Fluid - Dynamic - and - Heat - Transfer - Parameters in An Urban Canyon

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Available online at www.sciencedirect.

com

ScienceDirect
Solar Energy 99 (2014) 1–10
www.elsevier.com/locate/solener

Fluid dynamic and heat transfer parameters in an urban canyon


S. Bottillo ⇑, A. De Lieto Vollaro, G. Galli, A. Vallati
Sapienza University of Rome – DIAEE – Via Eudossiana 18, 00184 Rome, Italy

Received 28 June 2013; received in revised form 22 October 2013; accepted 24 October 2013
Available online 22 November 2013

Communicated by: Associate Editor Matheos Santamouris

Abstract

A microclimatic analysis in a typical urban configuration, has been carried out. Using a CFD method, a N-S oriented urban street
canyon, with a given H/W ratio, has been examined. The standard k–e turbulence model has been used to simulate a three-dimensional
flow field and to calculate the thermo-fluid dynamics parameters that characterize the street canyon. The aim of this study is to inves-
tigate the effect of solar radiation on the flow field and thermal parameters within the canyon. A comparison between transient and sta-
tionary simulations has been performed to evaluate the importance of considering the thermal inertia effects in an urban street canyon
study. The dynamic characteristics of the 3D flow in the canyon have been compared with other numerical simulations and experimental
results. Furthermore a thermo-fluid dynamic analysis of natural convection effects on the heat transfer coefficient and turbulent kinetic
energy, has been carried out.
Ó 2013 Elsevier Ltd. All rights reserved.

Keywords: Urban microclimate; Urban canyon; CFD; Solar radiation

1. Introduction a grid of buildings. By numerical tests characterized by a


2D spatial domain and with assigned surfaces tempera-
The landscape of dense urban areas can be described by tures, Lei et al. (2012) studied the impact of ground heating
units of street delimited by two continuous rows of build- on the flow fields in a street canyons, Xie et al. (2007) stud-
ings to form a “canyon”. This geometry is often described ied the effects of facßades and ground heating on the pollu-
by a single parameter, the canyon aspect ratio (H/W), tant dispersion, Saneinejad et al. (2011) investigated on the
which is defined as the ratio of the building height (H) to heat transfer coefficient in a street canyon, simulated as a
the width between buildings (W). As to the incoming solar cavity, using the low-Reynolds number modeling; they
radiation and the heating of canyon surfaces, the orienta- found a strong influence of thermal effect on the flow field.
tion of the canyon relative to the solar path is also critical Allegrini et al. (2012a) analyzed the convective heat trans-
in determining the timing and extent to which surfaces fer at building facßades in several urban configurations,
receive direct sunlight. Several studies have been performed using the adaptive wall function approach developed by
on different street canyons (Takebayashi and Moriyama, Defraeye et al. (2011) and Allegrini et al. (2012b); they con-
2012; Bozonnet et al., 2005; Lei et al., 2012; Xie et al., cluded that the AWF provides more accurate heat transfer
2007). An experimental validation of a 3D numerical sim- analysis in urban CFD studies. Offerle et al. (2007) used
ulation has been performed by Assimakopoulos et al. wind and temperature measurements to examine the ther-
(2006); they performed tests using a numerical model on mal structure within a street canyon. They found that
buoyancy effects were not seen to have as large an impact
on the measured flow field as has been shown in the numer-
⇑ Corresponding author. Tel.: +39 06 44 58 56 64; fax: +39 06 48 80 120. ical experiments. Kovar-Panskus et al. (2002) performed a
E-mail address: simone.bottillo@uniroma1.it (S. Bottillo).

0038-092X/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.solener.2013.10.031
2 S. Bottillo et al. / Solar Energy 99 (2014) 1–10

