You are on page 1of 18

Building and Environment 126 (2017) 355–372

Contents lists available at ScienceDirect

Building and Environment


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

CFD and wind-tunnel analysis of outdoor ventilation in a real compact T


heterogeneous urban area: Evaluation using “air delay”
Nestoras Antonioua,b,∗, Hamid Montazerib,c, Hans Wigod, Marina K.-A. Neophytoua,
Bert Blockenb,c, Mats Sandbergd
a
Department of Civil and Environmental Engineering, University of Cyprus, Nicosia, Cyprus
b
Department of the Built Environment, Eindhoven University of Technology, Eindhoven, The Netherlands
c
Department of Civil Engineering, KU Leuven, Leuven, Belgium
d
Department of Building, Energy and Environmental Engineering University of Gävle, Gavle, Sweden

A R T I C L E I N F O A B S T R A C T

Keywords: Outdoor urban ventilation in a real complex urban area is investigated by introducing a new ventilation in-
Computational Fluid Dynamics (CFD) dicator – the “air delay”. Computational Fluid Dynamics (CFD) simulations are performed using the 3D steady
Urban microclimate Reynolds-Averaged Navier-Stokes (RANS) and Large Eddy Simulation (LES) approaches. The up-to-date litera-
Large Eddy Simulations (LES) ture shows the lack of detailed evaluations of the two approaches for real compact urban areas. This study
Steady RANS
further presents a systematic evaluation of steady RANS and LES for the assessment of the ventilation conditions
City breathability
Air delay
in a dense district in Nicosia, Cyprus. The ventilation conditions within the urban area are investigated by
calculating the distribution of the age of air. To better assess the outdoor ventilation, a new indicator, the “air
delay” is introduced as the difference between the local mean age of air at an urban area and that in an empty
domain with the same computational settings, allowing the comparison of the results in different parts of the
domain, without impact of the boundary conditions. CFD results are validated using wind-tunnel measurements
of mean wind speed and turbulence intensity performed for the same urban area. The results show that LES can
accurately predict the mean wind speed and turbulence intensity with the average deviations of about 6% and
14%, respectively, from the wind-tunnel measurements while for the steady RANS, these are 8% and 31%,
respectively. The steady RANS simulations overestimate the local mean air delay. The deviation between the two
approaches is 52% at pedestrian level (2 m).

1. Introduction aspect ratio of urban canyons [8–12]. For example, it has been shown in
recent research work that building height variation can be beneficial in
More than 80% of the population living in urban areas is exposed to terms of breathability levels [7,11,13,14], while larger aspect ratios of
poor air quality levels [1]. In addition, the urban population has urban canyons can lead to higher pollutant concentrations inside the
reached 54% of the world's population, while it is expecting to increase street canyons [8,11]. The impact of meteorological parameters, such as
up to 66% by 2050 [2]. Consequently, without proper mitigation wind speed, wind direction, temperature and humidity is also extremely
measures, outdoor air pollution is expected to become the main reason important for urban ventilation. Furthermore, the actual pollutant
for premature mortality worldwide by 2050 [3]. In order to improve the concentration distribution in canyons was also found to be strongly
air quality in urban areas, strategies for improving urban ventilation linked to the source release characteristics [12] while compact cities
need to be evaluated and implemented. appear to have less well ventilated conditions [6].
Urban ventilation, also alternatively termed as urban breathability, While the contribution of laboratory and field experiments in the
is defined as the ability of an urban area to dilute pollutants, heat and progress of our understanding of urban fluid dynamics is well-estab-
moisture by exchanging air between inside and above of the urban lished, in the past decades, Computational Wind Engineering (CWE) has
canopy [4], while earlier research has shown that it can play a sig- become an increasingly established research method and field [15],
nificant role in urban air quality [5]. Urban ventilation depends on the covering a wide range of topics such as pedestrian-level wind comfort
combination of a wide variety of urban geometry parameters such as around buildings [7,16,17], pollutant dispersion [6,18,19], outdoor
frontal area density (λf), plan area density (λp) [6,7] as well as the thermal environment [5,20–22], convective heat transfer [23–25], etc.


Corresponding author. Department of Civil and Environmental Engineering, University of Cyprus, Nicosia, Cyprus.
E-mail address: antoniou.nestoras@ucy.ac.cy (N. Antoniou).

http://dx.doi.org/10.1016/j.buildenv.2017.10.013
Received 25 June 2017; Received in revised form 7 October 2017; Accepted 9 October 2017
Available online 13 October 2017
0360-1323/ © 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Table 1
Overview of CFD studies on urban wind flow and outdoor ventilation conditions in urban areas.

Study Ref. Scale Geometry Turbulence model Validation Focus

Juan et al. (2017) [26] Neighb. Real Steady RANS (RKE) Field a, c, d
Wen et al. (2017) [11] Neighb. Generic Steady RANS (SKE) WT a,c
Nazarian and Kleissl (2016) [27] Neighb. Generic LES (WMLES) WT b, c, d
Hang et al. (2015) [28] Neighb. Generic Steady RANS (SKE, RNG) WT a, b, d
Buccolieri et al. (2015) [29] Neighb. Generic Steady RANS (SKE) WT a, c, d
Ramponi et al. (2015) [10] Neighb. Generic Steady RANS (SKE) WT a, c, d
Lin et al. (2014) [30] Neighb. Generic Steady RANS (SKE) WT a, d
Panagiotou et al. (2013) [5] Neighb. Real Steady RANS (RSM) – a
Hang et al. (2013) [31] Neighb. Generic Steady RANS (SKE, RNG, RKE, RSM) WT a, d
Hang et al. (2012b) [32] Neighb. Generic Steady RANS (SKE, RNG) WT a, c
Hang et al. (2012a) [13] Neighb. Generic Steady RANS (SKE, RNG) WT a
Tominaga and Stathopoulos (2011) [18] Street Generic Steady RANS (RNG), LES (St.) WT –
Salim et al. (2011) [33] Street Generic Steady RANS (RSM, SKE) LES (Dyn.) WT –
Gousseau et al. (2011) [19] Neighb. Real Steady RANS (SKE), LES (Dyn.) WT c, d
Moonen et al. (2011) [34] Street Generic Steady RANS (RKE), LES (Dyn.) – a, d
Kim et al. (2010) [35] Street Real Steady RANS (SKE) WT a, d
Buccolieri et al. (2010) [6] Neighb. Generic Steady RANS (SKE) WT a, c
Hang and Li. (2010b) [36] Neighb. Generic Steady RANS (SKE) WT a, c
Hang and Li. (2010a) [7] Neighb. Generic Steady RANS (SKE) WT a, c, d
Hang et al. (2010b) [38] Neighb. Generic Steady RANS (SKE, RNG) WT a
Hang et al. (2010a) [37] Neighb. Generic Steady RANS (SKE) WT a
Hang et al. (2009b) [39] City Generic Steady RANS (SKE, RNG) WT a, d
Hang et al. (2009a) [40] City Generic Steady RANS (SKE, RNG) WT a, c, d
Yim et al. (2009) [41] Neighb. Generic Steady RANS (RKE) WT a, c, d
Bu et al. (2009) [9] Neighb. Generic Steady RANS (SKE) – a, d
Cheng et al. (2008) [42] Neighb. Generic Steady RANS (RNG) WT a, c
Blocken et al. (2008) [43] Street Generic Steady RANS (RKE) WT a, c, d
Bady et al. (2008) [44] Street Generic Steady RANS (SKE) – a, d
Yoshie et al. (2007) [45] Neighb. Real Steady RANS (SKE) WT, Field c, d
Blocken et al. (2007a) [46] Street Generic Steady RANS (RKE) WT a, c
Xie and Castro (2006) [47] Neighb. Generic Steady RANS (SKE, RSM, MKE) LES (Dyn.) WT –
Li et al. (2005) [48] Street Generic Steady RANS (RNG) WT a
Liu et al. (2005) [8] Street Generic LES (Dyn.) WT a
Skote et al. (2005) [49] City Generic Steady RANS (SSTKO) – a, c, d

Neighb. = Neighborhood scale, Real. = Realistic, WT = Wind-tunnel measurements, Field = Field measurements RANS = Reynolds-averaged Navier-Stokes, RKE = Realizable k-ε
model, MKE = Modified k-ε model, SKE = Standard k-ε model, RNG = Renormalization Group k-ε model, LES = Large Eddy Simulation, Dyn. = Dynamic Smagorinsky-Lilly SGS model,
St. = Standard Smagorinsky-Lilly SGS model, WMLES = Wall-Modeled SGS model, SSTKO = SST k-ω model. The entry “Focus” refers to different aspects that have been investigated in
each study, which are (a) geometrical characteristics (building height, street width, plan area density (λp), etc.), (b) heated surfaces, (c) grid resolution, (d) wind direction.

