Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Building and Environment 95 (2016) 299e313

Contents lists available at ScienceDirect

Building and Environment


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

CFD simulation of stratied indoor environment in displacement


ventilation: Validation and sensitivity analysis
Sara Gilani a, *, Hamid Montazeri b, Bert Blocken b, c
a
School of Architecture, University of Tehran, P.O. Box 14155-6458, Enghelab Avenue, Tehran, Iran
b
Building Physics and Services, Department of the Built Environment, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven,
The Netherlands
c
Building Physics Section, Department of Civil Engineering, Leuven University, Kasteelpark Arenberg 40 e Bus 2447, 3001 Leuven, Belgium

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

Article history: Knowledge of temperature stratication in indoor environments is important for occupant thermal
Received 13 June 2015 comfort and indoor air quality and for the design and evaluation of displacement ventilation systems.
Received in revised form This paper presents a detailed and systematic evaluation of the performance of 3D steady Reynolds-
11 August 2015
Averaged NaviereStokes (RANS) CFD simulations to predict the temperature stratication within a
Accepted 9 September 2015
Available online 11 September 2015
room with a heat source and two ventilation openings. The indoor air quality within the room is also
investigated by assessing the distribution of the age of air. The evaluation is based on validation with full-
scale measurements of air temperature. A sensitivity analysis is performed to investigate the impact of
Keywords:
Computational Fluid Dynamics (CFD)
computational grid resolution, turbulence model, discretization schemes and iterative convergence on
Displacement ventilation the predicted temperatures and age of air. The results show that steady RANS with the SST k-u model can
Buoyancy-driven ventilation accurately predict the temperature stratication in this particular indoor environment. However, of the
Temperature stratication ve commonly used turbulence models, only the SST k-u model and the standard k-u model succeed in
Turbulence modeling reproducing the thermal plume structure and the associated thermal stratication, while the three k-ε
Age of air models clearly fail in doing so. In addition, the iterative convergence criteria have a major impact on the
predicted age of air, where it is shown that very stringent criteria in terms of scaled residuals are required
to obtain accurate results. These required criteria are much more stringent than typical default settings.
This paper is intended to contribute to improved accuracy and reliability of CFD simulations for
displacement ventilation assessment.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction occupant thermal comfort and indoor air quality. This is especially
the case when buoyancy-driven ventilation, called displacement
Temperature stratication in indoor environments, generated ventilation, is incorporated into the design of a building. This mode
by heat sources such as heating systems, occupants, electronic of ventilation can efciently purge excess heat and pollutants from
equipment and solar radiation on interior surfaces, can affect interior spaces (e.g. Refs. [5e8]).
occupant thermal comfort. In addition, in higher building spaces, Research on temperature stratication in enclosed spaces can be
such as atria, temperature stratication can be considered as one of performed by different methods: full-scale measurements [9e15],
the main issues for providing thermal comfort in the lower and reduced-scale measurements [16,17], analytical methods
upper adjacent spaces. This phenomenon also affects the indoor air [16,18e21] and Computational Fluid Dynamics (CFD) [11,14,15,22].
quality that can be assessed by the local age of air [1e4], which can Building energy simulation is also another method to support de-
exhibit pronounced vertical gradients. Therefore, knowledge of signs and development of innovative building system concepts,
temperature stratication in building spaces is required to improve such as displacement ventilation, under a range of different dy-
namic operating conditions. However, for performance prediction
of stratied indoor environments, this method is generally
* Corresponding author. Present address: Department of Civil and Environmental considered too limited in terms of spatial resolution [23,24]. Full-
Engineering, Carleton University, 3432 Mackenzie Building, 1125 Colonel By Drive, scale measurements have the advantage of providing the capa-
Ottawa, Ontario K1S 5B6, Canada.
bility to study the real situations in their full complexity. However,
E-mail address: saragilani@cmail.carleton.ca (S. Gilani).

http://dx.doi.org/10.1016/j.buildenv.2015.09.010
0360-1323/© 2015 Elsevier Ltd. All rights reserved.
300 S. Gilani et al. / Building and Environment 95 (2016) 299e313

