Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Transportation Research Part D xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Transportation Research Part D


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

Influence of avenue trees on traffic pollutant dispersion in


asymmetric street canyons: Numerical modeling with empirical
analysis

Daniel(Jian) Suna,b,c, , Ying Zhangc
a
State Key Laboratory of Ocean Engineering, School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University, Shanghai
200240, China
b
China Institute of Urban Governance, Shanghai Jiao Tong University, Shanghai 200240, China
c
Center for ITS and UAV Applications Research, Shanghai Jiao Tong University, Shanghai 200240, China

AR TI CLE I NF O AB S T R A CT

Keywords: The dispersion of traffic-related pollutants in urban street canyons is of importance for the health
Urban traffic and quality of lives. To reveal the inherent principle, researchers have performed a lot of in-
Low-carbon transportation vestigations; many dispersion phenomena have also been assessed during recent years. However,
Traffic pollutant the presence of avenue trees in street canyons and their capacity for pollutant dispersion remains
Street canyons
partly addressed. In this study, we investigated the effects of avenue trees in urban street canyons
Computational Fluid Dynamics (CFD)
simulation
on traffic pollutant dispersion. The dispersion of CO concentration in asymmetric street canyons
was simulated under varied situations. The computational results showed a good agreement with
the experimental data, and the numerical model was validated to be adequate for investigating
the pollutant dispersion in street canyons. Then, the numerical simulations were extended to
explore the impacts of the effects of avenue trees on CO dispersion; the results indicated that
avenue trees generally increase CO concentrations in asymmetric street canyons. When the wind
direction is perpendicular to the street axis, a terraced building raises pollutant concentrations at
the windward wall and reduces concentration at the leeward wall on the pedestrian levels.
Findings of this study are expected to provide significant insight into urban road design and
strategy making for avenue tree planting, particularly under the existing worldwide sustainable
low-carbon urban development.

1. Introduction

Traffic-related pollutants have attracted increased attention during the last decades. The air pollutants, including inhalable
particles, black carbon (BC), carbon monoxide (CO), nitrogen oxide (NOx) and other volatile organic compounds have been proved to
be correlated with various diseases, especially cardio-respiratory diseases (Brunekreef and Holgate, 2002). Therefore, investigations
of the generation and dispersion of traffic pollutants on a pedestrian level are important and essential for the health of urban
residents.
Previous studies were conducted at a variety of scales, e.g., in an order, around buildings (Hefny and Ooka, 2009), in street
canyons, over building arrays (Carpentieri and Robins, 2015) and at a region or city-wide level (Jeanjean et al., 2015; Sun et al.,
2017). Among these research areas, street canyon is one of the hotspots. Widely existent in urban areas, a street canyon is defined as a


Corresponding author at: A829, Mulan Chu Chao Building, No. 800 Dongchuan Road, Shanghai, 200240, China.
E-mail addresses: danielsun@sjtu.edu.cn (D. Sun), haipingxian@sjtu.edu.cn (Y. Zhang).

http://dx.doi.org/10.1016/j.trd.2017.10.014

1361-9209/ © 2017 Elsevier Ltd. All rights reserved.

Please cite this article as: Sun, D., Transportation Research Part D (2017), http://dx.doi.org/10.1016/j.trd.2017.10.014
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