However, there is still need for further investigation of the performance our knowledge, the detailed evaluation of the two approaches for the
of CFD for different applications in wind engineering and urban phy- outdoor ventilation in complex urban areas has not yet been in-
sics. Table 1 provides a non-exhaustive overview of a number of CFD vestigated. Therefore, the new contribution of this paper is to provide
studies focusing on urban ventilation. The following observations can an investigation of the outdoor ventilation conditions for real complex
be made: urban areas and provide a quantitative evaluation of the accuracy of
LES and RANS (through the wind-tunnel study validation) for relevant
• Based on the spatial classification by Britter and Hanna [50], these metrics for such outdoor ventilation.
studies focus on either street scale (∼100 m), neighborhood scale It is well known that steady RANS is incapable of capturing the
(∼1 km) or city scale (∼10 km). It should be noted that at the city inherently transient behavior of separation and recirculation down-
scale, the focus is on the general shape of cities, while individual stream of buildings and of von Karman vortex shedding in the wake
buildings are not modeled explicitly in these studies [39,40]. [51,52,54,55]. Transition simulations with Large-Eddy Simulation
• The vast majority of the studies have been performed for generic (LES), on the other hand can provide accurate descriptions of the mean
urban configurations, with idealized geometry, while there is a clear and instantaneous flow field around bluff bodies, at the expense of
gap on information of ventilation conditions in real urban en- much larger computational requirements. However, steady RANS is still
vironments where the geometry is much more complex. the most frequently used method for urban breathability studies based
• In addition, 3D steady RANS is the most commonly used approach on the information provided in Table 1, since it can be very effective in
for such CFD studies. It should be noted that in almost all the studies terms of time and computational cost. It should be noted that earlier
using transient simulations, the simulations were performed for studies on urban ventilation have shown the significant contribution of
generic configurations, while for the study by Gousseau et al. [19], turbulence diffusion in the air mixing between inside and above of
in which transient simulations were used in a real urban area, the urban canopies [5,30,39,53,56], while turbulence can act as the main
focus was not on the ventilation conditions. mechanism controlling pollutant removal from an urban canopy layer
[12,57]. Given the important influence of turbulence on outdoor ven-
The comparison of RANS and LES has been a subject of many studies tilation, a detailed evaluation of steady RANS and LES for outdoor
[19,51–53]. However, based on the literature review that presents a ventilation of complex real urban areas is in order. Furthermore, using
sample of relevant studies, most of the formal cross-comparative in- the information provided in this paper in terms of accuracy and com-
vestigations deal with idealized urban geometries and the few dealing putational time, engineers, urban planners and researchers can easily
with real complex urban geometries do not address outdoor ventilation decide which of these turbulence modeling approaches would be more
as measured through specific quantitative metrics; in fact, to the best of beneficial.

356
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 1. Case study area and surroundings: (a) Top


view, (b) Aerial view taken from point 1 on
Fig. 1a and (c) taken from the Shakolas tower
indicated with point 2 on Fig. 1a. (d) Street view
at Ledras street. Black line on Fig. 1a indicates
borders of the case study area.

To gain insights into the performance of steady RANS compared to way. To cover this void, and to better assess the outdoor ventilation for
LES for predicting urban ventilation in a real compact heterogeneous the urban area, a new indicator called air delay (τd) is introduced,
urban area, this paper presents steady RANS and LES CFD simulations which is actually the difference between the local mean age of air inside
of the wind flow and the local mean age of air (τp ) distribution in a the area of interest, and the local mean age of air in an empty domain
dense district in the old city of Nicosia, Cyprus. Age of air represents the (when the buildings are not present) with the same computational
time it takes for the external air to reach a location after entering the settings and parameters. The air delay is independent of the boundary
urban canopy layer [58]. The evaluation is based on validation with conditions as it presents the amount of time delay caused by the pre-
wind-tunnel measurements of mean wind speed and turbulence in- sence of the urban geometry in each point inside the domain. In addi-
tensity for a reduced-scaled urban configuration representing this dis- tion, the air delay parameter can be used to easily assess and compare
trict. The wind-tunnel measurements, CFD simulations and ventilation ventilation conditions in different regions of an urban area, but also in
analysis are performed for a reduced-scale urban model. different urban areas.
The concept of age of air was firstly introduced as a ventilation In Section 2 the selected urban area and its surroundings are pre-
indicator for indoor environment, where it was used to characterize the sented. Section 3 describes the wind-tunnel measurements. Section 4
freshness of the air, representing the time that has passed since the air describes the computational settings and parameters for the CFD si-
entered the room [59]. This concept can be interpreted as the link mulations. In Section 5, the validation of the CFD results with the wind-
between a concentration level and a time scale [6]. For indoor en- tunnel measurements is presented, the performance of LES versus
vironments, the inlets and outlets are easily defined, according to the steady RANS is evaluated, and the limitations and further work are
position of the openings. However, in an open environment such as an discussed. The main conclusions are presented in section 6.
urban area, this is not straightforward. In order to address this, Hang
et al. [40] used the effective age of air, where the emission source is set
in the entire urban canopy. This was also applied in other studies with 2. Description of urban area and surroundings
generalized geometries [10]. For real complex urban environments
such the case of this study, where the urban canopy cannot be easily The complex heterogeneous urban area under study covers
defined, this method is less applicable. Other currently available me- 0.247 km2 and is composed of low-rise and medium-rise buildings
trics such as exchange velocity, pollutant exchange velocity [4,5,12], (Fig. 1), corresponding to Local Climate Zone 3 (compact low-rise
air exchange rate [48] and pollutant exchange rate [8] characterize buildings) based on the classification made by Steward and Oke [60]
ventilation using a process occurring over the rooftop level of canyons, and with hot-summer Mediterranean climate (Csa) based on the
not the process of dilution, removal and recirculation of contaminants Köppen-Geiger classification [61]. The average and maximum building
which occur inside the urban areas. This motivation arises from the height is about 8.1 m and 45.5 m (Shakolas Tower), respectively. The
interest that within large long canyons in highly complex real urban plan area density λp is 0.40, as derived from European Research Project
areas local ventilation processes are bound to occur in a non-uniform TOPEUM [20]. Most of the buildings have flat terrace-type roofs and
internal courtyards typical of urban textures in Mediterranean cities.

357
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 2. (a) Top view of the case study area with


buildings colored by building height (zHmax =
45.5 m). Results of Multi-Resolution Analysis: (b)
Aerodynamic surface roughness length z0 (z0max
= 0.5 m), (c) Plan area density λp (λpmax =
0.8 m), (d) Zero plane displacement height zd
(zdmax = 7 m).