in full-scale measurements it is generally not possible to perform


measurements in all points in the domain and to control all the
boundary conditions (e.g. Refs. [23,25,26]). In reduced-scale labo-
ratory measurements, controlling most boundary conditions is
possible, but similarity constraints can be the main concerns [23].
Analytical methods have the advantage of simplicity, potentially
providing reliable physical information, without need for powerful
computing resources. However, they are generally not reliable for
complicated cases concerning both geometry and thermo-uid
boundary conditions [23]. On the other hand, CFD provides
detailed ow eld data in the whole computational domain and
without similarity constraints because simulations can be per-
formed at full scale. It also allows parametric studies to be carried
out easily and efciently [14,23,25e29]. In addition, the application
of CFD for studying indoor air quality (e.g. Refs. [30e34]), natural
ventilation (e.g. Refs. [14,15,35e40]) and stratied indoor envi-
ronment (e.g. Ref. [41]) is increasing since these are difcult to
predict with other methods, as highlighted in some recent review
papers [23,25,29]. Despite all the mentioned advantages, the ac- Fig. 1. Geometry of the test room: position of the inlet and outlet openings, heat
curacy and reliability of CFD simulations remain important con- source, vertical pole, line-1 and line-2.
cerns and therefore, CFD validation and verication are imperative.
This is especially the case for CFD simulation of buoyancy-driven vertical walls 1, 2 and 3 (Fig. 1) were all equal to 0.36 W/m2K. For
ventilation, in which the ow is normally very complex and de- wall 4, the U-value was 0.15 W/m2K. Air was supplied into the test
pends on many different physical parameters. For this type of room through an inlet at oor level with dimensions
ventilation, several CFD validation studies have been performed width  height ¼ 0.45  0.5 m2, 50% of which was perforated,
and the performance of different turbulence models has been resulting in an effective opening area of 0.1125 m2. The air was
evaluated (e.g. Refs. [42e55]). However, the conclusions on which extracted from the test room through an outlet with dimensions
turbulence models perform best are not always consistent. For width  height ¼ 0.525  0.220 m2 that was located at a height of
example, the superior performance of the RNG k-ε model has been 2.39 m in wall 2. A prismatic porous heat source with dimensions
pointed out in some studies, e.g. Refs. [42,43,45e47,52,55], while width  length  height ¼ 0.3  0.4  0.3 m3 induced displacement
other studies show that the SST k-u model is superior ventilation in the test room. It was placed 0.1 m above the oor
[22,54,56e59]. This inconsistency can be related to physical char- level and 2.7 m from the inlet (Fig. 1). It incorporated 24 light bulbs
acteristics of the case studies, but also different computational of 25 W in total, which were controlled individually to provide
parameters and settings used in different studies. Therefore, there different levels of heat load up to 600 W. The main frame of the
is a strong need for detailed and systematic sensitivity analyses, prism was aluminum and it was lled with aluminum chips to
from which more detailed best practice guidelines can be derived. distribute the heat evenly and to avoid short-wave radiation to the
For these reasons, this paper presents a detailed evaluation of wall surfaces.
3D steady Reynolds-Averaged NaviereStokes (RANS) CFD simula- The air temperature was measured by 30 thermocouples located
tions to predict the temperature, velocity and air quality in a room along a single vertical pole inside the test room (Fig. 1). The ther-
with a heat source and two ventilation openings. The evaluation is mocouples were distributed along the vertical pole with a higher
based on validation with full-scale measurements of air tempera- concentration near the oor and ceiling. The interior surface tem-
ture by Li et al. [19,60]. The impact of the computational grid res- peratures were measured by 22 thermocouples, ve on each wall,
olution, turbulence model, discretization schemes and iterative one on the ceiling and one on the oor. Two thermocouples
convergence on the CFD results of air temperature, velocity and air measured the inlet and outlet air temperatures. Five thermocouples
quality is investigated. The concept of age of air introduced by also measured the exterior surface temperature of the walls and the
Sandberg [61] is used to assess the indoor air quality. ceiling. The measurement uncertainty was estimated to be ± 0.1  C.
In Section 2, the full-scale measurements of indoor air temper- In this experiment, the impacts of different parameters such as
ature by Li et al. [19,60] are described. Section 3 explains the heat load, wall emissivity, inlet ow rate and inlet air temperature
computational settings and parameters for the reference case and were investigated through different cases including B1 to B5 and A1
compares the results of the CFD simulations with those of the to A4, in which ‘B’ and ‘A’ stand for black-painted and aluminum-
experiment. In Section 4, the sensitivity analysis is performed. The covered cases, respectively. The experimental data for cases B1,
limitations of the study are discussed in Section 5 and the main B2, B3 and A2 were reported more completely than the other cases
conclusions are outlined in Section 6. in Refs. [19,60]. The focus of this paper is on case B1 for which the
largest amount of experimental data was provided. In addition, this
2. Description of full-scale experiment case had the lowest air change rate (n), which was 1 h1. Therefore,
the air was supplied with a lower velocity and as a result, the ow
Full-scale measurements of air temperature for a typical was not a jet-type ow, which is important in this paper to inves-
room subjected to displacement ventilation were performed by tigate the stratied properties of air in the room. For this case, the
Li et al. [19,60]. The test room had inside dimensions inlet (Ti) and outlet air temperatures (To) were 16.0  C and 27.3  C,
width  length  height ¼ 3.6  4.2  2.75 m3 (Fig. 1). Two sets of respectively. The heat load (E) was 300 W.
experiments were performed to investigate the impact of radiative
heat transfer on the air temperature in the room. In the rst set, the 3. CFD simulations: reference case
interior surfaces of the test-room walls were painted black, while in
the other one, they were covered with aluminum sheets. The In this section, the computational parameters of the reference
overall heat transmittance values (U-values) of the oor, ceiling and case are described. These parameters are modied systematically
S. Gilani et al. / Building and Environment 95 (2016) 299e313 301

for the sensitivity analysis in the next section. Table 1


Measured interior surface temperature of vertical walls at
different heights.
3.1. Computational grid
Height (z) [m] Temperature [ C]
A computational model was made for the room used in the full- 0.08 22.4
scale measurement. The computational grid had 451,248 hexahe- 0.73 23.4
dral cells generated using the surface-grid extrusion method pre- 1.39 24.0
2.04 24.5
sented by van Hooff and Blocken [14] (Fig. 2). The maximum 2.68 24.4
stretching ratio was 1.2. A total number of 6 and 25 cells were used
along the length and height of the inlet, respectively. There were 8
and 7 cells for the outlet. The distance from the center point of the 1
, characteristic length 2.75 m, temperature difference between
wall-adjacent cell to the wall for different surfaces of the room was
average room air and heat source ~200 K, dynamic viscosity
0.0005 m. This corresponds to a maximum y* value of 1.8. As low-
1.83  105 kg/ms and thermal conductivity 0.026 W/mK.
Reynolds number modeling was used in this study, this value
The inlet boundary condition was a uniform velocity (u) (m/s),
ensured that a few cells were placed inside the viscous sub-layer.
calculated based on the experimental data using Eq. (2):

3.2. Boundary conditions


Q nV
u¼ ¼ (2)
Ai Ai
Two separate zones were specied in the test room: a uid-type
zone for the room air and a solidetype zone for the heat source.
where Q is the inlet ow rate (m3/s), Ai the inlet opening area (m2),
Note that in the experiments, the heat source was a porous design.
n the number of test-room air changes (s1) and V the test-room
In our simulations, however, the heat source was modeled as a solid
volume (m3). The inlet air temperature (Ti) was 16.0  C according
body with a constant heat generation rate. A source term was
to the experimental data.
dened for the zone of the heat source with a constant volumetric
A xed temperature condition was applied for the test-room
heat generation (q) of about 8333 W/m3 according to Eq. (1):
surfaces. Note that in Ref. [60] the interior surface temperature of
E the vertical walls was provided as the average value of the
q¼ (1) measured temperatures of the walls at ve different heights
Vhs
(Table 1). In the CFD simulations, the same values were used to
where E is the heat load (W) and Vhs the heat-source volume (m3). dene the surface temperature of the walls assuming that the
The incompressible ideal gas law was used to estimate the air temperature in between these heights varied linearly with height
density as a function of T [62]. Other properties of the air were (Fig. 3).
determined at the average value of the measured air temperatures Although the interior surface temperature of the oor and the
along the vertical pole inside the test room (i.e. 24.5  C). In this case, ceiling were measured in the experiment, their values were not
the molecular Prandtl number (Pr) of air was 0.71. The operating mentioned in Ref. [60]. Therefore, in the present paper, the interior
pressure and operating density were 101,325 Pa and 1.17 kg/m3, surface temperatures of the oor and the ceiling had to be esti-
respectively [63]. For the outlet, zero static pressure was specied. mated using the procedure outlined below, although it should be
The Rayleigh number (Ra) was calculated to be about 4  1010 and acknowledged that this procedure entails quite some assumptions.
therefore, Ra/Pr was 5.6  1010 which is in the threshold for the It is expected that deviations between these estimates and the
occurrence of turbulent natural convection [62]. The computed Ra actual values will change the absolute values of temperature and
corresponds to the following values: density 1.17 kg/m3, specic
heat capacity 1006.95 J/kgK, thermal expansion coefcient 0.003 K-