relatively narrow street with tall, continuous buildings on both sides. If the two-side buildings are with significant height differences,
the canyon is asymmetric; otherwise, it is symmetric. Different wind flow patterns have been noticed in the two types of street
canyons (Santiago and Martin, 2005). Generally, an additional vortex rotating in the opposite direction appears in asymmetric street
canyons when the building height ratio (W/H) equals 1 (with W denoting the street-canyon width and H denoting the building
height).
Unfortunately, high pedestrian exposure was detected in both symmetric (Pavageau and Schatzmann, 1999) and asymmetric
street canyons (Santiago and Martin, 2005) by previous field measurements and wind tunnel experiments, attracting some extent of
attention. To understand the flow and transport mechanism within a street canyon, many experimental and numerical experiments
have been conducted with the various models developed. Some of the models are: Gaussian plume models (Boubel et al., 1994),
CALINE 4 models (Benson, 1992), Canyon Plume Box Model (CPBM) (Yamartino and Wiegand, 1986), Operational Street Pollution
Model (OSPM) (Berkowicz et al., 2008), and Computational Fluid Dynamics (CFD) codes (Xie et al., 2009; Meschini et al., 2014).
Various influencing factors were investigated to identify patterns and rules thus guiding city plannings. In general, emission con-
centrations in a street canyon are determined by the canyon geometry, the traffic flow characteristics, the wind velocity and di-
rection, and other meteorology information such as temperature and air pressure. Evidence has shown that avenue trees, although
previously believed to benefit the environment, may influence wind flow patterns and act as obstacles in street canyons thus in-
creasing the exposure of nearby humans to pollutants (Vardoulakis et al., 2003). With additional research conducted, this phe-
nomenon has become one of the most challenging topics within traffic pollutant dispersion studies.
Among the studies which focused on the avenue trees, Gromke and Ruck (2012) provided a basic frame of the influence of avenue
trees on traffic pollutant dispersion in symmetric street canyons when the wind is perpendicular to the street axis. During the study,
the numerical model MISCAM was used to reveal increasing concentrations in the vegetated symmetric street canyons (Balczo et al.,
2009), and the result was consistent with that of Ries and Eichhorn (2001). A strong increase of pollutant concentrations at the
leeward wall and a slight reduction at the windward wall were found in comparison to the treeless condition. From the series of wind
tunnel experiments and numerical investigations conducted during the past ten years, variations of crown diameter, crown per-
meability (or porosity), trunk height, and tree spacing (or stand density) were found to induce modifications, but with no substantial
changes (Gromke and Ruck, 2007; Gromke et al., 2008). Building height ratio W/H is believed to be the most important parameter,
compared with tree stand density and crown porosity (Buccolieri et al., 2009). However, another study showed slight differences
which indicated that crown porosity has a significant influence on pollutant concentrations for high degrees of porosity (Gromke and
Ruck, 2009). In summary, when the wind is perpendicular to the street axis, the negative effect of avenue trees in the symmetric
street canyons is confirmed. However, no agreement has been reached about the influence of varying parameters, especially that of
the crown porosity of avenue trees.
Researchers have conducted sevèral studies to account for the influence of different wind directions in symmetric street canyons.
When the wind flow is inclined by 45 degrees to the street axis, increases at both the leeward and windward walls were found by
wind tunnel experiments and CFD investigations (Buccolieri et al., 2011). Later on, different approaching wind directions were
analyzed with similar discoveries (Gromke and Ruck, 2012). However, serial experiments by Vos et al. (2012) using ENVI-met model
generated different results which implicated that pollutant concentration is highly related to wind direction and tree height, and the
leeward and windward rules may not be applicable under certain circumstances.
To this end, the influence of different wind conditions have not been fully investigated, and studies focused on the influence of
avenue trees on traffic pollutant dispersion in street canyons are still at preliminary stages. For example, previous research showed
that no agreement had been reached on the connections between pollutant concentrations and certain parameters (e.g., street
geometrics, crown porosities, and wind conditions), and there is still a need for further investigations. Hence, it is necessary and
significant to investigate the effects of avenues of trees in urban street canyons, particularly in asymmetric street canyons, on traffic
pollutant dispersion. The remainder of the paper is organized as follows. Section 2 presents the fundamental modeling approach,
dealing with the fluid flow in street canyons. Then, by using a typical asymmetric street canyon in Shanghai as a case study, the
simulation model of the canyon was created by FLUENT software package. The estimated CO concentrations were compared with the
field data collected at pedestrian levels to indicate the validity and reliability. In Section 3, four scenarios were analyzed with the
simulation models to assess the influence of avenue trees, wind velocities, and wind directions. Finally, the conclusion and future
research recommendations are summarized in Section 4.

2. Methodology and modeling

Computational Fluid Dynamics (CFD) is a branch of fluid mechanics that uses numerical analyses and algorithms by means of
computer-based simulation to solve and analyze problems that involve fluid flows, heat transfer and associated phenomena (Versteeg
and Malalasekera, 1995). Along with the rapid development of computing technologies, the use of CFD to predict internal and
external flows has risen dramatically during the past decades.
Fig. 1 shows the flowchart of this numerical modeling. First, with a representative asymmetrical street canyon in Shanghai
selected as a case study, field data about geometric information, meteorological data, and traffic pollutant data are obtained through
field monitoring. A mesh file composed of an unstructured grid volume and proper boundary conditions is then prepared based on
various input conditions. The mathematical model describes the principles of fluid movement. Afterwards, the simulation software
package FLUENT is used to assign an order to the cells, faces, and grid points in the mesh and to maintain contact between adjacent
cells. In the end, data including flow velocity and pollutant concentration is collected and then compared to that acquired by field
measurement to prove the reliability of the model. It should be noted that, in this study, wind velocity describes the natural

2
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Mathematical
Mathematical model
model
INPUT
INPUT
Fundamental
Fundamental assumptions
assumptions
Geometrical
Geometrical information
information Governing
Governing equations
equations
Building
Building Turbulence
Turbulence modeling
modeling
Street
Street
Avenue
Avenue trees
trees
Mesh
Mesh file
file
Meteorological Grid
Grid volume
volume COMPUTER
OMPUTER SIMULATION
SIMULATION
Meteorological data
data
Wind Boundary
Boundary conditions
conditions
Wind direction
direction
Wind
Wind velocity
velocity

OUTPUT
OUTPUT
Traffic
Traffic pollutant
pollutant data
data
Traffic
Traffic volume
volume Flow
Flow velocity
velocity
Emission
Emission factor
factor Pollutant
Pollutant concentration
concentration
Etc.
Etc.

Fig. 1. Flowchart of numerical simulation.

phenomenon which generates fluid movement, while flow velocity describes the consequence of fluid movement.

2.1. Model assumptions

This study was based on several assumptions to balance the computational cost with the simulation accuracy.