The area of interest is composed of narrow streets with widths ranging which they related the aerodynamic roughness parameters with planar
from 6 m to 10 m, where the aspect ratio of the street canyons (width to density and the average building height of the investigated neighbor-
height) ranges from 0.5 to 1. All streets are single direction for vehicles, hood. Then the MRA method by Mouzourides et al. [62] was used to
while the main street crossing the area in the south-north direction is provide scale-adaptive values of the z0 and zd at different resolutions
used exclusively by pedestrians (Ledras Street shown in Fig. 1a and d). relevant and appropriate to our calculations. Since z0 and zd are both
The lack of green areas in combination with the use of concrete and defined based on geometrical characteristics of the urban area such as
asphalt, as main building materials, lead to relatively high air tem- average building height and planar density, it would be reasonable to
peratures inside the street canyons during the summer period. investigate their relationship with the breathability levels. Moreover,
In this study, we also investigate the association of urban ventilation the analysis of the aerodynamic characteristics is considered extremely
with morphological and aerodynamic characteristics of the urban area. important in order to predict and describe the urban wind flow and
To do so, a scale-adaptive description of the morphological and aero- pollutant dispersion at different scales [50].
dynamic parameters is used, based on the Multi-Resolution Analysis
(MRA) method proposed by Mouzourides et al. [62]. The area is divided 3. Wind-tunnel measurements
into cells of horizontal surface area of 64 × 64 m2 and produces a
spatial varying map of the parameters. Fig. 2 presents the building The wind-tunnel measurements are conducted in the atmospheric
height (H) (Fig. 2a), aerodynamic roughness length (z0) (Fig. 2b), plan boundary layer wind tunnel at the University of Gävle, Sweden. The
area density (λp) (Fig. 2c) and the zero plane displacement height dis- closed-circuit wind-tunnel is 11 m long and has a test section of
tribution (zd) (Fig. 2d) of the area of interest. The calculation of z0 and 3 × 1.5 m2 (width × height). A reduced-scale model (1:250) of the
zd has been done using the relations proposed by Kastner-Klein and urban area is manufactured (Fig. 3). The blockage ratio is 2.7%. The
Rotach [63] derived after a series of wind-tunnel measurements in buildings are modeled with details of 2 m (full scale). The

358
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 3. (a) Top view of reduced-scale model used in wind-


tunnel measurements. (b) Close up of the model around the
Shakolas Tower.

measurements are performed for wind direction 290° from North, de- height of the highest building (45.5 m in full scale). The resulting
duced as prevailing wind direction from analyzing relevant associated vertical profile of mean wind speed at the middle upstream measuring
field measurements in the broader area [64]. The atmospheric location “2” (Fig. 4a) is represented by a log law with an aerodynamic
boundary layer is generated by roughness elements and spires. The roughness length z0 = 0.008 m (2 m in full scale) and a friction velocity

mean wind speed and turbulence intensity are measured using hot-wire uABL = 0.79 m/s. The results of the wind-tunnel measurements will be
anemometers at 1261 points along 38 vertical lines (Fig. 4). The mea- presented along with the CFD results in Section 5. In this study, due to
surement lines are located in different parts of the domain as shown in lack of high-resolution on-site and wind-tunnel data of age of air for
Fig. 4: (i) 3 lines in the upstream part of the area of interest to capture complex urban areas, the CFD validation study is performed based on
the approach-flow profiles (Fig. 4a); (ii) 3 lines downstream of the area wind-tunnel measurements of mean wind speed and turbulence in-
(Fig. 4b); (iii) 16 lines on top of the buildings (Fig. 4c) and (iv) 16 lines tensity.
inside the urban canyons (Fig. 4d). The heights of the measurement
points are distributed in a way to have more information near the
ground plane, with intervals as shown in Fig. 5, where Hmax is the

Fig. 4. Top view of the case study area with the


position of measurement lines: (a) Upstream of
the urban area (b) Downstream of the urban area
(c) at roof top levels (d) at street levels.

359
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 5. Height of measuring points corresponding to vertical


lines presented in Fig. 4 (Hmax = 0.182 m).

4. CFD simulations where the buildings are modeled explicitly, i.e. with their main shape,
the ground plane and building walls are modeled as smooth walls with
4.1. Computational geometry, domain and grid z0 = 0 m. At the outlet of the domain, zero static gauge pressure is
imposed.
The computational geometry of the urban area reproduces the Note that in the measurements, the lateral homogeneity of the ap-
geometry in wind-tunnel measurements at scale 1:250. The dimensions proach-flow profile is investigated measuring the mean wind speed and
of the computational domain are chosen based on the best practice turbulence intensity along the three lines (Fig. 4a), upstream of the area
guidelines by Refs. [65–67] (Fig. 6). The height of the domain is 6Hmax of interest, as presented in Fig. 8. The average deviation between the
where Hmax is the maximum building height. The upstream and line in the middle and the other two lines is within 4% for mean wind
downwind domain size is 5Hmax and 15Hmax, respectively. The width of speed and 6% for turbulence intensity. The turbulence kinetic energy k
the domain is set the same as the wind-tunnel width resulting in a is calculated from the measured U and Iu using Eq. (1) where “ɑ” is a
blockage ratio of 2.7%, which is below the maximum value of 3.0% constant parameter ranging between 0.5 and 1.5. In this study, ɑ = 1 is
recommended by the aforementioned guidelines. The computational taken, as recommended by Tominaga [66]. The turbulence dissipation
grid was created using the surface-grid extrusion technique by van rate ε at height z from the ground is given by Eq. (2) where uABL

is the
Hooff and Blocken [68]. The procedure was executed with the aid of the ABL friction velocity and κ the von Karman constant (0.42).
pre-processor Gambit 2.4.6, resulting in a grid with 13,408,998 hex-
k (z ) = ɑ (Iu (z ) u (z ))2 (1)
ahedral cells. The computational grid is shown in Fig. 7.
∗3
uABL
ε (z ) =
4.2. Boundary conditions κ (z + z 0) (2)

The boundary condition types are depicted in Fig. 6. The inlet


4.3. Computational settings
boundary conditions (mean wind speed U, turbulence kinetic energy k
and turbulence dissipation rate ε) are based on the measured incident
The commercial CFD code ANSYS/Fluent 15.0 based on the finite
profiles of mean wind speed and longitudinal turbulence intensity Iu
volume method is used to perform the simulations. The simulations are
along the middle line upstream of the model (line with label “2” in
performed on the HPC cluster (×86_64 architecture) at the Department
Fig. 4a). The aerodynamic roughness length for the surroundings is
of the Built Environment of Eindhoven University of Technology, the
determined based on the measured incident profiles of mean wind
Netherlands. For each simulation, a 22-core node (Intel(R) Xeon(R)
speed, as it was described in Section 3. For the inner part of the domain,
CPU - ×5650 @ 2.67 GHz), with 64 GB of system memory is used. The
computational settings and parameters for the steady RANS and LES
simulations are presented in Sections 4.3.1 and 4.3.2, respectively.

4.3.1. Steady RANS


The 3D steady RANS equations are solved in combination with
realizable k-ε turbulence model by Shih et al. [69] for closure for in-
compressible isothermal turbulent flow. The standard wall functions by
Launder and Spalding [70] are used with roughness modification by
Cebeci and Bradshaw [71]. The values of the roughness parameters, i.e.
the sand-grain roughness height ks and the roughness constant Cs, are
determined using their consistency relationship with the aerodynamic
roughness length z0 introduced by Blocken et al. [46,72]:
9.793z 0
ks =
Cs (3)
Note that up to this version of ANSYS/Fluent, the ks value cannot be
larger than yp (distance between the center point of the wall-adjacent
cell and the wall). Therefore, in this study, ks is set equal to yp and Cs is
chosen to satisfy Eq. (3). The SIMPLE algorithm is chosen for pressure-
Fig. 6. Perspective view of computational domain with indication of main dimensions velocity coupling and second-order discretization schemes are used for
and boundary conditions.
both the viscous and convection terms of the governing equations.