Fig. 2. Perspective view of high-resolution computational grid on three walls of the Fig. 3. Interior surface temperature of vertical walls at different heights imposed by
test room (total number of cells: 451,248). user-dened functions (UDFs) in the CFD simulations.
302 S. Gilani et al. / Building and Environment 95 (2016) 299e313

s), Tfa the near-oor air temperature (the nearest measured air
temperature to the oor, i.e. 22.8  C) and Ti the inlet air temperature
(16.0  C). hci is the interior convective heat transfer coefcient (5 W/
m2K) according to ISO 6946 [64] and Af the oor surface area (m2).
In this case, the interior surface temperature of the oor (Tif) ob-
tained by Eq. (3) was 24.0  C.
The interior surface temperature of the ceiling (Tic) was calcu-
lated by applying the energy conservation equation for the ceiling-
room interface as shown in Fig. 4. There are three heat transfer
mechanisms for this control surface: conduction from the ceiling to
00
the control surface (qcond ), convection from the ceiling to the air
00
inside the room (qconv ) and radiation from the ceiling to the sur-
00
roundings (qrad ) as shown in Eqs. (4) and (5):

Fig. 4. Energy balance at the interior surface of the ceiling.


00 00 00
qcond ¼ qconv þ qrad (4)
age of air in the room, but not the observed trends of the validation
and sensitivity analysis. Two reasons can support this expectation.
First, the comparison e albeit limited - of the numerically simu-
j¼25
X  
lated temperatures with the measured values shows quite a good k
ðTec  Tic Þ ¼ hi ðTic  T∞ Þ þ sεic Fic/j Tic4  Tsurf
4
(5)
agreement along the vertical pole, including the regions near the L ðjÞ
j¼1
oor and ceiling (will be explained in Section 3.4). Second, the
trends, e.g. concerning turbulence model performance (see Fig. 10) where k is the thermal conductivity of the ceiling material (W/mK),
are very pronounced, most likely beyond the limited inuence of L the ceiling thickness (m), Tec the exterior surface temperature of
potentially small deviations in the oor and ceiling temperatures. the ceiling (K) (20.3  C ¼ 293.45 K), Tic the interior surface tem-
The oor surface temperature was calculated using Eq. (3) [19]. perature of the ceiling (K) and hi the interior convective heat
This equation is for a room ventilated by displacement ventilation, transfer coefcient (10 W/m2K). T∞, the ambient air temperature
and three zones were assumed in the room: the zone just above the (K), was assumed to be equal to the average value of the measured
oor, the zone just below the ceiling and the region in-between. air temperatures along the vertical pole (24.5  C ¼ 297.65 K). s is
The temperature prole was assumed to vary linearly with height the StefaneBoltzmann constant (W/m2K4), εic the emissivity of the
in each zone. In addition, the supplied air and exhausted air were interior surface of the ceiling (i.e. 1.0 in the experiment), Tsurf(j) the
assumed to spread evenly in the region just above the oor and just temperature of each of the other surfaces (K) and Fic/j the view
below the ceiling, respectively. Eq. (3) expresses the heat balance in factor of the ceiling to other surfaces, including the oor and the
the zone just above the oor and represents that the supplied air is walls. The view factors of the ceiling to the other surfaces were
heated by convection near the oor surface. Note that this is an calculated based on equations presented in Ref. [63] for perpen-
approximation needed because of the lack of experimental data in dicular rectangles with a common edge (Table 2). Since different
Ref. [60]. This equation was developed and used by Li et al. [19]. values were measured for surface temperature at 5 different
heights of the walls, the view factors of the ceiling to the walls were
  calculated for 6 portions of walls for which different temperatures
rcp Q Tfa  Ti
were measured in the experiment. Therefore, in Eq. (5), j was 25
Tif ¼ þ Tfa (3)
hci Af including the oor and 6 portions for each wall. Tsurf(j) for each
portion of walls were calculated as the average of the temperatures
where Tif is the oor surface temperature, r the air density (kg/m3), of the bottom and top line of that portion. The conduction thermal
cp the specic heat capacity of air (J/kgK), Q the inlet ow rate (m3/ resistance (L/k) of the ceiling was calculated using Eq. (6):

Fig. 5. (a, b) Comparison of temperature scale (T eTi) by CFD simulation and experimental results.
S. Gilani et al. / Building and Environment 95 (2016) 299e313 303

Fig. 6. (a, b) Distribution of air temperature, (c, d) age of air and (e, f) velocity vector eld across two vertical planes.

(297.35 K).
1 1 L 1
¼ þ þ (6)
U he k hi
3.3. Solver settings
where U is the overall heat transmittance (W/m2K), L the ceiling
thickness (m), k the thermal conductivity of the ceiling material The commercial CFD code ANSYS/Fluent 12.1 was used to
(W/mK) and he and hi the exterior and interior surface heat transfer perform the simulations [62]. The 3D steady RANS equations were
coefcient, respectively. Note that he and hi were assumed to be solved in combination with the Shear-Stress Transport (SST) k-u
10 W/m2K as the combined radiation and convection coefcients model [65]. The SIMPLE algorithm was used for pressureevelocity
based on ISO 6946 [64]. coupling. Second order discretization schemes were used for both
By calculating the thermal resistance (L/k) of the ceiling based the convection and viscous terms of the governing equations. The
on Eq. (6) and knowing other parameters in Eq. (5), the interior PRESTO! scheme was applied for the pressure terms. Because of the
surface temperature of the ceiling was calculated to be 24.2  C temperature-dependent density of the air, the energy and
304 S. Gilani et al. / Building and Environment 95 (2016) 299e313

Fig. 7. Computational grids for grid-sensitivity analysis: (a) coarse grid (155,382 cells), (b) reference grid (451,248 cells) and (c) ne grid (1,268,736 cells).