Assumption 1: According to hydrodynamic theory, air and carbon monoxide (CO) involved in this study are treated as viscous
incompressible fluids.
Assumption 2:Three dimensional (3D) models obtain higher accuracy at the cost of much higher computation time. However,
conclusions have been drawn from previous adoptions of two dimensional (2D) models that the street canyon can be modeled as
an infinitely long canyon for a cross-wind condition; this allows simulations to use a 2D domain (Kumar et al., 2009). Meanwhile,
the wide oval shape of the parasol crown and the intensive stand density in the case study assure that majority of the cross sections
share similar shape features. As a result, a simplified 2D model was adopted and later verified.
Assumption 3: In the first stages after emission, pollutant dispersion is strongly dependent on the flow and turbulence fields
behind moving vehicles. However, since the procedure is completed in a very short timescale and causes no substantial influences
to pollutant concentration in larger spatial scale (Carpentieri et al., 2011; He et al., 2015), the vehicle turbulence was ignored in
the simulation model.
Assumption 4: Pollutant emission is fluctuant in the wake of traffic volumes. Due to the limits of computational capacities, a
successive and steady status of pollutant emission was postulated to simplify simulation.

2.2. Field measurement

Known for the roadside Chinese parasol trees, Hengshan Road in the downtown of Shanghai was selected as the typical re-
presentative street canyon for the field monitoring (Fig. 2a). It is one of the most famous commercial streets in Shanghai with a high
traffic volume, especially during peak hours.
The monitoring segment is between Wanping Road and Wuxing Road, about 250 m long. Fig. 2b and c present the two-dimen-
sional and three-dimensional street layouts, respectively. This sub-arterial street is oriented northeast-southwest with two motorways
(3.5 m wide for each lane) and one footpath (4m wide) in each direction. A total of six buildings lie along the road. Notably, a major
building on the left side of the street is made up of two parts, a lower main building and two higher detached wings, a common
architecture for the commercial buildings in Shanghai. As shown in Fig. 2d, Chinese parasol trees are located on the footpath for
every 6.8 m, with a crown diameter of about 6 m. The trunk is around 2.5 m high, with a diameter of 1 m.
Field monitoring was conducted during morning peak hours on two consecutive weekdays in June 2015 when the traffic flow had
a relatively steady volume and composition. All the environment-related instruments were professionally calibrated by the Shanghai
Environmental Monitoring Center (SEMC) before and after the experiments. The meteorological data obtained was also double-
checked by the information released by the nearby Xujiahui observatory.

2.2.1. CO concentration
During the experiment, three monitoring points were selected for each side of the street according to the normal pedestrian level
(Dadvand et al., 2015). As shown in Fig. 3a, Point A is 1 m above the ground close to the street, while Point B and Point C are 0.5 m
higher and 3 m far away from Point A, respectively. As presented in Fig. 3b, Langan Model T15n CO Measurers (the left picture) were

3
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Fig. 2. Geometrical information about the selected street canyon, (a) Location of Hengshan Road, (b) layout of the selected segment, (c) Three-dimensional sketch
map, and (d) detailed information about avenue trees.

Fig. 3. Location and equipment of field measurement, (a) Arrangement of the monitoring points and (b) equipment of field measurement.

used to detect and record CO concentrations every 1 s. On the southwest side of the street, an artificial greenbelt of 72.7 thousand
square meters is located, providing a perfect location for monitoring background concentrations. After roadside monitoring, the same
CO measurers were placed on the ground in the center of the greenbelt for one hour, with the triangular arrangement repeated.
The concentration acquired from the nearby greenbelt was then subtracted from the original raw data to eliminate the instru-
mental errors and minimize the influence of background pollutant sources. The treated average concentration was then calculated, as
introduced later in the model calibration.

4
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

2.2.2. Traffic flow and CO emission


To calculate the amount of the CO emissions, video recorders were set up by the road to capture traffic flow related data, while
samples of vehicle speed were collected with a portable radar speed detector. According to the video capture, the average traffic flow
within the street canyon is about 4000 vehicles per hour (vph) for the four lanes during the morning peak hours, with an average
speed of about 25 km/h. The data on the two monitoring dates are almost the same, resulting in less complex simulations.
If the CO emission factor is determined, the volume of CO emission V (l/h) and the average emission velocity vCO (m/s) can be
calculated using Eq. (1). Q (vph) is the average traffic volume, f (g/km) is the CO emission factor for a single vehicle, s (m) and w (m)
are the length and width of the monitoring segment, respectively, and ρ (kg/m3) is the CO density. α and β are the conversion
coefficient as 10−5 and 1/3600, respectively.
Since CO emission factors are largely dependent on speed, vehicle type, vehicle age and many other factors, estimation of the
factors is one major concern in traffic environment studies. Previous research has provided various integrated CO emission factors for
road segments, ranging from 2 g/km (Li et al., 2014) to 43 g/km (Hao et al., 2001). In Hao et al.’s study (2001), dynamometer tests
were conducted on 62 testing vehicles of the same type and similar speed on Hengshan Road. The average traffic volume during the
field monitoring in the study was about 4000 vph with small passenger cars only, while the traffic volume corresponding to 2 g/km
was approximately 3000 vph with 7% of the vehicles being buses. As a result, 43 g/km was chosen as the emission factor for this
study.
V = m /ρ = α (Q·f ·s )/ ρ = 10−5 × (4000 × 43 × 250)/1.1233 = 38.3lvCO = β·V /(s·w ) = 1/3600 × 38.3/(250 × 14) = 3 × 10−6 m/s
(1)

