2021 Zhang

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Theoretical and Applied Mechanics Letters 11 (2021) 100296

Contents lists available at ScienceDirect

Theoretical and Applied Mechanics Letters


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

Effects of a rooftop wind turbine on the dispersion of air pollutant


behind a cube-shaped building
Shuaibin Zhang a,b, Haoze Yang b, Bowen Du b, Mingwei Ge b,∗
a
State Grid Pingdingshan Power Supply Company, Pingdingshan 467001, China
b
State Key Laboratory of Alternate Electrical Power System with Renewable Energy Sources, North China Electric Power University, Beijing 102206, China

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

Article history: The concentration distribution of urban air pollutants is closely related to people’s health. As an impor-
Received 23 September 2021 tant utilization form of urban wind power, rooftop wind turbines have been widely used in cities. The
Accepted 23 September 2021
wake effect of the rooftop wind turbines will change the flow behind buildings and then affect the pollu-
Available online 2 October 2021
tant dispersion. To this end, the pollutant dispersion behind the building is studied via the computational
Keywords: fluid dynamics method. The actuator disk model and idealized cube are adopted to model the wind tur-
Urban wind power bine and the building, respectively. The study shows that the rooftop wind turbine can reduce the pollu-
Actuator disk model tant mass fraction near the ground and the pedestrian level. Due to the wake effect of the rooftop wind
Wind-turbine wake turbine, the turbulent fluctuation behind the building is weakened, and the spanwise pollutant disper-
Pollutant dispersion sion is suppressed. Besides, the rooftop wind turbine weakens the downwash movement of the building,
which enhances the vertical pollutant dispersion.
© 2021 The Authors. Published by Elsevier Ltd on behalf of The Chinese Society of Theoretical and
Applied Mechanics.
This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

In recent years, with the enhanced urbanization worldwide, the [10,11] used the RANS method and the wind tunnel experiments
dispersion of urban air pollutants, which is closely related to peo- to analyze the pollutant dispersion between buildings. They found
ple’s health, has attracted widespread concern [1,2]. The accelera- that the pollutant dispersion is highly sensitive to the height of
tion effect of the building rooftop provides the conditions for the the upstream building and much less sensitive to the width and
development of rooftop wind power [3–6]. Rooftop wind turbines length of the upstream building. Beside buildings, trees also have
can alleviate urban energy pressure and improve the urban envi- great effects on air pollutant dispersion in street canyons. Gromke
ronment. However, the wake effect of the rooftop wind turbine and Ruck [12,13] modeled avenue trees in urban street canyons
can significantly change the flow behind the building, which will through wind tunnel experiments and revealed the unfavorable ef-
affect the dispersion of air pollutants immersed in the building fects of trees on the pollutant dispersion in street canyons. Sun and
wake. Zhang [14] got similar conclusions used numerical simulation. Li
The effects of buildings on air pollutant dispersion have been and Wang [15] found that trees discourage the downward trans-
widely studied. Saathoff et al. [7] studied the influence of upstream port of high momentum, which weakens the pollutant dispersion
buildings of different heights on the pollutant dispersion through in street canyons. Similar to buildings and trees, wind turbines
wind tunnel experiments. They found that the height of buildings can significantly affect the wind conditions at the rooftop level,
and the width of rooftop structures can affect the pollutant dis- and thus also can have influences on the dispersion of pollutant
persion. Lateb et al. [8,9] used the Reynolds average Navier-Stokes sources in this height. However, there is still a lack of studies on
(RANS) method to investigate the pollutants dispersion exhausted this. To this end, the pollutant dispersion at different heights be-
from a building roof stack located in the wake of a tower, and the hind the building is studied via the computational fluid dynamics
results show that this pollutant dispersion mode is completely dif- (CFD) method. The actuator disk model and general cube are used
ferent from that of an isolated building. In addition, Chavez et al. to simulate the wind turbine and the building, respectively. The
results can be used to evaluate the comprehensive value of urban
rooftop wind power.

