You are on page 1of 11

Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

Contents lists available at ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

Air pollutant dispersion around high-rise buildings under different angles of MARK
wind incidence

Y. Yua,1, K.C.S. Kwokb,c, X.P. Liud,1, Y. Zhange,
a
School of Aerospace Engineering, Beijing Institute of Technology, China
b
Institute for Infrastructure Engineering, Western Sydney University, Australia
c
School of Civil and Environmental Engineering, Hong Kong University of Science and Technology,China
d
School of Civil Engineering, Hefei University of Technology, China
e
School of Clinical Medicine, Tsinghua University, China

A R T I C L E I N F O A BS T RAC T

Keywords: Gaseous pollutant originated from a building can disperse with ambient air flow and re-enter the same building
Air pollutant under different environmental conditions. In our previous studies, it was revealed that the near-wall dispersion
High-rise building characteristics in a windward emission are significantly different from that in leeward emission for a high-rise
Wind direction building due to the effect of the building on the wind flow. Since the atmospheric wind changes constantly in
Re-entry
both wind speed and wind direction, such that the wind rarely blows perpendicular to the front or the back of a
CFD
building, it is important to investigate what occurs under different angles of wind incidence. By means of both
physical wind tunnel measurement and simulations using Computational Fluid Dynamics (CFD), we firstly
studied the dispersion characteristics around a square-sectioned building. It was found that the building
influences on wind flow plays a significant role in the dispersion characteristics and an angle of wind incidence
of about 90 degrees is a transition angle for the pollutant dispersion pathway around the building. For angles of
wind incidence smaller than 90 degrees, air pollutant will migrate predominantly downward while for angles of
wind incidence greater than 90 degrees, the pollutant will migrate predominantly upward. Further tests
conducted on a more complicated crucifix-form building model with a re-entry along each building wing showed
a similar trend of pollutant dispersion at different angles of wind incidence, with the transition angle shifted to
approximately 75–80 degrees. The findings of this study show that building shape and the resultant wind-
structure interaction plays a significant role in the pollutant dispersion around a building, thus influencing the
air quality at different part of a building.

1. Introduction analysis and simulation studies supported the possibility of wind effect
enhancing an airborne spread of the SARS virus in a large community
In an urban environment, indoor air quality is strongly dependent outbreak (Brook et al., 2004). The airborne transmission over long
on urban air pollution. Indoor air quality in a building can be distances can also be an important mode of infectious disease
compromised by the re-ingestion of contaminated exhaust air origi- transmission (Nardell et al., 1991). These studies have increased the
nated from the same building (Gao et al., 2009). Thus the character- awareness of the scientific and engineering communities to micro-scale
istics of gaseous pollutants dispersion around a building are especially pollutant dispersion in the built environment.
important for building design and indoor air evaluation. Evidently, air ventilation of urban areas by wind flow is an
Near-field dispersion of air pollutant not only affects general health important process in urban pollutant drainage to maintain clean air.
issues, it may also promote the spreading of communicable diseases In particular, outdoor air quality can be improved by wind flow because
(Gao et al., 2009). It has been shown that several emerging and re- wind dilutes and removes pollutants. Urban wind flow is strongly
emerging respiratory infections including tuberculosis (TB) (Nardell related to urban morphology as a combination of building density,
et al., 1991) measles (Riley et al., 1978) and H1N1 influenza (Abd mutual arrangement of buildings and their individual shape and
Razak et al., 2013) can be spread by airborne route. Epidemiologic dimensions. Hence, wind-structure interaction, which controls air


Correspondence to: School of Clinical Medicine, Tsinghua University, Beijing 100084, China.
E-mail address: yuzhang2014@tsinghua.edu.cn (Y. Zhang).
1
Those authors have equal contributions to this paper.

http://dx.doi.org/10.1016/j.jweia.2017.04.006
Received 3 September 2016; Received in revised form 13 March 2017; Accepted 11 April 2017
0167-6105/ © 2017 Elsevier Ltd. All rights reserved.
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

(a) Crucifix-form building model in boundary layer wind tunnel

(b) Simplified square-sectioned building (c) Crucifix-form building model


model in a numerical wind tunnel in a numerical wind tunnel
(H:B:D=1:0.98:1.04) (scale 1:30)
Fig. 1. Physical and numerical models.