wind tunnel study of the influence of wall-heating on the and


flow regime in a simulated canyon. They found little evi-   
dence of thermal effects, except in a very thin layer near @ @ @ l @e
ðqeÞ þ ðqeui Þ ¼ lþ t
the heated wall, as Louka et al. (2002) noted from an exper- @t @xi @xj re @xj
imental campaign in Nantes, France. Most of numerical e e2
studies have been performed on infinitely long street can- þ C 1e ðGk þ C 3e Gb Þ  C 2e q ð5Þ
k k
yons and the prevailing wind direction has been assumed
perpendicular to them, so that the spatial domain has been where Gk is the generation of turbulence kinetic energy due
simplified from 3-D to 2-D. In this study, simulations have to the mean velocity gradients; Gb is the generation of turbu-
been performed on a 3-D domain, investigating the impacts lence kinetic energy due to buoyancy; C1e, C2e and C3e con-
of solar heating and ambient wind speed on the flow fields stants and the KT and lt expressions are reported in the
and surfaces temperatures in a street canyon with a fixed standard k–e model of Ansys Fluent 14.0, 2011; rk and re
H/W and L/W ratios. It has been chosen to study a street are the turbulent Prandtl numbers for k and e, respectively.
canyon isolated from the urban environment, as Blocken To evaluate the impact of thermal effects, the incompressible
et al. (2007) and Allegrini et al. (2012a) did, to evaluate ideal gas module has been used for air density.
how the buildings configuration affects a spatially homoge- The simulated urban canyon has the following charac-
neous dynamic and thermal field. Using the commercial teristics: it has an aspect ratio H/W = 1 and L/W = 5,
CFD code Ansys-Fluent, a series of numerical tests were the orientation is N–S, the buildings width and height are
performed to point out the differences between transient 20 m, the street width is 20 m and the street length is
and steady simulations and to investigate the effects of ther- 100 m. Based on the best practice guidelines by Franke
mal field on the surface temperature on the heat transfer et al. (2007) and Tominaga et al. (2008), the dimensions
coefficient and air velocity and turbulence. of the computational domain have been chosen in relation
to the buildings height H (Fig. 1): the distance between the
2. CFD numerical model side walls of the buildings and the north, east and south
planes is 5H = 100 m, instead the west plane is
In this section, the parameters of the computational 15H = 300 m from the westerly building. The distance
model and the boundary conditions are outlined. between the roofs of the buildings and the upper plane is
The simulations have been performed with the commer- 5H = 100 m. The buildings dimensions determine the dif-
cial CFD code Ansys Fluent 14.0, 3D double precision, ferent domain extension behind the built area. When the
pressure based version and the steady RANS equations flow direction is transversal to the canyon length, the
have been solved in combination with the standard k–e obstacle size is maximum and the flow re-development
model. The governing equations can be expressed as requires a distance of 15H from the buildings to the
follows: outflow bounds. Instead when the flow is longitudinal to
Momentum equation: the canyon direction, the obstacle size is minimum and
the distance behind the built area is 5H. Those dimensions
@ui p l @ 2 ui
1 @ @  allow to set the zero static pressure on the outlet plane and
uj ¼ þ  u0i u0j þ fi : ð1Þ
@xj q @xi q @xi @xj @xj zero gradients of all variables at the top and lateral sides of
the domain. The domain dimension over the buildings have
Continuity equation:
been chosen to take into account the blockage ratio,
@ui defined as the ratio of the area blocked by the buildings
¼ 0: ð2Þ
@xi to the total cross-section area. This parameter depends
on the obstacles size and wind direction: when the building
Heat conservation equation: obstacle area is minimum, the blockage ratio assumes the
value of 2%, instead when the wind impacts transversally
 
@T @ @T
ui þ KT ¼ 0; ð3Þ to the canyon direction, it assumes the maximum value
@xi @xi @xi
of 5.5%. To simulate the soil influence, the computational
where ui is the average speed of air flow; u0i u0j is the Rey- domain has been extended 5 m below the ground level.
nolds stress; q is the air density; l is the molecular viscosity; The soil has been simulated setting the following
fi is the thermal-induced buoyant force; T is the potential parameters: density = 1000 kg/m3; specific heat = 1000 J/
temperature; KT is the heat diffusivity. The standard k–e kgK; thermal conductivity = 2 W/mK; temperature
model has been used to solve the turbulence problem. at -5 m = 288 K; emissivity = 0.9; solar radiation absorp-
The turbulence kinetic energy, k, and its rate of dissipation, tivity (direct visible and infrared) = 0.8. The building walls
e, are obtained from the following transport equations: have: density = 1000 kg/m3; specific heat = 1000 J/kg K;
@ @ @
  
lt @k thermal conductivity = 0.15 W/m K; thickness = 0.30 m;
ðqkÞ þ ðqkui Þ ¼ lþ þ Gk þ Gb  qe internal air temperature = 299 K; emissivity = 0.9; solar
@t @xi @xj rk @xj
radiation absorptivity (direct visible and near infra-
ð4Þ red) = 0.8. The resulting domain size is: 17.25  106 m3.
S. Bottillo et al. / Solar Energy 99 (2014) 1–10 3

Fig. 1. Three-dimensional domain view and points and planes of interest.