Corresponding author. The RANS method is used in this study. With the advantages
E-mail address: gemingwei@ncepu.edu.cn (M. Ge). of short computation time and high computation accuracy, the k-ε

https://doi.org/10.1016/j.taml.2021.100296
2095-0349/© 2021 The Authors. Published by Elsevier Ltd on behalf of The Chinese Society of Theoretical and Applied Mechanics. This is an open access article under the
CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/)
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

turbulence model has been widely used in the flow simulation of


buildings and wind turbines [16–19]. The incompressible governing
equations are as follows:
continuity equation:
∂ ūi
= 0, (1)
∂ xi
momentum equation:
 
∂ ui ∂ ui 1 ∂p ∂ ∂ ui Su
+ uj =− + ν − ui u j + , (2)
∂t ∂xj ρ ∂ xi ∂ xi t ∂ x j ρ
k-ε equations:
    Fig. 1. Schematic of the computational domain.
∂k ∂k ∂ μt ∂ k ∂ ui ∂ u j ∂ ui
+ uj = + μt + − ε, (3)
∂t ∂ x j ∂ x i σk ∂ x j ∂xj ∂xj ∂xj Table 1
Parameters of cases.
  Cases Lx × Ly × Lz (m×m×m) h(m) ht (m) D(m) zs (m)
∂ε ∂ε ∂ μt ∂ε
+ uj =
∂t ∂ x j ∂ x i σε ∂ x j SC1 240 × 80 × 60 10 – – 2
    SC2 240 × 80 × 60 10 – – 5
ε ∂ ui ∂ u j ∂ ui SC3 240 × 80 × 60 10 – – 8
+ Cε1 μt + − Cε2 ε , (4) SCT1 240 × 80 × 60 10 15 5 2
k ∂xj ∂xj ∂xj SCT2 240 × 80 × 60 10 15 5 5
SCT3 240 × 80 × 60 10 15 5 8
species transport equation:
   
∂ u j ci ∂  μt ∂ c i
= Di,m + + S, (5) to describe the dispersion range of pollutant, which can be written
∂t ∂xj Sct ∂ x j
as:
where ūi is the time-averaged velocity, i = 1, 2, 3 denote the

y2 MdA
streamwise direction, the spanwise direction, and the vertical di- σy =

− my 2 , (10)
MdA
rection, respectively; ρ is the fluid density; p is the pressure; νt is
the kinematic viscosity; ui u j is the Reynolds stress tensor; Su is the