pollutant dispersion around buildings, needs to be properly considered have investigated the wind effect on the pollutant dispersion around a
and the results incorporated in the decision making process for urban typical Hong Kong high-rise residential (HRR) building with a crucifix
development programs. platform and a re-entry along each wing. A series of tests using a scaled
During the past three decades, a considerable amount of research model was studied in an atmospheric boundary layer wind tunnel
has been conducted regarding air pollutant dispersion in urban (Dimoudi and Nikolopoulou, 2003; Tominaga et al., 2008). Numerical
environments by wind action, including field measurements, wind simulations were also performed. The results obtained from both
tunnel tests and numerical simulations (Depaul and Sheih, 1985, 1986; experimental measurement and simulations highlighted the differences
Nakamura and Oke, 1988; Meroney et al., 1996; Baik et al., 2000; Lu in air pollutant dispersion pathway for a windward or leeward emission
et al., 2012). Researchers have found that, from a macroscopic point of source, and the dependence on the location of the emission source in
view, the dispersion of air pollution is generally determined by relation to the stagnation point. In the following discussion, the
meteorological factors such as ambient wind speed, wind direction, releasing point is below the stagnation. In the case of a windward
and atmospheric stability. In addition to these natural factors, urban emission, pollutant migrates predominantly downward due to
design, including the arrangement of building arrays and street downwash effect and spreads horizontally after reaching the ground.
canyons, also plays an important role in air pollutant dispersion. Yet, For a leeward emission, air pollutant migrates predominantly upward
in the neighborhood within the local turbulent boundary layer close to within the re-entry due to the pressure field and flow recirculation
the buildings where wind turbulence, vortex formation, exhaust gas around the building before discharging downstream. Obviously the air
emission and natural convection all play a role, the air pollutant pollutant dispersion process is sensitive to the wind direction; hence it
dispersion process is far more complex. These neighborhood regions is important to investigate how pollutant dispersion pathway responds
can be crucial as unexpected air flow may result in long residual time to a change in the angle of wind incidence from windward (0°) to
for air pollutant close to the residents which further promotes disease leeward (180 degrees). Furthermore, the effect of building shape on air
spreading. However, which factor amongst natural convection, wind pollution dispersion, particularly near-building dispersion remains
turbulence, and wind-structure interaction plays the most significant relatively under-explored.
role in such areas remains a controversial issue (Niu and Tung, 2008; The objectives of this research were to investigate air pollutant
Yip et al., 2007; Zhou and Jiang, 2004; Li et al., 2005). dispersion around high-rise buildings of different shapes under differ-
In previous work (Liu et al., 2010, 2011; Zhang et al., 2015), we ent angles of wind incidence. The emphasis of this research was to

52
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

investigate the mechanisms behind the transition of the pollutant


dispersion pathway from predominantly downward to predominantly
upward. The results will advance our understanding of the effect of
wind-structure interaction on air pollutant dispersion around high-rise
buildings. This advancement of knowledge will influence future high-
rise building design and urban planning to enhance air pollution
dispersion and minimize disease transmission around high-rise build-
ings.

2. Method

2.1. Wind tunnel model study

A 1:30 scale model was constructed to represent a 10 storey high-


(a) rise building 30 m in height in prototype with a crucifix shape typical of
residential buildings in Hong Kong, as shown in Fig. 1(a). Based on the
Hong Kong building code, each building wing must have a “re-entry” to
comply with building regulation that requires all rooms to be fitted with
a window/opening. The experimental study was carried out in a
boundary layer wind tunnel at the CLP Power Wind/Wave Tunnel
Facility (WWTF) at Hong Kong University of Science and Technology
(HKUST). The test section is 41 m in length, 4 m in height and 5 m in
width. Experiments were conducted to study the pollutant dispersion
around this complex building structure. The blockage ratio in the wind
tunnel is controlled at approximately 5% to minimize the requirements
for blockage correction. For wind tunnel modeling of flow and plumes
dispersion, similarity requirements between scale model and prototype
have been summarized by Meroney and others (Meroney, 2004).
Amongst these similarity conditions, Reynolds number (Re) indepen-
dence is one of the important requirements. When the Reynolds
(b)
number is sufficiently large, both flow and scalar dispersion can be
considered effectively independent of Re. During the experiment, the
Reynolds number calculated based on the building dimension and
approach wind velocity measured at building height was approximately
2.2×105, which is sufficiently high to ensure that the measurement
results are effectively independent of Reynolds number.
The tracer gas used was propane in air with a 99,000 ppm propane
concentration. The tracer gas was released through a flow-meter at a
constant flow rate of 58.5 ml/s. The emission exit opening was
enlarged to ensure the exit velocity of the source gas was below
0.4 m/s, which was about one order of magnitude lower than the wind
speed. Taking into consideration the density difference between the
source gas and ambient air, the local buoyancy Froude number is
approximately 2.69. The tracer gas was emitted at the same position
within the re-entry at 3rd floor, independent of the angle of wind
(c)
incidence.
A Cobra probe, with an accuracy of ± 0.3 m/s and ± 1° pitch and
k (m2/s2)
0.0 0.2 0.4 0.6 0.8 1.0
yaw up to about 30% turbulence intensity, was used to determine the
1.4
Wind profile in Exp
wind profile. The detection of tracer gas concentrations was achieved
1.2 Wind speed in CFD with a fast-response flame ionization detector (FID) (Cambustion
Turbulence kinetic energy in CFD
Model HFR400), which was carefully calibrated using synthetic air
1.0 Turbulence dissipation rate in CFD
and certified calibration gases of different concentrations (from
51.2 ppm to 99,000 ppm). The calibration was repeated to ensure that
Z (m)