The temperature of surfaces has been obtained as result of 9:793z0


ks ¼ ð6Þ
the heat transfers, setting up: the solar load module (longi- Cs
tude: 9.18, latitude: 45.47, UTC: +1), the temperature of
As the flow approaches the built area the velocity inlet
undisturbed air (303 K), the temperature of the internal
profile looks fully-developed before reaching the buildings,
air of the buildings (299 K). To ensure an high quality of
as it can be seen in Fig. 3, and it can be represented by Eq.
the computational grid, it is fully structured and the shape
(7), where u* is the friction velocity, k is the Von Karman
of the cells has been chosen hexahedral (Fig. 2): the normal
constant (0.4) and z is the height coordinate.
vector of a cell surface is parallel to the line connecting the
u
 
midpoints of neighboring cells. To simulate flow fields, in z þ z0
uðzÞ ¼ ln ð7Þ
the area of interest, 40 cells per cube root of the building k z0
volume has been used and 20 cells per building separation
The friction velocity value has been obtained by the cor-
(Franke et al., 2007). For the vertical resolution of the can-
relation with the calculated value of the turbulent kinetic
yon 20 cells have been used. Furthermore, the grid has been
energy (k) at the first node above the ground, as shown
arranged so that the evaluation height (1.5–5 m) and the
in Eq. (8) (Franke et al., 2007).
pedestrian wind speed at 1.5–2 m height is located higher
than the 3rd grid from ground surface. According to the u ¼ k 0:5 C 0:25
l ð8Þ
study of Ramponi and Blocken (2012), the velocity profile
has been set giving a uniform velocity magnitude of 2 m/s where Cl = 0.09.
at the velocity inlet boundary, the turbulence intensity at In Fig. 4 is shown the comparison between the velocity
10%. The aerodynamic roughness value z0 has been set in inlet profile before reaching the urban canyon and the pro-
relation to the roughness parameters in the ground surface file calculated with the logarithmic law (Eq. (7)).
boundary conditions (Ramponi and Blocken, 2012): the
sand-grain roughness height ks and the roughness constant 3. Validation of the CFD model by wind tunnel experiment
Cs. Setting ks = 1m and Cs = 0.5, the resulting z0 (Eq. (6)) data
is 0.05 m (Blocken et al., 2007; Ansys Fluent User’s Guide,
2011), that has been considered an appropriate value to The validation of the mathematical model used in our
represent the roughness of the outer region (Blocken study has been carried out through the comparison with
et al., 2007; Norris and Richards, 2010). A zero roughness the wind tunnel experiment performed by Uehara et al.
height has been used for the building surfaces. (2000). In order to describe the roughness elements of the
urban environment, 44 rows of blocks with a size of

Fig. 2. Views of the computational grid.


4 S. Bottillo et al. / Solar Energy 99 (2014) 1–10

Fig. 3. Wind velocity inlet profiles at different distances from the velocity inlet.