z2 MdA
momentum source, which is used to model the effect of rooftop σz =

− mz 2 . (11)
wind turbine; k is the turbulent kinetic energy; μt is the dynamic
MdA
viscosity; ε is the turbulent dissipation rate; σk and σε are Prandtl Following the Architectural Institute of Japan (AIJ) guidelines
numbers for k and ε , respectively; Cε1 and Cε2 are constants, here [26], the schematic of the computational domain is shown in Fig. 1.
we use 1.44 and 1.92, respectively; ci is the pollutant mass frac- The domain size is Lx × Ly × Lz = 240 m × 80 m × 60 m. With
tion; Di,m is the molecular diffusion rate of the species pollutant i; the height h = 10 m, the building is placed 80 m away from the
Sct = 0.7 is the turbulent Schmidt number; S is the strength of the inlet of the domain. The wind turbine, which has the hub height
pollutant source. ht = 15 m and the rotor diameter D = 5 m, is placed in the cen-
In the flow simulations of wind turbines and wind farms, the ter of the building’s rooftop. We select CO as the considered air
actuator disk model is often used to simulate the effect of the pollutant. The pollutant sources are at xs = 95 m, ys = 40 m,
wind turbine [20–24]. In this study, the momentum source term zs = 2 m, 5 m, and 8 m, which are denoted as S1, S2, and S3, re-
is added as the actuator force: spectively. The parameters of cases are summarized in Table 1. We
set the cases with the single building as SC, and the cases with
ρCTUin 2
Su = − , (6) the building and the rooftop wind turbine as SCT. The following
2dx
numbers represent different pollutant source heights. The grid is
where CT is the thrust coefficient, CT = 4a(1 − a ), a = 0.25 is the refined around the building and the wind turbine. 17 grid points
induction factor, Uin is the incoming velocity, and dx is the thick- are arranged around the building in each direction. This grid reso-
ness of the rotor disk. To simplify the simulation, the rotor disk lution has been validated sufficiently to obtain grid-insensitive re-
velocity is taken as the reference velocity, then the actuator force sults (see Appendix A). The inlet boundary condition is velocity-
can be written as: inlet, the outlet boundary condition is pressure-outlet, the sides
a and top boundary conditions are symmetric, the ground and build-
Su = −2ρ Ud 2 , (7)
(1 − a )dx ing boundary conditions are wall. In this study, numerical sim-
ulations are carried out with the regular SIMPLE (Semi-Implicit
where Ud = Uin (1 − a ) is the velocity at the position of the rotor Method for Pressure Linked Equations) solver, which has been
disk. adopted by many scholars [27,28]. The finite volume method with
To clearly describe the dispersion of air pollutants, similar to second-order upwind scheme is used for spatial discretization and
Philips et al. [25], the mass center of pollutant is defined as: the residual is set as 10−5 . The velocity and the turbulence inten-

yMdA sity of the inflow are obtained by fitting experimental data [26]. As
my =

, (8) shown in Fig. 2, the calculated inflow conditions agree well with
MdA
the experimental data.

zMdA Firstly, we analyze the flow structure around the building.


mz =

, (9) Figures 3a and 3b show the contours of the mean streamwise ve-
MdA
locity in SC and SCT cases, respectively. As shown in Fig. 3a, due to
where M is the pollutant mass fraction, and dA is the size of a the strong shear effect of the building rooftop, the flow above the
single grid. The spanwise and vertical standard deviation is used rooftop is significantly accelerated, and low-speed zones appear at

2
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

Fig. 4. Profiles of the mean (a) streamwise velocity Ux and (b) vertical velocity Uz
at the vertical slice of the building center, the numbers represent the streamwise
Fig. 2. Inflow profile of the (a) velocity and the (b) turbulent kinetic energy. position of the profile.