0.8

0.6
the calibration error is generally within 10% when it was used to
measure the standard gas with the lowest concentration, which is
0.4 51.2 ppm. It was found that there were some daily variations in the FID
0.2 calibration. Hence the calibration was performed before and after each
series of experiment to check the stability of the equipment. A pre-
0.0
0 1 2 3 4 tested analysis was performed to estimate the minimum sampling time
U (m/s), (m2/s2) from each specific test. The minimum averaging time to reach a
(d) converged (i.e. representative) standard deviation was between 60 s
Fig. 2. CFD settings. and 90 s at selected key measurement points. Therefore, at each
measurement position, sample air was collected over a period of
120 s at a data-acquisition rate of 150 Hz to obtain stable estimates
of mean concentration. The signal was then passed through a low-pass
filter which effectively eliminated the alias errors (Liu et al., 2010,

53
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

103 100

Normalized Concentration Kc
Normalized Concentration Kc
102 80

101 60

100 40
180 degree Experiment
10-1 180 degree k-e model 20 180 degree-coarsesca mesh
180 degree k-w sst model
180 degree LES 180 degree-medium mesh
180 degree-fine mesh
10-2
0 2 * 4 6 8 10
0
0 2 * 4 6 8 10
Building height: Floor Building height: Floor
(a) Simulations using different CFD models (b) Comparision of results from the
three grids

Fig. 3. CFD models and grid independence.

2011). Measurement of tracer gas concentration at each floor within independence, the baseline computational mesh is coarsened and
the re-entry area was carried out after the wind flow has stabilized. refined with twice about a factors 2. This produces three grids of
Details of the experimental work can be found in the previous reports successive refinement with 1,671,780 cells, 3,411,958 cells and
(Liu et al., 2010, 2011). 5,986,120 cells. The grid resolution resulted from a grid-sensitivity
analysis that will be outlined later.
2.2. Computational Fluid Dynamic (CFD) study
2.2.2. Boundary conditions
Two building models were simulated using CFD. One model As shown in Fig. 2(d), the measured approaching velocity in the
replicated the outer profile of the crucifix-form building to generate a wind tunnel test can be described by a logarithmic law (Eq. (1)), where
square-sectioned building model, which is shown in Fig. 1(b). This u*/ABL is determined based on the reference velocity at the building
method of simplification is consistent with the simulation of building height (3.27 m/s at 1 m) (Liu et al., 2010; Peren et al., 2015a) κ is the
groups in urban area (Dimoudi and Nikolopoulou, 2003; Tominaga von Karman constant (0.42) and z the height coordinate. Based on the
et al., 2008), which provides a considerable saving in computational curve-fitting, the aerodynamic roughness length z 0=0. 66mm (equiva-
time. The second model is a replica of the crucifix-form model which lent to 0.02 m in prototype, at a scale of 1:30). The turbulent kinetic
maintains the size and configuration identical to the physical model energy k was calculated from the mean wind speed and the measure
tested in the wind tunnel, as shown in Fig. 1(c). ANSYS Fluent CFD turbulence intensity using Eq. (2), where Iu is the stream wise
Software Package (version 14.0) was used for this study. turbulence intensity and α is a parameter assumed to be 1.0 here.
The turbulence dissipation rate ε was given by Eq. (3) and the specific
2.2.1. Computational domain and grid dissipation rate ω by Eq. (4), where Cμ is an empirical constant taken to
The distances from the buildings to the side, to the inlet and to the be equal to 0.09.
top of the domain are at least 5 H, with H is the height of the building.
u*/ ABL z + z0
The distance from the building to the outlet is 15 H. Therefore, the U (z ) = ln ( )
κ z (1)
requirements by Tominaga are satisfied (Tominaga et al., 2008).
The computational grid was non-uniform and fully structured. A k (z ) = α (Iu (z ) U (z ))2 (2)
maximum stretching ratio of 1.2 was applied to reduce the grid density
in areas far away from the region of interest. The grid system is u*3/ ABL
ε (z ) =
presented in Fig. 2 for both building models. In order to study the grid κ (z + z 0 ) (3)

54
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

2.2.3. Solver settings


The simulations were performed with the commercial CFD code
Fluent 14.0. The 3D steady RANS equation were solved using Standard
k-ε model. The SIMPLE algorithm was used for pressure-velocity
coupling, pressure interpolation was second order and QUICK dis-
cretization schemes were used for convection terms and second-order
center discretization schemes for viscous terms of all the governing
equations, including momentum, k and ε, species transportation
equations. Convergence was assumed to be obtained when all the
scaled residuals leveled off and reached a minimum of 10−6 for x, y and
z momentum, k, ε, and species.