360
N. Antoniou et al. Building and Environment 126 (2017) 355–372

velocity is obtained using the Smagorinsky-Lilly sub-grid scale model


[73,74]. Time discretization is second-order implicit and for pressure
interpolation second order is employed. For time advancement, the
iterative time step method is adopted [75–77], where all the equations
are solved iteratively for a given time-step and for a number of outer
iterations per time step [78]. The specification of the number of outer
iterations depends on the problem complexity [79]. The results of a
sensitivity analysis, performed in the present study, reveal that the use
of 10 outer iterations per time step is sufficient.
The Vortex Method (VM) [80] is used as turbulent inflow generator
technique with 190 2D vortices. The satisfactory performance of this
technique was already shown in several occasions in the past
[19,34,80].
The time-step value (Δt) was 0.0003 s, which corresponds to an
average CFL number ( = u × Δt/hcell), where u is the local velocity
value and hcell is the local cell size) of 0.2. The maximum CFL number is
4.8, which occurs at the top of the domain, far away from the area of
interest. Two initialization steps are performed. First, the simulations
are initialized with the solution from the steady RANS simulation on
which random noise is superimposed [78]. This initialization does not
affect the results, but it reduces the computational time. Second, an LES
initialization run is performed with an initialization period (Tinit) of 4 s,
corresponding to approximately 4 flow-through times (time needed for
wind flow to cover the distance from inlet to outlet with a reference
wind speed):
Lx
Tft =
Uh (4)
where Uh is the wind speed at a the height of the highest building in the
domain and Lx is the length of the computational domain [19,81]. After
initialization, LES simulations and sampling are performed for 20 s,
corresponding to 20 flow-through times.

4.3.3. Age of air and air delay


The ventilation is assessed using the local mean age of air (τp ), as-
Fig. 7. (a) Aerial view of the case study area and surrounding buildings, (b and c) suming that the air originating from outside the urban area is cleaner.
Corresponding view of the computational grid (13,408,998 cells). The homogeneous emission source method was used in which the local
mean age of air can be calculated from the local concentration C (kg/
Convergence is assumed to have been achieved when the scaled nor- m3) of each cell, based on the uniform release rate ṁ (kg/(m3s)) of the
malized residuals stabilize at a minimum of 10−6 for continuity, x, y source where the emission is released homogeneously in the entire
and z momentum, 10−8 for k and 10−11 for ε. domain:
C
τp =
ṁ (5)
4.3.2. LES
The three-dimensional time dependent Navier-Stokes equations are The turbulent Schmidt number (Sct = νt / Dt ) is assumed to be con-
solved with LES as a turbulence modeling approach, for incompressible stant in the whole domain (Sct = 0.7 where νt is the turbulent viscosity
isothermal turbulent flow. The closure of the equations for the filtered and Dt is the turbulent diffusivity of the species), in line with similar

Fig. 8. Measured profiles of (a) ratio of normal-


ized streamwise mean wind speed, U, to reference
wind speed at Shiakolas Tower height, Uref, and
(b) streamwise turbulence intensity in the wind-
tunnel measurements.

361
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 9. Comparison of wind-tunnel measurements


and CFD results of (a) normalized streamwise
mean wind speed and (b) streamwise turbulence
intensity along line with label “2” upstream of the
urban area (see Fig. 4a) for two cases: with and
without adjusted approach-flow profile.

previous studies with the passive homogeneous emission source method be seen that this adjustment significantly reduces the appearance of the
[40,44,82]. streamwise decay. In this case the decay of Iu drops to 25% and 19% for
For the steady RANS simulations, the local mean age of air is cal- steady RANS and LES, respectively. The deviations are about 6% and
culated using the mean concentration of the passive homogeneous 4% for U/Uref.
emission source. For LES, it is derived from the instantaneous local age Figs. 10 and 11 compare the wind-tunnel results and CFD results of
of air values for each time step using Eq. (6): normalized streamwise mean wind speed and turbulence intensity
n along the vertical lines positioned at the roof top levels (points with
∑i = 0 τ pinst
τ pLES = labels 2, 7 and 16 in Fig. 4c) and inside the street canyons (points with
n (6) labels 6, 8 and 14 in Fig. 4d), respectively. For LES, a good agreement is
where τ pLES is the local mean age of air, τ pinst is the instantaneous local observed for both the mean wind speed and the turbulence intensity.
mean age of air and n is the number of time steps during the averaging This is especially the case for the mean wind speed where the deviation
period. is about 4% for the six lines. For the turbulence intensity, higher de-
viations can be clearly observed in Figs. 10f and 11f, where LES tends to
5. CFD results and discussion overestimate the results, though it can still capture the trends. Note that
the line indicated by the point with label “16” in Fig. 10 is located in the
5.1. Comparison with wind-tunnel experiments wake of the Shakolas tower, while the line indicated by the point with
label “14” in Fig. 11 is located close to the ground inside a canyon.
In order to check the horizontal homogeneity of the approaching Steady RANS also provides a satisfactory agreement for the mean wind
flow, the appearance of streamwise gradients of mean wind speed and speed for most of the lines, though its deficiency in reproducing the
turbulence intensity between the inlet and the upstream measuring flow pattern around the tower can be clearly observed in Fig. 10e–f
locations is analyzed by comparing the CFD results of the normalized where the flow is complex. This can be explained by the incapability of
streamwise velocity, U/Uref, and streamwise turbulence intensity, Iu, RANS in predicting the periodic motions of the flow [51]. In this case,
with the measured data along the line indicated by the point with label the average deviation for the six lines presented in Figs. 10 and 11 is
“2” upstream of the urban area (Fig. 4a). The results are shown in about 20%. The average deviation increases to about 37% for the tur-
Fig. 9. Note that Hmax = 0.182 m (45.5 m in real scale) is the height of bulence intensity for these lines. This is in line with the results of earlier
the highest building in the domain and Uref = 6 m/s is the mean wind studies in which the deficiency of steady RANS in capturing inherently
speed at z = Hmax at the inlet. According to this figure, the average transient behavior of separation and recirculation downstream of bluff
deviation of U/Uref from the measured data is about 5% and 2% for bodies and of von Karman vortex shedding in the wake has been shown
steady RANS and LES, respectively. For Iu, however, this deviation is [18,19,34]. Note that the higher accuracy by LES is obtained at the
more pronounced along the line and it reaches its maximum value of expense of much larger computational costs (factor 15 compared to
35% and 69%, respectively. The roughness modeling is considered the steady RANS).
main reason to cause this decay. For LES a smooth ground surface is Table 2 presents the overall average deviation of mean wind speed
assumed, which leads to reduction of turbulence levels as the flow is and turbulence intensity for all 1261 points along the 38 lines. The
developing. This is less intense in RANS where wall functions are used overall deviations for the case if the inflow adjustment at the inlet
and roughness of wall boundaries is modeled using the roughness would not have been implemented is also reported in this table. The
parameters (details in section 4.3.1). This decay will be a factor con- average deviation of mean wind speed is 8% and 6% for steady RANS
tributing to possible discrepancies between the CFD results and the and LES, respectively. This is about 31% and 14% for turbulence in-
wind-tunnel measurements inside the area of interest, as also reported tensity. It can be seen that using the adjusted profiles at the inlet can
in previous studies [66,83]. As suggested by Blocken et al. [46], to significantly improve the agreement for the LES results. For steady
reduce the streamwise decay of turbulence intensity, the standard de- RANS, however, this adjustment has less impact on the accuracy of the
viation of the measured stream-wise wind speed is artificially increased results.
by a constant factor for all measuring heights, leading to higher values Fig. 12 presents the distribution of the normalized mean wind speed
of turbulence intensity along the line. The results show that a factor of at pedestrian level (2 m from the ground), and the normalized turbu-
0.3 is needed to achieve a satisfactory match between the LES results lence kinetic energy at the average canyon roof height (6.25 m from the
and the measured data. For the steady RANS simulations, this factor is ground) obtained by steady RANS and LES, given the high importance
0.1. No adjustment is applied for the inlet mean velocity profiles. It can of turbulence fluctuations on top of urban canyons [32,38]. Considering