the front and rear of the building, which agrees well with Hearst tion under half of the building height decreases, except pollutant
et al. [29]. As illustrated in Fig. 3b, as the result of the wake effect from source S1, which has an absence of pollutant reduction at
of the rooftop wind turbine, the acceleration of the flow above the 1.5h–4.5h behind the building.
rooftop weakened. A low-speed zone appears behind the wind tur- Furthermore, Fig. 8 shows the isolines of M = 0 at the span-
bine, and the wake deflects downward. wise slice of different pollutant source heights. The higher the
As illustrated in Fig. 4a and 4b, the rooftop wind turbine mainly pollutant source, the longer and narrower the M < 0 area in
affects the area below 2h. Due to the wind turbine wake, the shear the streamwise and spanwise direction, respectively. For pollu-
effect is weakened and the streamwise velocity is decreased. In tant from source S3, the M > 0 area even evolves downstream
addition, the downwash movement of the building is weakened, in a horn shape, and the pollutant mass fraction decreases out-
which leads to a lower vertical velocity below the building height. side this shape. As illustrated in Fig. 9, with the wake effect of
Figures 5a and 5b show the contours of the turbulent kinetic the rooftop wind turbine, the velocities behind the building sig-
energy in SC and SCT cases, respectively. Due to the strong shear nificantly decrease, which discourages the pollutant dispersion. As
effect, the maximum turbulent kinetic energy appears above the the flow evolves downstream, the range of velocity deficit expands
rooftop. As the flow evolves downstream, the turbulent kinetic en- spanwise. On both sides of the spanwise direction, the velocity is
ergy decreases due to the weakening of the shear effect and the slightly higher when there is a rooftop wind turbine, which leads
turbulent dissipation. With the rooftop wind turbine, the shear ef- to a lower pollutant mass fraction. This phenomenon is consistent
fect of the rooftop is weakened, and the maximum turbulent ki- with the change of pollutant mass fraction in Fig. 8.
netic energy decreases. However, the wind turbine enhances the Figures 10a–10f show the isosurfaces of the pollutant mass frac-
turbulence, and another maximum turbulent kinetic energy ap- tion M = 0.005 at different released pollutant heights. As shown in
pears behind the wind turbine rotor, which mixes with the flow Fig. 10a, 10c, and 10e, the higher the pollutant source, the longer
around the building. In addition, due to the lower wind shear in- and narrower the pollutant dispersion in the streamwise and span-
duced by the wind turbine, the turbulent kinetic energy in the wise direction, respectively. When the released source height is
building wake is significantly decreased. low, due to the downwash movement of the building, more pol-
Next, we discuss the dispersion of the air pollutant. As shown lutant is entrained at the bottom of the leeward side of the build-
in Figs. 6a and 6b, due to the weakened building downwash move- ing. The streamwise pollutant dispersion is mainly dominated by
ment induced by the rooftop wind turbine, the pollutant from the velocity behind the building. Therefore, the higher the released
higher source has lower mass fractions at the height of ground source height, the higher the velocity (see Fig. 3a), and the fur-
and pedestrian level. Meanwhile, with the increase of the pollu- ther the streamwise pollutant dispersion. In addition, when the re-
tant source height, the maximum decreased percent of pollutant leased source height is high, the entrainment effect of the build-
mass fraction are 3.97%, 4.47%, 6.63% and 1.21%, 3.95%, 6.22% at the ing is weak, which increases the vertical pollutant dispersion. With
height of the ground and the pedestrian level, respectively. While the wake effect of the rooftop wind turbine, the range of pollutant
the pollutant sources released close to the ground are less affected dispersion becomes shorter in streamwise direction and narrower
by the downwash movement, so they are either less affected by in spanwise direction, which is consistent with the influence of
the rooftop wind turbine. rooftop wind turbine on the velocity and the turbulent kinetic en-
Figure 7 shows the isolines of M = 0 at the vertical slice of ergy behind the building. Because the rooftop wind turbine weak-
the building center, where M = MSCT − MSC . Due to the weakened ens the downwash movement of the building and the turbulent
building downwash movement induced by the rooftop wind tur- kinetic energy behind the rooftop wind turbine is significantly in-
bine, more pollutants transport upwards. The pollutant mass frac- creased, which leads to stronger vertical pollutant dispersion.

Fig. 3. Contours of the mean streamwise velocity Ux at the vertical slice of the building center in (a) SC and (b) SCT cases.

3
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

Fig. 5. Contours of the mean turbulent kinetic energy k at the vertical slice of the building center in (a) SC and (b) SCT cases.

Fig. 6. Profiles of the pollutant mass fraction M at the height of (a) ground and (b) pedestrian level at the vertical slice of the building center.