2.2.4. Computational grid resolution and turbulence model choice


In order to select an appropriate turbulence model, the simulation
results of pollutant distribution using three widely accepted turbulence
(a) 3-D view of the model rotation
models: unsteady Large Eddy Simulation (LES), k−ω− sst, and k−ε
model, were compared. The comparison with experimental results
(Fig. 3a) showed that the differences between these models in the
prediction of pollutant dispersion were not significant, which is
consistent with the results of a recently published study by Yuan
θ
et al. The emission concentration can be more accurately modelled by a
LES model, but the computational cost of the LES model is several
times higher than that of the RANS model (Tominaga and
Stathopoulos, 2011). In this study, we focused on the influence of
(b) Top view of the model rotation
wind direction on the pathway of air pollutant, which are mainly macro
The wind direction is specified, only the building model is rotated to generate an
behaviour determined by the mean flow field, balancing the computa-
angle of wind incidence (θ)
tional cost and accuracy, thus k-ε model was selected to simulate the
Fig. 4. Angle of wind incidence. air pollutant transportation (Patankar, 1980).
Three grid sizes of approximately 1,671,780 cells, 3,411,958 cells
ε (z ) and 5,986,120 cells were used to study gird sensitivity through the
ω (z ) =
Cμ k (z ) (4) simulations of the leeward emission case for high-rise building. The
results for the three grid sizes are shown in Fig. 3(b). For the grid-
For the ground surface, the standard wall functions, proposed by convergence study, we adopted the concept of the grid-convergence
Launder and Spalding (Launder BE, 1974) with roughness modifica- index (GCI) by Roache (Roache, 1994, 1997). The calculated GCI is
tion by Cebeci and Bradshaw (Cebeci, 1977), were used. The values of shown in Fig. 3(c), which confirmed that the medium sized grid is
the roughness parameters, i.e. the sand-grain roughness height k s (m) sufficient for the numerical simulations reported in this paper. In order
and the roughness constant CS , were determined using their consis- to balance the simulation accuracy and computation time, we chose the
tency relationship with the aerodynamic roughness length z 0 derived by medium level of 3,411,958 cells for all the following simulations, in
Blocken et al. Blocken et al. (2007)(Eq. (5)) for FLUENT and CFX. which the smallest cell dimension is 5 mm×5 mm×5 mm in the x-, y-,
and z- directions at closed to the edges of the buildings, and
9. 793z 0 2.5 mm×2.5 mm×2.5 mm at close to the pollutant releasing location.
ks =
CS (5)

The value of k s is restricted in FLUENT; it should be smaller than 2.3. Rotation of building model in the CFD simulation
yP , which is the distance between the center point of the wall-adjacent
cell and the wall. Therefore, for the ground surface surrounding the A wind direction θ=0° is referenced to Y positive direction, as
modelled buildings, k s is equal to 0.93 mm and consequently CS should shown in Fig. 4(a). A change in wind direction was achieved by rotating
be 7 to satisfy Eq. (5) for z 0=0. 66mm . A user-defined function is used the building model about the Z direction in an anticlockwise direction
to set the value of the constant CS outside the standard allowable to create a corresponding angle of wind incidence, as shown in
interval of [0;1] in FLUENT 14.0. The building surfaces were set to Fig. 4(b). The pollutant emission point was fixed and rotated with
have zero roughness height (k s=0m ), as recommended by (Peren et al., the building model.
2015b). In order to avoid the interpolation of CFD results, which is
Zero static pressure is imposed at the outlet plane and slip wall necessary when a dynamic mesh is used to represent the building
conditions, i.e. zero normal velocity and zero normal gradients of all model rotation, a new set of grid was regenerated each time the
variables, at the top and lateral sides of the domain. building model was rotated. To reduce uncertainty introduced by every
A pollution source containing 10% propane with a 58.5 ml/s new grid for different wind directions, we separate the domain into two
volume flow rate was numerically emitted from within the re-entry at zones, one is the central region including building, which is 1 H away
3rd floor, which is located under the stagnation point, as marked by a from the building model side surface and 4 H from building model top
circle in Fig. 1(b) and (c). Mass flow rate boundary was used at the surface. Another region is the rest of the domain. For every new set of
pollution emission window. grid, the central region was rotated and the mesh distribution in this
The air pollutant concentrations at different locations were scaled central region including building remained unchanged, only the
CU H2
based on a non-dimensional concentration KC = QH where C is the meshes outside this central region were changed, but not too much.”
computed mean concentration, UH is the mean wind speed at building Thus, in total there were 26 sets of CFD grid for 2 models; 13 wind
model height H , and Q is the volumetric flow rate of the tracer gas (Liu directions for each model by rotating every 15 degrees from 0 to 180
et al., 2010). degrees.