momentum equations were solved simultaneously. Converged so- results along the vertical pole. The deviations between the mea-
lution was assumed to be achieved when the net heat ux imbal- surements and CFD are also depicted in Fig. 5b. The general
ance was less than 1% of the smallest heat ux through the domain agreement is quite good. CFD underestimates the temperature near
boundaries. In addition, the velocity magnitudes were monitored at the oor and the ceiling and slightly overestimates it along the
2 points in the test room until fully converged values were obtained other parts of the pole. The main reason for these deviations is not
[66]. In the remainder of the paper, the “heat ux criterion (ref. clear. However, it can be related to the assumptions made for the
case)” is referred to the criterion used for the reference case. uniform inlet velocity prole and for calculating the surface tem-
peratures of the ceiling and the oor. The average deviation and
3.4. Results and comparison with experiment maximum deviation of the absolute values of the temperature scale
between CFD and measurements are 4.2% and 11.9%, respectively.
The CFD results for the reference case, which was outlined in the Figs. 6a and b show the air temperature distribution across two
previous sub-sections, are compared with the full-scale measure- vertical planes located 2.1 m from wall 2 and 2.7 m from wall 1,
ments by Li et al. [60]. The difference between the air temperature respectively. The rst plane intersects the inlet opening and the
within the room (T) and the inlet air temperature (Ti) is given as the heat source volume. The second plane intersects the heat source
temperature scale. Fig. 5a compares the experimental and CFD volume and the outlet opening. The distribution of the age of air

Fig. 8. Impact of grid resolution on: (aec) prole of temperature scale and (def) age of air, along three vertical lines.
S. Gilani et al. / Building and Environment 95 (2016) 299e313 305

Fig. 9. Impact of grid resolution on: (aec) prole of temperature scale and (def) age of air, along three vertical lines, with error band of grid convergence index by Roache [67,68].

across the same planes is presented in Figs. 6c and d. Figs. 6e and f 4.1. Impact of computational grid resolution
display the velocity vector eld across the same planes. As shown in
these gures, a plume-type ow caused by the buoyancy effect can To perform a grid-sensitivity analysis, two additional grids were
be clearly seen above the heat source. This results in the develop- made: a ner grid and a coarser grid. Rening and coarsening were
ment of recirculation zones and vertical temperature distribution performed with an overall linear factor √2. The coarser grid had
outside the plume. The local age of air is also affected by the vertical 155,382 cells and the ner grid had 1,268,736 cells. The three grids
temperature stratication outside the plume. The temperature are shown in Fig. 7.
stratication leads to a vertical difference in the air density and The proles of the temperature scale and the age of air along the
therefore an air ow occurs from the lower parts to the upper parts pole and two other vertical lines, as shown in Fig. 1, for the three
of the room. This air ow impinges on the ceiling and is trapped in grids are presented in Fig. 8. The average deviations between the
the recirculation zones which lead to an increase in the local age of CFD results and measurements for the absolute values of the
air in the upper regions. The fresh air, entering from the inlet temperature scale along the pole are 4.2%, 4.2% and 4.0% for the
opening, is deected downward. As a result, the fresh air that has a coarse, reference and ne grid, respectively. In this case, the average
young age compared to the room air cannot propagate deeply in the deviation between the coarse and reference grid along the three
room. lines is 1.7% while it is about 0.9% between the ne and reference
grid. For the age of air, these deviations are about 2.7% and 1.6%,
respectively. The room average age of air for the coarse, reference
and ne grid is about 31.5, 31.3 and 31.0 min, respectively; which
4. CFD simulation: sensitivity analysis shows an absolute deviation of about 0.8% for both the coarse and
ne grid from the reference grid. The Grid-Convergence Index (GCI)
A systematic and detailed sensitivity analysis was performed by Roache [67,68] was used for uniform reporting of this grid
based on the reference case that was outlined in the previous convergence study (Fig. 9). The results show that the deviation is
section. To analyze the sensitivity of the results, a single parameter more noticeable for the lower parts of the pole, line-1 and line-2.
was varied while all other parameters were kept identical to those For the other parts of these lines, the deviation is negligible.
in the reference case. Then, the results were compared to the Therefore, the reference grid was retained for further analysis.
reference case to evaluate the impact of the change made on the
simulation results. The parameters tested are the resolution of the 4.2. Impact of turbulence model
computational grid (Section 4.1), the turbulence model (Section
4.2), the discretization scheme (Section 4.3) and the convergence 3D steady RANS CFD simulations were performed with different
criterion (Section 4.4). turbulence models including:
306 S. Gilani et al. / Building and Environment 95 (2016) 299e313

Fig. 10. Impact of turbulence model on: (aec) proles of temperature scale and (def) age of air, along three vertical lines.

 Standard k-ε model (Sk-ε) [69], the middle of the pole and underestimate it near the ceiling. Note
 Realizable k-ε model (Rk-ε) [70], that in several previous studies, the superior performance of the
 Renormalization Group k-ε model (RNG k-ε) [71], SST k-u model for displacement ventilation has also been pointed
 Standard k-u model (Sk-u) [72], and out [22,54,56e59]. The success of the prediction could depend on
 Shear-stress transport (SST) k-u model (SST k-u) [65]. the way in which the boundary layer is modeled. In this study, in
order to model the boundary layer accurately, wall functions were
Note that the Sk-ε, Rk-ε and RNG k-ε model were used in com- not applied but so-called near-wall modeling. All tested turbulence
bination with the low-Reynolds number Wolfshtein model [73]. models except the Sk-u and SST k-u models are so-called high-
The impact of the choice of turbulence model on the results of the Reynolds number models, which near the wall are replaced by the
temperature scale and the age of air along the vertical pole, line-1 one-equation Wolfshtein model [73], while the Sk-u and SST k-u
and line-2 is illustrated in Fig. 10. models are low-Reynolds number models and are applied all the
The average deviations of the absolute values of the temperature way down to the wall. However, since the Sk-u model does not
scale between CFD and measurements along the pole for Sk-ε, Rk-ε, show superior performance compared to the high-Re number
RNG k-ε, Sk-u and SST k-u are 7.0%, 7.9%, 5.6%, 4.9% and 4.2%, models, it appears that the near-wall treatment cannot be held
respectively. In this case, the maximum deviations of the absolute responsible for the superior performance of the SST k-u model and
values are 19.6%, 15.5%, 13.0%, 10.6% and 11.9%, respectively. It can be the inferior performance of the other four models. As a result, it can
seen that the SST k-u model shows the best performance, followed be concluded that the SST k-u model predicts the buoy-
by the Sk-u model, which also shows a fairly good performance. The ancyeturbulence interaction.
differences by the models are pronounced along the pole except Figs. 10def clearly show different results for the age of air ob-
near the ceiling. All models tend to overestimate the temperature in tained by different turbulence models along the three lines. The Sk-
u model provides very close results to the SST k-u model (reference
case) with the average deviation of 3.5% along the three lines. The
Table 2
other turbulence models, however, show higher deviations
View factors of ceiling to other surfaces.
compared to the reference case. In this case, the average deviations
Walls Height (z) [m] Wall 1, Wall 3 Wall 2, Wall 4 of the absolute values along the three lines are 9.0%, 5.7%, and 8.1%
0.00  z < 0.08 0.0025 0.0022
for Sk-ε, Rk-ε and RNG k-ε, respectively. In addition, the value for the
0.08  z < 0.73 0.0246 0.0213
0.73  z < 1.39 0.0350 0.0299 room average age of air for Sk-u is almost the same as that for the
1.39  z < 2.04 0.0488 0.0412 reference case (i.e. 31.3 min) while it is about 31.5, 30.9 and
2.04  z < 2.68 0.0688 0.0582 30.4 min for Sk-ε, Rk-ε and RNG k-ε, respectively.
2.68  z  2.75 0.0094 0.0080
Fig. 11 shows the velocity vector eld across two vertical planes
Floor 0.3004
S. Gilani et al. / Building and Environment 95 (2016) 299e313 307