4. Results

4.1. Comparison between steady and transient simulation

A comparison between the results obtained by the


steady simulation and the transient one has been carried
out. The steady simulation has been performed on 26 June
at 14:00, instead the transient simulation has been started
on 18 June at 00:00 and it has been stopped 9 days after,
on 26 June at 24:00. The ground temperature has been cho-
sen as parameter of comparison between the transient and
steady simulation, to highlight the thermal inertia effects,
because the ground is the element with higher thermal
capacity. The results of the steady simulation have been
compared with the last day results of the transient one.
In Fig. 6 are shown the ground temperature values of the
three points of interest (Point A and B within the canyon,
Fig. 4. Comparison between the simulated wind velocity inlet profile and placed at 1 m away from building, respectively from the
the profile calculated with the logarithmic law (Eq. (7)). easterly facßade and from the westerly one; Point Ext, exter-
nal to the canyon, as shown in Fig. 1). As it can been
100  100  50 mm3 have been used. Further, to represent noticed, the temperatures of the various points chosen to
street canyons, 14 rows of blocks with a size of study the urban canyon in transient case are very similar
100  100  100 mm3 have been set. The measured data to the stationary one. This does not mean that the temper-
have been carried out in the space between the 5th and ature of the soil inner layers are the same for both simula-
the 6th of the 14 rows. The experiment has been character- tions, but we are interested on the surface temperature and
ized by a flow direction transversal to rows of blocks and, to study its influence on the flow field. The ground temper-
at the measurement location, the flow appeared totally can- ature of the external point is the same for both simulations,
alized. A numerical validation test has been performed on instead, it can be noticed a difference of 3° for the point A
our street canyon model, which has been scaled by the and a smaller one for point B. Those differences are due to
reduction factor 1/200 and where the wind direction has the ground thermal inertia which determines a lower tem-
been considered totally transversal to the canyon direction. perature value as the point passes from shadow to sun
The air temperature Ta was set as 20 °C (293 K) and the exposition and viceversa. The same consideration with an
ground temperature Tf was set as 79 °C (352 K), the inflow opposite sign can be applied to the point B, which enter
wind speed u0 was set as 1.5 m/s in order to reproduce the into shadow shortly before 14:00. It can be concluded that
conditions of the experimental tests. The comparison the stationary case for a thermo-fluid dynamic simulation
between the data obtained with the wind tunnel experi- of a street canyon is representative of the physical phenom-
ments and the numerical test are illustrated in Fig. 5: the ena of the maximum solar load. For this reason, further
vertical profile of normalized horizontal velocity u/u0 and analysis will be performed on the stationary case, which
T T f
T a T f
, evaluated on a central vertical line within the street is much lighter as to the computational time. The analysis
canyon. As in Lei et al. (2012) study, it can be observed of hourly trends of ground temperatures shows that it is
that there is a good agreement between the mathematical higher within the canyon than outside, during the hours
model and the wind tunnel experiment within the canyon. of maximum solar radiation, while, during the rest of the