362
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 10. Comparison of wind-tunnel measure-


ments and CFD results of (a, c, e) normalized
streamwise mean wind speed and (b, d, f)
streamwise turbulence intensity along three lines
(with labels “2”, “7” and “16” in Fig. 4c) located
on top of buildings.

that LES outperforms RANS as shown in Figs. 10 and 11, Fig. 12 in- 5.2. Outdoor ventilation
dicates that RANS generally underestimates the local mean wind speed,
in line with previous studies on pedestrian-level wind conditions [84]. In this paper, the results will be presented as the difference between
The overall average deviation between steady RANS and LES is about the local mean age of air inside the area of interest τp , and the local
45%. Nevertheless, the channeling effect between parallel buildings mean age of air in an empty domain (when the buildings are not pre-
[85], lower wind velocities inside street canyons and courtyards [21] sent) with the same computational settings and parameters, τ pEMP . This
and wake regions are well predicted by both steady RANS and LES. On difference, which is called “air delay” in the remainder of this paper, is
the contrary, a significant difference in the distribution of the normal- given by Eq. (7):
ized turbulence kinetic energy can be clearly observed in Fig. 12c and d
τd = τp − τ pEMP (7)
where values are underestimated by RANS with about 70% deviation
from LES. Similar findings were also reported in previous CFD studies where τd is the local mean air delay, τp is the local mean age of air in
for wind flow around isolated buildings [51,54,56] and inside complex the domain when the urban geometry is present, and τ pEMP is the local
urban areas [19,34]. mean age of air in an empty domain. Note that in CFD simulations of

363
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 11. Comparison of wind-tunnel measure-


ments and CFD results of (a, c, e) normalized
streamwise mean wind speed and (b, d, f)
streamwise turbulence intensity along three lines
(with labels “6”, “8” and “14” in Fig. 4c) located
inside street canyons.

Table 2 largely dependent on the definition of the “inlet” based on which the
Average relative error from wind-tunnel measurements for all measuring points. calculations are performed. This can therefore be more problematic for
complex urban geometries, like that in the present study, as the local
Without inflow adjustment With inflow adjustment
mean age of air distribution inside the urban area is more sensitive to
eAVG(%) RANS LES RANS LES the definition of the initial values. The air delay is however in-
U 9 9 8 6 dependent of the initial values as it presents the amount of time delay
Iu 41 48 31 14 caused by the presence of the urban geometry in each point inside the
domain.
Fig. 13 shows contours of the normalized air delay at pedestrian
outdoor ventilation, the age of air distribution inside an urban geo-
height (2 m from the ground) obtained by steady RANS and LES. To
metry is normally assessed as the time that it takes for the external air
normalize the results, the maximum local mean air delay in the LES
to reach a certain location after entering the computational domain
simulations, τdmax = 0.5 s, is used. Fig. 14 presents the distribution of
[6,10,40]. Therefore, the age of air inside an area of interest can be
the difference between the local mean air delay obtained by the two

364
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 12. Contours of (a) normalized mean wind


speed at pedestrian level (2 m from the ground)
and (c) turbulence kinetic energy at canyon top
level (6.25 m from the ground) obtained by
steady RANS. (b, d) Same for LES (Umax = 6 m/s,
Uref = 6 m/s).

approaches. For easier comparison and generalization of the conclu- air delay are provided. Fig. 15 shows the location of the two planes,
sions, the results of mean air delay are normalized with the maximum Plane A and Plane B. Fig. 16 presents the results across plane A. It can
value in the area of study. The distributions presented in Fig. 13a and b be seen that the approaching wind diverges over and around the tower,
are fairly similar, though steady RANS tends to overestimate the air resulting in a substantial amount of fresh air to flow downwards in front
delay at pedestrian height. In this case, the overall average deviation (windward side) of the tower and to enter the Ledras Street. This leads
between steady RANS and LES is about 52%. In addition, the influence to a lower level of air delay in this region that is well predicted by both
of building height variations on the local mean air delay values is well steady RANS and LES. It should be noted that the turbulence levels
predicted by the two approaches. A closer look at this figure reveals across the roof level and inside the street canyon (with label “1” in
that steady RANS and LES predict a relatively low level of air delay in Fig. 16) are well predicted for both steady RANS and LES, though the
parts with high building height variations. For example, parts B6, B7, turbulence kinetic energy is overestimated by the steady RANS at the
C4, C5 and C6. The opposite trend is predicted by both approaches in stagnation point in front of the tower and especially near the leading
parts with relatively low building height variations, where the local air edge; this observation is in accordance with similar findings of previous
delays are relatively higher. This is especially the case in parts E5, E6, studies [51,54,56]. Behind the tower, a relatively large separation re-
E7, F5, F6, F7, G5, G6 and G7 (Fig. 2a). This is in agreement with findings gion is generated. The extent of the separation region is larger for
of previous studies performed for generic urban configurations in which steady RANS than for LES. As shown in Fig. 17, this leads to different
the positive influence of building height variation for urban ventilation flow fields inside canyons 2 and 3 downstream of the tower. In addi-
has been reported [7,13]. tion, less turbulence is swept around the leading edge and over the roof
In order to reveal more clearly differences and better understand the of the tower into the downstream region for steady RANS, leading to
performance of steady RANS and LES, two vertical planes at different lower levels of turbulence on top of street canyons with labels “2” and
locations inside the area of interest are selected, where contours of the “3” with deviation from LES of 65% and 85%, respectively. Figs. 16e
normalized mean wind speed, turbulence kinetic energy and local mean and f clearly indicate a significant difference in the local mean air delay

365
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 13. Contours of ratio of local mean air delay


to maximum mean air delay (τdmax = 0.5 s) at
pedestrian height (2 m) obtained by (a) steady
RANS and (b) LES.

Fig. 15. 3D map of buildings geometry with streamwise vertical bounded planes A and B.

across the roof level of street canyons with average deviation of 80%.
This overestimation can be considered the main reason for the generally
higher air delay values predicted by steady RANS, and it is consistent
with the results presented by previous studies [18,34]. In a more de-
tailed analysis inside the urban canyons with label “4” and 5 (Fig. 19),
the wind flow predicted by both RANS and LES presents skimming flow
characteristics as described by Oke [86]. Inside the urban canon with
label “4”, a region with underestimated mean wind speed by RANS is
Fig. 14. Distribution of ratio of difference between local mean air delay at pedestrian observed with 85% deviation from LES. At the same location, air delay
level (2 m) obtained by RANS and LES along with buildings colored by building height predicted by RANS is reaching up to 3 times higher than LES while in
(τdmax = 0.5 s and Hmax = 45.5 m).
canyon with label 5 this is around 1.5 times higher. Since turbulence is
overestimated by RANS at the top of canyons with label “4” and “5”
distributions obtained by the two approaches especially in the street within the same range of 80%, the under prediction of wind speed by
canyon with label “3”. The average deviation between steady RANS and RANS in canyon with label “4” seem to enhance the overestimation of
LES inside the street canyons with labels “2” and “3” is about 20% and mean air delay at the same location.
80%, respectively. Note that earlier research have shown that the ver- Similar morphological characteristics of the area under study can be
tical mean flows and turbulence fluctuations across the canyon roof found in several examples of historical towns mainly in the
level contribute significantly to air exchange between the street can- Mediterranean region. Hence the conclusions and any possible re-
yons and the external flow above them [7,38]. commendations for improving the breathability of the area under in-
Fig. 18 compares the steady RANS and LES results at a vertical plane vestigation are considered important, and can be useful for urban
in the region with a relatively low building height variation. It can be planners and decision makers of these regions.
seen that the wind flow is generally well predicted by both steady RANS The comparison of Figs. 13 and 2a shows a clear impact of building
and LES, showing relatively low wind velocities inside street canyons. height variability on the breathability at the pedestrian level. Areas that
However RANS, again underestimates the turbulence kinetic energy are poorly ventilated appear to be associated with lower building-