2.2.3. Meteorology information


Kestrel 4500 Pocket Weather Tracker (the right picture in Fig. 3b) was set in the center of the greenbelt on high and used to collect
data on wind and temperature every 10 s. Influenced by the mild monsoon climate, wind comes from the southeast direction in the
summer of Shanghai; the orientation is reversed during the winter. For the selected street, the wind direction agrees with the overall
environment and is almost perpendicular to the street axis. Field measurement showed an average wind velocity of 0.98 m/s and 3 m/
s for two different monitoring days, while the temperature and the pressure were 28 °C and 105 Pa, respectively.

2.3. Numerical modeling

After a variety of onsite measurements, a 2D simulation model of the street canyon was developed with the CFD code FLUENT 6.3.

2.3.1. Governing equations and turbulence modeling


To establish relationships among physical quantities like velocity, pressure, density, and temperature; governing equations, in-
cluding continuity equation, kinematic equation, and energy equation were created according to the conservation laws of physics,
Newton’s Second Law, and energy conservation equation, respectively. The gravitational force is regarded as a body force which does
not work on the fluid element as it moves through the gravity field.
A turbulence model is a computational procedure to close the system of mean flow equations so that a more or less wide variety of
flow problems can be calculated. Focusing on the mechanisms that affect the turbulent kinetic energy k and the turbulence dissipation
rate ε, the standard k-ε model, as in Eqs. (3) and (4), was proved to be effective in modeling pollutant dispersion in street canyons and
therefore selected for this study (So et al., 2005; Huang et al., 2008). In Eq. (1), k and ε are used to define velocity scale θ and length
scale ℓ. μt is the turbulence eddy viscosity, as calculated by Eq. (2), where the modeling constants include Cμ = 0.09, σk = 1.00,
σε = 1.30, C1ε = 1.44, and C2ε = 1.92.
θ = k1/2 ℓ = k 3/2/ ε (2)

μt = Cρθ ℓ = ρCμ k 2/ ε (3)

∂k ∂ ⎛ υt ∂k ⎞ ∂μ ∂υ ⎞ ∂μ
υ = ⎜ ⎟+ υt ⎛⎜ + ⎟ −ε
∂y ∂y ⎝ σk ∂y ⎠ ⎝ ∂y ∂y ⎠ ∂y (4)

∂ε ∂ ⎛ υt ∂ε ⎞ ε ∂μ ∂υ ⎞ ∂μ
υ = ⎜ ⎟ + ⎜⎛C1ε υt ⎛
⎜ + ⎟ −C2ε ε ⎞⎟
∂y ∂y ⎝ σε ∂y ⎠ k⎝ ⎝ ∂y ∂y ⎠ ∂y ⎠ (5)

2.3.2. Model Built-up


To imitate the cross-section where the field measurement was conducted exactly, a mesh file of over 235 thousand triangular cells
was developed through software package GAMBIT, with rather intensive cells arranged near each boundary. As shown in Fig. 4, the
shaded area is the computational domain filled with fluid. Field information obtained from monitoring was used as the input
variables, focusing on cross-wind conditions.

(1) Buildings: The surface of buildings were set as a no-slip wall boundary condition (black line in Fig. 4). Notably, the architectural
design of the main building and the detached wings still exists within the model, forming a terraced wall on the left side.

5
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Fig. 4. Layout of the simulation model.

(2) Trees: Trees were discussed separately as two different parts, tree canopies and tree trunks. Previous research has proven that
densely foliated vegetation can be modeled as solid barriers (Gallagher et al., 2015). As a result, the canopy was modeled as a
solid circle with no-slip wall boundary like buildings (Li et al., 2016). Although simulation accuracy can improve when the
influence of tree trunks is taken into consideration, existing research has proved that trunks cause no substantial changes to
pollutant dispersion (Gromke and Ruck, 2007; Gromke et al., 2008). Incorporating trunks into simulation may introduce addi-
tional operation time which consequently was not included in the simulation.
(3) Pollutant: Similar to previous research (Buccolieri et al., 2009), the pollution source was represented by a 14-meter line segment
(road width), which is about 0.3 m above the ground (terrain clearance of exhaust pipe). With successive and steady status
proposed (Assumption 4), velocity inlet boundary condition (red line in Fig. 4) was adopted, while emission velocity was set
perpendicular to the road surface according to Eq. (1).
(4) Environment: In the computational domain, three boundaries were connected with the external environment. Similar to the
pollutant source, the upwind boundary was chosen as velocity inlet condition, while wind velocity was determined based on the
field data. The downwind boundary was set as the pressure outlet (blue solid line in Fig. 4) to provide a better rate of con-
vergence, while the pressure was set at zero. With the sky distance from the interested area on a pedestrian level (Shyy, 1993), the
wind flow was considered fully developed at the upper boundary, and outflow boundary condition (blue dotted line in Fig. 4) was
therefore adopted. The safety distance in the simulation was later confirmed valid because no backflows appeared during the trial
simulation.