Instead the numerical tests overestimate the wind velocity day, their values are significantly lower due to the effect
and the air temperature above the canyon. of shadows.
S. Bottillo et al. / Solar Energy 99 (2014) 1–10 5

Fig. 5. Comparison between the simulated data and the observed data by Uehara et al. (2000). (a) u/u0 and (b) (T  Tf)/(Ta  Tf).

Fig. 6. Comparison between transient and steady simulation: hourly trends of ground temperature of the three points of interest and the values calculated
at 14:00 with the steady simulation.

4.2. Impact of thermal effects ity inlet magnitude of 2 m/s and a wind direction of 45°N,
so that the hottest facßade is the windward one. The result-
In order to evaluate the impact of thermal effects on our ing average Richardson number (Ri) for the windward facß-
street canyon model, a comparison between two simula- ade is 2.9 and for the leeward one is 1.0. The three
tions has been performed. The simulations have been per- dimensional effect of the flow is evident in the formation
formed in stationary case on 26 June at 11:00. In the first of a spiral flow, produced by the combination of the down-
simulation the natural convection has been excluded, ward vertical vortex and the longitudinal component of the
instead in the second one it has been considered. The flow wind velocity, as reported in other 3D simulations (Assi-
regime impacting on the built area is described by a veloc- makopoulos et al., 2006; Santamouris et al., 1999); they
6 S. Bottillo et al. / Solar Energy 99 (2014) 1–10

found that the wind field in urban areas is quite complex plane to the south opening, the flow pattern remains basi-
and the simulated wind speed intensities can be totally dif- cally constant. As it can be seen in Fig. 7(a1) and (d), in the
ferent from the measured data. Fig. 7 shows the velocity North plane the formation of a double vortex can be
vectors on three vertical planes of interest for both simula- observed. The vortices are generated by geometrical
tions. The figure shows the XZ velocity vectors for both discontinuities (the roof and the vertical corner of the east-
simulations and the XY velocity vectors for the simulation erly building) and they have two different rotation axis. The
with natural convection activated. On the XY plane, the upper vortex has its axis parallel to the canyon direction
flow pattern does not change significantly for the two sim- (Fig. 7(a1)), and the lower has a vertical one (Fig. 7(d)).
ulations; Fig. 7(d) shows that the flow within the canyon is From Central plane to the South plane, the aerodynamic
transported from the north opening to the south one. The vortex coming from the roof is fully-formed and it occupies
North and South planes are placed at 10 m from the all the space between buildings (Fig. 7(b1) and (c1)). In the
respective openings, instead the Central plane is in the mid- South plane (Fig. 7(c1)) the mass flow rate coming from the
dle of the canyon. The facßade of the easterly building, in roof is zero and the vortex is transported from the easterly
shadow, is on the right side of the figures, instead the facßade to the westerly one by the longitudinal flow. The
facßade of the westerly one, exposed to the sun radiation flow pattern of the North plane when the natural convec-
is on the left side. When the natural convection is deacti- tion is activated (Fig. 7(a2)), is double-vortex in structure
vated, the flow pattern changes significantly from the north as in the first simulation. In the Central and South plane
opening to the Central plane. Instead, from the Central (Fig. 7(b2) and (c2)) the flow pattern is divided in two