55
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

Fig. 5. Air pollutant dispersion around the square-sectioned model at different angles of wind incidence.

3. Result and discussion zone in the near wall region, which is dominated by the separation-
reattachment mechanism created by flow separation along the wind-
3.1. Pollutant concentration around square-sectioned model ward corner. Characteristically, the air flow, and hence pollutant
concentration, show a distinctive form of flow toward upstream,
Fig. 5 shows the pollutant concentration around the building at particularly close to the building surface.
different angles of wind incidence. At θ=0° where the emission is at the Evidently, θ=90° signifies a transition angle of wind incidence
windward face, pollutant dispersed predominantly downward due to beyond which the pollutant starts to migrate upwards. At angles of
downwash, as shown in Fig. 5(a). As the angle of wind incidence was wind incidence greater than θ=90°, the emission is at the leeward face
progressively shifted away from 0°, for example, at θ=30° as shown in where the air flow is dominated by the highly turbulent flow within the
Fig. 5(b), the flow regime around the building was controlled by an wake region. Furthermore, the wake flow associated with air flows over
increasing lateral flow component. At θ=90° where the emission is at and around the building is also characterized by a large flow recircula-
the side face, the pollutant became entrained within the recirculation tion with significant vertical flow components that creates an updraft

56
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

(a) θ=0° Air pollutant migrates downwards due to downwash

(b) θ=90° Air pollutant migrates horizontally upstream due to recirculating flow
underneath the separated shear layer

(c) θ=180° Air pollutant migrates upwards due to large recirculation bubble
Fig. 6. Flow pattern and directions of air pollutant dispersion around the square-sectioned model at different angles of wind incidence.

close to the leeward face of the building. This upward air flow along the separated shear layer originating from the windward edge which
leeward face results in the distribution features of building surface reattaches intermittently along the building side face (Saathoff and
pollutant concentration, as shown in Fig. 5(d) and (e) for θ=120° and Melbourne, 1989).
θ=150°, respectively. At θ=180°, the air flow, and hence the pollutant At θ=180° (leeward emission), in the near wake adjacent to the
concentration, display a nearly vertically upward pattern, as shown in leeward face, the air flow is predominantly upward, as shown in
Fig. 5(f). Fig. 6(c). This upward flow in the near wake is associated with the
To further illustrate the occurrence of the transition angle of wind flow and pressure fields in the wake of the building characterized by a
incidence at around θ=90°, Fig. 6 displays the streamlines around the large recirculating bubble stretches far downstream (Tsang et al.,
pollutant emission point under different angles of wind incidence. At 2012).
θ=0° (windward emission), since the emission point is below the Comparing the streamlines and air pollutant dispersion pathways,
stagnation point, the dominant downwash drives the pollutant down- it is clearly evident that the change in pollutant dispersion pathway
ward toward the ground, as shown in Fig. 6(a). corresponds to the air flow pattern driven by wind-structure interaction
At θ=90° (side wall emission), the air flow in the near wall region is at different angles of wind incidence. The near-wall pollutant disper-
predominantly toward upstream in the horizontal direction, as shown sion route changes from “vertical downward”, to “horizontal backward”
in Fig. 6(b). This migration of pollution upstream against the incident and then to “vertical upward” when the incident wind angles changes
wind is driven primarily by the recirculating flow underneath the from 0° to 90° and then to 180°.

57
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

Fig. 7. Air pollutant dispersion around the crucifix-form building with re-entries at different angles of wind incidence.

3.2. Pollutant concentration around crucifix-form building model the re-entry close to ground level and dispersed, as shown in Fig. 7(b).
At θ=90°, the entrapped pollutant migrated upward prior to exiting the
Fig. 7 shows the pollutant concentration around the crucifix-form re-entry, as shown in Fig. 7(c). Pollutant dispersion was dominated by
building model at different angles of wind incidence, which can be wake flow at angles of wind incidence greater than 90°, such as at
compared with that for the square-sectioned building model. Despite θ=120°, 150° and 180° shown in Fig. 7(d), (e) and (f) respectively.
differences associated with the different building shapes, the pollutant Evidently, not only was the pollutant driven upward, it remained
concentration distribution is predominantly governed by the air flow trapped within the re-entry until it exited at the roof and dispersed. It is
pattern produced by wind-structure interaction. At θ=0°, pollutant noteworthy that the re-entry of the crucifix-form building model acts as
migrated predominantly along the re-entry and downward due to a plenum where initial mixing and dilution of the emitted pollutant
downwash, before spilling outward at close to ground level and take place. However, the re-entry also acts as a well-protected conduit
dispersed, as shown in Fig. 7(a). At θ=30°, initially pollutant was which traps emitted pollutant and restricts its interaction with ambient
entrapped and migrated downward within the re-entry before escaping air, thus hinders the dilution process. Hence the entrapped pollutant

58
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

Fig. 8. Air pollutant dispersion around the crucifix-form building with re-entries at a transient angle of wind incidence.