The mesh file was then imported into CFD code FLUENT, with the temperature and pressure obtained as the basic input con-
ditions. For the inlet and outlet boundaries, turbulence parameters were calculated and applied as input conditions according to the
tutorial of FLUENT (Yu, 2008). Taking the upwind boundary as an example, Eqs. (6)–(13) showed the detailed computation with the
wind velocity ν = 1 m/s, where μ represents the viscosity coefficient. The upper boundary and the building roof are similar to two
parallel plates. As a result, hydraulic diameter DH is approximately considered as twice of the separation distance between parallel
plates. Similar calculations were carried out for the rest boundaries and velocities.
Hydraulic diameter DH ≈ 115 m (6)

Turbulence length scale l = 0.07DH = 8 m (7)

Reynolds number Re = ρνDH / μ = 7.873 × 106 (8)

Turbulence intensity I = 0.16Re−1/8 = 2.198 × 10−2 (9)

Turbulent kinetic energy k = 3/2·(μI )2 = 7.249 × 10−4 (10)

Turbulent dissipation rate ε = Cμ3/4·(k 3/2/l ) = 3.984 × 10−7, Cμ = 0.09 (11)

Specific dissipation rate ω = lk1/2c1/4 = 6.107 × 10−3, c = 0.09 (12)

Turbulent viscosity ratio μt /μ = 8127.088 (13)

2.4. Model calibration

After simulation, the model was subsequently converged (continuity and momentum residuals coming below 10–6, velocity
residuals coming below 10–3 and no obvious oscillation in residual curves). The volume fraction of CO estimated by CFD simulation
was then examined with the CO concentration acquired by field monitoring. As presented in Fig. 5, the overall trend of the estimated
CO concentrations among the monitoring points was found to agree with the field measurement results. Additionally, other findings
were obtained:

6
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Fig. 5. Comparison of CO concentration between simulation and field measurement.

(1) In the windward direction, Point C had the highest value, followed in descending order by Points B and A.
(2) In the leeward direction, Point B had the highest value, followed in descending order by Points A and C.

According to Fig. 5, disparities between estimated concentrations and field monitoring data were observed; this may be as a result
of experimental errors in the field measurement. Lateral vortexes, which cannot be discussed by two-dimensional simplifications in
simulation models, may appear in real situations, possibly leading to the inaccuracy of simulation results. Since the overall trend of
simulation results was consistent with the field monitoring data, the model was proved valid.

3. Comparative experiments and discussion

For further investigation of the influence of avenue trees under different wind directions and velocities, the simulation model was
altered to estimate various scenarios. The results are summarized as follows.

3.1. Influence of wind directions

Fig. 6 presents the influence of wind directions with no interference of avenue trees. The black arrows in the figure represent the
wind directions. The coordinate axis represents the spatial scale of the computational domain. The colors on the left illustrate volume
fraction of CO, while those on the right illustrate flow velocity which is defined by the method of the vector; the negative value
represents the opposite direction. Both figures denote high value with a red color; the rest figures within the paper follow the same
principle. For convenience, two sets of subareas are introduced in Fig. 7, along with two critical surfaces in bold dotted lines. In each
set, the street canyon is divided into two parts.
Under the symmetric circumstances, a single vortex is created in a street canyon. However, due to the altitude difference between
the main building and the detached wings, two vortexes are created in asymmetric canyons.

(1) Southeast wind

When the wind comes from the southeast direction, opposite to the terraced wall, two vortexes are developed around areas M and
N, respectively. Rather than approaching the ground, the fluid reaches Surface A in advance, resulting in a vortex in area M. Similar
to an intangible canopy, this vortex cuts off the connection between the fluid below the vortex and the outside, leading to the vortex
in area N.

(2) Northwest wind

When the wind comes from the northwest direction, the same side of the terraced wall, two vortexes are developed around areas J
and K, respectively. Rather than reaching the external wall of detached buildings, fluid in the lower part is blocked by Surface B
creating a vortex in area K. The higher part still reaches the detached buildings and creates a reverse vortex in area J.

3.2. Influence of wind velocities

The flow patterns under different wind velocities, when the wind comes from the northwest direction, were compared without the
interference of avenue trees. The velocity values were chosen according to the normal weather conditions in Shanghai.
As shown in Fig. 8, the fundamental patterns are immune to the wind velocities. However, scenarios with higher wind velocity
appear to have a clearer vortex boundary. With the increases of the wind velocities, the fluid flow and variation tendency become
intensified. Similar results were gotten when the wind comes from the southeast direction, as shown in Fig. 9. Since the wind

7
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

(a)

(b)
Fig. 6. Pollutant concentration and flow velocity, (a) Southeast wind, (b) Northwest wind. (Wind velocity at 1 m/s, without avenue trees).

