth roof building on cross-ventilation: passages between two long narrow perpendicular buildings, J. Aerosp. Eng.
straight, concave, hybrid or convex? J. Wind Eng. Ind. Aerodyn. 145 (2015) ASCE 21 (4) (2008) 280e287.
102e114. [82] T.H. Shih, W.W. Liou, A. Shabbir, Z.G. Yang, J. Zhu, A new kappa-epsilon eddy
[54] J.I. Peren, T. van Hooff, B.C.C. Leite, B. Blocken, Impact of eaves on cross- viscosity model for high Reynolds-number turbulent flows, Comput. Fluids 24
ventilation of a generic leeward sawtooth roof building: windward eaves, (3) (1995) 227e238.
leeward eaves and eaves inclination, Build. Environ. 92 (2015) 578e590. [83] V. Yakhot, S.A. Orszag, S. Thangam, T.B. Gatski, C.G. Speziale, Development of
[55] J.I. Peren, T. van Hooff, B.C.C. Leite, B. Blocken, CFD analysis of cross-ventilation turbulence models for shear flows by a double expansion technique, Phys.
of a generic isolated building with asymmetric opening positions: impact of Fluids A 4 (1992) 1510e1520.
roof angle and opening location, Build. Environ. 85 (2015) 263e276. [84] D. Choudhury, Introduction to the Renormalization Group Method and Tur-
[56] J.I. Peren, T. van Hooff, B.C.C. Leite, B. Blocken, CFD simulation of wind-driven bulence Modeling, 1993.
upward cross ventilation and its enhancement in long buildings: impact of [85] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering
single-span versus double-span leeward sawtooth roof and opening ratio, applications, AIAA J. 32 (1994) 1598e1605.
Build. Environ. 96 (2016) 142e156. [86] D.C. Wilcox, Turbulence Modeling for CFD, DCW Industries, Inc., La Canada,
[57] N.R. Martins, G. Carrilho da Graça, Validation of numerical simulation tools for California, 1998.
wind-driven natural ventilation design, Build. Sim 9 (2016) 75e87. [87] J. Smagorinsky, General circulation experiments with the primitive equations.
[58] Z. Tong, Y. Chen, A. Malkawi, G. Adamkiewicz, J.D. Spengler, Quantifying the I. The basic experiment, Mon. Weather Rev. 91 (1963) 99e164.
impact of traffic-related air pollution on the indoor air quality of a naturally [88] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, A dynamic subgrid-scale eddy
ventilated building, Environ. Int. 89e90 (2016) 138e146. viscosity model, Phys. Fluids A 3 (1991) 1760e1765.
[59] Y.S. Chen, S.W. Kim, Computation of Turbulent Flows Using an Extended K-ε [89] D.K. Lilly, A proposed modification of the Germano subgrid-scale closure
T. van Hooff et al. / Building and Environment 114 (2017) 148e165 165

method, Phys. Fluids A 4 (1992) 633e635. blades, J Turbomach. 124 (2002) 187e192.
[90] P.J. Roache, Quantification of uncertainty in computational fluid dynamics, [93] Y. Tominaga, Flow around a high-rise building using steady and unsteady
Annu. Rev. Fluid Mech. 29 (1997) 123e160. RANS CFD: effect of large-scale fluctuations on the velocity statistics, J. Wind
[91] S. Murakami, Overview of turbulence models applied in CWE-1997, J. Wind Eng. Ind. Aerodyn. 142 (2015) 93e103.
Eng. Ind. Aerodyn. 74e76 (1998) 1e24. [94] M. Wang, Q. Chen, On a hybrid RANS/LES approach for indoor airflow
[92] G. Medic, P.A. Durbin, Toward improved prediction of heat transfer on turbine modeling e RP1271, HVAC&R Res. 16 (2010) 731e747.

You might also like