Fig. 7. XZ velocity vectors on the North, Central and South planes, respectively (a), (b) and (c). No natural convection simulation (a1), (b1) and (c1);
simulation with natural convection activated (a2), (b2) and (c2). XY velocity vectors at 10 m height with natural convection activated (d).
S. Bottillo et al. / Solar Energy 99 (2014) 1–10 7

counter-rotating vortex: the upper one due to the geomet- 2D results of the heat transfer coefficient (Fig. 9) are con-
rical discontinuity and the second one, in the low corner gruent with the values reported in Saneinejad et al.
near the hot wall, due to the buoyancy effect. The convec- (2011), except for the upper part of the windward facßade,
tive vortex appears less developed than the simulations car- where our simulation shows higher values; this difference
ried out by Lei et al. (2012) and Xie et al. (2007), probably of heat transfer coefficient is due to the shape of our 2D
because of the 3D nature of our simulation. The impact of canyon model; Saneinejad et al. (2011) simulate the street
thermal effects have been studied also through the analysis canyon as a cavity in the ground, instead we simulate it
of velocity and turbulent kinetic energy near the canyon as two buildings over the ground level. When the domain
surfaces on the Central plane, for both simulations. For is 3D and the wind direction is 45°N, the heat transfer coef-
the building facßades a vertical line near the westerly one ficient values without buoyancy effects, shown in
exposed to the sun radiation, from ground level to building Fig. 8(b1), and the values obtained with buoyancy effects,
roof, has been considered. For the ground surface instead, Fig. 8(b2), are higher than the ones found in our 2D
an horizontal line, from the easterly facßade to the westerly (Fig. 9) and 3D transversal simulation, and the ones
one, has been taken. All those lines of interest are placed at reported by Saneinejad et al. (2011) and Allegrini et al.
0.40 m from the respective surfaces. In Fig. 8 are shown: (2012a). Those differences are probably due to the 3D
the vertical trends of Z velocity component and turbulent effects and to the longitudinal speed component within
kinetic energy near the sun exposed facßade, wall tempera- the canyon that, when the ambient wind speed is 2 m/s
ture and heat transfer coefficient on the facßade itself; veloc- and the direction is 45°N, is more than two times higher
ity magnitude and turbulent kinetic energy above the than the value of 0.5 m/s of the 2D simulations, reported
ground, temperature and heat transfer coefficient on the in Saneinejad et al. (2011). Allegrini et al. (2012b) showed
ground are shown in Fig. 8(d1) and (d2). The Z velocity that, when the Ri > 1, the standard wall function overesti-
component along a vertical line at 0.40 m from the sun- mates the value of heat transfer coefficient; our results
exposed wall (Fig. 8(a1) and (a2)) without natural convec- show that even when the natural convection is not consid-
tion is negative (downward) and it is affected only by the ered, the hc values are much higher than the 2D simula-
aerodynamic vortex, instead, when the natural convection tions. Natural convection increases remarkably (as shown
is activated, it is affected by buoyancy forces and it is posi- in Fig. 8) and the values seem to be similar to the measured
tive, from the ground to the half height of the building. On ones on building facßades (Defraeye et al., 2010). In this
the central plane, the buoyancy effect is maximum at 4 m study, the calculation of the heat transfer coefficient (hc)
height from the ground level and it extends up to 0.60 m is carried out at numerical level by the usual relations of
distance from the hot wall. It has been noticed that the the standard k–e model and it is strongly related to air ther-
temperatures of the facßade and of the ground that are mal conductivity (kT):
not exposed to the direct radiation, are 6° higher than dT
kT
the air temperature, but it does not seem to have relevant hc ¼
dy
ð9Þ
effects on the flow field. On the horizontal line the velocity T
magnitude is lower when the thermal effects are excluded. The air thermal conductivity is given by the following
The turbulent kinetic energy is higher near all the surfaces expression:
when the natural convection is activated and it strongly
affects the heat transfer coefficient trends. In particular, kT ¼ qcp K T ð10Þ
Fig. 8(b1), (b2), (d1) and (d2) shows that, activating the where the thermal turbulent diffusivity (KT) is related to the
natural convection, the turbulent kinetic energy increases kinematic turbulent viscosity (tT) through the turbulent
four times and the heat transfer coefficient doubles, from Prandtl number:
ground level to half height of the building. The natural
tT
convection effect determines also an increase of air circula- PrT ¼ ð11Þ
KT
tion within the canyon; as it can be seen in Fig. 8(c1) and
(c2), when the natural convection module is activated, the Which is taken equal to 0.85.
velocity magnitude reaches the value of 2 m/s, instead Thus, the kinematic turbulent viscosity is related to the
when it is deactivated, the maximum velocity magnitude turbulent kinetic energy (k) and to turbulent dissipation
value is 1.5 m/s. As it can be seen in Fig. 8(a1), (a2), (c1) rate (e) according to the following relationship:
and (c2), the surfaces temperatures are several degrees
lower when natural convection is activated. In order to k2
tT ¼ c l ð12Þ
evaluate the impact of 3D effects on the heat transfer coef- e
ficient, we have performed a 2D simulation and a 3D sim- where cl is 0.09.
ulation characterized by a transversal wind direction, The turbulent dissipation rate e is expressed by the fol-
excluding the natural convection. The results show that, lowing equation (Tominaga et al., 2008):
when the domain is tridimensional and the wind direction
is transversal to the canyon axis, the heat transfer coeffi- cl3=4 k 3=2
e¼ ð13Þ
cient values are very similar to the 2D simulation. The l
8 S. Bottillo et al. / Solar Energy 99 (2014) 1–10