Surface A
Surface B

Fig. 7. Two sets of sub-area.

velocities within a certain range resulted in no remarkable changes, this parameter was not involved in the following discussion.

3.3. Vegetated scenario V.S. treeless scenario

Based on the discussion above, four scenarios were created according to the combination of different wind directions and ve-
getation conditions as presented in Table 1, with the wind velocity at 1 m/s. Fig. 10 shows the volume fraction of CO and flow
velocities in each scenario.

(1) Southeast wind (Scenarios A & B)

8
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Fig. 8. Pollutant concentration and flow velocity at different wind velocities, (a) Wind velocity at 1 m/s, (b) Wind velocity at 2 m/s, (c) Wind velocity at 3 m/s.
(Northwest wind, without avenue trees).

Fig. 9. Pollutant concentration and flow velocity at different wind velocities, (a) Wind velocity at 1 m/s, (b) Wind velocity at 2 m/s, (c) Wind velocity at 3 m/s.
(Southeast wind, without avenue trees).

Table 1
Classification of four scenarios.

Wind Direction Avenue trees

With Without

Southeast A B
Northwest C D

9
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Scenario A Scenario B

(a)
Scenario C Scenario D

(b)
Fig. 10. Pollutant concentration and flow velocity in four scenarios, (a) Southwest wind and (b) Northeast wind.

When the wind comes from the southeast direction (Fig. 10a), the pollutant concentration in Scenario A is lower than that of
Scenario B in area M. However, the trend is reversed in area N, especially in the upwind direction. When avenue trees are planted, a
reduction of pollutants at both the windward wall and the leeward wall was detected in area M. However, an impress increase at the

10
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Fig. 11. Pollutant concentrations at pedestrian levels.

windward wall and a small reduction at the leeward wall were detected in area N, which is contrary to the previous result in
symmetric canyons (a strong increase of pollutant concentrations at the leeward wall and a slight reduction at the windward wall).
Due to the existence of avenue trees, the fluid flow in area N becomes weaker, and the air exchanges between the two areas are
somewhat reduced, resulting in pollutant accumulation in area N. These results fully verified the important role avenue trees play on
the dispersion of pollutant in street canyons.

(2) Northwest wind (Scenarios C & D)

When the wind comes from the northwest direction (Fig. 10b), the pollutant concentration in area J in scenario C is higher than
that in scenario D, while the trend in area K is complicated. This observation means that, when avenue trees are planted, a strong
increase of pollutants above Surface A and around the avenue trees is detected, along with an impressive reduction at the leeward
wall and a small increase at the windward wall within area J.
The phenomenon can be explained through low part figures in Fig. 10b. The high flow velocities between the trees assist in
creating a tunnel between fluid layers of different heights and intensifying the fluid flow in area K, thus forcing the pollutant
accumulated within area J.
Afterwards, the estimated CO concentrations at pedestrian levels (y = 1 m) for each scenario were compared. As shown in Fig. 11,
within the central canyon, the pollutant concentration increased in the upwind direction while it decreased in the downwind di-
rection. It can thus be concluded that avenue trees increase the CO concentrations at the pedestrian levels. Terraced buildings are not
considered as a wise design from the perspective of pollutant dispersion, as it may result in severe pollutant accumulations at
different locations.

4. Conclusion

This paper aimed to investigate the influence of avenue trees on traffic pollutant dispersion in asymmetric street canyons. FLUENT
software created a simulation model which was calibrated using field data, with different wind directions and velocities adopted, and
scenarios with and without avenue trees compared. The conclusions from this study are summarized as follows.

(1) Canyon geometry and wind direction play an important role in pollutant dispersion within asymmetric street canyons. The
terraced building on the left side made the flow patterns complicated in the street canyons and divided it into two parts with
different variation tendency in pollutant concentration. Also, avenue trees were found to increase the pollutant concentration in
the canyons in general.
(2) When the wind is perpendicular to the street axis, the terraced building increases the pollutant concentration at the windward
wall and reduces concentration at the leeward wall on the pedestrian levels.
(3) Higher wind velocity intensifies the variation tendency of pollutant concentration, rather than changes the fundamental flow
patterns.

Overall, this paper supplemented studies focused on traffic pollutant dispersion in street canyons by incorporating terraced
buildings and asymmetric street geometry. While the results are promising, there is a need for further studies to improve the per-
formance of the model. Further analysis on similar street geometrics may be conducted to explain the phenomena discovered in this
study. Different crown shapes may also be incorporated as additional variables, with 3D simulation model adopted if necessary.
Furthermore, future work may take crown porosity and large-scaled field measurement into consideration.

11
D. Sun, Y. Zhang Transportation Research Part D xxx (xxxx) xxx–xxx

Acknowledgements

The authors would like to express their appreciation to Drs. Zhong-Ren Peng from School of Naval Architecture Ocean and Civil
Engineering, Shanghai Jiao Tong University, and Hong-Di He from Logistics Research Center, Shanghai Maritime University for their
valuable suggestions and assistance in this study. The research was supported in part by the Major Project of National Social Science
Foundation of China [16ZDA048], the Shanghai Municipal Natural Science Foundation [17ZR1445500], China, and the Humanities
and Social Science Research Project, Ministry of Education [15YJCZH148], China. Any opinions, findings and conclusions or re-
commendations expressed in this paper are those of the authors and do not necessarily reflect the views of the sponsors.