Fig. 11. Impact of turbulence model on velocity vector eld across two vertical planes: (a, b) SST k-u, (c, d) Sk-u, (e, f) Sk-ε, (g, h) Rk-ε, (i, j) RNG k-ε.
308 S. Gilani et al. / Building and Environment 95 (2016) 299e313

Fig. 12. Impact of order of discretization scheme on: (aec) proles of temperature scale and (def) age of air, along three vertical lines.

Fig. 13. Comparison of the rst-order discretization scheme on the reference and coarse grid: (aec) proles of temperature scale and (def) age of air, along three vertical lines.
S. Gilani et al. / Building and Environment 95 (2016) 299e313 309

located 2.1 m from wall 2 and 2.7 m from wall 1, respectively, as For the age of air, the average deviations from the reference case,
obtained by the different turbulence models. The plume-type ow along the pole, line-1 and line-2 are about 5.6%, 3.3% and 3.2%,
and recirculation zones outside the plume are clearly reproduced respectively. And, the room average age of air for the rst-order
by the SST k-u and Sk-u model, but not by the other models. This discretization scheme on the coarse grid is about 31.5 min, which
difference in performance is the reason why the latter models show is only 0.6% higher than the reference case (Fig. 13). It can be
larger discrepancies along the pole in Fig. 10. They poorly reproduce concluded that even on a relatively coarse grid (155,382 cells), the
the air ow pattern in the test room. impact of the order of the discretization schemes is minor.

4.3. Impact of order of discretization scheme 4.4. Impact of level of iterative convergence

The impact of the order of the discretization scheme on the In literature, there is no general agreement on the required level
temperature scale and age of air along the three lines is shown in of iterative convergence for an adequately-converged solution. Too
Fig. 12. For the current grid resolution, the impact is rather small: lenient convergence criteria should be avoided. For instance, the
the average deviations of the absolute values of the temperature default setting of scaled residuals in Fluent 6.3 is 103, which is a
scale between CFD and measurements along the pole are 5.0% and rather high and lenient value [29], although the ANSYS training
4.2% for rst-order and second-order discretization schemes, manual has presented tighter criteria to check whether a converged
respectively. The deviation between the two schemes for line-2 is solution has been achieved [66]. In some previous studies, much
less than for the pole and line-1. more stringent convergence criteria were imposed [14,28,74e76].
For the age of air, the average deviations between the two In this section, the impact of three different minimum thresholds of
schemes along the pole, line-1 and line-2 are also small: about 3.1%, scaled residuals on the results is investigated. For these three cases,
2.2% and 1.6%, respectively. The room average age of air for the rst- a converged solution was assumed to be obtained when all the
order discretization scheme is about 31.1 min, which is 0.6% lower scaled residuals leveled off and reached a minimum of the pre-
than the one with the second-order discretization scheme (refer- scribed value (i.e. 103, 104 and 108).
ence case). Fig. 14 shows the proles of the temperature scale and age of air
The small impact of the order of the discretization scheme is along the lines for the three thresholds. It can be seen that the level of
attributed to the relatively high grid resolution, which reduces the iterative convergence has a signicant impact on the results of the
error by numerical diffusion by the rst-order discretization age of air. This is especially the case for the thresholds of 103 and
scheme. Even for the rst-order calculation on the coarse grid, the 104, where the average deviations from the reference case are 14.9%
results show that the average absolute deviation of the temperature and 12.9%, respectively. For the temperature scale, however, these
scale between CFD and measurements along the pole is only 4.9%. deviations are less, i.e. 1.4% and 0.4%. Fig. 14 also conrms that the

Fig. 14. Impact of iterative convergence limit for the reference case and minimum thresholds for scaled residuals of 103, 104 and 108 on: (aec) proles of temperature scale and
(def) age of air, along three vertical lines.
310 S. Gilani et al. / Building and Environment 95 (2016) 299e313

Fig. 15. Impact of iterative convergence limit on distribution of age of air across two vertical planes for: (a, b) reference case, (c, d) minimum thresholds for scaled residuals of 103,
(e, f) 104 and (g, h) 108.
S. Gilani et al. / Building and Environment 95 (2016) 299e313 311