Fig. 8. Impact of thermal effects on a vertical central plane within the canyon, comparison between parameter evaluated without natural convection
(subscript 1) and with natural convection (subscript 2): Z velocity component along a vertical line at 0.40 m from the windward facßade and its temperature
on a parallel line on the facßade itself (a1 and a2); turbulent kinetic energy and heat transfer coefficient along the same lines (b1 and b2); velocity magnitude
along an horizontal line at 0.40 m from the ground surface and it is temperature (c1 and c2), turbulent kinetic energy and heat transfer coefficient along the
same lines (d1 and d2).
S. Bottillo et al. / Solar Energy 99 (2014) 1–10 9

related to the square root of turbulent kinetic, as shown


in Fig. 8 and in Eq. (14), that in turn depends on the local
difference between the surfaces and air temperature.

5. Conclusions

In this study a numerical simulation method has been


used to investigate the physical phenomena that character-
ize a street canyon with a certain H/W ratio, N-S oriented,
during a summer day. A fully 3D model has been simu-
lated, considering the effects of solar radiation and radia-
tive exchange on canyon air and surfaces temperature. In
the first part, a comparison between a steady and a tran-
sient simulation has been carried out, and it can be con-
cluded that the stationary case is representative of
thermo-fluid dynamics parameters within the canyon dur-
ing the hottest hours of the day. In the second part of
the study the impact of thermal effects on the flow field
has been analyzed and the difference between a 2D simula-
Fig. 9. Heat transfer coefficient on the windward (WW) facßade and on the
leeward (LW) one for the 2D simulation of the street canyon.
tion and the fully tridimensional case has been shown.
Even when the flow field is characterized by a strong veloc-
ity component parallel to the canyon direction, it has been
where l is the mixing length. noticed that the buoyancy forces determine a thermal
So the correlation between the heat transfer coefficient induced vortex near the hottest facßade. The obtained
and turbulent kinetic energy is: results are coherent with the studies on numerical simula-
tions performed by Lei et al. (2012) and Xie et al. (2007),
dT
qcp c1=4
l l dy 1=2 even if their simulation has been carried out on 2D
hc ¼ k ð14Þ domains and the thermal effects are more significant. The
PrT T
importance of considering a 3D domain has been shown,
because the flow field within the canyon changes signifi-
In literature there are several empirical relationships cantly as the air passes through the canyon itself and the
between heat transfer coefficient over the external surfaces heat transfer coefficient values are remarkable higher than
of the buildings (hc) and air velocity (Defraeye et al., 2010), the 2D and the 3D transversal case. It can be seen that the
but a few authors (Oleson et al., 2008; Masson, 2000) con- wind direction strongly affects the thermal processes within
sider the effect of turbulent kinetic energy on hc. In those the canyon. So that further analysis must be performed
studies the heat transfer coefficient is related to the friction considering not only different wind velocity intensities but
velocity value, through the correlations reported in Mas- also different directions. Besides, buoyancy forces affect
cart et al. (1995), so that the heat transfer coefficient the flow field, determining thermal induced vortices and a
depends on the atmospheric stability considering average higher value of turbulent kinetic energy, affecting the heat
values on all the surfaces and for every wind speed direc- transfer coefficient and consequently the surface tempera-
tions. In our analysis, instead, the heat transfer coefficient tures. Taking into account that the standard k–e seems to
is related to the square root of turbulent kinetic, as shown overestimate the production of turbulent kinetic energy
in Fig. 8 and in Eq. (14). It has been noticed a direct cor- (Assimakopoulos et al., 2006; Lakehal, 1998; Meroney
relation of the heat transfer coefficient and the local ther- et al., 1999), further numerical analysis and measurement
mal stratification: for example, the thermal effects on the campaigns must be performed to investigate the correlation
heat transfer coefficient on the leeward wall in shadow between the heat transfer coefficient and the characteristics
are less marked than on the sun-exposed wall, but an (magnitude and direction) of flow field in an urban canyon,
increase of the average heat transfer coefficient can be in order to determine the local heat exchange processes and
observed (from 7.1 W/m2 K without thermal effects to their effects on the thermal comfort in outdoor spaces and
10.5 W/m2 K with natural convection module activated). on the buildings energy needs for cooling and heating.
It seems that the buoyancy induced vortex affects the over-
all canyon flow field. These variations of heat exchange References
processes among the various surfaces of the canyon, can
have an effect on the surface temperatures, as seen in Allegrini, J., Dorer, V., Carmeliet, J., 2012a. Analysis of convective heat
transfer at building facßades in street canyons and its influence on the
Fig. 8, and then a direct influence on thermal comfort indi-
predictions of space cooling demand in buildings. Journal of Wind
ces like the mean radiant temperature (De Lieto Vollaro Engineering and Industrial Aerodynamics 104–106, 464–473.
et al., 2013), in particular the heat transfer coefficient is
10 S. Bottillo et al. / Solar Energy 99 (2014) 1–10