Appendix A. Supplementary material

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.trd.2017.10.
014.

References

Balczo, M., Gromke, C., Ruck, B., 2009. Numerical modeling of flow and pollutant dispersion in street canyons with tree planting. Meteorol. Zeitschrift 18 (2), 197–206. http://dx.
doi.org/10.1127/0941-2948/2009/0361.
Benson, P.E., 1992. A review of the development and application of the Caline3 and Caline4 models. Atmos. Environ. Part B Urban Atmos. 26, 379–390. http://dx.doi.org/10.
1016/0957-1272(92)90013-I.
Berkowicz, R., Ketzel, M., Jensen, S.S., Hvidberg, M., Raaschou-Nielsen, O., 2008. Evaluation and application of OSPM for traffic pollution assessment for a large number of street
locations. Environ. Model. Softw. 23 (3), 296–303. http://dx.doi.org/10.1016/j.envsoft.2007.04.007.
Boubel, R.W., Stern, A.C., Fox, Donald L., Bruce, T., Vallero, D., 1994. Fundamentals of Air Pollution. Academic Presshttp://dx.doi.org/10.1016/B978-012373615-4/50001-7.
Brunekreef, B., Holgate, S.T., 2002. Air pollution and health. Lancet 360, 1233–1242. http://dx.doi.org/10.1016/S0140-6736(02)11274-8.
Buccolieri, R., Gromke, C., Di Sabatino, S., Ruck, B., 2009. Aerodynamic effects of trees on pollutant concentration in street canyons. Sci. Total Environ. 407 (19), 5247–5256.
http://dx.doi.org/10.1016/j.scitotenv.2009.06.016.
Buccolieri, R., Salim, S.M., Leo, L.S., Di Sabatino, S., Chan, A., Ielpo, P., de Gennaro, G., Gromke, C., 2011. Analysis of local scale tree-atmosphere interaction on pollutant
concentration in idealized street canyons and application to a real urban junction. Atmos. Environ. 45 (9), 1702–1713. http://dx.doi.org/10.1016/j.atmosenv.2010.12.058.
Carpentieri, M., Robins, A.G., 2015. Influence of urban morphology on air flow over building arrays. J. Wind Eng. Ind. Aerodyn. 145, 61–74. http://dx.doi.org/10.1016/j.jweia.
2015.06.001.
Carpentieri, M., Kumar, P., Robins, A., 2011. An overview of experimental results and dispersion modelling of nanoparticles in the wake of moving vehicles. Environ. Pollut. 159
(3), 685–693. http://dx.doi.org/10.1016/j.envpol.2010.11.041.
Dadvand, P., Rivas, I., Basagaña, X., Alvarez-Pedrerol, M., Su, J., Pascual, M.D.C., Amato, F., Jerret, M., Querol, X., Sunyer, J., Nieuwenhuijsen, M.J., 2015. The association
between greenness and traffic-related air pollution at schools. Sci. Total Environ. 523, 59–63. http://dx.doi.org/10.1016/j.scitotenv.2015.03.103.
Gallagher, J., Baldauf, R., Fuller, C.H., Kumar, P., Gill, L.W., Mcnabola, A., 2015. Passive methods for improving air quality in the built environment: a review of porous and solid
barriers. Atmos. Environ. 120, 61–70. http://dx.doi.org/10.1016/j.atmosenv.2015.08.075.
Gromke, C., Ruck, B., 2007. Influence of trees on the dispersion of pollutants in an urban street canyon-experimental investigation of the flow and concentration field. Atmos.
Environ. 41 (16), 3287–3302. http://dx.doi.org/10.1016/j.atmosenv.2006.12.043.
Gromke, C., Buccolieri, R., Di Sabatino, S., Ruck, B., 2008. Dispersion study in a street canyon with tree planting by means of wind tunnel and numerical investigations –
evaluation of CFD data with experimental data. Atmos. Environ. 42 (37), 8640–8650. http://dx.doi.org/10.1016/j.atmosenv.2008.08.019.
Gromke, C., Ruck, B., 2009. On the impact of trees on dispersion processes of traffic emissions in street canyons. Boundary-Layer Meteorol. 131, 19–34. http://dx.doi.org/10.
1007/s10546-008-9301-2.
Gromke, C., Ruck, B., 2012. Pollutant concentrations in street canyons of different aspect ratio with avenues of trees for various wind directions. Boundary-Layer Meteorol. 144
(1), 41–64. http://dx.doi.org/10.1007/s10546-012-9703-z.
Hao, J.M., Fu, L.X., He, K.B., Wu, Y., 2001. City Vehicle Emission Control Technology. China Environment Science Press, Beijing (in Chinese).
He, H.D., Zhang, C.Y., Wang, W.L., Hao, Y.Y., Ding, Y., 2015. Feedback control scheme for traffic jam and energy consumption based on two-lane traffic flow model. Transp. Res.