θ= 75°

θ= 80°
Fig. 9. Flow pattern within reentry at the transient angle of wind incidence.

remains high in concentration within the re-entry. Since the re-entry is different angles of wind incidence. As shown in Fig. 10(a), for angles of
connected to existing areas within the building so that building wind incidence within 0° to 60°, the concentration distribution within
occupants may encounter pollutant within the re-entry, this presents the re-entry indicates a similar trend of dispersion where pollutant
a grave health hazard if the pollutant contains a high concentration of predominantly migrates downwards. Experimental measurement, ta-
harmful substances. ken in the case with a windward emission, confirmed the simulation
It is likewise notable that although the flow pattern and the results with a reasonable agreement. When the angle of wind incidence
pollutant concentration distribution of both the square-sectioned and is within 105° to 180°, the pollutant predominantly migrates upwards,
crucifix-form building models underwent a transition from predomi- which is again supported by the experimental measurement of the
nantly downward to predominantly upward, the angle of wind inci- event with a leeward emission (θ=180°), as shown in Fig. 10(b). In the
dence at which the transition took place and the characteristics of the transition range, i.e. angle of wind incidence is within 75° to 90°, the
transition differ between the two building models. In order to further maximum air pollutant concentration occurred at the emission point
reveal the transition process when the angles of wind incidence is (3rd floor), hence air pollutant concentration firstly increases and then
between θ=75° and θ=90°, additional simulations were conducted at decreases as the floor number increases from 1st floor to 10th floor,
θ=5° interval for further analysis. These differences are highlighted in which indicates the suspension of the pollutant around the emission
Fig. 8 which shows the pollutant concentration distribution at θ=75° point. Experiment conducted with a side wall emission (90 degrees)
and 80°. A close examination of pollutant concentration in the vicinity also recorded a high concentration of air pollutant within the re-entry,
of the emission point suggests that the pollutant remained suspended particularly around the emission point at 3rd floor, as presented in
within the re-entry and the concentration value fluctuated slowly. The Fig. 10(c). Overall, despite the variability of tracer gas concentration
flow pattern at θ=75° shown in Fig. 9, particularly the streamlines close measurements taken under experimental conditions, and the assump-
to the emission point, clearly shows a clear downdraft at θ=75° but tions and simplifications made during numerical simulations, the
updraft at θ=80°, which means between 75° and 80° is the transient experimental results generally validated the simulation work and the
angle of wind incidence to change pollutant concentration value flow characteristic and pollutant concentration distribution at different
distribution trend inside re-entry. angles of wind incidence, including the transition angle of wind
Fig. 10 shows the computed and measured normalized tracer gas incidence.
concentration along the building height within the re-entry under The results presented in this paper clearly demonstrated that the

59
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

103 Exp-00 degree 102


00 degree
Normalized Concentration Kc

Normalized Concentration Kc
00 degree 30 degree
15 degree 101 60 degree
102 30 degree
45 degree 100
60 degree
101 10-1

10-2
100
10-3
-1
10
10-4

10-2
0 2 * 4 6 8 10
10-5
0 2 4 6 8 * 10
Building height: Floor Building height: Floor
(a) θ = 0° to 60° (a) θ = 0° to 60°

103 103
90 degree

Normalized Concentration Kc
Normalized Concentration Kc

120 degree
150 degree
102 102 180 degree

101 101

105 degree
100
100 120 degree
135 degree
150 degree 10-1
10-1 165 degree
180 degree
Kc-Exp-180degree 10-2 *
10-2
0 2 * 4 6 8 10
0 2 4 6 8
Building height: Floor
10