108 threshold results in almost the same results as those of the  Although the amount of experimental data is limited, the vali-
reference case for both the temperature scale and the age of air. Note dation study shows that the 3D steady RANS approach with the
that further reduction in the minimum threshold of scaled residuals SST k-u can accurately predict the vertical temperature prole
has no signicant impact on the results (not shown in this gure). and stratication.
Fig. 15 shows the distribution of the age of air in the vertical  The use of three different computational grids (155,382 cells,
planes for the reference case and for the 103, 104 and 108 451,248 cells and 1,268,736 cells) indicates that the grid reso-
thresholds. It shows that 103 and 104 thresholds result in sub- lution only has a minor inuence on the resulting vertical pro-
stantial deviations from the reference case. However, the 108 les of temperature and age of air.
threshold yields the same distribution of the age of air as the  The impact the turbulence model however is very large. Out of
reference case. The room average age of air for the 103 and 104 the ve commonly used turbulence models tested, the Sk-ε, Rk-
threshold are about 26.8 and 27.8 min, which shows a deviation of ε, RNG k-ε, Sk-u and SST k-u model, only the two k-u models
14.3% and 11.2% from the reference case. The 108 threshold pro- succeed in reproducing the thermal plume structure and the
vides the same value for the room average age of air as the refer- associated thermal stratication, while the three k-ε models
ence case (i.e. 31.3 min). clearly fail in doing so.
 The impact of the order of the discretization schemes is minor,
5. Discussion even for the coarsest grid (155,382), which indicates that nu-
merical diffusion is limited and/or does not substantially affect
It is important to mention some limitations of this study: the results in terms of temperature and age of air.
 Finally, the iterative convergence criteria have a very large
 In this study, a room with a heat source and two ventilation impact on the results of the age of air, where it is shown that
openings was considered. Earlier studies have shown the very stringent criteria in terms of scaled residuals are required
importance of different physical parameters on the performance to obtain accurate results. These required criteria are much more
of displacement ventilation systems, such as supply air velocity stringent than typical default settings.
and temperature [10,43], room geometry [16,46,77], intensity
[10,16], location [78], size [79,80] and number of heat sources Nomenclature
[81,82], size [16,77,81], number [43,77,81] and location of
openings [44] and air change rate [43,79]. Further research is s StefaneBoltzmann constant (W/m2K4)
therefore needed to evaluate the performance of steady RANS εi surface emissivity ()
CFD simulations to predict the performance of this type of conduction heat ux (W/m2)
00
qcond
ventilation under different physical characteristics. convection heat ux (W/m2)
00
qconv
 This study was performed for a single isolated building space radiation heat ux (W/m2)
00
qrad
and the inuence of connected interior spaces was not included. u inlet air velocity (m/s)
Further research is needed on the validation, verication and Af oor surface area (m2)
sensitivity analysis of steady RANS CFD simulations for the Ai inlet opening area (m2)
stratied indoor environment in more complex spaces. cp specic heat capacity of air (J/kgK)
 The study only considered steady RANS CFD simulations, as the E heat load (W)
purpose was to investigate how well steady RANS CFD would be hci interior convective heat transfer coefcient (W/m2K)
able to predict the temperature stratication within a room with he exterior surface heat transfer coefcient (W/m2K)
a heat source and two ventilation openings. In spite of the well- hi interior surface heat transfer coefcient (W/m2K)
known deciencies of steady RANS, a good agreement was ob- k thermal conductivity (W/mK)
tained between CFD simulations and measurements. Obtaining L thickness (m)
a better agreement here would necessitate the use of Large Eddy n number of test-room air changes (1/s)
Simulation (LES), which however is much more computationally Pr molecular Prandtl number ()
expensive than steady RANS. Q inlet ow rate (m3/s)
 Further research is necessary on exploring the impact of radia- q volumetric heat generation (W/m3)
tive heat transfer between the interior surfaces and conductive Ra Rayleigh number ()
heat transfer through the walls. Tfa near-oor air temperature ( C)
T∞ ambient air temperature (K)
Tec exterior surface temperature of the ceiling (K)
6. Conclusions Ti inlet air temperature ( C)
Tic interior surface temperature of the ceiling (K)
Knowledge of temperature stratication in indoor environ- Tif oor surface temperature ( C)
ments is required for occupants' local thermal comfort. In addition, To outlet air temperature ( C)
in higher building spaces, such as atriums, temperature stratica- Tsurf surface temperature (K)
tion is important for providing thermal comfort in the lower and U overall heat transmittance (W/m2K)
upper adjacent spaces. This phenomenon also affects the indoor air V test-room volume (m3)
quality by causing a vertical difference in local age of air. Vhs heat-source volume (m3)
In this paper, a detailed and systematic assessment of 3D steady r air density (kg/m3)
RANS CFD simulations was performed for determining the air
temperature distribution and for evaluating the air quality caused References
by temperature stratication within a space with a heat source and
two ventilation openings. This paper is intended to contribute to [1] M. Sandberg, M. Sjo € berg, The use of moments for assessing air quality in
improved accuracy and reliability of CFD simulations for displace- ventilated rooms, Build. Environ. 18 (1983) 181e197.
[2] D. Etheridge, Natural Ventilation of Buildings: Theory, Measurement and
ment ventilation assessment. Design, John Wiley & Sons, Ltd, 2012.
The following conclusions are obtained: [3] C. Buratti, R. Mariani, E. Moretti, Mean age of air in a naturally ventilated
312 S. Gilani et al. / Building and Environment 95 (2016) 299e313