366
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 16. Distribution of normalized (a) mean velocity, (c)


turbulence kinetic energy and (e) local mean air delay in a
vertical plane obtained by steady RANS. (b, d and f) same
for LES (Umax = 6 m/s, Uref = 6 m/s, τdmax = 0.8 s).

height variation. Therefore, an increase in the variability of building to have less well ventilated conditions due to the stronger recirculation
height in a given neighborhood or urban area can improve the zones [6]. According to Fig. 2a, regions with the highest planar density
breathability in complex urban areas. This in line with the results of can be identified in and around square C5 and regions with the lowest
previous studies for generic urban areas [11,13,14,41,58]. planar density in squares B2, B6, D6 and F2. However, based on the
The planar density of an urban area has been proven to significantly results presented in Fig. 13, a clear impact of planar density on air delay
affect the breathability levels [14], where more compact cities appear levels cannot be identified. The reason for this is considered to be the

367
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 17. Distribution of normalized (a) mean ve-


locity and velocity streamlines and (c) turbulence
kinetic energy obtained by steady RANS. (b and
d) Same for LES (Umax = 6 m/s, Uref = 6 m/s).

complexity of the urban area, which creates difficulties in identifying ventilation conditions in the selected area.
these relations, while previous findings were based on generic config- • In order to reduce computational cost, buildings were modeled with
urations. The same conclusions can be made for the relations of aero- simplified geometry. However geometrical details of buildings have
dynamic characteristics z0 and zd where their impact on breathability proven to affect significantly the flow pattern around them [88],
levels cannot be identified clearly. Consequently, the creation of formal and a further analysis using a more detailed model could provide
connections between the morphological parameters and breathability more information on the wind flow conditions of the case study
in a real urban geometry can be the scope and objective of further re- area.
search. • The temperature distribution inside the urban canopy layer can play
an important role in air mass transfer and flow exchange. In addi-
5.3. Limitations and future work tion, compact urban cities with lack of open spaces and green areas
are expected to have more chance in experiencing excessive heat
Some important points of this study can be further discussed: stress periods and urban heat island effect. Hence it would be of high
interest to evaluate thermal comfort and microclimatic conditions in
• Based on the complexity of the specific problem, and the high the selected area.
computational cost, only one wind direction has been considered. • As shown in this study, there is a clear impact of morphological
Hence, conclusions made for analysis of local mean age of air, are characteristics like building height variability, on the ventilation
valid for the specific wind direction. However as it has been stated conditions inside real urban areas. However, more investigation on
in Section 2, the wind direction was chosen after analysis of the different urban configurations must be performed, in order to arrive
prevailing wind conditions of the same area, derived from a field at more generalized conclusions.
measurement campaign [87]. In a next step, other wind directions • In this study, the CFD validation is performed for mean wind speed
can be considered in order to investigate in more depth the and turbulence intensity. Further research is required to (i) obtain

368
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 18. Distribution of normalized (a) mean velocity, (c) turbu-


lence kinetic energy and (e) local mean air delay in a vertical plane
obtained by steady RANS. (b, d and f) Same for LES (Umax = 6 m/s,
Uref = 6 m/s, τdmax = 0.8 s).

high-resolution experimental data of age of air for real complex a satisfactory agreement with the experiments only for mean wind
urban areas and (ii) perform detailed CFD validation studies for air speed. For LES, the average deviations in wind speed and turbulence
delay in complex urban areas. intensity from the experiments are about 6% and 14%, respectively,
while for steady RANS they increase to 8% and 31%, respectively.
6. Conclusions • In order to achieve satisfactory CFD results and fair agreement with
experimental data, it is often needed to adjust the inflow conditions.
The main conclusions of this study are: • Based on the validation, the LES turbulence approach is suggested
for calculating local mean age of air in real complex urban en-
• LES can accurately predict the mean wind speed and turbulence vironments.
intensity in a complex urban area. Steady RANS, however, indicates • To better assess the outdoor ventilation for the urban area, a new

369
N. Antoniou et al. Building and Environment 126 (2017) 355–372

Fig. 19. Distribution of normalized (a) mean ve-


locity and velocity streamlines and (c) turbulence
kinetic energy obtained by steady RANS. (b and
d) same for LES (Umax = 6 m/s, Uref = 6 m/s).

indicator called “air delay” is introduced as the difference between References


the local mean age of air at an urban area and that in an empty
domain with the same computational settings. [1] WHO, Ambient Air Pollution: a Global Assessment of Exposure and Burden of

• RANS overestimates the local mean air delay with an average de- Disease, (2016).
[2] United Nations, World urbanization prospects: the 2014 revision, highlights,
viation from LES of 52% at pedestrian level. The main reason for this Department of Economic and Social Affairs, 2014.
is the overestimation of turbulence by RANS while maximum de- [3] OECD, Environmental Outlook to 2050: the Consequences of Inaction, (2012).
viation is observed in regions where RANS underestimates both [4] M.K.-A. Neophytou, R.E. Britter, Modelling of atmospheric dispersion in complex
urban topographies: a computational fluid dynamics study of the central London
wind speed and turbulence intensity.

area, 5th GRACM Int. Congr. Comput. Mech. Limassol, 2 2005, pp. 967–974.
Complex wind flow structures occurring in wake regions of tall [5] I. Panagiotou, M.K.A. Neophytou, D. Hamlyn, R.E. Britter, City breathability as
buildings and narrow span wise streets inside complex urban areas quantified by the exchange velocity and its spatial variation in real inhomogeneous
urban geometries: an example from central London urban area, Sci. Total Environ.
cannot be captured well with RANS, resulting in overestimation of 442 (2013) 466–477.
local mean air delay at these locations. [6] R. Buccolieri, M. Sandberg, S. Di Sabatino, City breathability and its link to pol-
lutant concentration distribution within urban-like geometries, Atmos. Environ. 44
(2010) 1894–1903.
[7] J. Hang, Y. Li, Ventilation strategy and air change rates in idealized high-rise
Acknowledgements compact urban areas, Build. Environ. 45 (2010) 2754–2767.
[8] C. Liu, D.Y.C. Leung, M.C. Barth, On the prediction of air and pollutant exchange
Hamid Montazeri is currently a postdoctoral fellow of the Research rates in street canyons of different aspect ratios using large-Eddy simulation, Atmos.
Environ. 39 (2005) 1567–1574.
Foundation – Flanders (FWO) and is grateful for its financial support [9] Z. Bu, S. Kato, Y. Ishida, H. Huang, New criteria for assessing local wind environ-
(project FWO 12M5316N). The authors gratefully acknowledge the ment at pedestrian level based on exceedance probability analysis, Build. Environ.
partnership with ANSYS CFD. The TOPEUM project was funded by 44 (2009) 1501–1508.
[10] R. Ramponi, B. Blocken, L.B. de Coo, W.D. Janssen, CFD simulation of outdoor
ERA-NET (Urban-Net Call) through Cyprus Research Promotion ventilation of generic urban configurations with different urban densities and equal
Foundation with contract number ΔΙΕΘΝΗ/URBAN-NET/0308/02. and unequal street widths, Build. Environ. 92 (2015) 152–166.