Building height: Floor


(b) θ = 90 to 180
(b) θ = 105° to 180°
Fig. 11. Normalized concentration distributions within the re-entry at different angles
103 of wind incidence for pollutant emission (* show pollutant emission point).
Normalized Concentration Kc

and a re-entry along each building wing. Despite the noticeable


102
differences in the building shape, the pollutant dispersion pathway
around the two buildings was found to show a similar tendency.
101 Generally, when the angle of wind incidence positions the pollutant
emission point at a windward location, the pollutant will migrate
downward due to downwash. When the angle of wind incidence
100
Exp-90degree positions the pollutant emission point at a leeward location, the
75 degree-revised pollutant will migrate upward due to the highly turbulent flow and
10-1 80 degree-revised flow recirculation within the wake region in the lee of the building.
85 degree-revised
90 degree-revised However, the local air flow and pollutant dispersion characteristics are
highly dependent on the local building features, for example, within
10-2
0 2
* 4 6 8 10 and in close proximity to the re-entry of the crucifix-form building
Building height: Floor model. The near-wall pollutant dispersion route changes from “vertical
(c) θ = 75° to 90° downward”, “horizontal backward” to “vertical upward” for the square-
sectioned building when the angle of wind incidence changes from 0° to
Fig. 10. Normalized concentration distributions within the re-entry at different angles 180°. However, for the crucifix-form building model, the results show
of wind incidence for 3rd pollutant emission (* show pollutant emission point).
competing downdraft and updraft which give rise to an oscillating flow
and a corresponding fluctuation in pollutant concentration value.
angle of wind incidence and the building shape play an important role
Furthermore, the transition angle at which the pollutant dispersion
in the pollutant dispersion process and thus influence the air quality at
pathway changes from predominantly downward to predominantly
different part of a building.
upward also differs slightly, from around 90° for the square-sectioned
building model to around 75° for the crucifix-form building model.
4. Conclusion It should be noted that the results presented above are dependent
on the actual location of the emission source on the building face. They
CFD simulations were performed to study the effect of angle of wind are indicative of the pollution dispersion pathways for different angles
incidence on air pollutant dispersion around a simplified square- of wind incidence for an emission location at the 3rd floor of a 10 storey
sectioned model and a similar height prototype building model of a building, which is below the stagnation point. To validate the effect of
typical Hong Kong high-rise residential building with a crucifix-form the emission location, we further tested a case in which the emission