ofce: experimental data and simulations, Energy Build. 43 (2011) [35] C.F. Gao, W.L. Lee, Evaluating the inuence of openings conguration on
2021e2027. natural ventilation performance of residential units in Hong Kong, Build.
[4] A. Meiss, J. Feijo-Mun ~ oz, M.A. García-Fuentes, Age-of-the-air in rooms ac- Environ. 46 (2011) 961e969.
cording to the environmental condition of temperature: a case study, Energy [36] M.Z.I. Bangalee, S.Y. Lin, J.J. Miau, Wind driven natural ventilation through
Build. 67 (2013) 88e96. multiple windows of a building: a computational approach, Energy Build. 45
[5] Y. Wang, F.-Y. Zhao, J. Kuckelkorn, D. Liu, J. Liu, J.-L. Zhang, Classroom energy (2012) 317e325.
efciency and air environment with displacement natural ventilation in a [37] R. Ramponi, B. Blocken, CFD simulation of cross-ventilation for a generic
passive public school building, Energy Build. 70 (2014) 258e270. isolated building: impact of computational parameters, Build. Environ. 53
[6] S. Hussain, P.H. Oosthuizen, Numerical investigations of buoyancy-driven (2012) 34e48.
natural ventilation in a simple three-storey atrium building and thermal [38] H. Montazeri, Experimental and numerical study on natural ventilation per-
comfort evaluation, Appl. Therm. Eng. 57 (2013) 133e146. formance of various multi-opening wind catchers, Build. Environ. 46 (2011)
[7] S. Hussain, P.H. Oosthuizen, Numerical investigations of buoyancy-driven 370e378.
natural ventilation in a simple atrium building and its effect on the thermal [39] H. Montazeri, F. Montazeri, R. Azizian, S. Mostafavi, Two-sided wind catcher
comfort conditions, Appl. Therm. Eng. 40 (2012) 358e372. performance evaluation using experimental, numerical and analytical
[8] Z. Lin, T.T. Chow, C.F. Tsang, Effect of door opening on the performance of modeling, Renew. Energy 35 (2010) 1424e1435.
displacement ventilation in a typical ofce building, Build. Environ. 42 (2007) [40] H. Montazeri, R. Azizian, Experimental study on natural ventilation perfor-
1335e1347. mance of a two-sided wind catcher, Proc. Inst. Mech. Eng. Part A J. Power
[9] M.N.A. Saïd, R.A. MacDonald, G.C. Durrant, Measurement of thermal strati- Energy 223 (2009) 387e400.
cation in large single-cell buildings, Energy Build. 24 (1996) 105e115. [41] S.A. Howell, I. Potts, On the natural displacement ow through a full-scale
[10] M. Xu, T. Yamanaka, H. Kotani, Vertical proles of temperature and contam- enclosure, and the importance of the radiative participation of the water
inant concentration in rooms ventilated by displacement with heat loss vapour content of the ambient air, Build. Environ. 37 (2002) 817e823.
through room envelopes, Indoor Air 11 (2001) 111e119. [42] M.J. Cook, K.J. Lomas, Buoyancy-driven displacement ventilation ows: eval-
[11] M.P. Wan, C.Y. Chao, Numerical and experimental study of velocity and uation of two eddy viscosity turbulence models for prediction, Build. Serv.
temperature characteristics in a ventilated enclosure with underoor venti- Eng. Res. Technol. 19 (1998) 15e21.
lation systems, Indoor Air 15 (2005) 342e355. [43] N. Kobayashi, Q. Chen, Floor-supply displacement ventilation in a small ofce,
[12] C. Crouzeix, J.L. Le Moue €l, F. Perrier, P. Richon, Thermal stratication induced Indoor Built Environ. 12 (2003) 281e291.
by heating in a non-adiabatic context, Build. Environ. 41 (2006) 926e939. [44] Z. Lin, T.T. Chow, C.F. Tsang, K.F. Fong, L.S. Chan, CFD study on effect of the air
[13] A.S. Awad, R.K. Calay, O.O. Badran, A.E. Holdo, An experimental study of supply location on the performance of the displacement ventilation system,
stratied ow in enclosures, Appl. Therm. Eng. 28 (2008) 2150e2158. Build. Environ. 40 (2005) 1051e1067.
[14] T. van Hooff, B. Blocken, Coupled urban wind ow and indoor natural venti- [45] Y. Ji, M.J. Cook, Numerical studies of displacement natural ventilation in multi-
lation modelling on a high-resolution grid: a case study for the Amsterdam storey buildings connected to an atrium, Build. Serv. Eng. Res. Technol. 28
ArenA stadium, Environ. Model. Softw. 25 (2010) 51e65. (2007) 207e222.
[15] T. van Hooff, B. Blocken, CFD evaluation of natural ventilation of indoor en- [46] Y. Ji, M.J. Cook, V. Hanby, CFD modelling of natural displacement ventilation in
vironments by the concentration decay method: CO2 gas dispersion from a an enclosure connected to an atrium, Build. Environ. 42 (2007) 1158e1172.
semi-enclosed stadium, Build. Environ. 61 (2013) 1e17. [47] P.-C. Liu, H.-T. Lin, J.-H. Chou, Evaluation of buoyancy-driven ventilation in
[16] P. Cooper, P.F. Linden, Natural ventilation of an enclosure containing two atrium buildings using computational uid dynamics and reduced-scale air
buoyancy sources, J. Fluid Mech. 311 (1996) 153e176. model, Build. Environ. 44 (2009) 1970e1979.
[17] R. Tovar, P.F. Linden, L.P. Thomas, Hybrid ventilation in two interconnected [48] Y. Cheng, J. Niu, N. Gao, Stratied air distribution systems in a large lecture
rooms with a buoyancy source, Sol. Energy 81 (2007) 683e691. theatre: a numerical method to optimize thermal comfort and maximize
[18] P.V. Nielsen, A. Restivo, J.H. Whitelaw, Buoyancy-affected ows in ventilated energy saving, Energy Build. 55 (2012) 515e525.
rooms, Numer. Heat. Transf. Part A 2 (1979) 115e127. [49] Y. Jiang, Q. Chen, Buoyancy-driven single-sided natural ventilation in build-
[19] Y. Li, M. Sandberg, L. Fuchs, Vertical temperature proles in rooms ventilated ings with large openings, Int. J. Heat Mass Transf. 46 (2003) 973e988.
by displacement: full-scale measurement and nodal modelling, Indoor Air 2 [50] C. Gladstone, A.W. Woods, On buoyancy-driven natural ventilation of a room
(1992) 225e243. with a heated oor, J. Fluid Mech. 441 (2001) 293e314.
[20] Y. Li, Buoyancy-driven natural ventilation in a thermally stratied one-zone [51] G. Gan, Simulation of buoyancy-driven natural ventilation of buil-
building, Build. Environ. 35 (2000) 207e214. dingsdimpact of computational domain, Energy Build. 42 (2010) 1290e1300.
[21] C. Crouzeix, J.L. Le Moue €l, F. Perrier, P. Richon, Non-adiabatic boundaries and [52] C. Walker, G. Tan, L. Glicksman, Reduced-scale building model and numerical
thermal stratication in a conned volume, Int. J. Heat Mass Transf. 49 (2006) investigations to buoyancy-driven natural ventilation, Energy Build. 43 (2011)
1974e1980. 2404e2413.
[22] A. Stamou, I. Katsiris, Verication of a CFD model for indoor airow and heat [53] H. Xing, A. Hatton, H.B. Awbi, A study of the air quality in the breathing zone
transfer, Build. Environ. 41 (2006) 1171e1181. in a room with displacement ventilation, Build. Environ. 36 (2001) 809e820.
[23] Q. Chen, Ventilation performance prediction for buildings: a method overview [54] T. Zhang, K. Lee, Q. Chen, A simplied approach to describe complex diffusers
and recent applications, Build. Environ. 44 (2009) 848e858. in displacement ventilation for CFD simulations, Indoor Air 19 (2009)
[24] J.L.M. Hensen, M.J.H. Hamelinck, Energy simulation of displacement ventila- 255e267.
tion in ofces, Build. Serv. Eng. Res. Technol. 16 (1995) 77e81. [55] Z. Lin, T.T. Chow, Q. Wang, K.F. Fong, L.S. Chan, Validation of CFD model for
[25] B. Blocken, 50 years of computational wind engineering: past, present and research into displacement ventilation, Archit. Sci. Rev. 48 (2005) 305e316.
future, J. Wind Eng. Industrial Aerodyn. 129 (2014) 69e102. [56] Z.J. Zhai, Z. Zhang, W. Zhang, Q.Y. Chen, Evaluation of various turbulence
[26] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa, et al., models in predicting airow and turbulence in enclosed environments by
AIJ guidelines for practical applications of CFD to pedestrian wind environ- CFD: part 1dsummary of prevalent turbulence models, HVAC&R Res. 13
ment around buildings, J. Wind Eng. Industrial Aerodyn. 96 (2008) (2007) 853e870.
1749e1761. [57] Z. Zhang, W. Zhang, Z.J. Zhai, Q.Y. Chen, Evaluation of various turbulence
[27] B. Blocken, T. Stathopoulos, J. Carmeliet, J.L.M. Hensen, Application of models in predicting airow and turbulence in enclosed environments by
computational uid dynamics in building performance simulation for the CFD: part 2dcomparison with experimental data from literature, HVAC&R
outdoor environment: an overview, J. Build. Perform. Simul. 4 (2011) Res. 13 (2007) 871e886.
157e184. [58] S. Hussain, P.H. Oosthuizen, Validation of numerical modeling of conditions in
[28] H. Montazeri, B. Blocken, CFD simulation of wind-induced pressure co- an atrium space with a hybrid ventilation system, Build. Environ. 52 (2012)
efcients on buildings with and without balconies: validation and sensitivity 152e161.
analysis, Build. Environ. 60 (2013) 137e149. [59] S. Hussain, P.H. Oosthuizen, A. Kalendar, Evaluation of various turbulence
[29] B. Blocken, Computational uid dynamics for urban physics: importance, models for the prediction of the airow and temperature distributions in
scales, possibilities, limitations and ten tips and tricks towards accurate and atria, Energy Build. 48 (2012) 18e28.
reliable simulations, Build. Environ. 91 (2015) 219e245. [60] Y. Li, M. Sandberg, L. Fuchs, Effects of thermal radiation on airow with
[30] C. Yang, P. Demokritou, Q. Chen, J. Spengler, Experimental validation of a displacement ventilation: an experimental investigation, Energy Build. 19
computational uid dynamics model for IAQ applications in ice rink arenas, (1993) 263e274.
Indoor Air 11 (2001) 120e126. [61] M. Sandberg, What is ventilation efciency? Build. Environ. 16 (1981)
[31] T. Hayashi, Y. Ishizu, S. Kato, S. Murakami, CFD analysis on characteristics of 123e135.
contaminated indoor air ventilation and its application in the evaluation of [62] ANSYS, ANSYS Fluent 12.0 User's Guide, ANSYS, Inc., 2009.
the effects of contaminant inhalation by a human occupant, Build. Environ. 37 [63] T.L. Bergman, A.S. Lavine, F.P. Incropera, D.P. Dewitt, Fundamentals of Heat
(2002) 219e230. and Mass Transfer, seventh ed., John Wiley and Sons, Inc., 2011.
[32] D.N. Sørensen, P.V. Nielsen, Quality control of computational uid dynamics in [64] ISO, Building Components and Building Elements e Thermal Resistance and
indoor environments, Indoor Air 13 (2003) 2e17. Thermal Transmittance e Calculation Method (ISO 6946:2007), second ed.,
[33] P.V. Nielsen, Computational uid dynamics and room air movement, Indoor International Organization for Standardization, 2007.
Air 14 (2004) 134e143. [65] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering
[34] T. van Hooff, B. Blocken, G.J.F. van Heijst, On the suitability of steady RANS CFD applications, AIAA J. 32 (1994) 1598e1605.
for forced mixing ventilation at transitional slot reynolds numbers, Indoor Air [66] ANSYS, Introductory FLUENT Training e Chapter 5: Solver Settings, ANSYS,
23 (2013) 236e249. Inc., 2009.
S. Gilani et al. / Building and Environment 95 (2016) 299e313 313