370
N. Antoniou et al. Building and Environment 126 (2017) 355–372

[11] C. Wen, Y. Juan, A. Yang, Enhancement of city breathability with half open spaces [42] W.C. Cheng, C. Liu, D.Y.C. Leung, Computational formulation for the evaluation of
in ideal urban street canyons, Build. Environ. 112 (2017) 322–336. street canyon ventilation and pollutant removal performance, Atmos. Environ. 42
[12] A. Kubilay, M.K. Neophytou, S. Matsentides, M. Loizou, J. Carmeliet, Urban Climate (2008) 9041–9051.
the pollutant removal capacity of urban street canyons as quanti fi ed by the pol- [43] B. Blocken, P. Moonen, T. Stathopoulos, F. Asce, J. Carmeliet, Numerical study on
lutant exchange velocity, Urban Clim. 21 (2017) 136–153. the existence of the Venturi effect in passages between perpendicular buildings, J.
[13] J. Hang, Y. Li, M. Sandberg, R. Buccolieri, S. Di, The influence of building height Eng. Mech. 134 (2008) 1021–1028.
variability on pollutant dispersion and pedestrian ventilation in idealized high-rise [44] M. Bady, S. Kato, H. Huang, Towards the Application of Indoor Ventilation
urban areas, Build. Environ. 56 (2012) 346–360. Efficiency Indices to Evaluate the Air Quality of Urban Areas 43 (2008), pp.
[14] W. Wang, E. Ng, C. Yuan, S. Raasch, Urban climate large-Eddy simulations of 1991–2004.
ventilation for thermal comfort — A parametric study of generic urban configura- [45] R. Yoshie, A. Mochida, Y. Tominaga, H. Kataoka, Cooperative Project for CFD
tions with perpendicular approaching winds, Urban Clim. 20 (2017) 202–227. Prediction of Pedestrian Wind Environment in the Architectural Institute of Jpn 95
[15] B. Blocken, 50 years of computational wind engineering: past, present and future, J. (2007), pp. 1551–1578.
Wind Eng. Ind. Aerodyn. 129 (2014) 69–102. [46] B. Blocken, J. Carmeliet, T. Stathopoulos, CFD evaluation of wind speed conditions
[16] T. Stathopoulos, Pedestrian level winds and outdoor human comfort, J. Wind Eng. in passages between parallel buildings-effect of wall-function roughness modifica-
Ind. Aerodyn. 94 (2006) 769–780. tions for the atmospheric boundary layer flow, J. Wind Eng. Ind. Aerodyn. 95
[17] H. Montazeri, B. Blocken, W.D. Janssen, T. van Hooff, CFD evaluation of new (2007) 941–962.
second-skin facade concept for wind comfort on building balconies: case study for [47] Z. Xie, I.P. Castro, LES and RANS for Turbulent Flow over Arrays of Wall- Mounted
the Park Tower in Antwerp, Build. Environ. 68 (2013) 179–192. Obstacles, (2006), pp. 1–30.
[18] Y. Tominaga, T. Stathopoulos, CFD modeling of pollution dispersion in a street [48] X. Li, C. Liu, D.Y.C. Leung, Development of a k-ε Model for the Determination of Air
canyon: comparison between LES and RANS, J. Wind Eng. Ind. Aerodyn. 99 (2011) Exchange Rates for Street Canyons 39 (2005), pp. 7285–7296.
340–348. [49] M. Skote, M. Sandberg, U. Westerberg, L. Claesson, A.V. Johansson, Numerical and
[19] P. Gousseau, B. Blocken, T. Stathopoulos, G.J.F. van Heijst, CFD simulation of near- experimental studies of wind environment in an urban morphology, Atmos.
field pollutant dispersion on a high-resolution grid: a case study by LES and RANS Environ. 39 (2005) 6147–6158.
for a building group in downtown Montreal, Atmos. Environ. 45 (2011) 428–438. [50] R.E. Britter, S.R. Hanna, Flow and dispersion in urban areas, Annu. Rev. Fluid
[20] M.K.-A. Neophytou, P. Fokaides, I. Panagiotou, I. Ioannou, M. Petrou, M. Sandberg, Mech. 35 (2003) 469–496.
H. Wigo, E. Linden, E. Batchvarova, P. Videnov, B. Dimitroff, A. Ivanov, Towards [51] W. Rodi, Comparison of LES and RANS calculations of the flow around bluff bodies,
optimization of urban planning and architectural parameters for energy use mini- J. Wind Eng. Ind. Aerodyn. 69–71 (1997) 55–75.
mization in mediterranean cities, World Renew. Energy Congr, 2011 Sweden. [52] K. Hanjalic, Will RANS survive LES? A view of perspectives, J. Fluids Eng. 127
[21] Y. Toparlar, B. Blocken, P. Vos, G.J.F. Van Heijst, W.D. Janssen, T. van Hooff, (2005) 831–839.
H. Montazeri, H.J.P. Timmermans, CFD simulation and validation of urban mi- [53] M. Neophytou, A. Gowardhan, M. Brown, An inter-comparison of three urban wind
croclimate: a case study for Bergpolder Zuid, Rotterdam, Build. Environ. 83 (2015) models using Oklahoma City Joint Urban 2003 wind field measurements, Jnl. Wind
79–90. Eng. Ind. Aerodyn. 99 (2011) 357–368.
[22] H. Montazeri, Y. Toparlar, B. Blocken, J.L.M. Hensen, Simulating the cooling effects [54] S. Murakami, Comparison of various turbulence models applied to a bluff body, J.
of water spray systems in urban landscapes: a computational fluid dynamics study Wind Eng. Ind. Aerodyn. 46–47 (1993) 21–36.
in Rotterdam, The Netherlands, Landsc. Urban Plan. 159 (2017) 85–100. [55] Y. Tominaga, T. Stathopoulos, Numerical simulation of dispersion around an iso-
[23] B. Blocken, T. Defraeye, D. Derome, J. Carmeliet, High-resolution CFD simulations lated cubic building: model evaluation of RANS and LES, Build. Environ. 45 (2010)
for forced convective heat transfer coefficients at the facade of a low-rise building, 2231–2239.
Build. Environ. 44 (2009) 2396–2412. [56] T. Stathopoulos, Y. Tominaga, Numerical simulation of dispersion around an iso-
[24] H. Montazeri, B. Blocken, D. Derome, J. Carmeliet, J.L.M. Hensen, Journal of Wind lated building, Atmos. Environ. (2007) 20–24.
Engineering CFD analysis of forced convective heat transfer coefficients at wind- [57] C. Liu, C.C.C. Wong, On the pollutant removal, dispersion, and entrainment over
ward building facades: Influ. Build. geometry 146 (2015) 102–116. two-dimensional idealized street canyons, Atmos. Res. 135–136 (2014) 128–142.
[25] H. Montazeri, B. Blocken, New generalized expressions for forced convective heat [58] J. Hang, Y. Li, Age of air and air exchange efficiency in high-rise urban areas and its
transfer coefficients at building facades and roofs, Build. Environ. 119 (2017) link to pollutant dilution, Atmos. Environ. 45 (2011) 5572–5585.
153–168. [59] M. Sandberg, What is ventilation efficiency? Build. Environ. 16 (1981) 123–135.
[26] Y. Juan, A. Yang, C. Wen, Y. Lee, P. Wang, Optimization procedures for enhance- [60] D. Steward, T.R. Oke, Local climate zones for urban temperature studies, Bull. Am.
ment of city breathability using arcade design in a realistic high-rise urban area, Meteorol. Soc. 93 (2012) 1879–1900.
Build. Environ. 121 (2017) 247–261. [61] A.N. Srahler, A.H. Strahler, Elements of Physical Geography, John Wiley & Sons Ltd,
[27] N. Nazarian, J. Kleissl, Realistic solar heating in urban Areas: air exchange and 1984.
street-canyon ventilation, Build. Environ. 95 (2016) 75–93. [62] P. Mouzourides, A. Kyprianou, M.J. Brown, B. Carissimo, R. Choudhary, M.K.-
[28] J. Hang, Q. Wang, X. Chen, M. Sandberg, W. Zhu, R. Buccolieri, S. Di, City A. Neophytou, Searching for the distinctive signature of a city in atmospheric
breathability in medium density urban-like geometries evaluated through the pol- modelling: could the Multi-Resolution Analysis (MRA) provide the DNA of a city?
lutant transport rate and the net escape velocity, Build. Environ. 94 (2015) Urban Clim 10 (2014) 447–475.
166–182. [63] P. Kastner-klein, M.W. Rotach, Mean Flow and Turbulence Characteristics in an
[29] R. Buccolieri, P. Salizzoni, L. Soulhac, V. Garbero, S. Di, Urban Climate the Urban Roughness Sublayer, (2004), pp. 55–84.
breathability of compact cities, Urban Clim. 13 (2015) 73–93. [64] M.K.-A. Neophytou, E. Tryphonos, P. Fokaides, M. Sandberg, E. Batchvarova,
[30] M. Lin, J. Hang, Y. Li, Z. Luo, M. Sandberg, Quantitative ventilation assessments of H.J.S. Fernando, J. Lelieveld, G. Zittis, Toward designing strategies for urban heat
idealized urban canopy layers with various urban layouts and the same building island mitigation based on multi scale flow considerations, 34thInt. Conf. Energy
packing density, Build. Environ. 79 (2014) 152–167. Conserv. Technol. Mitig. Adapt. Built Environ. Role Vent. Strateg. Smart Mater,
[31] J. Hang, Z. Luo, M. Sandberg, J. Gong, Natural ventilation assessment in typical 2013, pp. 1–10. Athens.
open and semi-open urban environments under various wind directions, Build. [65] J. Franke, A. Hellsten, H. Schlünzen, B. Carissimo, Best practice guideline for the
Environ. 70 (2013) 318–333. CFD simulation of flows in the urban environment, COST Action 732, 2007, pp.
[32] J. Hang, Y. Li, R. Buccolieri, M. Sandberg, S. Di Sabatino, On the contribution of 1–52.
mean flow and turbulence to city breathability: the case of long streets with tall [66] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa,
buildings, Sci. Total Environ. 416 (2012) 362–373. T. Shirasawa, AIJ guidelines for practical applications of CFD to pedestrian wind
[33] S.M. Salim, S.C. Cheah, A. Chan, Numerical simulation of dispersion in urban street environment around buildings, J. Wind Eng. Ind. Aerodyn 96 (2008) 1749–1761.
canyons with avenue-like tree plantings: comparison between RANS and LES, Build. [67] B. Blocken, Computational Fluid Dynamics for urban physics: importance, scales,
Environ. 46 (2011) 1735–1746. possibilities, limitations and ten tips and tricks towards accurate and reliable si-
[34] P. Moonen, V. Dorer, J. Carmeliet, Evaluation of the ventilation potential of mulations, Build. Environ 91 (2015) 219–245.
courtyards and urban street canyons using RANS and LES, Jnl. Wind Eng. Ind. [68] T. van Hooff, B. Blocken, Coupled urban wind flow and indoor natural ventilation
Aerodyn. 99 (2011) 414–423. modelling on a high-resolution grid: a case study for the Amsterdam ArenA stadium,
[35] T. Kim, K. Kim, B. Sean, A wind tunnel experiment and CFD analysis on airflow Environ. Model. Softw 25 (2010) 51–65.
performance of enclosed-arcade markets in Korea, Build. Environ. 45 (2010) [69] T.-H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k-ε eddy viscosity model
1329–1338. for high reynolds number turbulent flows, Comput. Fluids 24 (1995) 227–238.
[36] J. Hang, Y. Li, Wind conditions in idealized building clusters: macroscopic simu- [70] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows,
lations using a porous turbulence model, Bound. Layer. Meteorol. (2010) 129–159. Comput. Methods Appl. Mech. Eng 3 (1974) 269–289.
[37] J. Hang, M. Sandberg, Y. Li, L. Claesson, Flow Mechanisms and Flow Capacity in [71] T. Cebeci, P. Bradshaw, Momentum Transfer in Boundary Layers, (1977)
Idealized Long-Street City Models 45 (2010), pp. 1042–1044. Washington.
[38] J. Hang, Y. Li, M. Sandberg, L. Claesson, Wind conditions and ventilation in high- [72] B. Blocken, T. Stathopoulos, J. Carmeliet, CFD simulation of the atmospheric
rise long street models, Build. Environ. 45 (2010) 1353–1365. boundary layer: wall function problems, Atmos. Environ 41 (2007) 238–252.
[39] J. Hang, M. Sandberg, Y. Li, Effect of urban morphology on wind condition in [73] J. Smagorinsky, General circulation experiments with the primitive equations, Mon.
idealized city models, Atmos. Environ. 43 (2009) 869–878. Weather Rev. 91 (1963) 99–164.
[40] J. Hang, M. Sandberg, Y. Li, Age of air and air exchange efficiency in idealized city [74] D.K. Lilly, A proposed modification of the Germano subgrid-scale closure method A
models, Build. Environ. 44 (2009) 1714–1723. proposed modification of the Germano subgrid-scale closure method, Phys. Fluids
[41] S.H.L. Yim, J.C.H. Fung, A.K.H. Lau, S.C. Kot, Air ventilation impacts of the wall 633 (1992) 633–635.
effect resulting from the alignment of high-rise buildings, Atmos. Environ. 43 [75] J.B. Perot, An analysis of the fractional step method, J. Comput. Phys. 108 (1993)
(2009) 4982–4994. 51–58.