60
Y. Yu et al. Journal of Wind Engineering & Industrial Aerodynamics 167 (2017) 51–61

location is at the 9th floor, which is above the stagnation point and very cross-unit contamination around high-rise building due to wind effect: mean
concentration and infection risk assessment. J. Hazard Mater. 192, 160–167.
close to the roof. The results show that for a wide range of angle of wind Lu, C., Deng, Q.H., Liu, W.W., Huang, B.L., Shi, L.Z., 2012. Characteristics of ventilation
incidence, the pollutant migrates upwards within the re-entry, as coefficient and its impact on urban air pollution. J. Cent. South Univ. T 19, 615–622.
shown in Fig. 11. Meroney, R.N., 2004. Wind tunnel and numerical simulation of pollution dispersion: a
hybrid approach. Invited Lecture (in: Croucher Advanced Study Insitute on Wind
Tunnel Modeling). Hong Kong University of Science and Technology.
Acknowledgments Meroney, R.N., Pavageau, M., Rafailidis, S., Schatzmann, M., 1996. Study of line source
characteristics for 2-D physical modelling of pollutant dispersion in street canyons.
J. Wind Eng. Ind. Aerod 62, 37–56.
This research was supported by a special fund of the State Key Joint Nakamura, Y., Oke, T.R., 1988. Wind, temperature and stability conditions in an East
Laboratory of Environment Simulation and Pollution Control West Oriented Urban Canyon. Atmos. Environ. 22, 2691–2700.
(2015)−15K09ESPCT, Tsinghua University, China, and the Research Nardell, E.A., Keegan, J., Cheney, S.A., Etkind, S.C., 1991. Airborne infection -
theoretical limits of protection achievable by building ventilation. Am. Rev. Respir.
Grant Scheme (RGS) 2014, Institute for Infrastructure Engineering,
Dis. 144, 302–306.
Western Sydney University, Australia. Niu, J., Tung, T.C., 2008. On-site quantification of re-entry ratio of ventilation exhausts
in multi-family residential buildings and implications. Indoor Air 18, 12–26.
References Patankar, S.V., 1980. Numerical Heat Transfer and Fluid Flow. Hemisphere Publishing
Corporation, US.
Peren, J.I., van Hooff, T., Leite, B.C.C., Blocken, B., 2015a. Impact of eaves on cross-
Abd Razak, A., Hagishima, A., Ikegaya, N., Tanimoto, J., 2013. Analysis of airflow over ventilation of a generic isolated leeward sawtooth roof building: windward eaves,
building arrays for assessment of urban wind environment. Build. Environ. 59, leeward eaves and eaves inclination. Build. Environ. 92, 578–590.
56–65. Peren, J.I., van Hooff, T., Leite, B.C.C., Blocken, B., 2015b. CFD analysis of cross-
Baik, J.J., Park, R.S., Chun, H.Y., Kim, J.J., 2000. A laboratory model of urban street- ventilation of a generic isolated building with asymmetric opening positions: impact
canyon flows. J. Appl. Meteorol. 39, 1592–1600. of roof angle and opening location. Build. Environ. 85, 263–276.
Blocken, B., Stathopoulos, T., Carmeliet, J., 2007. CFD simulation of the atmospheric Riley, R.L., Riley, E.C., Murphy, G., 1978. Airborne spread of measles in a suburban
boundary layer: wall function problems. Atmos. Environ. 41, 238–252. elementary-school. Am. Rev. Respir. Dis. 117, (255-255).
Brook, R.D., Franklin, B., Cascio, W., Hong, Y., Howard, G., Lipsett, M., Luepker, R., Roache, P.J., 1994. Perspective - a method for uniform reporting of grid refinement.
Mittleman, M., Samet, J., Smith, S.C., Jr., Tager, I., Expert Panel on, P., Prevention Stud., J. Fluid Eng.-T Asme 116, 405–413.
Science of the American Heart, A., 2004. Air pollution and cardiovascular disease: a Roache, P.J., 1997. Quantification of uncertainty in computational fluid dynamics. Annu
statement for healthcare professionals from the Expert Panel on Population and Rev. Fluid Mech. 29, 123–160.
Prevention Science of the American Heart Association. Circulation 109, 2655–2671. Saathoff, P.J., Melbourne, W.H., 1989. The generation of peak pressures in separated
Cebeci, T., B.P., 1977. Momentum Transfer in Boundary Layers. Hemisphere Publishing reattaching flows. J. Wind Eng. Ind. Aerod 32, 121–134.
Corporation, New York. Tominaga, Y., Stathopoulos, T., 2011. CFD modeling of pollution dispersion in a street
Depaul, F.T., Sheih, C.M., 1985. A tracer study Of dispersion Inan urban street canyon. canyon: comparison between LES and RANS. J. Wind Eng. Ind. Aerod 99, 340–348.
Atmos. Environ. 19, 555–559. Tominaga, Y., Mochida, A., Yoshie, R., Kataoka, H., Nozu, T., Yoshikawa, M., Shirasawa,
Depaul, F.T., Sheih, C.M., 1986. Measurements of wind velocities in a street canyon. T., 2008. AIJ guidelines for practical applications of CFD to pedestrian wind
Atmos. Environ. 20, 455–459. environment around buildings. J. Wind Eng. Ind. Aerod 96, 1749–1761.
Dimoudi, A., Nikolopoulou, M., 2003. Vegetation in the urban environment: Tsang, C.W., Kwok, K.C.S., Hitchcock, P.A., 2012. Wind tunnel study of pedestrian level
microclimatic analysis and benefits. Energ. Build. 35, 69–76. wind environment around tall buildings: effects of building dimensions, separation
Gao, N.P., Niu, J.L., Perino, M., Heiselberg, P., 2009. The airborne transmission of and podium. Build. Environ. 49, 167–181.
infection between flats in high-rise residential buildings: particle simulation. Build. Yip, C., Chang, W.L., Yeung, K.H., Yu, I.T., 2007. Possible meteorological influence on
Environ. 44, 402–410. the severe acute respiratory syndrome (SARS) community outbreak at Amoy
Launder BE, S.D., 1974. The numerical computation of turbulent flows. Comput. Meth Gardens, Hong Kong. J. Environ. Health 70, 39–46.
Appl. Mech. Eng. 3, 21. Zhang, Y., Kwok, K.C., Liu, X.P., Niu, J.L., 2015. Characteristics of air pollutant
Li, Y., Duan, S., Yu, I.T.S., Wong, T.W., 2005. Multi-zone modeling of probable SARS dispersion around a high-rise building. Environ. Pollut. 204, 280–288.
virus transmission by airflow between flats in Block E. Amoy Gard., Indoor Air 15, Zhou, Z.X., Jiang, C.Q., 2004. [Effect of environment and occupational hygiene factors of
96–111. hospital infection on SARS outbreak], Zhonghua lao dong wei sheng zhi ye bing za
Liu, X.P., Niu, J.L., Kwok, K.C.S., Wang, J.H., Li, B.Z., 2010. Investigation of indoor air zhi = Zhonghua laodong weisheng zhiyebing zazhi = Chinese journal of industrial
pollutant dispersion and cross-contamination around a typical high-rise residential hygiene and occupational diseases, 22 261-263.
building: wind tunnel tests. Build. Environ. 45, 1769–1778.
Liu, X.P., Niu, J.L., Kwok, K.C.S., Wang, J.H., Li, B.Z., 2011. Local characteristics of

61

You might also like