[67] P.J. Roache, Perspective: A method for uniform reporting of grid renement comfort and wind safety in urban areas: general decision framework and case
studies, J. Fluids Eng. 116 (1994) 405e413. study for the Eindhoven University campus, Environ. Model. Softw. 30 (2012)
[68] P.J. Roache, Quantication of uncertainty in computational uid dynamics, 15e34.
Annu. Rev. Fluid Mech. 29 (1997) 123e160. [76] H. Montazeri, B. Blocken, W.D. Janssen, T. van Hooff, CFD evaluation of new
[69] B.E. Launder, D.B. Spalding, Lectures in Mathematical Models of Turbulence, second-skin facade concept for wind comfort on building balconies: case
Academic Press, London, England, 1972. study for the Park Tower in Antwerp, Build. Environ. 68 (2013) 179e192.
[70] T.-H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k-ε eddy viscosity [77] Z.D. Chen, Y. Li, Buoyancy-driven displacement natural ventilation in a single-
model for high reynolds number turbulent ows, Comput. Fluids 24 (1995) zone building with three-level openings, Build. Environ. 37 (2002) 295e303.
227e238. [78] H.-J. Park, D. Holland, The effect of location of a convective heat source on
[71] V. Yakhot, S. Orszag, Renormalization group analysis of turbulence. I. Basic displacement ventilation: CFD study, Build. Environ. 36 (2001) 883e889.
theory, J. Sci. Comput. 1 (1986) 3e51. [79] O. Auban, F. Lemoine, P. Vallette, J.R. Fontaine, Simulation by solutal con-
[72] D.C. Wilcox, Turbulence Modeling for CFD, DCW Industries, Inc, La Canada, vection of a thermal plume in a conned stratied environment: application
California, 1998. to displacement ventilation, Int. J. Heat Mass Transf. 44 (2001) 4679e4691.
[73] M. Wolfshtein, The velocity and temperature distribution in one-dimensional [80] A. Bouzinaoui, P. Vallette, F. Lemoine, J.R. Fontaine, R. Devienne, Experimental
ow with turbulence augmentation and pressure gradient, Int. J. Heat Mass study of thermal stratication in ventilated conned spaces, Int. J. Heat Mass
Transf. 12 (1969) 301e318. Transf. 48 (2005) 4121e4131.
[74] T. van Hooff, B. Blocken, L. Aanen, B. Bronsema, A venturi-shaped roof for [81] P.F. Linden, G.F. Lane-Serff, D.A. Smeed, Emptying lling boxes: the uid
wind-induced natural ventilation of buildings: wind tunnel and CFD evalua- mechanics of natural ventilation, J. Fluid Mech. 212 (1990) 309e335.
tion of different design congurations, Build. Environ. 46 (2011) 1797e1807. [82] P.F. Linden, P. Cooper, Multiple sources of buoyancy in a naturally ventilated
[75] B. Blocken, W.D. Janssen, T. van Hooff, CFD simulation for pedestrian wind enclosure, J. Fluid Mech. 311 (1996) 177e192.

You might also like