ent released sources tend to be identical. With the wake effect of


the rooftop wind turbine, the vertical mass centers of pollutant be-
come higher, especially for the pollutant far away from the build-
ing.
Figure 12a–12d show the profiles of the width and height of
Fig. 7. Isolines of M = 0 at the vertical slice of the building center, M = MSCT − the pollutant plume calculated by Eqs. (10) and (11), respectively.
MSC , M < 0 below the isoline, and M > 0 above the isoline.
The higher the pollutant source, the narrower the pollutant plume,
which agrees well with Fig. 10a–10f. The rooftop wind turbine re-
duces the width of pollutant plumes, which proves the correla-
tion between pollutant below the building height and the spanwise
pollutant dispersion. Within 1h after the pollutant source, as the
height of the pollutant source increases, the dispersion rate of the
pollutant plume increases from about 1/8. After passing through
the vortex reattachment zone of the building wake, the pollutant
plume disperses faster, and the dispersion rate increases to about
1/3. This characteristic of pollutant dispersion, which is negatively
Fig. 8. Isolines of M = 0 at the spanwise slice of different source heights, M > 0 related to the height of pollutant sources, originates from the mix-
at the left side of S1 and S2, and the inner side of S3.
ing length scale of the turbulent fluctuation. The rooftop wind tur-
bine reduces the turbulent fluctuation and weakens the mixing ef-
fect, which leads to a decrease in the spanwise dispersion rate of
pollutant plume.
As illustrated in Fig. 12c, the pollutant plumes from different
sources have the same height at 1h after the sources. After this
position, due to the vertical increasement of the turbulent kinetic
energy, the pollutant plume from a higher source has a larger dis-
persion rate. As the rooftop wind turbine weaken the downwash
movement of the building, the entrainment effect of the building is
suppressed. Beside, the increased turbulent kinetic energy behind
the rooftop wind turbine also accelerates the pollutant dispersion
in the vertical direction. Finally, due to the equilibrium effect, the
vertical dispersion rates of pollutant plumes from different sources
all reach about 0.27. In addition, since the vertical dispersion is af-
Fig. 9. Profiles of the mean streamwise velocity Ux at z = 8 m.
fected by the building downwash movement, the vertical disper-
sion rate is generally lower than the spanwise dispersion rate.
Figure 13 shows the profiles of the maximum mass fraction of
Figures 11a and 11b show the profiles of the spanwise and ver-
pollutant. The higher the pollutant source, the less the pollutant
tical mass centers of pollutant calculated by Eqs. (8) and (9), re-
is affected by the building wake vortex, which makes the initial
spectively. The mass centers of pollutant from different released
pollutant mass fraction and the concentration decay rate larger. At
sources have no deflection in spanwise. Behind the building, the
about 2h after the pollutant source, the maximum mass fractions
higher the pollutant source, the higher the vertical mass center
of pollutant from different sources are almost the same. Besides,
of the pollutant. However, at about 10h downstream of the pol-
the effect of the rooftop wind turbine on the concentration decay
lutant source, the vertical mass centers of pollutant from differ-

4
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

Fig. 10. Isosurfaces of the pollutant mass fraction M = 0.005 at different released pollutant heights in (a) SC1, (b) SCT1, (c) SC2, (d) SCT2, (e) SC3, and (f) SCT3 cases.

Fig. 11. Profiles of the (a) spanwise and (b) vertical mass centers of pollutant.

5
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

Fig. 12. Profiles of the (a) width and (c) height of the pollutant plume, with (b) and (d) in log-log coordinates.

studied via the RANS method. The main conclusions are as


follows:
• Due to the wake effect of the rooftop wind turbine, the velocity
and turbulent kinetic energy decrease behind the building. The
pollutant mass fraction under half of the building height de-
creases. Furthermore, with the increase of the pollutant source
height, the maximum decreased percent of pollutant mass frac-
tion are 3.97%, 4.47%, 6.63% and 1.21%, 3.95%, 6.22% at the
height of the ground and the pedestrian level, respectively.
• Under the wake effect of the rooftop wind turbine, the turbu-
lent fluctuation behind the building is reduced, and the mixing
effect is weakened, resulting in the weakened spanwise pollu-
tant dispersion. Besides, the rooftop wind turbine weakens the
downwash movement of the building, which enhances the pol-
lutant dispersion in the vertical direction.
• The higher the released pollutant source, the less the pollutant
is affected by the building wake vortex. Therefore, for high pol-
lutant sources, the initial pollutant mass fraction and the con-
centration decay rate of pollutant are larger. In addition, the
effect of the rooftop wind turbine on the concentration decay
rate of pollutant sources at different heights in the streamwise
Fig. 13. Profiles of the maximum mass fraction of pollutant. direction is almost negligible.