Allegrini, J., Dorer, V., Defraeye, T., Carmeliet, J., 2012b. An adaptive ness length values for heat and momentum. Boundary-Layer Meteo-
temperature wall function for mixed convective flows at exterior rology 72, 331–344.
surfaces of buildings in street canyons. Building and Environment 49, Masson, V., 2000. A physically-based scheme for the urban energy budget
55–66. in atmospheric models. Boundary-Layer Meteorology 94, 357–397.
Ansys Fluent version 14.0.0, 2011. User’s Guide. Meroney, R.N., Leitl, B.M., Rafailidis, S., Schatzmann, M., 1999. Wind-
Assimakopoulos, V.D., Georgakis, C., Santamouris, M., 2006. Experi- tunnel and numerical modeling of flow and dispersion about several
mental validation of a computational fluid dynamics code to predict building shapes. Journal of Wind Engineering and Industrial Aerody-
the wind speed in street canyons for passive cooling purposes. Solar namics 81, 333–345.
Energy 80, 423–434. Offerle, B., Eliasson, I., Grimmond, C.S.B., Holmer, B., 2007. Surface
Blocken, B., Stathopoulos, T., Carmeliet, J., 2007. CFD simulation of the heating in relation to air temperature, wind and turbulence in an urban
atmospheric boundary layer: wall function problems. Atmospheric street canyon. Boundary-Layer Meteorology 122, 273–292.
Environment 41, 238–252. Oleson, K.W., Bonan, G.B., Feddema, J., Vertenstein, M., Grimmond,
Bozonnet, E., Belarbi, R., Allard, F., 2005. Modeling solar effects on the C.S.B., 2008. An urban parameterization for a global climate model.
heat and mass transfer in a street canyon, a simplified approach. Solar Part I: Formulation and evaluation for two cities. Journal of Applied
Energy 79, 10–24. Meteorology and Climatology 47, 1038–1060.
Defraeye, T., Blocken, B., Carmeliet, J., 2010. Convective heat transfer Ramponi, R., Blocken, B., 2012. CFD simulation of cross-ventilation for
coefficients for exterior building surfaces: existing correlations and a generic isolated building: impact of computational parameters.
CFD modeling. Energy Conversion and Management. Building and Environment 53, 34–48.
Defraeye, T., Blocken, B., Carmeliet, J., 2011. An adjusted temperature Richards, P.J., Norris, S.E., 2010. Appropriate boundary conditions for
wall function for turbulent forced convective heat transfer for bluff computational wind engineering models revisited. Journal of Wind
bodies in the atmospheric boundary layer. Building and Environment Engineering and Industrial Aerodynamics 99, 257–266.
46 (11), 2130–2141. Saneinejad, S., Moonen, P., Defraeye, T., Carmeliet, J., 2011. Analysis of
De Lieto Vollaro, R., Vallati, A., Bottillo, S., 2013. Different methods to convective heat and mass transfer at the vertical walls of a street
estimate the mean radiant temperature in an urban canyon. Advanced canyon. Journal of Wind Engineering and Industrial Aerodynamics
Materials Research 650, 647–651. 99, 424–433.
Franke, J., Hellsten, A., Schlünzen, H., Carissimo, B., 2007. Best practice Santamouris, M., Papanikolaou, N., Koronakis, I., Livada, I., Asimok-
guideline for the CFD simulation of flows in the urban environment. opoulos, D., 1999. Thermal and air flow characteristics in a deep
COST Action 732. pedestrian canyon under hot weather conditions. Atmospheric Envi-
Kovar–Panskus, A., Moulinneuf, L., Savory, E., Abdelqari, A., Sini, J.F., ronment 33, 4503–4521.
Rosant, J.M., Robins, A., Toy, N., 2002. A wind tunnel investigation Takebayashi, H., Moriyama, M., 2012. Relationships between the
of the influence of solar-induced wall-heating on the flow regime within properties of an urban street canyon and its radiant environment:
a simulated urban street canyon. Water, Air, and Soil Pollution: Focus introduction of appropriate urban heat island mitigation technologies.
2, 555–571. Solar Energy 86, 2255–2262.
Lakehal, D., 1998. Application of the k–e model to flow over a building Tominaga, Y., Mochidab, A., Yoshiec, R., Kataokad, H., Nozue, T.,
placed in different roughness sublayers. Journal of Wind Engineering Yoshikawaf, M., Shirasawac, T., 2008. AIJ guidelines for practical
and Industrial Aerodynamics 73, 59–77. applications of CFD to pedestrian wind environment around build-
Lei, L., Lin, Y., Li-Jie, Z., Yin, J., 2012. Numerical study on the impact of ings. Journal of Wind Engineering and Industrial Aerodynamics 96,
ground heating and ambient wind speed on flow fields in street 1749–1761.
canyons. Advances in Atmospheric Sciences 29, 1227–1237. Uehara, K., Murakami, S., Oikawa, S., Wakamatsu, S., 2000. Wind
Louka, P., Vachon, G., Sini, J.F., Mestayer, P.G., Rosant, J.M., 2002. tunnel experiments on how thermal stratification affects how in and
Thermal effects on the airflow in a street canyon – Nantes’99 above urban street canyons. Atmospheric Environment 34, 1553–1562.
experimental results and model simulations. Water, Air, Soil Pollution: Xie, X., Liu, C.H., Leung, D.Y.C., 2007. Impact of building facades and
Focus 2, 351–364. ground heating on wind flow and pollutant transport in street canyons.
Mascart, P., Noilhan, J., Giordani, H., 1995. A modified parameterization Atmospheric Environment 41, 9030–9049.
of flux-profile relationship in the surface layer using different rough-

You might also like