Part D. http://dx.doi.org/10.1016/j.trd.2015.11.005.
Hefny, M.M., Ooka, R., 2009. CFD analysis of pollutant dispersion around buildings: effect of cell geometry. Build. Environ. 44 (8), 1699–1706. http://dx.doi.org/10.1016/j.
buildenv.2008.11.010.
Huang, H., Ooka, R., Chen, H., Kato, S., Takahashi, T., Watanabe, T., 2008. CFD analysis on traffic-induced air pollutant dispersion under non-isothermal condition in a complex
urban area in winter. J. Wind Eng. Ind. Aerodyn. 96 (10–11), 1774–1788. http://dx.doi.org/10.1016/j.jweia.2008.02.010.
Jeanjean, A.P.R., Hinchliffe, G., McMullan, W.A., Monks, P.S., Leigh, R.J., 2015. A CFD study on the effectiveness of trees to disperse road traffic emissions at a city scale. Atmos.
Environ. 120, 1–14. http://dx.doi.org/10.1016/j.atmosenv.2015.08.003.
Kumar, P., Garmory, A., Ketzel, M., Berkowicz, R., Britter, R., 2009. Comparative study of measured and modelled number concentrations of nanoparticles in an urban street
canyon. Atmos. Environ. 43 (4), 949–958. http://dx.doi.org/10.1016/j.atmosenv.2008.10.025.
Li, R.K., Zhao, T., Li, Z.P., Ding, W.J., Cui, X.Y., Xu, Q., Song, X.F., 2014. Estimation of average traffic emission factor based on synchronized incremental traffic flow and air
pollutant concentration. Environ. Sci. 35 (4), 1245–1249. http://dx.doi.org/10.13227/j.hjkx.2014.04.005. In Chinese.
Li, X.B., Lu, Q.C., Lu, S.J., He, H.D., Peng, Z.R., Gao, Y., Wang, Z.Y., 2016. The impacts of roadside vegetation barriers on the dispersion of gaseous traffic pollution in urban street
canyons. Urban Forest. Urban Green. 17, 80–91. http://dx.doi.org/10.1016/j.ufug.2016.03.006.
Meschini, D., Busini, V., Ratingen, S.W.V., Rota, R., 2014. Modeling of pollutant dispersion in street canyon by means of CFD. J. Chem. Ind. Eng. 58 (8), 1967–1972.
Pavageau, M., Schatzmann, M., 1999. Wind tunnel measurements of concentration fluctuations in an urban street canyon. Atmos. Environ. 33 (24–25), 3961–3971. http://dx.doi.
org/10.1016/S1352-2310(99)00138-7.
Ries, K., Eichhorn, J., 2001. Simulation of effects of vegetation on the dispersion of pollutants in street canyons. Meteorol. Zeitschrift 10 (4), 229–233. http://dx.doi.org/10.1127/
0941-2948/2001/0010-0229.
Santiago, J.L., Martin, F., 2005. Modelling the air flow in symmetric and asymmetric street canyons. Int. J. Environ. Pollut. 25, 156–160. http://dx.doi.org/10.1504/IJEP.2005.
007662.
Shyy, W., 1993. Computational Modeling for Fluid Flow and Interfacial Transport. Elsevier, Amsterdam.
So, E.S.P., Chan, A.T.Y., Wong, A.Y.T., 2005. Large-eddy simulations of wind flow and pollutant dispersion in a street canyon. Atmos. Environ. 39 (20), 3573–3582. http://dx.doi.
org/10.1016/j.atmosenv.2005.02.044.
Sun, D.J., Zhang, Y., Xue, R., Zhang, Y., 2017. Modeling carbon emissions from urban traffic system using mobile monitoring. Sci. Total Environ. 599–600, 944–951.
Vardoulakis, S., Fisher, B.E.A., Pericleous, K., Gonzalez-Flesca, N., 2003. Modelling air quality in street canyons: a review. Atmos. Environ. 37 (2), 155–182. http://dx.doi.org/10.
1016/S1352-2310(02)00857-9.
Versteeg, H.K., Malalasekera, W., 1995. An Introduction to Computational Fluid Dynamics: The Finite Volume Method Longman Scientific and Technical, New York.
Vos, P.E.J., Maiheu, B., Vankerkom, J., Janssen, S., 2012. Improving local air quality in cities: to tree or not to tree? Environ. Pollut. 183, 113–122. http://dx.doi.org/10.1016/j.
envpol.2012.10.021.
Xie, X.M., Wang, J.S., Huang, Z., 2009. Traffic emission transportation in street canyons. J. Hydrodyn. 21 (1), 108–117. http://dx.doi.org/10.1016/S1001-6058(08)60125-0.
Yamartino, R.J., Wiegand, G., 1986. Development and evaluation of simple models for the flow, turbulence and pollutant concentration fields within an urban street canyon.
Atmos. Environ. 20 (11), 2137–2156. http://dx.doi.org/10.1016/0004-6981(86)90307-0.
Yu, Y., 2008. FLUENT: Introductory and Advanced Tutorial. Beijing Institute of Technology Press, Beijing (in Chinese).

12

You might also like