Declaration of Competing Interest


rate of pollutant sources at different heights in the streamwise di-
rection is almost negligible. The authors declare that they have no known competing finan-
In summary, the effects of the rooftop wind turbine on the cial interests or personal relationships that could have appeared to
dispersion of air pollutant immersed in the building wake are influence the work reported in this paper.

6
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

Acknowledgments Appendix A. Grid independence study

The research is supported by the National Natural Science Foun- Figure A1a-A1d show the results of the grid independence
dation of China (Nos. 11772128 and 11772266) and the State Key study, and the schematic of the validation case is shown in Fig. 1.
Laboratory for Alternative Electrical Power System with Renewable Four sets of grids, with 9 grid nodes, 17 grid nodes, 26 grid nodes,
Energy Sources (No. LAPS202107). and 33 grid nodes arranged around the building in each direction,
are employed to study the grid resolution. As shown in Fig. A1a-
A1d, the velocity profile at different streamwise locations of the
case with 17 grid nodes around the building is the same as the ve-

Fig. A1. Profiles of the velocity at different streamwise locations of cases with different grid nodes around the building. (a) Inlet, (b) 0.5h in front of the building’s leeward
side, (c) 0.5h behind the building’s leeward side, and (d) 1.5h behind the building’s leeward side.

7
S. Zhang, H. Yang, B. Du et al. Theoretical and Applied Mechanics Letters 11 (2021) 100296

locity profiles with other finer grids, while the 9 grid nodes may [14] D.J. Sun, Y. Zhang, Influence of avenue trees on traffic pollutant dispersion
be not enough. This indicates that it is sufficient to arranged 17 in asymmetric street canyons: numerical modeling with empirical analysis,
Transp. Res. Part D Transp. Environ. 65 (2018) 784–795.
grid nodes around the building in the present study to obtain grid- [15] Q. Li, Z. Wang, Large-eddy simulation of the impact of urban trees on momen-
insensitive results. tum and heat fluxes, Agric. For. Meteorol. 255 (2) (2018) 44–56.
[16] P. Gousseau, B. Blocken, T. Stathopoulos, et al., CFD simulation of near-field
References pollutant dispersion on a high-resolution grid: a case study by LES and RANS
for a building group in downtown montreal, Atmos. Environ. 45 (2) (2010)
428–438.
[1] X. Li, C. Liu, D.Y.C. Leung, et al., Recent progress in CFD modelling of wind
[17] Y. Tominaga, T. Stathopoulos, Numerical simulation of dispersion around an
field and pollutant transport in street canyons, Atmos. Environ. 40 (2006)
isolated cubic building: model evaluation of RANS and LES, Build. Environ. 45
5640–5658.
(2010) 2231–2239.
[2] Y. Tominaga, T. Stathopoulos, CFD simulation of near-field pollutant dispersion
[18] F. Toja, C. Peralta, O. Lopez-Garcia, et al., Roof region dependent wind potential
in the urban environment: a review of current modeling techniques, Atmos.
assessment with different RANS turbulence models, J. Wind Eng. Ind. Aerodyn.
Environ. 79 (2013) 716–730.
142 (2015) 258–271.
[3] M.A. Heath, J. Walshe, S. Watson, Estimating the potential yield of small build-
[19] M.P. van der Laan, N.N. Sørensen, P. Réthoré, et al., An improved k-epsilon
ing-mounted wind turbines, Wind Energy 10 (2007) 271–287.
model applied to a wind turbine wake in atmospheric turbulence, Wind En-
[4] I. Abohela, N. Hamza, S. Dudek, Effect of roof shape, wind direction, building