371
N. Antoniou et al. Building and Environment 126 (2017) 355–372

[76] G.E.M. Karniadakis, S.A. Orszag, High-order splitting methods for the in- building: impact of computational parameters, Build. Environ 53 (2012) 34–48.
compressible Navier-stokes equations, J. Comput. Phys. 97 (1991) 414–443. [84] B. Blocken, T. Stathopoulos, J.P.A.J. Van Beeck, Pedestrian-level wind conditions
[77] S. Armfield, R. Street, The Fractional-Step Method for the Navier – Stokes Equations around buildings: review of wind-tunnel and CFD techniques and their accuracy for
on Staggered Grids The Accuracy of Three Variations 665 (2006) 660–665. wind comfort assessment, Build. Environ 100 (2016) 50–81.
[78] ANSYS FLUENT User's Guide, ANSYS Inc, 2013. [85] M. Princevac, J. Baik, X. Li, H. Pan, S. Park, Lateral channeling within rectangular
[79] F. Menter, Best Practice: Scale-resolving Simulations in ANSYS CFD, ANSYS arrays of cubical obstacles, J. Wind Eng. Ind. Aerodyn. 46–47 (2010) 21–36.
Germany GmbH, 2015. [86] T.R. Oke, Street design and urban canopy layer climate, Energy Build 11 (1988)
[80] F. Mathey, D. Cokljat, K.-P. Bertoglio, E. Sergent, Assessment of the vortex method 103–113.
for large eddy simulation inlet condition, Prog. Comput. Fluid Dyn 6 (2006) 58–67. [87] M.K.-A. Neophytou, H.J.S. Fernando, E. Batchvarova, M. Sandberg, J. Lelieveld,
[81] S. Iousef, H. Montazeri, B. Blocken, P.J.V. Van Wesemael, On the use of non-con- E. Tryphonos, A scaling law for the urban heat island phenomenon: deductions from
formal grids for economic LES of wind flow and convective heat transfer for a wall- field measurements and comparisons with existing results from laboratory experi-
mounted cube 119 (2017) 44–61. ments, Proc. 4th US-European Fluids Eng. Div. Symp, Urban Fluid Mech, Chicago,
[82] S. Di Sabatino, R. Buccolieri, B. Pulvirenti, R. Britter, Simulations of pollutant IL, USA, 2014.
dispersion within idealised urban-type geometries with CFD and integral models, [88] H. Montazeri, B. Blocken, CFD simulation of wind-induced pressure coefficients on
Atmos. Environ 41 (2007) 8316–8329. buildings with and without balconies: validation and sensitivity analysis, Build.
[83] R. Ramponi, B. Blocken, CFD simulation of cross-ventilation for a generic isolated Environ 60 (2013) 137–149.

372

You might also like