ergy 18 (5) (2014) 889–907.
height and urban configuration on the energy yield and positioning of roof
[20] M. Ge, S. Zhang, H. Meng, et al., Study on interaction between the wind-tur-
mounted wind turbines, Renew. Energy 50 (2013) 1106–1118.
bine wake and the urban district model by large eddy simulation, Renew. En-
[5] T.F. Ishugah, Y. Li, R.Z. Wang, et al., Advances in wind energy resource exploita-
ergy 157 (23) (2020) 941–950.
tion in urban environment: a review, Renew. Sustain. Energy Rev. 37 (2014)
[21] M. Ge, H. Yang, H. Zhang, et al., A prediction model for vertical turbulence
613–626.
momentum flux above infinite wind farms, Phys. Fluids 33 (5) (2021) 055108.
[6] D. Micallef, T. Sant, C.S. Ferreira, The influence of a cubic building on a rooftop
[22] M. Calaf, C. Meneveau, J. Meyers, Large eddy simulation study of fully devel-
mounted wind turbine, J. Phys. Conf. Ser. 753 (2) (2016) 022044.
oped wind-turbine array boundary layers, Phys. Fluids 22 (1) (2010) 46–56.
[7] P. Saathoff, A. Gupta, T. Stathopoulos, et al., Contamination of fresh air intakes
[23] X. Yang, S. Kang, F. Sotiropoulos, Computational study and modeling of tur-
due to downwash from a rooftop structure, J. Air Waste Manag. Assoc. 59 (3)
bine spacing effects in infinite aligned wind farms, Phys. Fluids 24 (11) (2012)
(2009) 343–353.
403–430.
[8] M. Lateb, C. Masson, T. Stathopoulos, et al., Numerical simulation of pollutant
[24] Z. Li, X. Yang, Evaluation of actuator disk model relative to actuator surface
dispersion around a building complex, Build. Environ. 45 (8) (2010) 1788–1798.
model for predicting utility-scale wind turbine wakes, Energies 13 (2020).
[9] M. Lateb, C. Masson, T. Stathopoulos, et al., Effect of stack height and exhaust
[25] D. Philips, R. Rossi, G. Iaccarino, Large-eddy simulation of passive scalar dis-
velocity on pollutant dispersion in the wake of a building, Atmos. Environ. 45
persion in an urban-like canopy, J. Fluid Mech. 723 (2013) 404–428.
(29) (2011) 5150–5163.
[26] Architectural Institute of Japan Guidebook For CFD Predictions of Urban Wind
[10] M. Chavez, B. Hajra, T. Stathopoulos, et al., Near-field pollutant dispersion in
Environment, Architectural Institute of Japan, 2017, https://www.aij.or.jp/jpn/
the built environment by CFD and wind tunnel simulations, J. Wind Eng. Ind.
publish/cfdguide/index_e.htm.
Aerodyn. 99 (4) (2011) 330–339.
[27] Y. Huang, R. Hou, Z. Liu, et al., Effects of wind direction on the airflow and
[11] M. Chavez, B. Hajra, T. Stathopoulos, et al., Assessment of near-field pollu-
pollutant dispersion inside a long street canyon, Aerosol. Air Qual. Res. 19 (5)
tant dispersion: effect of upstream buildings, J. Wind Eng. Ind. Aerodyn. 104
(2019) 1152–1171.
(104–106) (2012) 509–515.
[28] Z. Li, H. Zhang, C. Wen, et al., Effects of height-asymmetric street canyon con-
[12] C. Gromke, A vegetation modeling concept for building and environmental
figurations on outdoor air temperature and air quality, Build. Environ. 183
aerodynamics wind tunnel tests and its application in pollutant dispersion
(2020) 107195.
studies, Environ. Pollut. 159 (8–9) (2010) 2094–2099.
[29] R.J. Hearst, G. Gomit, B. Ganapathisubramani, Effect of turbulence on the wake
[13] C. Gromke, B. Ruck, Pollutant concentrations in street canyons of different as-
of a wall-mounted cube, J. Fluid Mech. 804 (2016) 513–530.
pect ratio with avenues of trees for various wind directions, Bound. Layer Me-
teorol. 144 (2012) 41–64.

You might also like